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

On the Ξ 21\Pi^{1}_{2} consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}

YUDAI SUZUKI [email protected] KEITA YOKOYAMA [email protected]
Abstract

In this paper, we introduce a hierarchy dividing the set {ΟƒβˆˆΞ 21:Ξ 11βˆ’π–’π– 0βŠ’Οƒ}\{\sigma\in\Pi^{1}_{2}:\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\sigma\}. Then, we give some characterizations of this set using weaker variants of some principles equivalent to Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}: leftmost path principle, Ramsey’s theorem for Ξ£n0\Sigma^{0}_{n} classes of [β„•]β„•[\mathbb{N}]^{\mathbb{N}} and determinacy for (Ξ£10)n(\Sigma^{0}_{1})_{n} classes of β„•β„•\mathbb{N}^{\mathbb{N}}.

1 Introduction

Reverse mathematics is a program to classify theorems in various fields of mathematics according to their logical strength. In the most typical study of this area, one uses the big five of axiomatic systems of second-order arithmetic, 𝖱𝖒𝖠0\mathsf{RCA}_{0}, 𝖢π–ͺ𝖫0\mathsf{WKL}_{0}, 𝖠𝖒𝖠0\mathsf{ACA}_{0}, 𝖠𝖳𝖱0\mathsf{ATR}_{0} and Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. Undeniably, many theorems in the β€œcore-of-mathematics” are shown to be equivalent to one of the big five (see, e.g., for [13, 2]).

Among the big five systems, the strongest system, Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, is considerably different from others; all other systems are axiomatized by Ξ 21\Pi^{1}_{2} sentences while Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is never implied from a true Ξ 21\Pi^{1}_{2} sentence. In the study of reverse mathematics, there are several mathematical theorems which are formalized by Ξ 21\Pi^{1}_{2} sentences and known to be provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} such as Kruskal’s tree theorem, Nash-Williams’ theorem, Menger’s theorem, FraΓ―ssé’s conjecture and Caristi-Kirk fixed point theorem (see [6, 10, 12, 7, 4, 3] for the recent progress on these studies111It is shown that Nash-Williams’ theorem is provable from 𝖠𝖳𝖱0\mathsf{ATR}_{0} in a recent paper by Pakhomov and SoldΓ  [10].). Then a natural question arises: how can we calibrate the strength of consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} which are of complexity Ξ 21\Pi^{1}_{2}?

In [18], Towsner introduced the relative leftmost path principles to give a new upper bound for theorems located between 𝖠𝖳𝖱0\mathsf{ATR}_{0} and Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. He focused on an equivalent of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} called the leftmost path principle which states that any ill-founded tree has a leftmost path. His argument is based on the following idea. Since the leftmost path principle is a Ξ 31\Pi^{1}_{3} statement, the actual leftmost path is not needed to prove a Ξ 21\Pi^{1}_{2} sentence. Instead of the leftmost path, it is enough to use a path which behaves as a leftmost path in a certain range. Such a path is called a relative leftmost path. Indeed, the hierarchy of the relative leftmost path principles already capture many of the Ξ 21\Pi^{1}_{2} consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} which are mentioned above. On the other hand, this hierarchy does not cover the whole Ξ 21\Pi^{1}_{2} consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}.

In this paper, we extend the idea of relativization of Ξ 31\Pi^{1}_{3} statements to the level of nn-th hyperjump. We define Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) as the assertion that for each set XX, there is a coded Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\mathsf{ACA}_{0} such that Xβˆˆβ„³X\in\mathcal{M} and β„³βŠ§βˆƒY​(Y=HJn⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{n}(X)). Then a witness YY works as a good approximation of the actual nn-th hyperjump. We show that a Ξ 21\Pi^{1}_{2} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is already provable from Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) for some nn.

Intuitively, a sentence Οƒ\sigma is provable from Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) means that there is a proof of Οƒ\sigma from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} such that the use of the hyperjump operator in it is up to nn-times. In this sense, Towsner’s transfinite leftmost path principle 𝖳𝖫𝖯𝖯\mathsf{TLPP} is in the level of single use of the hyperjump operator. Specifically, we show that there is a variant of the relative leftmost path principle which is equivalent to Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1), and a variant of Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is enough to prove 𝖳𝖫𝖯𝖯\mathsf{TLPP}. We also introduce an iterated form of the relative leftmost path principles which is in the level of Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n). We note that the first author shows that both of Towsner’s relative leftmost path principles and Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) (including some variants of it) can be characterized as a variant of the Ο‰\omega-model reflection of transfintie induction in a forthcoming paper [14]222A characterization of Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) via the Ο‰\omega-model reflection of transfinite induction is independently given by Freund [4].. In this sense, our work can be seen as a natural extension of Towsner’s work reaching the whole set of the Ξ 21βˆ’\Pi^{1}_{2}\mathchar 45\relaxconsequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}.

We apply the idea of relativization to certain descriptive set theoretic principles: Ramsey’s theorem for [β„•]β„•[\mathbb{N}]^{\mathbb{N}} (known as Galvin-Prikry’s theorem) and the determinacy for β„•β„•\mathbb{N}^{\mathbb{N}}. It is known that both of Ramsey’s theorem for Ξ£n0\Sigma^{0}_{n} classes and the determinacy for (Ξ£10)n(\Sigma^{0}_{1})_{n} classes are equivalent to Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} for n>1n>1. In this paper, we show that the Ξ 21\Pi^{1}_{2} consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} coincides with the theories 𝖠𝖒𝖠0+{rel⁑(Ξ£n0βˆ’π–±π–Ίπ—†):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\operatorname{rel}(\Sigma^{0}_{n}\mathchar 45\relax\mathsf{Ram}):n\in\omega\} or 𝖠𝖒𝖠0+{rel⁑((Ξ£10)nβˆ’π–£π–Ύπ—):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\operatorname{rel}((\Sigma^{0}_{1})_{n}\mathchar 45\relax\mathsf{Det}):n\in\omega\}, where rel⁑(Οƒ)\operatorname{rel}(\sigma) denotes the relativization of Οƒ\sigma.

Acknowledgements

The authors thank Anton Freund for helpful discussions and comments. The first author’s work is supported by JST, the establishment of university fellowships towards the creation of science technology innovation, Grant Number JPMJFS2102. The second author’s work is partially supported by JSPS KAKENHI grant numbers JP19K03601, JP21KK0045 and JP23K03193.

2 Preliminaries

This section introduces the basic notions of reverse mathematics. For the details, see also Simpson’s textbook [13]. As usual, we denote the set of natural numbers by Ο‰\omega, and use β„•\mathbb{N} to represent the range of number variables in the language β„’2\mathcal{L}_{2} of second-order arithmetic.

Reverse mathematics is a research program that classifies mathematical theorems according to their logical strength. To achieve this, we identify the logical strength of a theorem with the weakest axiom needed to prove it. Most research in reverse mathematics is based on second-order arithmetic333Sometimes the phrase β€˜second-order arithmetic’ means a specific axiomatic system 𝖹2\mathsf{Z}_{2}. However, in this paper, we use this phrase as a general term for axiomatic systems of the language of second-order arithmetic.. Specifically, for a given formula TT in second-order arithmetic representing a mathematical theorem, we seek an axiomatic system Ξ“\Gamma and an axiom AA such that:

  • β€’

    Ξ“\Gamma is strong enough to interpret TT, but may not be strong enough to prove it,

  • β€’

    AA is equivalent to TT over Ξ“\Gamma.

We begin with introducing some axiomatic systems of second-order arithmetic.

Definition 2.1.

Let θ​(y,xβ†’,Xβ†’)\theta(y,\vec{x},\vec{X})444Here, xβ†’\vec{x} and Xβ†’\vec{X} are abbreviations of x1,…,xnx_{1},\ldots,x_{n} and X1,…,XmX_{1},\ldots,X_{m}. When the numbers nn and mm of variables are not important, we use these notations. be a formula with exactly displayed free variables. Then, ΞΈ\theta-comprehension axiom is the formula

βˆ€xβ†’,Xβ†’βˆƒYβˆ€y(y∈Y↔θ(y,xβ†’,Xβ†’)).\displaystyle\forall\vec{x},\vec{X}\exists Y\forall y(y\in Y\leftrightarrow\theta(y,\vec{x},\vec{X})).

Intuitively, comprehension for ΞΈ\theta states that the set {y:θ​(y)}\{y:\theta(y)\} exists.

Some β„’2\mathcal{L}_{2}-theories are characterized by comprehension schemas.

  • β€’

    𝖱𝖒𝖠0\mathsf{RCA}_{0} consists of β€˜(0,1,+,β‹…,<)(0,1,+,\cdot,<) forms a discrete ordered semi-ring’, induction for Ξ£10\Sigma^{0}_{1} formulas and comprehension for Ξ”10\Delta^{0}_{1} formulas, i.e., ΞΈ\theta-comprehension is applicable when ΞΈ\theta is a Ξ£10\Sigma^{0}_{1} formula which is equivalent to a Ξ 10\Pi^{0}_{1} formula.

  • β€’

    𝖠𝖒𝖠0\mathsf{ACA}_{0} consists of 𝖱𝖒𝖠0\mathsf{RCA}_{0} and comprehension for Ξ£01\Sigma^{1}_{0} formulas.

  • β€’

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} consists of 𝖱𝖒𝖠0\mathsf{RCA}_{0} and comprehension for Ξ 11\Pi^{1}_{1} formulas.

In 𝖱𝖒𝖠0\mathsf{RCA}_{0}, the Turing jump operator and its iteration can be defined. For a given (countable) ordinal Ξ±\alpha, we write X(Ξ±)X^{(\alpha)} to mean the Ξ±\alpha-times iteration of Turing jump at XX. For the details, see section VIII.1 in [13].

Definition 2.2.

We define 𝖠𝖒𝖠0β€²\operatorname{\mathsf{ACA}}_{0}^{\prime} and 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} as follows.

  • β€’

    𝖠𝖒𝖠0β€²\operatorname{\mathsf{ACA}}_{0}^{\prime} consists of 𝖱𝖒𝖠0\mathsf{RCA}_{0} and βˆ€nβ€‹βˆ€Xβ€‹βˆƒY​(Y=X(n))\forall n\forall X\exists Y(Y=X^{(n)}).

  • β€’

    𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} consists of 𝖱𝖒𝖠0\mathsf{RCA}_{0} and βˆ€Xβ€‹βˆƒY​(Y=X(Ο‰))\forall X\exists Y(Y=X^{(\omega)}).

  • β€’

    𝖠𝖳𝖱0\mathsf{ATR}_{0} consists of 𝖱𝖒𝖠0\mathsf{RCA}_{0} and the statement β€˜for any ordinal Ξ±\alpha, Y=X(Ξ±)Y=X^{(\alpha)} exists’.

It is known that 𝖠𝖒𝖠0<𝖠𝖒𝖠0β€²<𝖠𝖒𝖠0+<𝖠𝖳𝖱0\mathsf{ACA}_{0}<\operatorname{\mathsf{ACA}}_{0}^{\prime}<\operatorname{\mathsf{ACA}}_{0}^{+}<\mathsf{ATR}_{0}.

2.1 Trees

We introduce some notions on trees. Then we give a characterization of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} by trees.

Definition 2.3.

Let XX be a set. We write X<β„•X^{<\mathbb{N}} for the set of finite sequences of elements of XX and Xβ„•X^{\mathbb{N}} for the set of infinite sequences of elements of XX. More precisely, each ΟƒβˆˆX<β„•\sigma\in X^{<\mathbb{N}} is a function whose domain is {0,1,…,nβˆ’1}\{0,1,\ldots,n-1\} for some nn and σ​(i)\sigma(i) is an element of XX for any i<ni<n . Similarly, each f∈Xβ„•f\in X^{\mathbb{N}} is a function whose domain is β„•\mathbb{N} and whose codomain is XX.

For each ΟƒβˆˆX<β„•\sigma\in X^{<\mathbb{N}}, define the length |Οƒ||\sigma| by the number nn such that dom⁑(Οƒ)={0,1,…,nβˆ’1}\operatorname{dom}(\sigma)=\{0,1,\ldots,n-1\}.

Intuitively, we identify ΟƒβˆˆX<β„•\sigma\in X^{<\mathbb{N}} and the sequence βŸ¨Οƒβ€‹(0),…,σ​(|Οƒ|βˆ’1)⟩\langle\sigma(0),\ldots,\sigma(|\sigma|-1)\rangle, f∈Xβ„•f\in X^{\mathbb{N}} and the sequence ⟨f​(n)⟩nβˆˆβ„•\langle f(n)\rangle_{n\in\mathbb{N}}.

Definition 2.4.

Let Οƒ,Ο„\sigma,\tau be finite sequences.

  • β€’

    Define the concatenation Οƒβˆ—Ο„\sigma\ast\tau by βŸ¨Οƒβ€‹(0),…,σ​(|Οƒ|βˆ’1),τ​(0),…,τ​(|Ο„|βˆ’1)⟩\langle\sigma(0),\ldots,\sigma(|\sigma|-1),\tau(0),\ldots,\tau(|\tau|-1)\rangle.

  • β€’

    We say Οƒ\sigma is an initial segment of Ο„\tau (write Οƒβͺ―Ο„\sigma\preceq\tau) if |Οƒ|≀|Ο„||\sigma|\leq|\tau| and βˆ€i<|Οƒ|​(σ​(i)=τ​(i))\forall i<|\sigma|(\sigma(i)=\tau(i)). We say Οƒ\sigma is a proper initial segment of Ο„\tau (write Οƒβ‰ΊΟ„\sigma\prec\tau) if Οƒβͺ―Ο„\sigma\preceq\tau and Οƒβ‰ Ο„\sigma\neq\tau.

  • β€’

    Let ff be an infinite sequence. We say Οƒ\sigma is an initial segment of ff (write Οƒβ‰Ίf\sigma\prec f) if βˆ€i<|Οƒ|​(σ​(i)=f​(i))\forall i<|\sigma|(\sigma(i)=f(i)).

Definition 2.5.

Let TT be a subset of X<β„•X^{<\mathbb{N}}. We say TT is a tree on XX if TT is closed under taking initial segments.

Let TβŠ†X<β„•T\subseteq X^{<\mathbb{N}} be a tree. A function f∈Xβ„•f\in X^{\mathbb{N}} is called a path of TT if (βˆ€n)​(f​[n]∈T)(\forall n)(f[n]\in T). Here, f​[n]f[n] denotes the initial segment of ff with length nn. The set of all paths of TT is denoted by [T][T]. We say TT is ill-founded if [T]β‰ βˆ…[T]\neq\varnothing.

Lemma 2.6.

([13], III.7.2) 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves the following KΕ‘nig lemma. Let TT be an infinite tree. If βˆ€ΟƒβˆˆTβ€‹βˆƒmβ€‹βˆ€i​(Οƒβˆ—βŸ¨i⟩∈Tβ†’i<m)\forall\sigma\in T\exists m\forall i(\sigma\ast\langle i\rangle\in T\to i<m), then TT is ill-founded.

Lemma 2.7.

[13, VI.1.1] The following are equivalent over 𝖱𝖒𝖠0\mathsf{RCA}_{0}.

  1. 1.

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0},

  2. 2.

    for any sequence ⟨Tk⟩k\langle T_{k}\rangle_{k} of trees, there is a set X={k:[Tk]β‰ βˆ…}X=\{k:[T_{k}]\neq\varnothing\}.

Following the above lemma, we give a new characterization of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} by using the notion of leftmost path.

Definition 2.8.

(𝖱𝖒𝖠0\mathsf{RCA}_{0}) Let f,gβˆˆβ„•β„•f,g\in\mathbb{N}^{\mathbb{N}}. We write f<lgf<_{l}g if βˆƒn​(f​[n]=g​[n]∧f​(n)<g​(n))\exists n(f[n]=g[n]\land f(n)<g(n)) and f≀lgf\leq_{l}g if f<lg∨f=gf<_{l}g\lor f=g. This <l<_{l} forms a total order on β„•β„•\mathbb{N}^{\mathbb{N}} and is called the lexicographical order.

Let TT be a tree and f∈[T]f\in[T]. We say ff is the leftmost path of TT if βˆ€g∈[T]​(f≀lg)\forall g\in[T](f\leq_{l}g).

In the remaining of this section, we show that the equivalence of the existence of a leftmost path and Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. The key is that a sequence of trees can be coded by a tree.

Definition 2.9.

(𝖱𝖒𝖠0\mathsf{RCA}_{0}) Let Οƒβˆˆβ„•<β„•\sigma\in\mathbb{N}^{<\mathbb{N}}. For each l<|Οƒ|l<|\sigma|, define nln_{l} as the maximum nn such that (n,l)<|Οƒ|(n,l)<|\sigma|. Then we define Οƒl=βŸ¨Οƒβ€‹((0,l)),…,σ​((nl,l))⟩\sigma_{l}=\langle\sigma((0,l)),\ldots,\sigma((n_{l},l))\rangle. Thus, each Οƒβˆˆβ„•<β„•\sigma\in\mathbb{N}^{<\mathbb{N}} is regarded as a sequence βŸ¨Οƒl⟩l<|Οƒ|\langle\sigma_{l}\rangle_{l<|\sigma|} of sequences.

Conversely, for each sequence βŸ¨Οƒl⟩l<L\langle\sigma_{l}\rangle_{l<L} of sequences, define ⨁l<LΟƒl\bigoplus_{l<L}\sigma_{l} by

(⨁l<LΟƒl)​(n,l)=Οƒl​(n).\displaystyle(\bigoplus_{l<L}\sigma_{l})(n,l)=\sigma_{l}(n).

For each sequences ⟨fl⟩lβˆˆβ„•\langle f_{l}\rangle_{l\in\mathbb{N}} of functions, define ⨁lfl\bigoplus_{l}f_{l} by

(⨁l<Lfl)​(n,l)=fl​(n).\displaystyle(\bigoplus_{l<L}f_{l})(n,l)=f_{l}(n).

Finally, for each ⟨Tl⟩lβˆˆβ„•\langle T_{l}\rangle_{l\in\mathbb{N}} of trees, define ⨁lTl\bigoplus_{l}T_{l} by

{⨁l<LΟƒl:βŸ¨Οƒl⟩l<L∈T0Γ—β‹―Γ—TLβˆ’1}.\displaystyle\{\bigoplus_{l<L}\sigma_{l}:\langle\sigma_{l}\rangle_{l<L}\in T_{0}\times\cdots\times T_{L-1}\}.
Lemma 2.10.

(𝖱𝖒𝖠0)(\mathsf{RCA}_{0}) Let ⟨Tl⟩l\langle T_{l}\rangle_{l} be a sequence of trees. Then, the operator ⨁l\bigoplus_{l} is a bijection between ∏l[Tl]\prod_{l}[T_{l}] and [⨁lTl][\bigoplus_{l}T_{l}]. Moreover, this operator preserves the lexicographical order in the sense that for any ⟨fi⟩i,⟨gi⟩i∈∏i[Ti]\langle f_{i}\rangle_{i},\langle g_{i}\rangle_{i}\in\prod_{i}[T_{i}],

  1. 1.

    βˆ€i​(fi≀lgi)→⨁ifi≀l⨁igi\forall i(f_{i}\leq_{l}g_{i})\to\bigoplus_{i}f_{i}\leq_{l}\bigoplus_{i}g_{i},

  2. 2.

    ⨁ifi\bigoplus_{i}f_{i} is the leftmost path of ⨁iTi\bigoplus_{i}T_{i} if and only if each fif_{i} is the leftmost path of TiT_{i}.

Definition 2.11.

Let TT be a tree. We say TT is pruned if βˆ€ΟƒβˆˆTβ€‹βˆƒn​(Οƒβˆ—βŸ¨n⟩∈T)\forall\sigma\in T\exists n(\sigma\ast\langle n\rangle\in T).

Lemma 2.12.

(𝖱𝖒𝖠0\mathsf{RCA}_{0}) Let TT be a pruned tree and ΟƒβˆˆT\sigma\in T. Then there is a TT-computable path ff such that Οƒβ‰Ίf\sigma\prec f. Moreover, we can take this ff to be the leftmost of all paths extending Οƒ\sigma.

Proof.

It is immediate from the definition. ∎

Theorem 2.13.

[6, Theorem 6.5.] The following are equivalent over 𝖱𝖒𝖠0\mathsf{RCA}_{0}.

  1. 1.

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0},

  2. 2.

    Every ill-founded tree has the leftmost path.

We review the proof of this theorem with the notions defined above since we will refine the argument here in the later discussions.

Proof.

(1β†’2)(1\to 2). Let TT be an ill-founded tree. Define S={ΟƒβˆˆT:βˆƒf∈[T]​(Οƒβ‰Ίf)}S=\{\sigma\in T:\exists f\in[T](\sigma\prec f)\}. Then SS is a pruned tree such that [S]=[T][S]=[T]. Since SS is pruned, SS has the leftmost path. Clearly this is also the leftmost path of [T][T].

(2β†’1)(2\to 1) It is enough to show that the second clause implies for any sequence ⟨Tk⟩k\langle T_{k}\rangle_{k} of trees, there is a set X={k:[Tk]β‰ βˆ…}X=\{k:[T_{k}]\neq\varnothing\}.

Assume that each ill-founded tree has the leftmost path. Let ⟨Tk⟩k\langle T_{k}\rangle_{k} be a sequence of trees. For each kk, define SkS_{k} by {⟨0βŸ©βˆ—Οƒ:ΟƒβˆˆTk}βˆͺ{1l:lβˆˆβ„•}\{\langle 0\rangle\ast\sigma:\sigma\in T_{k}\}\cup\{1^{l}:l\in\mathbb{N}\} where 1l1^{l} is the sequence with length =l=l such that each element of it is 11. Then, each SkS_{k} has a path 1∞=⟨1,1,…,⟩1^{\infty}=\langle 1,1,\ldots,\rangle. Thus, ⨁kSk\bigoplus_{k}S_{k} is also ill-founded.

By the assumption, take the leftmost path ff of ⨁kSk\bigoplus_{k}S_{k} and put X={k:f​(0,k)=0}X=\{k:f(0,k)=0\}. It is easy to check that X={k:[Tk]β‰ βˆ…}X=\{k:[T_{k}]\neq\varnothing\}. ∎

2.2 Coded Ο‰\omega-models

We introduce the definition of coded Ο‰\omega-models in 𝖱𝖒𝖠0\mathsf{RCA}_{0} and give some basic properties.

Definition 2.14.

(𝖱𝖒𝖠0\mathsf{RCA}_{0}) For a set AA and a sequence X=⟨Xi⟩iX=\langle X_{i}\rangle_{i}, we say AA is in ⟨Xi⟩i\langle X_{i}\rangle_{i} or ⟨Xi⟩i\langle X_{i}\rangle_{i} contains AA (write A∈⟨Xi⟩iA\in\langle X_{i}\rangle_{i}) if βˆƒi​(A=Xi)\exists i(A=X_{i}), and thus XX is identified with an β„’2\mathcal{L}_{2}-structure (β„•,⟨Xi⟩i,0,1,+,β‹…,∈)(\mathbb{N},\langle X_{i}\rangle_{i},0,1,+,\cdot,\in). In this sense, we call XX a coded Ο‰\omega-model.

Remark 2.15.

Let β„³\mathcal{M} be a coded Ο‰\omega-model. For a sentence Οƒ\sigma with parameters from β„³\mathcal{M}, the satisfaction β„³βŠ§Οƒ\mathcal{M}\models\sigma is defined by the existence of an evaluating function for Οƒ\sigma which represents Tarski’s truth definition and ensures Οƒ\sigma is true. If we work in 𝖠𝖒𝖠0\mathsf{ACA}_{0}, then β„³βŠ§Οƒ\mathcal{M}\models\sigma if and only if the relativization Οƒβ„³\sigma^{\mathcal{M}} is true. In particular, if Οƒ\sigma is arithmeical, then β„³βŠ§Οƒ\mathcal{M}\models\sigma if and only if Οƒ\sigma is true.

Definition 2.16.

Let Οƒ\sigma be an β„’2\mathcal{L}_{2}-sentence. Define the assertion Ο‰\omega-model reflection of Οƒ\sigma by

βˆ€Xβ€‹βˆƒβ„³:Ο‰-model ​(Xβˆˆβ„³βˆ§β„³βŠ§Οƒ).\forall X\exists\mathcal{M}:\text{$\omega$-model }(X\in\mathcal{M}\land\mathcal{M}\models\sigma).
Lemma 2.17.

Let Οƒ\sigma be a Ξ 21\Pi^{1}_{2} sentence. Then, over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, the Ο‰\omega-model reflection of Οƒ\sigma implies Οƒ+π–’π—ˆπ—‡β‘(Οƒ)\sigma+\operatorname{\mathsf{Con}}(\sigma) where π–’π—ˆπ—‡β‘(Οƒ)\operatorname{\mathsf{Con}}(\sigma) is the Ξ 10\Pi^{0}_{1} formula stating that β€˜Οƒ\sigma is consistent’. Therefore, if π–’π—ˆπ—‡β‘(Οƒ)\operatorname{\mathsf{Con}}(\sigma) is not provable from 𝖠𝖒𝖠0\mathsf{ACA}_{0}, then the Ο‰\omega-model reflection of Οƒ\sigma is strictly stronger than Οƒ\sigma.

Proof.

Write Οƒβ‰‘βˆ€Xβ€‹βˆƒY​θ​(X,Y)\sigma\equiv\forall X\exists Y\theta(X,Y) by an arithmetical formula ΞΈ\theta. Assume 𝖠𝖒𝖠0\mathsf{ACA}_{0} and the Ο‰\omega-model reflection of Οƒ\sigma. We show that Οƒ+π–’π—ˆπ—‡β‘(Οƒ)\sigma+\operatorname{\mathsf{Con}}(\sigma) holds. Let AA be a set. By the Ο‰\omega-model reflection, take an Ο‰\omega-model β„³\mathcal{M} such that Aβˆˆβ„³βˆ§β„³βŠ§βˆ€Xβ€‹βˆƒY​θ​(X,Y)A\in\mathcal{M}\land\mathcal{M}\models\forall X\exists Y\theta(X,Y). Then, β„³βŠ§βˆƒY​θ​(A,Y)\mathcal{M}\models\exists Y\theta(A,Y) and hence θ​(A,B)\theta(A,B) holds for some Bβˆˆβ„³B\in\mathcal{M}. In addition, since β„³\mathcal{M} is a (weak-)model of Οƒ\sigma, π–’π—ˆπ—‡β‘(Οƒ)\operatorname{\mathsf{Con}}(\sigma) holds. ∎

The following characterization of 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} is well-known.

Theorem 2.18.

Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, the following are equivalent.

  1. 1.

    𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+},

  2. 2.

    the Ο‰\omega-model reflection of 𝖠𝖒𝖠0\mathsf{ACA}_{0}; any set is contained in an Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

2.3 Hyperjump and Ξ²\beta-model reflection

In this subsection, we introduce the notion of hyperjump and Ξ²\beta-models. Hyperjump is a Ξ£11\Sigma^{1}_{1} analogue of Turing jump. As 𝖠𝖒𝖠0\mathsf{ACA}_{0} is characterized by Turing jumps, Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is characterized by hyperjumps.

Definition 2.19.

(Universal Ξ£11\Sigma^{1}_{1} formula) Let 𝖲𝖺𝗍Π00\mathsf{Sat}_{\Pi^{0}_{0}} be a universal Ξ 00\Pi^{0}_{0} formula. We define the Ξ£11\Sigma^{1}_{1} universal formula Οƒ11\sigma^{1}_{1} by

Οƒ11​(e,x,X)β‰‘βˆƒfβ€‹βˆ€y​𝖲𝖺𝗍Π00​(e,x,X,f​[y]).\displaystyle\sigma^{1}_{1}(e,x,X)\equiv\exists f\forall y\mathsf{Sat}_{\Pi^{0}_{0}}(e,x,X,f[y]).

.

Definition 2.20.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let β„³\mathcal{M} be a coded Ο‰\omega-model. We say β„³\mathcal{M} is a coded Ξ²\beta-model if

βˆ€Xβˆˆβ„³βˆ€e,x(Οƒ11(e,x,X)β†”β„³βŠ§Οƒ11(e,x,X)).\forall X\in\mathcal{M}\forall e,x(\sigma^{1}_{1}(e,x,X)\leftrightarrow\mathcal{M}\models\sigma^{1}_{1}(e,x,X)).

Intuitively, a coded Ξ²\beta-model is a Ξ£11\Sigma^{1}_{1} absolute Ο‰\omega-model. In fact, we can show the following.

Lemma 2.21.

Let θ​(x,X)\theta(x,X) be a Ξ£11\Sigma^{1}_{1} formula. Then, 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that for any Ξ²\beta-model β„³\mathcal{M}, βˆ€Xβˆˆβ„³βˆ€x(ΞΈ(x,X)β†”β„³βŠ§ΞΈ(x,X))\forall X\in\mathcal{M}\forall x(\theta(x,X)\leftrightarrow\mathcal{M}\models\theta(x,X)).

Proof.

Immediate from the definition of β\beta-models. ∎

We next introduce the notion of hyperjump. Intuitively, the hyperjump operator is an operator to make a Ξ£11,X\Sigma^{1,X}_{1} complete set from a set XX.

Definition 2.22.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let XX be a set. Define the hyperjump HJ⁑(X)\operatorname{HJ}(X) by HJ⁑(X)={(e,x):Οƒ11​(e,x,X)}\operatorname{HJ}(X)=\{(e,x):\sigma^{1}_{1}(e,x,X)\}. We write HJn⁑(X)\operatorname{HJ}^{n}(X) for the nn-times iteration of the hyperjump.

Lemma 2.23.

Let φ​(x,y,X)\varphi(x,y,X) be a Ξ£11\Sigma^{1}_{1} formula with exactly displayed free variables. Then, 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves

βˆ€y,X,Y​(Y=HJ⁑(X)β†’βˆƒZ≀TY​(Z={x:φ​(x,y,X)})).\forall y,X,Y(Y=\operatorname{HJ}(X)\to\exists Z\leq_{\mathrm{T}}Y(Z=\{x:\varphi(x,y,X)\})).

That is, any Σ11,X\Sigma^{1,X}_{1}-definable set is computable from HJ⁑(X)\operatorname{HJ}(X). Hence, any Π11,X\Pi^{1,X}_{1}-definable set is also computable from HJ⁑(X)\operatorname{HJ}(X).

Proof.

Immediate from the definition of HJ\operatorname{HJ}. ∎

Lemma 2.24 ([13] Lemma VII.2.9).

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, for any set XX, the existence of HJ⁑(X)\operatorname{HJ}(X) is equivalent to the existence of a coded Ξ²\beta-model containing XX.

Combining the above two lemmas, we have that

Theorem 2.25.

The following assertions are equivalent over 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

  1. 1.

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0},

  2. 2.

    βˆ€Xβ€‹βˆƒY​(Y=HJ⁑(X))\forall X\exists Y(Y=\operatorname{HJ}(X)),

  3. 3.

    Ξ²\beta-model reflection : βˆ€Xβ€‹βˆƒβ„³β€‹Β :Β Ξ²-model ​(Xβˆˆβ„³)\forall X\exists\mathcal{M}\text{ : $\beta$-model }(X\in\mathcal{M}).

We note that all of these assertions are Ξ 31\Pi^{1}_{3} statements and any of them is never equivalent to a Ξ 21\Pi^{1}_{2} statement. In this paper, we introduce Ξ 21\Pi^{1}_{2} variants of the second and third assertions to consider the structure of Ξ 21\Pi^{1}_{2} consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}.

At last, we see that a hyperjump in the ground model behaves a hyperjump in a coded Ο‰\omega-model. The following two lemmas are refinements of [13, VII.1.12].

Lemma 2.26.

Let β„³=(β„•β„³,SM)\mathcal{M}=(\mathbb{N}^{\mathcal{M}},S^{M}) and β„³β€²=(β„•β„³β€²,Sβ„³β€²)\mathcal{M}^{\prime}=(\mathbb{N}^{\mathcal{M}^{\prime}},S^{\mathcal{M}^{\prime}}) be models of 𝖠𝖒𝖠0\mathsf{ACA}_{0} such that β„•β„³=β„•β„³β€²\mathbb{N}^{\mathcal{M}}=\mathbb{N}^{\mathcal{M}^{\prime}} and Sβ„³β€²βŠ†Sβ„³S^{\mathcal{M}^{\prime}}\subseteq S^{\mathcal{M}}.

Let A,Bβˆˆβ„³β€²A,B\in\mathcal{M}^{\prime} such that β„³βŠ§B=HJ⁑(A)\mathcal{M}\models B=\operatorname{HJ}(A). Then, β„³β€²βŠ§B=HJ⁑(A)\mathcal{M}^{\prime}\models B=\operatorname{HJ}(A). Moreover, if β„³β€²\mathcal{M}^{\prime} is a coded Ο‰\omega-model in β„³\mathcal{M}, then β„³βŠ§(B=HJ⁑(A))β„³β€²\mathcal{M}\models(B=\operatorname{HJ}(A))^{\mathcal{M}^{\prime}}.

Proof.

Take β„³,β„³β€²,A\mathcal{M},\mathcal{M}^{\prime},A and BB be as above. We first show the following claim.

Claim 2.26.1.

Let θ​(X)\theta(X) be a Ξ£11\Sigma^{1}_{1} formula having exactly one set variable. Then, β„³βŠ§ΞΈβ€‹(A)\mathcal{M}\models\theta(A) if and only if β„³β€²βŠ§ΞΈβ€‹(A)\mathcal{M}^{\prime}\models\theta(A).

Proof of Claim.

It is enough to show that β„³βŠ§ΞΈβ€‹(A)\mathcal{M}\models\theta(A) implies β„³β€²βŠ§ΞΈβ€‹(A)\mathcal{M}^{\prime}\models\theta(A). Write θ​(X)β‰‘βˆƒZ​φ​(X,Z)\theta(X)\equiv\exists Z\varphi(X,Z) by an arithmetical formula Ο†\varphi. Assume β„³βŠ§βˆƒZ​φ​(A,Z)\mathcal{M}\models\exists Z\varphi(A,Z). We will show that β„³β€²βŠ§βˆƒZ≀TB​φ​(A,B)\mathcal{M}^{\prime}\models\exists Z\leq_{\mathrm{T}}B\varphi(A,B). Now there is a Ξ£00\Sigma^{0}_{0} formula ΞΈ\theta such that both of β„³,β„³β€²\mathcal{M},\mathcal{M}^{\prime} satisfies βˆ€Z(Ο†(A,Z)β†”βˆƒfβˆ€nΞΈ(A,Z[n],f[n]))\forall Z(\varphi(A,Z)\leftrightarrow\exists f\forall n\theta(A,Z[n],f[n])). Define trees T,Tβ€²βŠ†(2Γ—β„•)<β„•T,T^{\prime}\subseteq(2\times\mathbb{N})^{<\mathbb{N}} by

T\displaystyle T ={(s,Οƒ):βˆ€(t,Ο„)βͺ―(s,Οƒ)​θ​(A,t,Ο„)},\displaystyle=\{(s,\sigma):\forall(t,\tau)\preceq(s,\sigma)\theta(A,t,\tau)\},
Tβ€²\displaystyle T^{\prime} ={(s,Οƒ):βˆƒ(Z,f)∈[T]​((s,Οƒ)β‰Ί(Z,f))}.\displaystyle=\{(s,\sigma):\exists(Z,f)\in[T]((s,\sigma)\prec(Z,f))\}.

Then, TT is Ξ”10\Delta^{0}_{1} in AA and hence Tβ€²T^{\prime} is Ξ£11\Sigma^{1}_{1} in AA. Therefore, T′≀THJ⁑(A)T^{\prime}\leq_{\mathrm{T}}\operatorname{HJ}(A). Moreover, since βˆƒZ​φ​(X,Z)\exists Z\varphi(X,Z) holds, Tβ€²T^{\prime} is a nonempty pruned tree. Thus Tβ€²T^{\prime} has a Tβ€²T^{\prime}-computable path (Z,f)(Z,f). Now Z≀TT′≀THJ⁑(X)Z\leq_{\mathrm{T}}T^{\prime}\leq_{\mathrm{T}}\operatorname{HJ}(X), so this completes the proof. ∎

Now B={(e,x)βˆˆβ„•β„³:β„³βŠ§Οƒ11​(e,x,A)}={(e,x)βˆˆβ„•β„³β€²:β„³β€²βŠ§Οƒ11​(e,x,A)}B=\{(e,x)\in\mathbb{N}^{\mathcal{M}}:\mathcal{M}\models\sigma^{1}_{1}(e,x,A)\}=\{(e,x)\in\mathbb{N}^{\mathcal{M}^{\prime}}:\mathcal{M}^{\prime}\models\sigma^{1}_{1}(e,x,A)\}. Therefore β„³β€²βŠ§B=HJ⁑(A)\mathcal{M}^{\prime}\models B=\operatorname{HJ}(A). ∎

Lemma 2.27.

Let β„³=(β„•β„³,SM)\mathcal{M}=(\mathbb{N}^{\mathcal{M}},S^{M}) and β„³β€²=(β„•β„³β€²,Sβ„³β€²)\mathcal{M}^{\prime}=(\mathbb{N}^{\mathcal{M}^{\prime}},S^{\mathcal{M}^{\prime}}) be such that β„³\mathcal{M} is a model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, β„•β„³=β„•β„³β€²\mathbb{N}^{\mathcal{M}}=\mathbb{N}^{\mathcal{M}^{\prime}}, Sβ„³β€²βŠ†Sβ„³S^{\mathcal{M}^{\prime}}\subseteq S^{\mathcal{M}} and Sβ„³β€²S^{\mathcal{M}^{\prime}} is closed under Turing sum βŠ•\oplus and lower closed under Turing reduction. If for any Aβˆˆβ„³β€²A\in\mathcal{M}^{\prime}, there exists Bβˆˆβ„³β€²B\in\mathcal{M}^{\prime} such that β„³βŠ§B=HJ⁑(A)\mathcal{M}\models B=\operatorname{HJ}(A), then β„³β€²\mathcal{M}^{\prime} is a model of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} and a Ξ²\beta-submodel of β„³\mathcal{M}.

Proof.

Since Sβ„³β€²S^{\mathcal{M}^{\prime}} is closed under Turing sum and lower closed under Turing reduction, β„³β€²\mathcal{M}^{\prime} is a model of 𝖱𝖒𝖠0\mathsf{RCA}_{0}. Moreover, by the definition of hyperjumps, it is easy to see that β„³β€²\mathcal{M}^{\prime} is closed under Ξ 10\Pi^{0}_{1}-comprehension. Therefore, β„³β€²\mathcal{M}^{\prime} is a model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

By the previous lemma, β„³β€²βŠ§βˆ€Xβ€‹βˆƒY​(Y=HJ⁑(X))\mathcal{M}^{\prime}\models\forall X\exists Y(Y=\operatorname{HJ}(X)). Therefore, β„³β€²\mathcal{M}^{\prime} is a model of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}.

To show β„³β€²\mathcal{M}^{\prime} is a Ξ²\beta-submodel of β„³\mathcal{M}, it is enough to show that for any Ξ£11\Sigma^{1}_{1} sentence with parameters from β„³β€²\mathcal{M}^{\prime}, if it is true in β„³\mathcal{M}, then it is also true in β„³β€²\mathcal{M}^{\prime}. Let θ​(X,Y)\theta(X,Y) be an arithmetical formula and Aβˆˆβ„³β€²A\in\mathcal{M}^{\prime}. Assume β„³βŠ§βˆƒY​θ​(A,Y)\mathcal{M}\models\exists Y\theta(A,Y). We show that β„³β€²βŠ§βˆƒY​θ​(A,Y)\mathcal{M}^{\prime}\models\exists Y\theta(A,Y). Let Bβˆˆβ„³B\in\mathcal{M} such that β„³βŠ§B=HJ⁑(A)\mathcal{M}\models B=\operatorname{HJ}(A). Then, β„³βŠ§βˆƒY≀TB​θ​(A,Y)\mathcal{M}\models\exists Y\leq_{\mathrm{T}}B\theta(A,Y) and Bβˆˆβ„³β€²B\in\mathcal{M}^{\prime}. Thus, β„³β€²βŠ§βˆƒY≀TB​θ​(A,Y)\mathcal{M}^{\prime}\models\exists Y\leq_{\mathrm{T}}B\theta(A,Y). ∎

Lemma 2.28.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let X,YX,Y be sets such that Y=HJ⁑(X)Y=\operatorname{HJ}(X). Then, for any coded Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, if Yβˆˆβ„³Y\in\mathcal{M}, then Xβˆˆβ„³βˆ§β„³βŠ§Y=HJ⁑(X)X\in\mathcal{M}\land\mathcal{M}\models Y=\operatorname{HJ}(X).

Proof.

We reason in 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Take a pair of sets X,YX,Y such that Y=HJ⁑(X)Y=\operatorname{HJ}(X). Let β„³\mathcal{M} be a coded Ο‰\omega-model such that Yβˆˆβ„³Y\in\mathcal{M}.

By Lemma 2.26, we only need to show Xβˆˆβ„³X\in\mathcal{M}. Take a number ee such that

βˆ€Z(x∈Z↔σ11(e,x,Z)).\forall Z(x\in Z\leftrightarrow\sigma^{1}_{1}(e,x,Z)).

Then, X={x:(e,x)∈Y}X=\{x:(e,x)\in Y\}. Since X≀TYX\leq_{\mathrm{T}}Y, Xβˆˆβ„³X\in\mathcal{M}. ∎

Remark 2.29.

In general, even if (Y=HJ⁑(X))β„³(Y=\operatorname{HJ}(X))^{\mathcal{M}} holds for a coded Ο‰\omega-model β„³\mathcal{M}, Y=HJ⁑(X)Y=\operatorname{HJ}(X) may not hold. This is immediate from some results in the next section.

3 Approximation of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}

In this section, we introduce Ξ 21\Pi^{1}_{2} variants of the Ξ²\beta-model reflection Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n). Then we prove that the theories 𝖠𝖒𝖠0+{Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} and {ΟƒβˆˆΞ 21:Ξ 11βˆ’π–’π– 0βŠ’Οƒ}\{\sigma\in\Pi^{1}_{2}:\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\sigma\} proves the same sentences.

As we have seen in Theorem 2.25, Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is characterized by the hyperjump operator or the Ξ²\beta-model reflection. This means that in Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, we can take any (standard) finite iteration of hyperjump or Ξ²\beta-models. More precisely, the following holds.

Observation 3.1.

For any nβˆˆΟ‰,n>0n\in\omega,n>0, the following assertions are equivalent over 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

  • β€’

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0},

  • β€’

    βˆ€Xβ€‹βˆƒY​(Y=HJn⁑(X))\forall X\exists Y(Y=\operatorname{HJ}^{n}(X)),

  • β€’

    βˆ€Xβ€‹βˆƒβ„³0,…,β„³nβˆ’1:Ξ²-models ​(Xβˆˆβ„³0βˆˆβ‹―βˆˆβ„³nβˆ’1)\forall X\exists\mathcal{M}_{0},\ldots,\mathcal{M}_{n-1}:\text{$\beta$-models }(X\in\mathcal{M}_{0}\in\cdots\in\mathcal{M}_{n-1}).

From this observation, we can conjecture that theorems provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} are classified by the number of the hyperjump operator used to proved those theorems. The following theorem ensures this conjecture.

Theorem 3.2.

Let θ​(X,Y,Z)\theta(X,Y,Z) be an arithmetical formula such that Ξ 11βˆ’π–’π– 0βŠ’βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z)\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\forall X\exists Y\forall Z\theta(X,Y,Z). Then, there exists nβˆˆΟ‰n\in\omega such that

𝖠𝖒𝖠0βŠ’βˆ€X,W(W=HJn(X)β†’βˆƒY≀TWβˆ€ZΞΈ(X,Y,Z))).\mathsf{ACA}_{0}\vdash\forall X,W(W=\operatorname{HJ}^{n}(X)\to\exists Y\leq_{\mathrm{T}}W\forall Z\theta(X,Y,Z))).
Proof.

Assume θ​(X,Y,Z)\theta(X,Y,Z) be an arithmetical formula such that Ξ 11βˆ’π–’π– 0βŠ’βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z)\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\forall X\exists Y\forall Z\theta(X,Y,Z). For the sake of contradiction, suppose that for any nβˆˆΟ‰n\in\omega, 𝖠𝖒𝖠0+βˆƒX,W​(W=HJn⁑(X)βˆ§βˆ€Y≀TWβ€‹βˆƒZ​¬θ​(X,Y,Z))\mathsf{ACA}_{0}+\exists X,W(W=\operatorname{HJ}^{n}(X)\land\forall Y\leq_{\mathrm{T}}W\exists Z\lnot\theta(X,Y,Z)) is consistent.

Let CC and DD be new constant set symbols. Let β„’2​(C,D)\mathcal{L}_{2}(C,D) be the language extending β„’2\mathcal{L}_{2} by adding C,DC,D. Define an β„’2​(C,D)\mathcal{L}_{2}(C,D) theory TT by

T=𝖠𝖒𝖠0\displaystyle T=\mathsf{ACA}_{0} +{D0=C∧D1=HJ⁑(D0)βˆ§β‹―βˆ§Dk=HJ⁑(Dkβˆ’1):kβˆˆΟ‰}\displaystyle+\{D_{0}=C\land D_{1}=\operatorname{HJ}(D_{0})\land\cdots\land D_{k}=\operatorname{HJ}(D_{k-1}):k\in\omega\}
+{βˆ€Y≀TDkβ€‹βˆƒZ​¬θ​(C,Y,Z):kβˆˆΟ‰}.\displaystyle+\{\forall Y\leq_{\mathrm{T}}D_{k}\exists Z\lnot\theta(C,Y,Z):k\in\omega\}.

Then, by compactness theorem, TT has a model β„³=(β„•β„³,S~,C,D)\mathcal{M}=(\mathbb{N}^{\mathcal{M}},\widetilde{S},C,D). Define S=⋃nβˆˆΟ‰{X∈S~:β„³βŠ§X≀TDn}S=\bigcup_{n\in\omega}\{X\in\widetilde{S}:\mathcal{M}\models X\leq_{\mathrm{T}}D_{n}\}.

By the definition, for any X∈SX\in S, there exists Y∈SY\in S such that β„³βŠ§Y=HJ⁑(X)\mathcal{M}\models Y=\operatorname{HJ}(X). Thus, by Lemma 2.27, (β„•,S)(\mathbb{N},S) is a model of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} and a Ξ²\beta-submodel of (β„•,S~)(\mathbb{N},\widetilde{S}). We show that (β„•β„³,S)(\mathbb{N}^{\mathcal{M}},S) is a model of Β¬βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z)\lnot\forall X\exists Y\forall Z\theta(X,Y,Z). In particular, (β„•β„³,S)βŠ§βˆ€Yβ€‹βˆƒZ​¬θ​(C,Y,Z)(\mathbb{N}^{\mathcal{M}},S)\models\forall Y\exists Z\lnot\theta(C,Y,Z). Let Y∈SY\in S. Then, β„³βŠ§Y≀TDk\mathcal{M}\models Y\leq_{\mathrm{T}}D_{k} for some kk and thus β„³βŠ§βˆƒZ​¬θ​(C,Y,Z)\mathcal{M}\models\exists Z\lnot\theta(C,Y,Z). Hence (β„•,S)βŠ§βˆƒZ​¬θ​(C,Y,Z)(\mathbb{N},S)\models\exists Z\lnot\theta(C,Y,Z). ∎

Remark 3.3.

We note that this theorem generalizes a result by Montalban and Shore [8, Theorem 6.8.]. They proved that if

Ξ 11βˆ’π–’π– 0βŠ’βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z),\displaystyle\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\forall X\exists Y\forall Z\theta(X,Y,Z),

then there exists nβˆˆΟ‰n\in\omega such that

Ξ 11βˆ’π–’π– 0βŠ’βˆ€XβˆƒY≀THJn(X)βˆ€ZΞΈ(X,Y,Z))).\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\forall X\exists Y\leq_{\mathrm{T}}\operatorname{HJ}^{n}(X)\forall Z\theta(X,Y,Z))).

We give a Ξ 21\Pi^{1}_{2} variant of the statement that says any set is contained in a coded Ξ²\beta-model. For this purpose, we consider Ξ²\beta-submodels of a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} instead of the acutual Ξ²\beta-models.

Definition 3.4.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let β„³0\mathcal{M}_{0} and β„³1\mathcal{M}_{1} be coded Ο‰\omega-models such that β„³0βˆˆβ„³1\mathcal{M}_{0}\in\mathcal{M}_{1}. We say β„³0\mathcal{M}_{0} is a coded Ξ²\beta-submodel of β„³1\mathcal{M}_{1} if

βˆ€Xβˆˆβ„³0βˆ€e,x(β„³0βŠ§Οƒ11(e,x,X)↔ℳ1βŠ§Οƒ11(e,x,X))\displaystyle\forall X\in\mathcal{M}_{0}\forall e,x(\mathcal{M}_{0}\models\sigma^{1}_{1}(e,x,X)\leftrightarrow\mathcal{M}_{1}\models\sigma^{1}_{1}(e,x,X))

holds. We write β„³0βˆˆΞ²β„³1\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1} to mean β„³0\mathcal{M}_{0} is a coded Ξ²\beta-submodel of β„³1\mathcal{M}_{1}.

Definition 3.5.

For each nβˆˆΟ‰n\in\omega, define Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) as the following assertion:

βˆ€Xβ€‹βˆƒβ„³0,…,β„³n:codedΒ Ο‰-models ​(Xβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³nβˆ§β„³nβŠ§π– π–’π– 0).\forall X\exists\mathcal{M}_{0},\ldots,\mathcal{M}_{n}:\text{coded $\omega$-models }(X\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n}\land\mathcal{M}_{n}\models\mathsf{ACA}_{0}).
Remark 3.6.

This form of Ο‰\omega-model reflection is first introduced in [9] as their ψ0​(1,n)\psi_{0}(1,n). They mainly considered any (nonstandard) length of sequences of Ξ²\beta-submodels, but we mainly consider standard length of those.

It is easy to see that the implications Ξ²01​𝖱π–₯𝖭⁑(0)←β01​𝖱π–₯𝖭⁑(1)←⋯\beta^{1}_{0}\operatorname{\mathsf{RFN}}(0)\leftarrow\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1)\leftarrow\cdots holds over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. More precisely, these implications are strict.

Lemma 3.7.

Let nβˆˆΟ‰n\in\omega such that n>0n>0. Then, over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) proves the Ο‰\omega-model reflection of Ξ²01​𝖱π–₯𝖭⁑(nβˆ’1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n-1). In particular, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is strictly stronger than Ξ²01​𝖱π–₯𝖭⁑(nβˆ’1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n-1).

Proof.

Assume 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n+1)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1). Take a sequence of Ο‰\omega-models such that β„³0βˆˆΞ²β„³1β€‹β‹―βˆˆΞ²β„³nβˆ§β„³nβŠ§π– π–’π– 0\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\cdots\in_{\beta}\mathcal{M}_{n}\land\mathcal{M}_{n}\models\mathsf{ACA}_{0}. We show that β„³0⊧β01​𝖱π–₯𝖭⁑(n)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n). For any Xβˆˆβ„³0X\in\mathcal{M}_{0}, β„³n\mathcal{M}_{n} satisfies

βˆƒπ’©0,β‹―,𝒩nβˆ’1​(Xβˆˆπ’©0βˆˆΞ²β‹―βˆˆΞ²π’©nβˆ’1βˆ§π’©nβˆ’1βŠ§π– π–’π– 0)\displaystyle\exists\mathcal{N}_{0},\cdots,\mathcal{N}_{n-1}(X\in\mathcal{N}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{N}_{n-1}\land\mathcal{N}_{n-1}\models\mathsf{ACA}_{0})

via taking (𝒩0,…,𝒩nβˆ’1)=(β„³0,…,β„³nβˆ’1)(\mathcal{N}_{0},\ldots,\mathcal{N}_{n-1})=(\mathcal{M}_{0},\ldots,\mathcal{M}_{n-1}). Since this is a Ξ£11\Sigma^{1}_{1} formula and β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³n\mathcal{M}_{n}, β„³0\mathcal{M}_{0} also satisfies

βˆƒπ’©0,β‹―,𝒩nβˆ’1​(Xβˆˆπ’©0βˆˆΞ²β‹―βˆˆΞ²π’©nβˆ’1βˆ§π’©nβˆ’1βŠ§π– π–’π– 0).\displaystyle\exists\mathcal{N}_{0},\cdots,\mathcal{N}_{n-1}(X\in\mathcal{N}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{N}_{n-1}\land\mathcal{N}_{n-1}\models\mathsf{ACA}_{0}).

Since XX is arbitrary, β„³0⊧β01​𝖱π–₯𝖭⁑(n)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n). ∎

Lemma 3.8.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is equivalent to the following assertion:

βˆ€Xβ€‹βˆƒβ„³:codedΒ Ο‰-model ​(Xβˆˆβ„³βˆ§β„³βŠ§π– π–’π– 0+βˆƒY​(Y=HJn⁑(X))).\forall X\exists\mathcal{M}:\text{coded $\omega$-model }(X\in\mathcal{M}\land\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}^{n}(X))).
Proof.

Immediate from the fact the existence of HJ⁑(X)\operatorname{HJ}(X) is equivalent to the existence of a Ξ²\beta-model containing XX over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. ∎

Theorem 3.9.

Let θ​(X,Y)\theta(X,Y) be an arithmetical formula such that βˆ€Xβ€‹βˆƒY​θ​(X,Y)\forall X\exists Y\theta(X,Y) is provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. Then, there exists nβˆˆΟ‰n\in\omega such that 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) proves βˆ€Xβ€‹βˆƒY​θ​(X,Y)\forall X\exists Y\theta(X,Y).

Proof.

Let Οƒn\sigma_{n} be the following sentence.

βˆ€Xβ€‹βˆƒβ„³:codedΒ Ο‰-model ​(Xβˆˆβ„³βˆ§β„³βŠ§π– π–’π– 0+βˆƒY​(Y=HJn⁑(X))).\displaystyle\forall X\exists\mathcal{M}:\text{coded $\omega$-model }(X\in\mathcal{M}\land\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}^{n}(X))).

Then, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is equivalent to Οƒn\sigma_{n} over 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

We show that βˆ€Xβ€‹βˆƒY​θ​(X,Y)\forall X\exists Y\theta(X,Y) is provable from the theory 𝖠𝖒𝖠0+{Οƒn:nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\sigma_{n}:n\in\omega\}. Since βˆ€Xβ€‹βˆƒY​θ​(X,Y)\forall X\exists Y\theta(X,Y) is provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, there exists nβˆˆΟ‰n\in\omega such that

𝖠𝖒𝖠0βŠ’βˆ€X,W​(W=HJn⁑(X)β†’βˆƒY≀TW​θ​(X,Y))\mathsf{ACA}_{0}\vdash\forall X,W(W=\operatorname{HJ}^{n}(X)\to\exists Y\leq_{\mathrm{T}}W\theta(X,Y))

by Thereom 3.2. Take such an nn. Assume 𝖠𝖒𝖠0+Οƒn\mathsf{ACA}_{0}+\sigma_{n}. Let XX be an arbitrary set. Then, there is a coded Ο‰\omega-model β„³\mathcal{M} such that

Xβˆˆβ„³βˆ§β„³βŠ§π– π–’π– 0+βˆƒW​(W=HJn⁑(X)).X\in\mathcal{M}\land\mathcal{M}\models\mathsf{ACA}_{0}+\exists W(W=\operatorname{HJ}^{n}(X)).

Take such β„³\mathcal{M} and WW. Then β„³βŠ§βˆƒY≀TW​θ​(X,Y)\mathcal{M}\models\exists Y\leq_{\mathrm{T}}W\theta(X,Y) and hence β„³βŠ§βˆƒY​θ​(X,Y)\mathcal{M}\models\exists Y\theta(X,Y). Since ΞΈ\theta is arithmetical, we have βˆƒY​θ​(X,Y)\exists Y\theta(X,Y). This completes the proof. ∎

Since each Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is a Ξ 21\Pi^{1}_{2} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, the theory 𝖠𝖒𝖠0+{Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} identifies the family of all Ξ 21\Pi^{1}_{2}-consequences of Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, {ΟƒβˆˆΞ 21:Ξ 11βˆ’π–’π– 0βŠ’Οƒ}\{\sigma\in\Pi^{1}_{2}:\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}\vdash\sigma\}. In Sections 4–7, we see the relationship between the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy and certain Ξ 21\Pi^{1}_{2} theorems provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. We note that Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} may prove Ξ 21\Pi^{1}_{2} theorems faster than 𝖠𝖒𝖠0+{Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\}. More precisely, Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} has iterated exponential speedup over 𝖠𝖒𝖠0+{Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} when Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is expressed by using the nn-th numeral. This can be seen as follows. Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} proves that Ξ²01​𝖱π–₯𝖭⁑(x)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(x) is inductive (i.e., βˆ€x​(Ξ²01​𝖱π–₯𝖭⁑(x)β†’Ξ²01​𝖱π–₯𝖭⁑(x+1))\forall x(\beta^{1}_{0}\operatorname{\mathsf{RFN}}(x)\to\beta^{1}_{0}\operatorname{\mathsf{RFN}}(x+1)) holds). Then, by Solovay’s shortening method (see 3.4–3.5 of [11]), we get a poly(n)(n)-size proof of Ξ²01​𝖱π–₯𝖭⁑(2n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2_{n}) (where 20=12_{0}=1 and 2n+1=22n2_{n+1}=2^{2_{n}}, and note that 2x2_{x} can be expressed as a Ξ£1\Sigma_{1} formula of size O​(log⁑n)O(\log n)). Since Ξ²01​𝖱π–₯𝖭⁑(2n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2_{n}) implies the consistency of 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(m)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(m) for all m<2nm<2_{n}, the proof of Ξ²01​𝖱π–₯𝖭⁑(2n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2_{n}) from 𝖠𝖒𝖠0+{Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\mathsf{ACA}_{0}+\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} exactly needs the 2n2_{n}-th axiom of Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}} whose size is bigger than 2n2_{n}.

In the remaining of this section, we see some basic properties of Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n). We note that Ξ²01​𝖱π–₯𝖭⁑(0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(0) is the same as the Ο‰\omega-model reflection of 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Thus, Ξ²01​𝖱π–₯𝖭⁑(0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(0) is equivalent to 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} over 𝖱𝖒𝖠0\mathsf{RCA}_{0}. Therefore, the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy is above 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}. Moreover, it is easy to prove that even Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is stronger than 𝖠𝖳𝖱0\mathsf{ATR}_{0}.

Lemma 3.10 ([13], Theorem VII.2.7.).

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, any coded Ξ²\beta-model satisfies 𝖠𝖳𝖱0\mathsf{ATR}_{0}.

Theorem 3.11.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) implies the Ο‰\omega-model reflection of 𝖠𝖳𝖱0\mathsf{ATR}_{0}.

Proof.

We reason in 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(1)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1). Let XX be a set. By Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1), take coded Ο‰\omega-models β„³0\mathcal{M}_{0} and β„³1\mathcal{M}_{1} such that

Xβˆˆβ„³0βˆˆΞ²β„³1βˆ§β„³1βŠ§π– π–’π– 0.\displaystyle X\in\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\land\mathcal{M}_{1}\models\mathsf{ACA}_{0}.

Then, β„³1⊧[β„³0βŠ§π– π–³π–±0]\mathcal{M}_{1}\models[\mathcal{M}_{0}\models\mathsf{ATR}_{0}]. Thus β„³0βŠ§π– π–³π–±0\mathcal{M}_{0}\models\mathsf{ATR}_{0} actually holds. ∎

In the following section, we will show that there is a very large gap between 𝖠𝖳𝖱0\mathsf{ATR}_{0} and Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1).

4 Relativized leftmost path principle and Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}

As we have noted in introduction, Towsner introduced a family of Ξ 21\Pi^{1}_{2} statements called relative leftmost path principles to give an upper bound of some Ξ 21\Pi^{1}_{2} theorems provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} [18]. Indeed, he proved that Kruskal’s theorem for trees, Nash-Williams’ theorem in bqo theory and Menger’s theorem for countable graphs are provable from one of the relative leftmost path principles.

In this section, we compare the relative leftmost path principles and the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy. More precisely, we introduce a variant of relative leftmost path principle which we call arithmetical relative leftmost path principle (𝖠𝖫𝖯𝖯\mathsf{ALPP}) and show that 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1), and we show that the strongest form of relative leftmost path principle, the transfinite relativized leftmost path principle 𝖳𝖫𝖯𝖯\mathsf{TLPP}, is weaker than Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2). Moreover, we introduce the nn-th iteration of 𝖠𝖫𝖯𝖯\mathsf{ALPP} denoted 𝖨𝗍n​(𝖠𝖫𝖯𝖯)\mathsf{It}^{n}(\mathsf{ALPP}), which is equivalent to Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n).

To define 𝖠𝖫𝖯𝖯\mathsf{ALPP}, we introduce the notion of arithmetical reducibility.

Definition 4.1 (Arithmetical reducibility, 𝖱𝖒𝖠0\mathsf{RCA}_{0}).

Let X,YX,Y be sets. We say XX is arithmetically reducible to YY (write X≀TaYX\leq_{\mathrm{T}}^{\mathrm{a}}Y) if

βˆƒnβ€‹βˆƒZ​(Z=Y(n)∧X≀TZ).\exists n\exists Z(Z=Y^{(n)}\land X\leq_{\mathrm{T}}Z).

Let φ​(X)\varphi(X) be a formula. We write βˆ€X≀TaY​φ​(X)\forall X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) to mean βˆ€X​(X≀TaY→φ​(X))\forall X(X\leq_{\mathrm{T}}^{\mathrm{a}}Y\to\varphi(X)) and write βˆƒX≀TaY​φ​(X)\exists X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) to mean βˆƒX​(X≀TaYβˆ§Ο†β€‹(X))\exists X(X\leq_{\mathrm{T}}^{\mathrm{a}}Y\land\varphi(X)).

Remark 4.2.

The notation ≀Ta\leq_{\mathrm{T}}^{\mathrm{a}} follows from the notation ≀Wa\leq_{\mathrm{W}}^{\mathrm{a}} of arithmetical Weihrauch reducibility.

Remark 4.3.

We note that arithmetical reducibility is an extension of the notion of arithmetical definability. For A,BβŠ‚Ο‰A,B\subset\omega, we say AA is arithmetically definable from BB if there is an arithmetical formula θ​(x,X)\theta(x,X) with exactly displayed free variables such that A={x:θ​(x,B)}A=\{x:\theta(x,B)\}. Thus, AA is arithmetically definable from BB is equivalent to the condition that for any Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\operatorname{\mathsf{ACA}}_{0} containing AA and BB, β„³βŠ§A≀TaB\mathcal{M}\models A\leq^{\mathrm{a}}_{\mathrm{T}}B.

In contrast to this remark, for a non Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\mathsf{ACA}_{0} and A,Bβˆˆβ„³A,B\in\mathcal{M}, β„³βŠ§A≀TaB\mathcal{M}\models A\leq^{\mathrm{a}}_{\mathrm{T}}B does not mean that there is an arithmetical formula ΞΈ\theta such that β„³βŠ§A={x:θ​(x,B)}\mathcal{M}\models A=\{x:\theta(x,B)\}. However, the connection between arithmetical reducibility and arithmetical definability still holds in the following sense.

Lemma 4.4.

(𝖠𝖒𝖠0)(\mathsf{ACA}_{0}) Let β„³\mathcal{M} be a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Then, β„³\mathcal{M} is closed under arithmetical reduction. Hence, for any arithmetical formula ΞΈ\theta and Yβˆˆβ„³Y\in\mathcal{M},

βˆƒX≀TaY​θ​(X)\displaystyle\exists X\leq^{\mathrm{a}}_{\mathrm{T}}Y\theta(X) β†’[β„³βŠ§βˆƒX≀TaY​θ​(X)]​ and\displaystyle\to[\mathcal{M}\models\exists X\leq^{\mathrm{a}}_{\mathrm{T}}Y\theta(X)]\text{ and }
[β„³βŠ§βˆ€X≀TaY​θ​(X)]\displaystyle[\mathcal{M}\models\forall X\leq^{\mathrm{a}}_{\mathrm{T}}Y\theta(X)] β†’βˆ€X≀TaY​θ​(X).\displaystyle\to\forall X\leq^{\mathrm{a}}_{\mathrm{T}}Y\theta(X).
Proof.

Let β„³\mathcal{M} be an Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Then, β„³\mathcal{M} satisfies Ξ£11\Sigma^{1}_{1} induction and βˆ€Yβ€‹βˆƒZ​(Z=TJ​(Y))\forall Y\exists Z(Z=\mathrm{TJ}(Y)) where TJ\mathrm{TJ} denotes the Turing jump operator. Thus β„³βŠ§βˆ€Yβ€‹βˆ€nβ€‹βˆƒZ​(Z=Y(n))\mathcal{M}\models\forall Y\forall n\exists Z(Z=Y^{(n)}).

To show that β„³\mathcal{M} is closed under arithmetical reduction, take Yβˆˆβ„³Y\in\mathcal{M} and XX such that X≀TaYX\leq_{\mathrm{T}}^{\mathrm{a}}Y. Then, there exist n,Zn,Z such that Z=Y(n)∧X≀TZZ=Y^{(n)}\land X\leq_{\mathrm{T}}Z. We note that Zβˆˆβ„³Z\in\mathcal{M} because the condition Z=Y(n)Z=Y^{(n)} is absolute for β„³\mathcal{M}. Thus, Xβˆˆβ„³X\in\mathcal{M}.

The latter part immediately follows from the former part. ∎

Lemma 4.5.

Let φ​(X)\varphi(X) be an arithmetical formula. Then, over 𝖠𝖒𝖠0+\mathsf{ACA}_{0}^{+}, βˆ€X≀TaY​φ​(X)\forall X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) is Ξ”11\Delta^{1}_{1}. Similarly, βˆƒX≀TaY​φ​(X)\exists X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) is Ξ”11\Delta^{1}_{1}.

Proof.

Within 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, βˆ€X≀TaY​φ​(X)\forall X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) can be written as follows.

βˆ€X≀TaY​φ​(X)\displaystyle\forall X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) β†”βˆ€Z​(Z=Y(Ο‰)β†’βˆ€nβ€‹βˆ€X≀TZn​φ​(X))\displaystyle\leftrightarrow\forall Z(Z=Y^{(\omega)}\to\forall n\forall X\leq_{\mathrm{T}}Z_{n}\varphi(X))
β†”βˆƒZ​(Z=Y(Ο‰)βˆ§βˆ€nβ€‹βˆ€X≀TZn​φ​(X))\displaystyle\leftrightarrow\exists Z(Z=Y^{(\omega)}\land\forall n\forall X\leq_{\mathrm{T}}Z_{n}\varphi(X))

where ZnZ_{n} denotes the nn-th segment of ZZ as introduced in Section 2. Thus, in this case, ZnZ_{n} is Y(n)Y^{(n)}. From the above equuivalences, βˆ€X≀TaY​φ​(X)\forall X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) is Ξ”11\Delta^{1}_{1}. The case for βˆƒX≀TaY​φ​(X)\exists X\leq_{\mathrm{T}}^{\mathrm{a}}Y\varphi(X) is the same. ∎

Lemma 4.6 (Folklore).

Assume 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}. Then, for any XX, the coded model β„³={Y:Y≀TaX}\mathcal{M}=\{Y:Y\leq_{\mathrm{T}}^{\mathrm{a}}X\} exists. Moreover, for any coded Ο‰\omega-model β„³β€²\mathcal{M}^{\prime} of 𝖠𝖒𝖠0\mathsf{ACA}_{0} containing XX,

βˆ€Y​(Yβˆˆβ„³β†’Yβˆˆβ„³β€²)\displaystyle\forall Y(Y\in\mathcal{M}\to Y\in\mathcal{M}^{\prime})

holds. In this sense, β„³\mathcal{M} is the smallest coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} containing XX.

Definition 4.7 (Relative leftmost path principle, See [18]).

Let kβˆˆΟ‰k\in\omega. We define Ξ”k0​𝖫𝖯𝖯\Delta^{0}_{k}\operatorname{\mathsf{LPP}} as the assertion that for any ill-founded tree TT, there is a path g∈[T]g\in[T] such that

βˆ€fβˆˆΞ”k0​(TβŠ•g)​(f∈[T]β†’g≀lf).\forall f\in\Delta^{0}_{k}(T\oplus g)(f\in[T]\to g\leq_{l}f).

We define 𝖠𝖫𝖯𝖯\mathsf{ALPP} as the assertion that for any ill-founded tree TT, there is a path g∈[T]g\in[T] such that

βˆ€f≀Ta(TβŠ•g)​(f∈[T]β†’g≀lf).\forall f\leq^{\mathrm{a}}_{\mathrm{T}}(T\oplus g)(f\in[T]\to g\leq_{l}f).

We call a witness gg of 𝖠𝖫𝖯𝖯\mathsf{ALPP} an arithmetical leftmost path.

Finally, over 𝖠𝖳𝖱0\mathsf{ATR}_{0}, define 𝖳𝖫𝖯𝖯\mathsf{TLPP} as the following assertion: For any ill-founded tree TT and any well-order Ξ±\alpha, there is a path g∈[T]g\in[T] such that

βˆ€f≀T(TβŠ•g)(Ξ±)​(f∈[T]β†’g≀lf).\forall f\leq_{\mathrm{T}}(T\oplus g)^{(\alpha)}(f\in[T]\to g\leq_{l}f).

We call such gg a Δα+10\Delta^{0}_{\alpha+1} leftmost path.

Remark 4.8.

In the definition of Ξ”k0​𝖫𝖯𝖯,𝖠𝖫𝖯𝖯\Delta^{0}_{k}\operatorname{\mathsf{LPP}},\mathsf{ALPP} and 𝖳𝖫𝖯𝖯\mathsf{TLPP}, we can replace (TβŠ•g)(T\oplus g) by gg if we work in 𝖱𝖒𝖠0\mathsf{RCA}_{0}.

From now on, we see the relationship between Ξ²01​𝖱π–₯𝖭⁑(1),Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1),\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2) and 𝖠𝖫𝖯𝖯,𝖳𝖫𝖯𝖯\mathsf{ALPP},\mathsf{TLPP}. For this comparison, we prefer the base theory to be 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} rather than 𝖠𝖒𝖠0\mathsf{ACA}_{0} to simplify the discussion by Lemma 4.5 and 4.6. However, the following lemma take the difference of the choice of the base theory away.

Lemma 4.9.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ”00​𝖫𝖯𝖯\Delta^{0}_{0}\operatorname{\mathsf{LPP}} implies 𝖠𝖳𝖱0\mathsf{ATR}_{0}.

Proof.

See [18], Theorem 4.2. ∎

Our strategy to prove Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) from 𝖠𝖫𝖯𝖯\mathsf{ALPP} is essentially the same as for the proof of Theorem 2.13; to make a hyperjump, it is sufficient to make the set {i:[Ti]β‰ βˆ…}\{i:[T_{i}]\neq\varnothing\} from given sequence of trees ⟨Ti⟩i\langle T_{i}\rangle_{i}.

Lemma 4.10.

There are a Ξ£10\Sigma^{0}_{1} formula φ​(n,X)\varphi(n,X) and a Ξ 10\Pi^{0}_{1} formula Οˆβ€‹(n,X)\psi(n,X) such that

  • β€’

    Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, βˆ€n,X(Ο†(n,X)β†”Οˆ(n,X))\forall n,X(\varphi(n,X)\leftrightarrow\psi(n,X)).

  • β€’

    Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, for any XX, T={n:φ​(n,X)}T=\{n:\varphi(n,X)\} is a tree such that 1∞∈[T]1^{\infty}\in[T].

  • β€’

    Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, for any XX, if {n:φ​(n,X)}\{n:\varphi(n,X)\} has a leftmost path, then HJ⁑(X)\operatorname{HJ}(X) is computable from it. Moreover, this reduction is uniform.

We may assume that there is a Turing functional Ξ¦T​(X)\Phi_{T}(X) defined by Ξ¦T​(X)={n:φ​(n,X)}\Phi_{T}(X)=\{n:\varphi(n,X)\}.

Proof.

Recall that there is a Ξ£00\Sigma^{0}_{0} formula ΞΈ\theta such that over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, HJ⁑(X)={(e,x):βˆƒfβ€‹βˆ€z​θ​(e,x,X,f​[z])}\operatorname{HJ}(X)=\{(e,x):\exists f\forall z\theta(e,x,X,f[z])\} for any XX.

We give a construction of Ξ¦T​(X)\Phi_{T}(X) over 𝖱𝖒𝖠0\mathsf{RCA}_{0}. For given XX, define a sequence ⟨Te,x⟩e,x\langle T_{e,x}\rangle_{e,x} of trees by ΟƒβˆˆTe,x↔θ​(e,x,X,Οƒ)\sigma\in T_{e,x}\leftrightarrow\theta(e,x,X,\sigma). Then put Se,x={⟨0βŸ©βˆ—Οƒ:ΟƒβˆˆTe,x}βˆͺ{Οƒ:βˆ€n<|Οƒ|​σ​(n)=1}S_{e,x}=\{\langle 0\rangle\ast\sigma:\sigma\in T_{e,x}\}\cup\{\sigma:\forall n<|\sigma|\sigma(n)=1\}. Finally, put Ξ¦T​(X)=⨁e,xSe,x\Phi_{T}(X)=\bigoplus_{e,x}S_{e,x} where ⨁\bigoplus is the operator defined in Definition 2.9.

We show that this construction satisfies desired conditions. It is easy to see that Ξ¦T​(X)\Phi_{T}(X) is uniformly computable from XX over 𝖱𝖒𝖠0\mathsf{RCA}_{0} and its has an infinite path 1∞1^{\infty}. Thus, it is enough to show that HJ⁑(X)\operatorname{HJ}(X) is uniformly computable from the leftmost path of Ξ¦T​(X)\Phi_{T}(X) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}.

We reason in 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Let ff be the leftmost path of Ξ¦T​(X)\Phi_{T}(X). Then, for each ee and xx, f​((0,(e,x)))=0f((0,(e,x)))=0 iff Te,xT_{e,x} has a path. Thus, {(e,x):f​((0,(e,x)))=0}={(e,x):βˆƒfβ€‹βˆ€z​θ​(e,x,X,f​[z],z)}=HJ⁑(X)\{(e,x):f((0,(e,x)))=0\}=\{(e,x):\exists f\forall z\theta(e,x,X,f[z],z)\}=\operatorname{HJ}(X). ∎

We next see that we can modify the second condition as follows :

  • β€’

    Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, for any XX, Ξ¦T​(X)={n:φ​(n,X)}\Phi_{T}(X)=\{n:\varphi(n,X)\} is an ill-founded tree which has a path uniformly computable from XX, and any path computes XX.

For this condition, consider the following transformation.

Definition 4.11.

Let Οƒβˆˆβ„•<β„•\sigma\in\mathbb{N}^{<\mathbb{N}}. Put

Οƒeven=βŸ¨Οƒβ€‹(0),σ​(2),…,σ​(le)⟩,\displaystyle\sigma_{\mathrm{even}}=\langle\sigma(0),\sigma(2),\ldots,\sigma(l_{e})\rangle,
Οƒodd=βŸ¨Οƒβ€‹(1),σ​(3),…,σ​(lo)⟩,\displaystyle\sigma_{\mathrm{odd}}=\langle\sigma(1),\sigma(3),\ldots,\sigma(l_{o})\rangle,

where lel_{e} is the maximum even number such that le<|Οƒ|l_{e}<|\sigma| and lol_{o} is the maximum odd number such that lo<|Οƒ|l_{o}<|\sigma|. Similarly, for a function fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}}, define feven=⟨f​(2​n)⟩nβˆˆβ„•f_{\mathrm{even}}=\langle f(2n)\rangle_{n\in\mathbb{N}} and fodd=⟨f​(2​n+1)⟩nβˆˆβ„•f_{\mathrm{odd}}=\langle f(2n+1)\rangle_{n\in\mathbb{N}}.

Let XX be a set and TT be a tree. We identify XX and its characteristic function. Thus, XX is regarded as an infinite binary sequence. Define a tree ST,XS_{T,X} by

ΟƒβˆˆST,X↔σeven∈Tβˆ§Οƒoddβ‰ΊX.\sigma\in S_{T,X}\leftrightarrow\sigma_{\mathrm{even}}\in T\land\sigma_{\mathrm{odd}}\prec X.
Lemma 4.12.

Let XX be a set and TT be a tree. Then, over 𝖱𝖒𝖠0\mathsf{RCA}_{0},

  • β€’

    ST,XS_{T,X} is uniformly computable from XX and TT.

  • β€’

    For any path ff of ST,XS_{T,X}, fevenf_{\mathrm{even}} is a path of TT and fodd=Xf_{\mathrm{odd}}=X. Conversely, for f∈[T]f\in[T] and XX, the sequence f⊎X=⟨f​(0),X​(0),f​(1),X​(1),β€¦βŸ©f\uplus X=\langle f(0),X(0),f(1),X(1),\ldots\rangle is a path of ST,XS_{T,X}.

  • β€’

    If ff and gg are paths of TT such that f≀lgf\leq_{l}g, then f⊎X≀lg⊎Xf\uplus X\leq_{l}g\uplus X. In particular, if ff is the leftmost path of TT, then the corresponding path f⊎Xf\uplus X is the leftmost path of ST,XS_{T,X}.

Proof.

This is immediate from the definition of ST,XS_{T,X}. ∎

Therefore, by replacing Ξ¦T​(X)\Phi_{T}(X) with SΞ¦T​(X),XS_{\Phi_{T}(X),X}, we have

Lemma 4.13.

There is a Turing functional Ξ¦T​(βˆ™)\Phi_{T}(\bullet) defined in 𝖱𝖒𝖠0\mathsf{RCA}_{0} such that

  • β€’

    Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, for any XX, Ξ¦T​(X)\Phi_{T}(X) is an ill-founded tree which has a path uniformly computable from XX, and any path computes XX.

  • β€’

    Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, for any XX, if Ξ¦T​(X)\Phi_{T}(X) has a leftmost path, then HJ⁑(X)\operatorname{HJ}(X) is computable from that path. Moreover, this reduction is uniform.

Theorem 4.14.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1).

Proof.

First, we show that 𝖠𝖫𝖯𝖯\mathsf{ALPP} implies Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1). Since Ξ”00​𝖫𝖯𝖯\Delta^{0}_{0}\operatorname{\mathsf{LPP}} implies 𝖠𝖳𝖱0\mathsf{ATR}_{0}, it is enough to show that 𝖠𝖒𝖠0++𝖠𝖫𝖯𝖯\operatorname{\mathsf{ACA}}_{0}^{+}+\mathsf{ALPP} proves Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1).

Let XX be a set. By 𝖠𝖫𝖯𝖯\mathsf{ALPP}, take an arithmetical leftmost path gg of Ξ¦T​(X)\Phi_{T}(X). Let β„³={Y:Y≀Tag}\mathcal{M}=\{Y:Y\leq_{\mathrm{T}}^{\mathrm{a}}g\}. Then X,Ξ¦T​(X)βˆˆβ„³X,\Phi_{T}(X)\in\mathcal{M}, β„³βŠ§π– π–’π– 0\mathcal{M}\models\mathsf{ACA}_{0} and gg is the leftmostM path of Ξ¦T​(X)\Phi_{T}(X). Thus, β„³\mathcal{M} satisfies HJ⁑(X)\operatorname{HJ}(X) exists.

We next show that Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) implies 𝖠𝖫𝖯𝖯\mathsf{ALPP}. Let TT be an ill-founded tree. By Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1), take an Ο‰\omega-model β„³\mathcal{M} such that Tβˆˆβ„³T\in\mathcal{M} and β„³βŠ§π– π–’π– 0+βˆƒY​(Y=HJ⁑(T))\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}(T)).

Now, β„³\mathcal{M} satisfies [TT has a leftmost path] because its hyperjump exists. Take the leftmostM path gg of TT. We claim that gg is an arithmetical leftmost path of TT. Let f≀Tagf\leq_{\mathrm{T}}^{\mathrm{a}}g and f∈[T]f\in[T]. Since gβˆˆβ„³g\in\mathcal{M} and β„³βŠ§π– π–’π– 0\mathcal{M}\models\mathsf{ACA}_{0}, fβˆˆβ„³f\in\mathcal{M}. Since gg is the leftmostM path, g≀lfg\leq_{l}f. ∎

From now on, for theories TT and Tβ€²T^{\prime}, we write T<Tβ€²T<T^{\prime} to mean Tβ€²T^{\prime} proves T+π–’π—ˆπ—‡β‘(T)T+\operatorname{\mathsf{Con}}(T) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. It is shown in [18] that 𝖠𝖫𝖯𝖯<𝖳𝖫𝖯𝖯\mathsf{ALPP}<\mathsf{TLPP}. More precisely, a general form of relative leftmost path principle Σα0​𝖫𝖯𝖯\Sigma^{0}_{\alpha}\operatorname{\mathsf{LPP}} for a well order Ξ±\alpha is introduced, and it is proved that Σα+50​𝖫𝖯𝖯\Sigma^{0}_{\alpha+5}\operatorname{\mathsf{LPP}} proves the consistency of Σα+10​𝖫𝖯𝖯\Sigma^{0}_{\alpha+1}\operatorname{\mathsf{LPP}}. Since 𝖠𝖫𝖯𝖯\mathsf{ALPP} is provable from Σω+10​𝖫𝖯𝖯\Sigma^{0}_{\omega+1}\operatorname{\mathsf{LPP}}, the consistency of 𝖠𝖫𝖯𝖯\mathsf{ALPP} is provable from 𝖳𝖫𝖯𝖯\mathsf{TLPP}. Thus we have

Theorem 4.15.

Ξ²01​𝖱π–₯𝖭⁑(1)<𝖳𝖫𝖯𝖯\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1)<\mathsf{TLPP}.

On the other hand, 𝖳𝖫𝖯𝖯\mathsf{TLPP} is strictly weaker than Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2). To prove this result, we introduce a variant of Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n).

Definition 4.16.

Let Οƒ\sigma be a Ξ 21\Pi^{1}_{2} formula. We define Ξ²01​𝖱π–₯𝖭⁑(n;Οƒ)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\sigma) as the following assertion:

βˆ€Xβˆƒβ„³0,…,β„³n:codedΒ Ο‰-modelsΒ (\displaystyle\forall X\exists\mathcal{M}_{0},\ldots,\mathcal{M}_{n}:\text{coded $\omega$-models }( Xβˆˆβ„³0βˆˆΞ²β„³1β€‹β‹―βˆˆΞ²β„³n∧\displaystyle X\in\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\cdots\in_{\beta}\mathcal{M}_{n}\,\land
β„³nβŠ§π– π–’π– 0+Οƒ).\displaystyle\mathcal{M}_{n}\models\mathsf{ACA}_{0}+\,\sigma).
Lemma 4.17.

Over 𝖠𝖳𝖱0\mathsf{ATR}_{0}, any Ξ²\beta-model satisfies 𝖳𝖫𝖯𝖯\mathsf{TLPP}.

Proof.

Let β„³\mathcal{M} be a Ξ²\beta-model. Let T,Ξ±βˆˆβ„³T,\alpha\in\mathcal{M} such that

β„³βŠ§\displaystyle\mathcal{M}\models\ T​ is an ill-founded tree ∧\displaystyle T\text{ is an ill-founded tree }\land
α​ is a well order.\displaystyle\alpha\text{ is a well order}.

We show that β„³\mathcal{M} has a Δα0\Delta^{0}_{\alpha} leftmost path.

Let S={ΟƒβˆˆT:βˆƒfβˆˆβ„³β€‹(Οƒβ‰Ίf∧f∈[T])}S=\{\sigma\in T:\exists f\in\mathcal{M}(\sigma\prec f\land f\in[T])\}. Since β„³\mathcal{M} is a Ξ²\beta-model, S={ΟƒβˆˆT:βˆƒf​(Οƒβ‰Ίf∈[T])}S=\{\sigma\in T:\exists f(\sigma\prec f\in[T])\}. Now, SS is a nonempty pruned tree and hence its leftmost path gg is computable from SS. Moreover, gg is the leftmost path of TT.

Now we have βˆƒgβ€‹βˆ€f∈[T]​(g≀lf)\exists g\forall f\in[T](g\leq_{l}f) and hence

βˆƒgβ€‹βˆ€f≀Tg(Ξ±)​(f∈[T]β†’g≀lf).\exists g\forall f\leq_{\mathrm{T}}g^{(\alpha)}(f\in[T]\to g\leq_{l}f).

The above condition is Ξ£11\Sigma^{1}_{1}. Therefore,

β„³βŠ§βˆƒgβ€‹βˆ€f≀Tg(Ξ±)​(f∈[T]β†’g≀lf).\mathcal{M}\models\exists g\forall f\leq_{\mathrm{T}}g^{(\alpha)}(f\in[T]\to g\leq_{l}f).

This completes the proof. ∎

Theorem 4.18.

Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}) proves the Ο‰\omega-model reflection of 𝖳𝖫𝖯𝖯\mathsf{TLPP} over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Therefore, 𝖳𝖫𝖯𝖯<Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\mathsf{TLPP}<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}).

Proof.

Assume 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}). Let XX be a set. Take coded Ο‰\omega-models β„³0,β„³1\mathcal{M}_{0},\mathcal{M}_{1} such that

Xβˆˆβ„³0βˆˆΞ²β„³1βˆ§β„³1βŠ§π– π–³π–±0.\displaystyle X\in\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\land\mathcal{M}_{1}\models\mathsf{ATR}_{0}.

Then, since β„³0\mathcal{M}_{0} is a coded Ξ²\beta-submodel of β„³1\mathcal{M}_{1}, β„³0βŠ§π–³π–«π–―π–―\mathcal{M}_{0}\models\mathsf{TLPP}. This completes the proof. ∎

Theorem 4.19.

Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2) proves the Ο‰\omega-model reflection of Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}). Therefore, Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)<Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0})<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2).

Proof.

Let XX be a set. By Ξ²01​𝖱π–₯𝖭⁑(2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(2), take coded Ο‰\omega-models β„³0,β„³1,β„³2\mathcal{M}_{0},\mathcal{M}_{1},\mathcal{M}_{2} such that

Xβˆˆβ„³0βˆˆΞ²β„³1βˆˆΞ²β„³2βˆ§β„³2βŠ§π– π–’π– 0.X\in\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\in_{\beta}\mathcal{M}_{2}\land\mathcal{M}_{2}\models\mathsf{ACA}_{0}.

Then β„³2⊧[β„³1βŠ§π– π–³π–±0]\mathcal{M}_{2}\models[\mathcal{M}_{1}\models\mathsf{ATR}_{0}] by Lemma 3.10. Thus, for any Yβˆˆβ„³0Y\in\mathcal{M}_{0},

β„³2βŠ§βˆƒπ’©0,𝒩1​[Yβˆˆπ’©0βˆˆΞ²π’©1βˆ§π’©1βŠ§π– π–³π–±0].\mathcal{M}_{2}\models\exists\mathcal{N}_{0},\mathcal{N}_{1}[Y\in\mathcal{N}_{0}\in_{\beta}\mathcal{N}_{1}\land\mathcal{N}_{1}\models\mathsf{ATR}_{0}].

Since β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³2\mathcal{M}_{2}, we have

β„³0βŠ§βˆ€Yβ€‹βˆƒπ’©0,𝒩1​[Yβˆˆπ’©0βˆˆΞ²π’©1βˆ§π’©1βŠ§π– π–³π–±0].\mathcal{M}_{0}\models\forall Y\exists\mathcal{N}_{0},\mathcal{N}_{1}[Y\in\mathcal{N}_{0}\in_{\beta}\mathcal{N}_{1}\land\mathcal{N}_{1}\models\mathsf{ATR}_{0}].

Hence, Xβˆˆβ„³0X\in\mathcal{M}_{0} and β„³0⊧β01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}). This completes the proof. ∎

Next, let us consider iterated versions of 𝖠𝖫𝖯𝖯\mathsf{ALPP}.

Definition 4.20.

Let nβˆˆΟ‰n\in\omega. We define Ω​(n+1)\Omega(n+1) as the following assertion: for any XX there are f0,…,fnf_{0},\ldots,f_{n} such that

f0∈[Ξ¦T​(X)]βˆ§β‹€i<nfi+1∈[Ξ¦T​(fi)]∧\displaystyle f_{0}\in[\Phi_{T}(X)]\land\bigwedge_{i<n}f_{i+1}\in[\Phi_{T}(f_{i})]\,\land
βˆ€g≀Tafn​[(g∈[Ξ¦T​(X)]β†’f0≀lg)βˆ§β‹€i<n(g∈[Ξ¦T​(fi)]β†’fi+1≀lg)].\displaystyle\forall g\leq_{\mathrm{T}}^{\mathrm{a}}f_{n}[(g\in[\Phi_{T}(X)]\to f_{0}\leq_{l}g)\land\bigwedge_{i<n}(g\in[\Phi_{T}(f_{i})]\to f_{i+1}\leq_{l}g)].

We also define 𝖨𝗍n+1​(𝖠𝖫𝖯𝖯)\mathsf{It}^{n+1}(\mathsf{ALPP}) as follows. Define Ξ˜β€‹(T,f,g)\Theta(T,f,g) be the following formula:

Ξ˜β€‹(T,f,g)≑TΒ is a tree andΒ g∈[T]Β and ​T,g≀Taf.\displaystyle\Theta(T,f,g)\equiv\text{$T$ is a tree and $g\in[T]$ and }T,g\leq_{\mathrm{T}}^{\mathrm{a}}f.

For each m<nm<n, define ψmn\psi^{n}_{m} by

ψ0n​(T0,…,Tn,f0,…,fn)β‰‘βˆ€g≀Taf0βŠ•β‹―βŠ•fn​⋀i≀n(g∈[Ti]β†’fi≀lg),\displaystyle\psi^{n}_{0}(T_{0},\ldots,T_{n},f_{0},\ldots,f_{n})\equiv\forall g\leq^{\mathrm{a}}_{\mathrm{T}}f_{0}\oplus\cdots\oplus f_{n}\bigwedge_{i\leq n}(g\in[T_{i}]\to f_{i}\leq_{l}g),
ψm+1n​(T0,…,Tnβˆ’mβˆ’1,f0,…,fnβˆ’mβˆ’1)≑\displaystyle\psi^{n}_{m+1}(T_{0},\ldots,T_{n-m-1},f_{0},\ldots,f_{n-m-1})\equiv
βˆ€Tnβˆ’m,gnβˆ’m​(Ξ˜β€‹(Tnβˆ’m,gnβˆ’m,fnβˆ’mβˆ’1)β†’βˆƒfnβˆ’m∈[Tnβˆ’m]β€‹Οˆmn​(T0,…,Tnβˆ’m,f0,…,fnβˆ’m))\displaystyle\forall T_{n-m},g_{n-m}(\Theta(T_{n-m},g_{n-m},f_{n-m-1})\to\exists f_{n-m}\in[T_{n-m}]\psi^{n}_{m}(T_{0},\ldots,T_{n-m},f_{0},\ldots,f_{n-m}))

Then, define 𝖨𝗍n+1​(𝖠𝖫𝖯𝖯)\mathsf{It}^{n+1}(\mathsf{ALPP}) as follows.

𝖨𝗍n+1​(𝖠𝖫𝖯𝖯)β‰‘βˆ€T0​ : ill-founded treeΒ β€‹βˆƒf0∈[T0]β€‹Οˆnn​(T0,f0).\displaystyle\mathsf{It}^{n+1}(\mathsf{ALPP})\equiv\forall T_{0}\text{ : ill-founded tree }\exists f_{0}\in[T_{0}]\psi^{n}_{n}(T_{0},f_{0}).

For example, 𝖨𝗍1​(𝖠𝖫𝖯𝖯)\mathsf{It}^{1}(\mathsf{ALPP}) is equivalent to 𝖠𝖫𝖯𝖯\mathsf{ALPP} and 𝖨𝗍2​(𝖠𝖫𝖯𝖯)\mathsf{It}^{2}(\mathsf{ALPP}) states that for any ill-founded tree T0T_{0}, there exists a path f0∈[T0]f_{0}\in[T_{0}] such that

βˆ€T1,g1​(Ξ˜β€‹(T1,g1,f0)β†’βˆƒf1∈[T1]β€‹βˆ€g≀Taf0βŠ•f1​⋀i≀1(g∈[Ti]β†’fi≀lg)).\displaystyle\forall T_{1},g_{1}(\Theta(T_{1},g_{1},f_{0})\to\exists f_{1}\in[T_{1}]\forall g\leq_{\mathrm{T}}^{\mathrm{a}}f_{0}\oplus f_{1}\bigwedge_{i\leq 1}(g\in[T_{i}]\to f_{i}\leq_{l}g)).

For simplicity, we write (T,g)βˆˆΞ˜β€‹(f)(T,g)\in\Theta(f) instead of Ξ˜β€‹(T,f,g)\Theta(T,f,g). Then, 𝖨𝗍n+1​(𝖠𝖫𝖯𝖯)\mathsf{It}^{n+1}(\mathsf{ALPP}) states that for any ill-founded tree T0T_{0}, there exists a path f0∈[T0]f_{0}\in[T_{0}] such that

βˆ€(T1,g1)βˆˆΞ˜β€‹(f0)β€‹βˆƒf1∈[T1]β€‹β‹―β€‹βˆ€(Tn,gn)βˆˆΞ˜β€‹(fnβˆ’1)β€‹βˆƒfn∈[Tn]\displaystyle\forall(T_{1},g_{1})\in\Theta(f_{0})\exists f_{1}\in[T_{1}]\cdots\forall(T_{n},g_{n})\in\Theta(f_{n-1})\exists f_{n}\in[T_{n}]
βˆ€g≀Taf0βŠ•β‹―βŠ•fn​⋀i≀n(g∈[Ti]β†’fi≀lg).\displaystyle\forall g\leq_{\mathrm{T}}^{\mathrm{a}}f_{0}\oplus\cdots\oplus f_{n}\bigwedge_{i\leq n}(g\in[T_{i}]\to f_{i}\leq_{l}g).
Theorem 4.21.

For nβˆˆΟ‰n\in\omega, the following assertions are equivalent over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}.

  1. 1.

    Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1),

  2. 2.

    Ω​(n+1)\Omega(n+1),

  3. 3.

    𝖨𝗍n+1​(𝖠𝖫𝖯𝖯)\mathsf{It}^{n+1}(\mathsf{ALPP}).

Proof.

(1β†’31\to 3) Let T0T_{0} be an ill-founded tree and hh be a path of T0T_{0}. By Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1) take a sequence of Ο‰\omega-models such that

T0,hβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³n+1βˆ§β„³n+1βŠ§π– π–’π– 0.T_{0},h\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n+1}\land\mathcal{M}_{n+1}\models\mathsf{ACA}_{0}.

Now β„³1\mathcal{M}_{1} satisfies β€œT0T_{0} has a leftmost path”. Take the leftmostβ„³1{}^{\mathcal{M}_{1}} path f0f_{0}. Then, β„³n+1\mathcal{M}_{n+1} also satisfies β€œf0f_{0} is the leftmost path of T0T_{0}”. Let T1,g1≀Taf0T_{1},g_{1}\leq_{\mathrm{T}}^{\mathrm{a}}f_{0} such that T1T_{1} is a tree and g1∈[T1]g_{1}\in[T_{1}]. Then, β„³2\mathcal{M}_{2} satisfies β€œT1T_{1} has a leftmost path”. Let f1f_{1} be the leftmostβ„³2{}^{\mathcal{M}_{2}} path. Then, β„³n+1\mathcal{M}_{n+1} satisfies β€œf1f_{1} is the leftmost path of T1T_{1}”. Continuing this argument, we have that

βˆ€(T1,g1)βˆˆΞ˜β€‹(f0)β€‹βˆƒf1∈[T1]β€‹β‹―β€‹βˆ€(Tn,gn)βˆˆΞ˜β€‹(fnβˆ’1)β€‹βˆƒfn∈[Tn]\displaystyle\forall(T_{1},g_{1})\in\Theta(f_{0})\exists f_{1}\in[T_{1}]\cdots\forall(T_{n},g_{n})\in\Theta(f_{n-1})\exists f_{n}\in[T_{n}]
β‹€i≀n(β„³n+1⊧`​`​fi​ is the leftmost path of ​Ti​").\displaystyle\bigwedge_{i\leq n}(\mathcal{M}_{n+1}\models``f_{i}\text{ is the leftmost path of }T_{i}").

Since β„³n+1\mathcal{M}_{n+1} is an Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, it is closed under arithmetical reduction. This completes the proof.

(3β†’23\to 2) This is trivial.

(2β†’12\to 1) Let XX be a set. We show that there is an Ο‰\omega-model β„³\mathcal{M} such that β„³βŠ§π– π–’π– 0+βˆƒY​(Y=HJn+1⁑(X))\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}^{n+1}(X)). Let f0,…,fnf_{0},\ldots,f_{n} be witnesses of Ω​(n)\Omega(n) for XX. Then, f0≀T⋯≀Tfnf_{0}\leq_{\mathrm{T}}\cdots\leq_{\mathrm{T}}f_{n}. By 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, take the smallest Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\mathsf{ACA}_{0} containing fnf_{n}. Then, β„³\mathcal{M} satisfies β€œf0f_{0} is the leftmost path of Ξ¦T​(X)\Phi_{T}(X)” and hence f0f_{0} computes HJ⁑(X)\operatorname{HJ}(X) in β„³\mathcal{M}. Similarly, fif_{i} is the leftmost path of Ξ¦T​(fiβˆ’1)\Phi_{T}(f_{i-1}) in β„³\mathcal{M} and hence it computes HJ⁑(fi)\operatorname{HJ}(f_{i}) in β„³\mathcal{M}. Thus, β„³βŠ§βˆƒY​(Y=HJn+1⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{n+1}(X)). ∎

As shown in [15], Ξ”n0​𝖫𝖯𝖯\Delta^{0}_{n}\operatorname{\mathsf{LPP}}s and variants of Ξ²\beta-model reflection called Ξ”n0​β\Delta^{0}_{n}\beta-model refections are deeply related. In the following, we see the relationship of variants of leftmost path principles and variants of Ξ²\beta-models reflection.

Remark 4.22.

We can show that

  • β€’

    for any nβˆˆΟ‰n\in\omega, 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that any Ξ²\beta-model satisfies Ξ”n0​𝖫𝖯𝖯\Delta^{0}_{n}\operatorname{\mathsf{LPP}},

  • β€’

    𝖠𝖒𝖠0β€²\operatorname{\mathsf{ACA}}_{0}^{\prime} proves that for any nβˆˆβ„•n\in\mathbb{N}, any Ξ²\beta-model satisfies Ξ”n0​𝖫𝖯𝖯\Delta^{0}_{n}\operatorname{\mathsf{LPP}},

  • β€’

    𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} proves that any Ξ²\beta-model satisfies 𝖠𝖫𝖯𝖯\mathsf{ALPP}.

by the same proof of Lemma 4.17.

In contrast to this result, even 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} does not prove that any Ξ²\beta-model satisfies 𝖳𝖫𝖯𝖯\mathsf{TLPP}. To prove this, we can use the same proof as in the following lemma.

Lemma 4.23.

𝖳𝖫𝖯𝖯\mathsf{TLPP} is equivalent to

βˆ€Ξ±:WOβˆ€Xβˆƒβ„³(\displaystyle\forall\alpha:\operatorname{WO}\forall X\exists\mathcal{M}( Ξ±,Xβˆˆβ„³,\displaystyle\alpha,X\in\mathcal{M},
β„³βŠ§π– π–’π– 0+βˆƒY(Y=HJ(X))),\displaystyle\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}(X))),
β„³Β is closed underΒ Ξ±-jump).\displaystyle\text{$\mathcal{M}$ is closed under $\alpha$-jump}).
Proof.

First, assume 𝖳𝖫𝖯𝖯\mathsf{TLPP}. Take a well order Ξ±\alpha. Then, Ξ±β‹…Ο‰\alpha\cdot\omega is also a well-ordering 555Here, Ξ±β‹…Ο‰\alpha\cdot\omega denotes the order type of lexicographical ordering of β„•Γ—Ξ±\mathbb{N}\times\alpha.. We note that Ξ±β‹…Ο‰\alpha\cdot\omega is closed under +Ξ±+\alpha in the following sense: Recall that |L||L| denotes the field {i:(i,i)∈L}\{i:(i,i)\in L\} of LL for a linear order LL. For each (n,i)∈|Ξ±β‹…Ο‰|(n,i)\in|\alpha\cdot\omega|, there is a certain embedding from Ξ±\alpha to the interval [(n,i),(n+2,0)][(n,i),(n+2,0)] defined by f​(j)=(n+1,j)f(j)=(n+1,j). Since 𝖠𝖳𝖱0\mathsf{ATR}_{0} holds, there is an embedding gg from Ξ±\alpha to an initial segment of [(n,i),(n+2,0)][(n,i),(n+2,0)]. Then, supg\sup g is regarded as (n,i)+Ξ±(n,i)+\alpha.

Take an arbitrary XX. We show that

βˆƒβ„³(\displaystyle\exists\mathcal{M}( Ξ±,Xβˆˆβ„³,\displaystyle\alpha,X\in\mathcal{M},
β„³βŠ§π– π–’π– 0+βˆƒY(Y=HJ(X))),\displaystyle\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}(X))),
β„³Β is closed underΒ Ξ±-jump).\displaystyle\text{$\mathcal{M}$ is closed under $\alpha$-jump}).

Recall that there is a total Turing functional Ξ¦T\Phi_{T} which outputs an ill-founded tree for any input (see Lemma 4.13). By 𝖳𝖫𝖯𝖯\mathsf{TLPP}, take a Δα⋅ω0\Delta^{0}_{\alpha\cdot\omega}-leftmost path ff of Ξ¦T​(Ξ±βŠ•X)\Phi_{T}(\alpha\oplus X). Define

β„³={A:A≀Tf(n,i)​ for some ​(n,i)∈|Ξ±β‹…Ο‰|}.\mathcal{M}=\{A:A\leq_{\mathrm{T}}f^{(n,i)}\text{ for some }(n,i)\in|\alpha\cdot\omega|\}.

Since β„³βŠ§[fΒ is a leftmost path ofΒ Ξ¦T​(Ξ±βŠ•X)]\mathcal{M}\models[\text{$f$ is a leftmost path of $\Phi_{T}(\alpha\oplus X)$}], β„³βŠ§βˆƒY​(Y=HJ⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}(X)). Moreover, by construction, β„³βŠ§π– π–’π– 0\mathcal{M}\models\mathsf{ACA}_{0} and closed under Ξ±\alpha-jump.

The converse is trivial because the leftmost path in β„³\mathcal{M} is an Δα0\Delta^{0}_{\alpha}-leftmost path in the ground model. ∎

Recall that a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} satisfies 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} if it is closed under taking the Ο‰\omega-jump of a set. Therefore, by the same proof as above, we have

Lemma 4.24.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Δω20​𝖫𝖯𝖯\Delta^{0}_{\omega^{2}}\operatorname{\mathsf{LPP}} proves Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\operatorname{\mathsf{ACA}}_{0}^{+}).

Theorem 4.25.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, 𝖳𝖫𝖯𝖯\mathsf{TLPP} is strictly stronger than Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\operatorname{\mathsf{ACA}}_{0}^{+}).

Proof.

It is immediate from the fact that 𝖳𝖫𝖯𝖯\mathsf{TLPP} is strictly stronger than Δω20​𝖫𝖯𝖯\Delta^{0}_{\omega^{2}}\operatorname{\mathsf{LPP}}. ∎

Corollary 4.26.

𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} does not prove that any Ξ²\beta-model satisfies 𝖳𝖫𝖯𝖯\mathsf{TLPP}.

Proof.

For the sake of contradiction, assume 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} proves that any Ξ²\beta-model satisfies 𝖳𝖫𝖯𝖯\mathsf{TLPP}.

We work in 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖒𝖠0+)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\operatorname{\mathsf{ACA}}_{0}^{+}) and show that π–’π—ˆπ—‡β‘(𝖳𝖫𝖯𝖯)\operatorname{\mathsf{Con}}(\mathsf{TLPP}) holds. Take coded Ο‰\omega-models β„³0,β„³1\mathcal{M}_{0},\mathcal{M}_{1} such that β„³0βˆˆΞ²β„³1\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1} and β„³1βŠ§π– π–’π– 0+\mathcal{M}_{1}\models\operatorname{\mathsf{ACA}}_{0}^{+}. Then, β„³1⊧[β„³0βŠ§π–³π–«π–―π–―]\mathcal{M}_{1}\models[\mathcal{M}_{0}\models\mathsf{TLPP}] by assumption and hence β„³1βŠ§π–’π—ˆπ—‡β‘(𝖳𝖫𝖯𝖯)\mathcal{M}_{1}\models\operatorname{\mathsf{Con}}(\mathsf{TLPP}). Since π–’π—ˆπ—‡β‘(𝖳𝖫𝖯𝖯)\operatorname{\mathsf{Con}}(\mathsf{TLPP}) is arithmetical, π–’π—ˆπ—‡β‘(𝖳𝖫𝖯𝖯)\operatorname{\mathsf{Con}}(\mathsf{TLPP}) holds. However, this is impossible because 𝖳𝖫𝖯𝖯\mathsf{TLPP} implies Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\operatorname{\mathsf{ACA}}_{0}^{+}) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. ∎

Theorem 4.27.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n;𝖠𝖫𝖯𝖯)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{ALPP}) implies the Ο‰\omega-model reflection of Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1).

Proof.

We reason in 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n;𝖠𝖫𝖯𝖯)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{ALPP}). Take an arbitrary set AA. By Ξ²01​𝖱π–₯𝖭⁑(n;𝖠𝖫𝖯𝖯)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{ALPP}), take coded Ο‰\omega-models β„³0,…,β„³n\mathcal{M}_{0},\ldots,\mathcal{M}_{n} such that

Aβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³nβˆ’1βˆˆΞ²β„³nβˆ§β„³nβŠ§π– π–’π– 0+𝖠𝖫𝖯𝖯.\displaystyle A\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n-1}\in_{\beta}\mathcal{M}_{n}\land\mathcal{M}_{n}\models\mathsf{ACA}_{0}+\mathsf{ALPP}.

Since 𝖠𝖫𝖯𝖯\mathsf{ALPP} and Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is equivalent over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, β„³n⊧β01​𝖱π–₯𝖭⁑(1)\mathcal{M}_{n}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1).

We show that β„³0⊧β01​𝖱π–₯𝖭⁑(n+1)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1). By Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) in β„³n\mathcal{M}_{n}, there are coded Ο‰\omega-model 𝒩0,𝒩1βˆˆβ„³n\mathcal{N}_{0},\mathcal{N}_{1}\in\mathcal{M}_{n} such that

β„³n⊧[β„³nβˆ’1βˆˆπ’©0βˆˆΞ²π’©1βˆ§π’©1βŠ§π– π–’π– 0].\displaystyle\mathcal{M}_{n}\models[\mathcal{M}_{n-1}\in\mathcal{N}_{0}\in_{\beta}\mathcal{N}_{1}\land\mathcal{N}_{1}\models\mathsf{ACA}_{0}].

Since β„³nβˆ’1\mathcal{M}_{n-1} is a Ξ²\beta-submodel of β„³n\mathcal{M}_{n} and 𝒩0\mathcal{N}_{0} is an Ο‰\omega-submodel of β„³n\mathcal{M}_{n}, for any Ξ 11\Pi^{1}_{1} sentence Οƒ\sigma with parameters from β„³nβˆ’1\mathcal{M}_{n-1},

[β„³nβˆ’1βŠ§Οƒ]\displaystyle[\mathcal{M}_{n-1}\models\sigma] β†’[β„³nβŠ§Οƒ]\displaystyle\to[\mathcal{M}_{n}\models\sigma]
β†’[𝒩0βŠ§Οƒ].\displaystyle\to[\mathcal{N}_{0}\models\sigma].

Thus, β„³nβˆ’1\mathcal{M}_{n-1} is a Ξ²\beta-submodel of 𝒩0\mathcal{N}_{0}. Therefore, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³n⊧[Xβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³nβˆ’1βˆˆΞ²π’©0βˆˆΞ²π’©1βˆ§π’©1βŠ§π– π–’π– 0].\displaystyle\mathcal{M}_{n}\models[X\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n-1}\in_{\beta}\mathcal{N}_{0}\in_{\beta}\mathcal{N}_{1}\land\mathcal{N}_{1}\models\mathsf{ACA}_{0}].

Therefore, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³nβŠ§βˆƒβ„‹0,…,β„‹n+1​[Xβˆˆβ„‹0βˆˆΞ²β‹―βˆˆΞ²β„‹nβˆ’1βˆˆΞ²β„‹nβˆˆΞ²β„‹n+1βˆ§β„‹n+1βŠ§π– π–’π– 0].\displaystyle\mathcal{M}_{n}\models\exists\mathcal{H}_{0},\ldots,\mathcal{H}_{n+1}[X\in\mathcal{H}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{H}_{n-1}\in_{\beta}\mathcal{H}_{n}\in_{\beta}\mathcal{H}_{n+1}\land\mathcal{H}_{n+1}\models\mathsf{ACA}_{0}].

Since β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³n\mathcal{M}_{n}, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³0βŠ§βˆƒβ„‹0,…,β„‹n+1​[Xβˆˆβ„‹0βˆˆΞ²β‹―βˆˆΞ²β„‹nβˆ’1βˆˆΞ²β„‹nβˆˆΞ²β„‹n+1βˆ§β„‹n+1βŠ§π– π–’π– 0].\displaystyle\mathcal{M}_{0}\models\exists\mathcal{H}_{0},\ldots,\mathcal{H}_{n+1}[X\in\mathcal{H}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{H}_{n-1}\in_{\beta}\mathcal{H}_{n}\in_{\beta}\mathcal{H}_{n+1}\land\mathcal{H}_{n+1}\models\mathsf{ACA}_{0}].

Thus, β„³0⊧β01​𝖱π–₯𝖭⁑(n+1)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1). ∎

The point of the above proof is that 𝖠𝖫𝖯𝖯\mathsf{ALPP} implies Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) and hence β„³n\mathcal{M}_{n} satisfies Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1). Since 𝖳𝖫𝖯𝖯\mathsf{TLPP} implies Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\operatorname{\mathsf{ACA}}_{0}^{+}), we can also show the following theorem by the same way as the above proof.

Theorem 4.28.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n;𝖳𝖫𝖯𝖯)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{TLPP}) implies the Ο‰\omega-model reflection of Ξ²01​𝖱π–₯𝖭⁑(n+1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1;\operatorname{\mathsf{ACA}}_{0}^{+}).

Theorem 4.29.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1) proves the Ο‰\omega-model reflection of Ξ²01𝖱π–₯𝖭(n;βˆ€k.Ξ”k0𝖫𝖯𝖯)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}). Here, βˆ€k.Ξ”k0​𝖫𝖯𝖯\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}} is the assertion that

βˆ€kβ€‹βˆ€T:ill-founded treeΒ β€‹βˆƒg∈[T]β€‹βˆ€h≀Tg(k)​(h∈[T]β†’g≀lh).\displaystyle\forall k\forall T:\text{ill-founded tree }\exists g\in[T]\forall h\leq_{\mathrm{T}}g^{(k)}(h\in[T]\to g\leq_{l}h).
Proof.

We reason in 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n+1)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1). Take an arbitrary set AA. By Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1), take coded Ο‰\omega-models β„³0,…,β„³n+1\mathcal{M}_{0},\ldots,\mathcal{M}_{n+1} such that

Aβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³nβˆˆΞ²β„³n+1βˆ§β„³n+1βŠ§π– π–’π– 0.\displaystyle A\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n}\in_{\beta}\mathcal{M}_{n+1}\land\mathcal{M}_{n+1}\models\mathsf{ACA}_{0}.

We show that β„³0⊧β01𝖱π–₯𝖭(n;βˆ€k.Ξ”k0𝖫𝖯𝖯)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}). Since β„³n+1βŠ§π– π–’π– 0β€²\mathcal{M}_{n+1}\models\operatorname{\mathsf{ACA}}_{0}^{\prime}, β„³n+1⊧[β„³nβŠ§βˆ€k.Ξ”k0𝖫𝖯𝖯]\mathcal{M}_{n+1}\models[\mathcal{M}_{n}\models\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}]. Therefore, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³n+1⊧[Xβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³nβˆ§β„³nβŠ§βˆ€k.Ξ”k0𝖫𝖯𝖯].\displaystyle\mathcal{M}_{n+1}\models[X\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n}\land\mathcal{M}_{n}\models\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}].

Hence, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³n+1βŠ§βˆƒβ„‹0β‹―β„‹n[Xβˆˆβ„‹0βˆˆΞ²β‹―βˆˆΞ²β„‹nβˆ§β„‹nβŠ§βˆ€k.Ξ”k0𝖫𝖯𝖯].\displaystyle\mathcal{M}_{n+1}\models\exists\mathcal{H}_{0}\cdots\mathcal{H}_{n}[X\in\mathcal{H}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{H}_{n}\land\mathcal{H}_{n}\models\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}].

Since β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³n+1\mathcal{M}_{n+1}, for any Xβˆˆβ„³0X\in\mathcal{M}_{0},

β„³0βŠ§βˆƒβ„‹0β‹―β„‹n[Xβˆˆβ„‹0βˆˆΞ²β‹―βˆˆΞ²β„‹nβˆ§β„‹nβŠ§βˆ€k.Ξ”k0𝖫𝖯𝖯].\displaystyle\mathcal{M}_{0}\models\exists\mathcal{H}_{0}\cdots\mathcal{H}_{n}[X\in\mathcal{H}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{H}_{n}\land\mathcal{H}_{n}\models\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}].

Therefore, β„³0⊧β01𝖱π–₯𝖭(n,βˆ€k.Ξ”k0𝖫𝖯𝖯)\mathcal{M}_{0}\models\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n,\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}}). ∎

Recall that any coded Ξ²\beta-model satisfies 𝖠𝖫𝖯𝖯\mathsf{ALPP} over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}. Thus, by the same proof as above, we have

Theorem 4.30.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n+1;𝖠𝖒𝖠0+)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1;\operatorname{\mathsf{ACA}}_{0}^{+}) proves the Ο‰\omega-model reflection of Ξ²01​𝖱π–₯𝖭⁑(n;𝖠𝖫𝖯𝖯)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{ALPP}).

Summarizing previous theorems, we have a hierarchy between Ξ²01​𝖱π–₯𝖭⁑(n+1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1) and Ξ²01​𝖱π–₯𝖭⁑(n+2)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+2) as follows.

Corollary 4.31.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, the following holds.

Ξ²01​𝖱π–₯𝖭⁑(n+1)\displaystyle\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1) <Ξ²01​𝖱π–₯𝖭⁑(n;𝖠𝖫𝖯𝖯)<Ξ²01​𝖱π–₯𝖭⁑(n+1;𝖠𝖒𝖠0+)\displaystyle<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{ALPP})<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1;\operatorname{\mathsf{ACA}}_{0}^{+})
<Ξ²01​𝖱π–₯𝖭⁑(n;𝖳𝖫𝖯𝖯)<Ξ²01​𝖱π–₯𝖭⁑(n+1;𝖠𝖳𝖱0)\displaystyle<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\mathsf{TLPP})<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1;\mathsf{ATR}_{0})
<Ξ²01𝖱π–₯𝖭(n+1;βˆ€k.Ξ”k0𝖫𝖯𝖯)<Ξ²01𝖱π–₯𝖭(n+2)\displaystyle<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1;\forall k.\Delta^{0}_{k}\operatorname{\mathsf{LPP}})<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+2)
Remark 4.32.

In a forthcoming paper [14], the first author introduced a characterization of 𝖠𝖫𝖯𝖯,𝖳𝖫𝖯𝖯,Ξ”k0​𝖫𝖯𝖯\mathsf{ALPP},\mathsf{TLPP},\Delta^{0}_{k}\operatorname{\mathsf{LPP}} and Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) in terms of the Ο‰\omega-model reflection of transfinite inductions as follows. Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, we have

  • β€’

    𝖠𝖫𝖯𝖯↔𝖱π–₯𝖭⁑(Π∞1βˆ’π–³π–¨)\mathsf{ALPP}\leftrightarrow\operatorname{\mathsf{RFN}}(\Pi^{1}_{\infty}\mathchar 45\relax\operatorname{\mathsf{TI}})666This equivalence is independently proved by Freund [4].,

  • β€’

    𝖳𝖫𝖯𝖯↔β01​𝖱π–₯𝖭⁑(1;Ξ 11βˆ’π–³π–¨)\mathsf{TLPP}\leftrightarrow\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\Pi^{1}_{1}\mathchar 45\relax\operatorname{\mathsf{TI}}),

  • β€’

    Ξ”k0​𝖫𝖯𝖯↔𝖱π–₯𝖭⁑(Ξ k+11βˆ’π–³π–¨)\Delta^{0}_{k}\operatorname{\mathsf{LPP}}\leftrightarrow\operatorname{\mathsf{RFN}}(\Pi^{1}_{k+1}\mathchar 45\relax\operatorname{\mathsf{TI}}) for k>1k>1,

  • β€’

    Ξ²01​𝖱π–₯𝖭⁑(n+1)↔β01​𝖱π–₯𝖭⁑(n;Π∞1βˆ’π–³π–¨)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n+1)\leftrightarrow\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n;\Pi^{1}_{\infty}\mathchar 45\relax\operatorname{\mathsf{TI}}).

Here, 𝖱π–₯𝖭⁑(T)\operatorname{\mathsf{RFN}}(T) denotes the Ο‰\omega-model reflection of a theory TT. These characterization may be helpful to understand the structure of the hierarchy in Corollary 4.31.

Remark 4.33.

We note that the gap between each pair of the hierarchy in Corollary 4.31 is relatively large. The proofs for theses inequalities actually show that the stronger one implies not only the omega-model reflection of the lower one, but also the omega-model reflection of [the lower one and transfinite induction for Ξ 11\Pi^{1}_{1} formulas (Ξ 11βˆ’π–³π–¨\Pi^{1}_{1}\mathchar 45\relax\operatorname{\mathsf{TI}})]. Since, for any Ξ 21\Pi^{1}_{2} sentence Οƒ\sigma, the omega-model reflection of [Οƒ+Ξ 11βˆ’π–³π–¨]\sigma+\Pi^{1}_{1}\mathchar 45\relax\operatorname{\mathsf{TI}}] proves any time iteration of the omega-model reflection of Οƒ\sigma as in [15, Section 5], we conclude that any time iterations of the omega-model reflection of the lower one is weaker than the stronger one.

In the remaining of this section, we give another characterization of 𝖠𝖫𝖯𝖯\mathsf{ALPP}.

Definition 4.34.

Let β„³\mathcal{M} be a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}. We say β„³\mathcal{M} is an 𝖠​β\mathsf{A}\beta-models if for any Ξ 20\Pi^{0}_{2} formula θ​(X)\theta(X),

βˆƒX≀Taℳ​θ​(X)β†”β„³βŠ§βˆƒX​θ​(X).\exists X\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}\theta(X)\leftrightarrow\mathcal{M}\models\exists X\theta(X).
Lemma 4.35.

Let φ​(X)\varphi(X) be an arithmetical formula with exactly displayed free variable. Then, 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that for any 𝖠​β\mathsf{A}\beta-model β„³\mathcal{M},

βˆƒX≀Taℳ​θ​(X)β†”β„³βŠ§βˆƒX​θ​(X).\exists X\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}\theta(X)\leftrightarrow\mathcal{M}\models\exists X\theta(X).
Proof.

Let φ​(X)\varphi(X) be an arithmetical formula. Then there exists a number kk and a Ξ 20\Pi^{0}_{2} formula θ​(X,f)\theta(X,f) such that 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves

βˆ€X(Ο†(X)β†”βˆƒf≀TX(k)ΞΈ(X,f)).\displaystyle\forall X(\varphi(X)\leftrightarrow\exists f\leq_{\mathrm{T}}X^{(k)}\theta(X,f)).

We work in 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Assume β„³\mathcal{M} is an 𝖠​β\mathsf{A}\beta-model and βˆƒX≀Taℳ​φ​(X)\exists X\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}\varphi(X). Since Ο†\varphi is arithmetical, φ​(X)\varphi(X) actually holds. Thus, there exists f≀TX(k)f\leq_{\mathrm{T}}X^{(k)} such that θ​(X,f)\theta(X,f). Since β„³\mathcal{M} is a model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, β„³βŠ§βˆƒf≀TX(k)​θ​(X,f)\mathcal{M}\models\exists f\leq_{\mathrm{T}}X^{(k)}\theta(X,f). Hence β„³βŠ§Ο†β€‹(X)\mathcal{M}\models\varphi(X). ∎

Definition 4.36.

Define the 𝖠​β\mathsf{A}\beta-model reflection as the assertion that for any set XX there is an 𝖠​β\mathsf{A}\beta-model containing XX.

Theorem 4.37.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is equivalent to 𝖠​β\mathsf{A}\beta-model reflection.

Proof.

First, assume Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1). Let XX be a set. By Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1), take a coded Ο‰\omega-models β„³0,β„³1\mathcal{M}_{0},\mathcal{M}_{1} such that Xβˆˆβ„³0βˆˆΞ²β„³1βˆ§β„³1βŠ§π– π–’π– 0X\in\mathcal{M}_{0}\in_{\beta}\mathcal{M}_{1}\land\mathcal{M}_{1}\models\mathsf{ACA}_{0}.

We claim that β„³0\mathcal{M}_{0} is an 𝖠​β\mathsf{A}\beta-model. Let θ​(Y)\theta(Y) be an arithmetical formula. Assume βˆƒZ≀Taβ„³0​θ​(Z)\exists Z\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}_{0}\theta(Z). Then, there exists nβˆˆβ„•n\in\mathbb{N} such that βˆƒZ≀T(β„³0)(n)​θ​(Z)\exists Z\leq_{\mathrm{T}}(\mathcal{M}_{0})^{(n)}\theta(Z). Since β„³0βˆˆβ„³1\mathcal{M}_{0}\in\mathcal{M}_{1} and β„³1βŠ§π– π–’π– 0β€²\mathcal{M}_{1}\models\operatorname{\mathsf{ACA}}_{0}^{\prime}, β„³1βŠ§βˆƒZ​θ​(Z)\mathcal{M}_{1}\models\exists Z\theta(Z). Since β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³1\mathcal{M}_{1}, β„³0βŠ§βˆƒZ​θ​(Z)\mathcal{M}_{0}\models\exists Z\theta(Z).

We next assume 𝖠​β\mathsf{A}\beta-model reflection. Then, 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} holds. Let XX be a set. By 𝖠​β\mathsf{A}\beta-model reflection, take an 𝖠​β\mathsf{A}\beta-model β„³0\mathcal{M}_{0} such that Xβˆˆβ„³0X\in\mathcal{M}_{0}. By 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, take β„³1={Y:Y≀Taβ„³0\mathcal{M}_{1}=\{Y:Y\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}_{0}}.

Now Xβˆˆβ„³0βˆˆβ„³1βˆ§β„³1βŠ§π– π–’π– 0X\in\mathcal{M}_{0}\in\mathcal{M}_{1}\land\mathcal{M}_{1}\models\mathsf{ACA}_{0}. Thus, it is enough to show that β„³0\mathcal{M}_{0} is a Ξ²\beta-submodel of β„³1\mathcal{M}_{1}. Let θ​(Y)\theta(Y) be an arithmetical formula. Assume β„³1βŠ§βˆƒY​θ​(Y)\mathcal{M}_{1}\models\exists Y\theta(Y). Then, βˆƒY≀Taβ„³0​θ​(Y)\exists Y\leq_{\mathrm{T}}^{\mathrm{a}}\mathcal{M}_{0}\theta(Y) and hence β„³0βŠ§βˆƒY​θ​(Y)\mathcal{M}_{0}\models\exists Y\theta(Y). This completes the proof. ∎

Corollary 4.38.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}.) 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to 𝖠​β\mathsf{A}\beta-model reflection.

Proof.

Since 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) and Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is equivalent to 𝖠​β\mathsf{A}\beta-model reflection, 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to 𝖠​β\mathsf{A}\beta-model reflection. ∎

For each n>0n>0, define 𝖠​β​𝖱π–₯𝖭⁑(n)\mathsf{A}\beta\operatorname{\mathsf{RFN}}(n) as the assertion that

βˆ€Xβ€‹βˆƒβ„³0​⋯​ℳnβˆ’1​(XβˆˆΞ²β‹―βˆˆΞ²β„³nβˆ’1βˆ§β„³nβˆ’1​ is an 𝖠​β-model).\displaystyle\forall X\exists\mathcal{M}_{0}\cdots\mathcal{M}_{n-1}(X\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{n-1}\land\mathcal{M}_{n-1}\text{ is an $\mathsf{A}\beta$-model}).

Then, by the same argument as in Theorem 4.37, we have

Theorem 4.39.

Over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is equivalent to 𝖠​β​𝖱π–₯𝖭⁑(nβˆ’1)\mathsf{A}\beta\operatorname{\mathsf{RFN}}(n-1) for any n>0n>0.

5 The arithmetical relativization

As we have noted, the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy reveals that how many times the hyperjump operator is used to prove a Ξ 21\Pi^{1}_{2} sentence. In addition, we can classify Ξ 31\Pi^{1}_{3} sentences provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} by comparing their Ξ 21\Pi^{1}_{2} approximations and the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy. In this section, we summarize the idea of Ξ 21\Pi^{1}_{2} approximations of Ξ 31\Pi^{1}_{3} sentences.

Definition 5.1.

Let Οƒβ‰‘βˆ€X(ΞΈ(X)β†’βˆƒYβˆ€ZΞ·(X,Y,Z)\sigma\equiv\forall X(\theta(X)\to\exists Y\forall Z\eta(X,Y,Z) be a Ξ 31\Pi^{1}_{3} sentence. Define rel⁑(Οƒ)\operatorname{rel}(\sigma) by βˆ€X(ΞΈ(X)β†’βˆƒYβˆ€Z≀Ta(XβŠ•Y)Ξ·(X,Y,Z)\forall X(\theta(X)\to\exists Y\forall Z\leq_{\mathrm{T}}^{\mathrm{a}}(X\oplus Y)\eta(X,Y,Z). We call rel⁑(Οƒ)\operatorname{rel}(\sigma) the arithmetical relativization of Οƒ\sigma.

For example, 𝖠𝖫𝖯𝖯\mathsf{ALPP} and 𝖠​β​𝖱π–₯𝖭⁑(n)\mathsf{A}\beta\operatorname{\mathsf{RFN}}(n) are instances of rel⁑(Οƒ)\operatorname{rel}(\sigma). Indeed, if we take θ​(X)\theta(X) to be β€˜XX is an ill-founded tree’ and η​(X,Y,Z)\eta(X,Y,Z) to be (Y∈[X])∧(Z∈[X]β†’Y≀lX)(Y\in[X])\land(Z\in[X]\to Y\leq_{l}X), then 𝖠𝖫𝖯𝖯\mathsf{ALPP} is equivalent to rel⁑(βˆ€X​(θ​(X)β†’βˆƒYβ€‹βˆ€Z​η​(X,Y,Z)))\operatorname{rel}(\forall X(\theta(X)\to\exists Y\forall Z\eta(X,Y,Z))).

Remark 5.2.

In the case of 𝖠𝖫𝖯𝖯\mathsf{ALPP}, the following rel′⁑(Οƒ)\operatorname{rel}^{\prime}(\sigma) is equivalent to rel⁑(Οƒ)\operatorname{rel}(\sigma).

rel′⁑(Οƒ)β‰‘βˆ€X​(θ​(X)β†’βˆƒYβ€‹βˆ€Z≀TaY​η​(X,Y,Z)).\operatorname{rel}^{\prime}(\sigma)\equiv\forall X(\theta(X)\to\exists Y\forall Z\leq_{\mathrm{T}}^{\mathrm{a}}Y\eta(X,Y,Z)).

In general, rel⁑(Οƒ)\operatorname{rel}(\sigma) and rel′⁑(Οƒ)\operatorname{rel}^{\prime}(\sigma) are not equivalent. The equivalence of rel⁑(Οƒ)\operatorname{rel}(\sigma) and rel′⁑(Οƒ)\operatorname{rel}^{\prime}(\sigma) is related to the notion of cylinder in Weihrauch reduction.

We note that if Οƒ\sigma is a Ξ 31\Pi^{1}_{3} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, then rel⁑(Οƒ)\operatorname{rel}(\sigma) is a Ξ 21\Pi^{1}_{2} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. Thus, rel⁑(Οƒ)\operatorname{rel}(\sigma) is provable from 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) for some nn. Let us write nΟƒn_{\sigma} for the smallest nn such that 𝖠𝖒𝖠0+Ξ²01​𝖱π–₯𝖭⁑(n)⊒rel⁑(Οƒ)\mathsf{ACA}_{0}+\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n)\vdash\operatorname{rel}(\sigma) for such a Οƒ\sigma. Then, we can classify Ξ 31\Pi^{1}_{3} sentences provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} according to nΟƒn_{\sigma}.

We then give a lemma which is useful to evaluate nσn_{\sigma}.

Definition 5.3.

Let Ο†β‰‘βˆ€Xβ€‹βˆƒY​θ\varphi\equiv\forall X\exists Y\theta and Οˆβ‰‘βˆ€Xβ€‹βˆƒY​η\psi\equiv\forall X\exists Y\eta. For a theory TT, we say TT proves the lightface implication of Ο†\varphi to ψ\psi if

TβŠ’βˆ€X​(βˆƒYβ€‹ΞΈβ†’βˆƒZ​η).T\vdash\forall X(\exists Y\theta\to\exists Z\eta).
Lemma 5.4.

Let Ο†β‰‘βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z)\varphi\equiv\forall X\exists Y\forall Z\theta(X,Y,Z) and Οˆβ‰‘βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​η​(X,Y,Z)\psi\equiv\forall X\exists Y\forall Z\eta(X,Y,Z) be Ξ 31\Pi^{1}_{3} sentences such that 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves the lightface implication of Ο†\varphi to ψ\psi. Then, 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} proves rel⁑(Ο†)β†’rel⁑(ψ)\operatorname{rel}(\varphi)\to\operatorname{rel}(\psi).

Proof.

We reason in 𝖠𝖒𝖠0++βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z≀TaXβŠ•Y​θ​(X,Y,Z)\operatorname{\mathsf{ACA}}_{0}^{+}+\forall X\exists Y\forall Z\leq_{\mathrm{T}}^{\mathrm{a}}X\oplus Y\theta(X,Y,Z). Let X,YX,Y be such that βˆ€Z≀TaXβŠ•Y​θ​(X,Y,Z)\forall Z\leq_{\mathrm{T}}^{\mathrm{a}}X\oplus Y\theta(X,Y,Z). Take the smallest Ο‰\omega-model β„³\mathcal{M} of 𝖠𝖒𝖠0\operatorname{\mathsf{ACA}}_{0} containing XβŠ•YX\oplus Y. Then, β„³βŠ§βˆƒUβ€‹βˆ€Z​θ​(X,U,Z)\mathcal{M}\models\exists U\forall Z\theta(X,U,Z) via U=YU=Y. Since β„³\mathcal{M} is a model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, β„³βŠ§βˆƒVβ€‹βˆ€Z​η​(X,V,Z)\mathcal{M}\models\exists V\forall Z\eta(X,V,Z). Take such a VV. Thus, βˆƒWβ€‹βˆ€Z≀TaXβŠ•W​η​(X,W,Z)\exists W\forall Z\leq_{\mathrm{T}}^{\mathrm{a}}X\oplus W\eta(X,W,Z) via W=VW=V. ∎

For example, if 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves βˆ€X​(βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z)β†’βˆƒW​(W=HJ⁑(X)))\forall X(\exists Y\forall Z\theta(X,Y,Z)\to\exists W(W=\operatorname{HJ}(X))), then 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} proves rel⁑(βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ)β†’Ξ²01​𝖱π–₯𝖭⁑(1)\operatorname{rel}(\forall X\exists Y\forall Z\theta)\to\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1). Conversely, if 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves βˆ€X​(βˆƒW​(W=HJ⁑(X))β†’βˆƒYβ€‹βˆ€Z​θ​(X,Y,Z))\forall X(\exists W(W=\operatorname{HJ}(X))\to\exists Y\forall Z\theta(X,Y,Z)), then 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} (and hence even 𝖱𝖒𝖠0\mathsf{RCA}_{0}) proves Ξ²01​𝖱π–₯𝖭⁑(1)β†’rel⁑(βˆ€Xβ€‹βˆƒYβ€‹βˆ€Z​θ)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1)\to\operatorname{rel}(\forall X\exists Y\forall Z\theta).

6 Pseudo Ramsey’s theorem and β​𝖱π–₯𝖭\beta\operatorname{\mathsf{RFN}}

In this section, we introduce Ξ 21\Pi^{1}_{2} variants of Ramsey’s theorem for [β„•]β„•[\mathbb{N}]^{\mathbb{N}}. Our variants may be seen as an instance of the arithmetical relativization of usual Ramsey’s theorem. In this section, we see the relationship between them and the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy.

Let [X]β„•[X]^{\mathbb{N}} denotes the class of all infinite subsets of XX. Then, Ramsey’s theorem for [β„•]β„•[\mathbb{N}]^{\mathbb{N}}, also known as Galvin-Prikry’s theorem, claims that certain subclasses of [β„•]β„•[\mathbb{N}]^{\mathbb{N}} has Ramsey’s property in the sense that a class π’œβŠ†[β„•]β„•\mathcal{A}\subseteq[\mathbb{N}]^{\mathbb{N}} has Ramsey’s property if there is an infinite set HβŠ†β„•H\subseteq\mathbb{N} such that [H]β„•[H]^{\mathbb{N}} is a subset of π’œ\mathcal{A} or its complement.

We first recall how to formalize Ramsey’s theorem in second-order arithmetic and some existing results. For the details, see also [13].

Definition 6.1.

Let [β„•]β„•[\mathbb{N}]^{\mathbb{N}} denote the class of all strictly increasing sequence of natural numbers. Let φ​(f,X)\varphi(f,X) be a formula for f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} and XβŠ†β„•X\subseteq\mathbb{N}. We say Ο†\varphi has Ramsey’s property if

βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​RPφ​(h,X),\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}\,\mathrm{RP}_{\varphi}(h,X),

where RPφ​(h,X)\mathrm{RP}_{\varphi}(h,X) is

βˆ€g∈[β„•]ℕ​φ​(h∘g,X)βˆ¨βˆ€g∈[β„•]ℕ​¬φ​(h∘g,X).\forall g\in[\mathbb{N}]^{\mathbb{N}}\varphi(h\circ g,X)\lor\forall g\in[\mathbb{N}]^{\mathbb{N}}\lnot\varphi(h\circ g,X).

In this definition, we identify f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} and its range, φ​(f,X)\varphi(f,X) and the class {f∈[β„•]β„•:φ​(f,X)}\{f\in[\mathbb{N}]^{\mathbb{N}}:\varphi(f,X)\}. In particular, h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}} corresponds to an infinite subset HH of β„•\mathbb{N}, h∘gh\circ g corresponds to an infinite subset of HH.

Definition 6.2.

Let Ξ“\Gamma be a class of formulas. Then, Ξ“\Gamma-Ramsey’s theorem (write Ξ“βˆ’π–±π–Ίπ—†)\Gamma\mathchar 45\relax\mathsf{Ram}) is the assertion that any Ο†βˆˆΞ“\varphi\in\Gamma has Ramsey’s property.

Theorem 6.3.

[13, VI.6.4] Over 𝖱𝖒𝖠0\mathsf{RCA}_{0}, Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is equivalent to Ξ£01βˆ’π–±π–Ίπ—†\Sigma^{1}_{0}\mathchar 45\relax\mathsf{Ram}.

We give a refined analysis for this theorem by using the Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}-hierarchy. We start with giving Ξ 21\Pi^{1}_{2} variants of Ramsey’s property which we call pseudo-Ramsey’s property.

Definition 6.4.

Let φ​(f,X)\varphi(f,X) be a formula for f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} with exactly displayed free variables. We say φ​(f,X)\varphi(f,X) has pseudo-Ramsey’s property if

βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​(pRPφ​(h,X))\displaystyle\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}(\mathrm{pRP}_{\varphi}(h,X))

where pRP\mathrm{pRP} is

βˆ€g∈[β„•]ℕ​(g≀TafβŠ•X→φ​(f∘g))\displaystyle\forall g\in[\mathbb{N}]^{\mathbb{N}}(g\leq^{\mathrm{a}}_{\mathrm{T}}f\oplus X\to\varphi(f\circ g))
∨\displaystyle\lor\,\, βˆ€g∈[β„•]ℕ​(g≀TafβŠ•X→¬φ​(f∘g)).\displaystyle\forall g\in[\mathbb{N}]^{\mathbb{N}}(g\leq_{\mathrm{T}}^{\mathrm{a}}f\oplus X\to\lnot\varphi(f\circ g)).

The condition Ο†\varphi has pseudo-Ramsey’s property is the same as rel⁑(βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​RPΟ†)\operatorname{rel}(\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}\mathrm{RP}_{\varphi}) holds. Therefore, we write rel⁑(Ξ“βˆ’π–±π–Ίπ—†)\operatorname{rel}(\Gamma\mathchar 45\relax\mathsf{Ram}) for the set of sentences {βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​(pRPφ​(h,X)):Ο†βˆˆΞ“}\{\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}(\mathrm{pRP}_{\varphi}(h,X)):\varphi\in\Gamma\}.

We note that for each n>0n>0, there exists a formula Ο†βˆˆΞ£n0\varphi\in\Sigma^{0}_{n} such that Ξ£n0βˆ’π–±π–Ίπ—†\Sigma^{0}_{n}\mathchar 45\relax\mathsf{Ram} is axiomatizable by βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​RPφ​(h,X)\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}\,\mathrm{RP}_{\varphi}(h,X) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, and hence rel⁑(Ξ£n0βˆ’π–±π–Ίπ—†)\operatorname{rel}(\Sigma^{0}_{n}\mathchar 45\relax\mathsf{Ram}) is axiomatizable by rel⁑(βˆ€Xβ€‹βˆƒh∈[β„•]ℕ​RPφ​(h,X))\operatorname{rel}(\forall X\exists h\in[\mathbb{N}]^{\mathbb{N}}\,\mathrm{RP}_{\varphi}(h,X)) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. Since Ξ£n0βˆ’π–±π–Ίπ—†\Sigma^{0}_{n}\mathchar 45\relax\mathsf{Ram} is provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, rel⁑(Ξ£n0βˆ’π–±π–Ίπ—†)\operatorname{rel}(\Sigma^{0}_{n}\mathchar 45\relax\mathsf{Ram}) is provable from Ξ²01​𝖱π–₯𝖭⁑(p​(n))\beta^{1}_{0}\operatorname{\mathsf{RFN}}(p(n)) for some p​(n)p(n). We show that p​(n)=np(n)=n. It follows from the following lemma.

Lemma 6.5.

([13], VI.6.2.) (𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let β„³0,…,β„³kβˆ’1\mathcal{M}_{0},\ldots,\mathcal{M}_{k-1} be coded Ξ²\beta models such that β„³0βˆˆβ‹―βˆˆβ„³kβˆ’1\mathcal{M}_{0}\in\cdots\in\mathcal{M}_{k-1}. Then, for any Ξ£k0\Sigma^{0}_{k} formula φ​(f)\varphi(f) with parameters from β„³0\mathcal{M}_{0}, φ​(f)\varphi(f) has Ramsey’s property via some hβˆˆβ„³kβˆ’1h\in\mathcal{M}_{k-1}.

Theorem 6.6.

rel⁑(Ξ£k0​𝖱𝖺𝗆)\operatorname{rel}(\Sigma^{0}_{k}\mathsf{Ram}) is provable from Ξ²01​𝖱π–₯𝖭⁑(k)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k).

Proof.

Assume Ξ²01​𝖱π–₯𝖭⁑(k)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k). Let φ​(f,X)\varphi(f,X) be a Ξ£k0\Sigma^{0}_{k} formula for f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} with exactly displayed set variables. By Ξ²01​𝖱π–₯𝖭⁑(k)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k), take coded Ο‰\omega-models β„³0,…,β„³k\mathcal{M}_{0},\ldots,\mathcal{M}_{k} such that Xβˆˆβ„³0βˆˆΞ²β‹―βˆˆΞ²β„³kβˆ’1βˆˆΞ²β„³kX\in\mathcal{M}_{0}\in_{\beta}\cdots\in_{\beta}\mathcal{M}_{k-1}\in_{\beta}\mathcal{M}_{k} and β„³kβŠ§π– π–’π– 0\mathcal{M}_{k}\models\mathsf{ACA}_{0}. Now, by the previous lemma, there exists fβˆˆβ„³kβˆ’1f\in\mathcal{M}_{k-1} such that

β„³kβŠ§βˆ€g​φ​(f∘g)βˆ¨βˆ€g​¬φ​(f∘g).\mathcal{M}_{k}\models\forall g\varphi(f\circ g)\lor\forall g\lnot\varphi(f\circ g).

Since β„³k\mathcal{M}_{k} is a model of 𝖠𝖒𝖠0\mathsf{ACA}_{0}, Ο†\varphi has pseudo-Ramsey’s property via ff. ∎

In the remaining of this section, we show that 𝖠𝖒𝖠0++{rel⁑(Ξ£n0​𝖱𝖺𝗆):nβˆˆΟ‰}\operatorname{\mathsf{ACA}}_{0}^{+}+\{\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}):n\in\omega\} proves Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) for all nn. In [16], it is shown that over 𝖠𝖒𝖠0\mathsf{ACA}_{0}, (lightface-​Δ20,X)β€‹π–±π–Ίπ—†β†’βˆƒY​(Y=HJ⁑(X))(\text{lightface-}\Delta^{0,X}_{2})\mathsf{Ram}\to\exists Y(Y=\operatorname{HJ}(X)). Here lightface-​Δ20,X\text{lightface-}\Delta^{0,X}_{2} is the set of Ξ”20\Delta^{0}_{2} formulas having no set parameters other than XX. Moreover, it is commented without proofs that this result can be extended to the transfinite level: 𝖠𝖳𝖱0⊒(lightface-Ξ”11𝖱𝖺𝗆)β†’βˆ€Ξ±: recursive well orderΒ βˆƒY(Y=HJΞ±(βˆ…)))\mathsf{ATR}_{0}\vdash(\text{lightface-}\Delta^{1}_{1}\mathsf{Ram})\to\forall\alpha\text{: recursive well order }\exists Y(Y=\operatorname{HJ}^{\alpha}(\varnothing))). In this section, we give an explicit proof for finite level.

Definition 6.7.

Let TT be a tree. We say f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} majorizes TT if βˆƒg∈[T]β€‹βˆ€n​(g​(n)≀f​(n))\exists g\in[T]\forall n(g(n)\leq f(n)).

Definition 6.8.

Let ff be a function and nβˆˆβ„•n\in\mathbb{N}. Define f+nf_{+n} by f+n​(x)=f​(x+n)f_{+n}(x)=f(x+n).

Lemma 6.9.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let TT be a tree and f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}}. Then, ff majorizes TT if and only if βˆ€n>0β€‹βˆƒΟ„βˆˆT​(|Ο„|=nβˆ§βˆ€m<n​(τ​(m)≀f​(m)))\forall n>0\exists\tau\in T(|\tau|=n\land\forall m<n(\tau(m)\leq f(m))). In particular, ff majorizes TT is an arithmetical condition.

Proof.

It is easy by KΓΆnig’s lemma. For the details, see the proof of Lemma V.9.6. in [13]. ∎

Definition 6.10.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) Let θ​(n,A)\theta(n,A) be the Ξ£11\Sigma^{1}_{1} formula defining HJ⁑(A)\operatorname{HJ}(A). Write θ​(n,A)β‰‘βˆƒfβˆˆβ„•β„•β€‹βˆ€y​R​(n,f​[y],A​[y])\theta(n,A)\equiv\exists f\in\mathbb{N}^{\mathbb{N}}\forall yR(n,f[y],A[y]) by a Ξ£00\Sigma^{0}_{0} formula RR.

For each nβˆˆβ„•n\in\mathbb{N} and AβŠ†β„•A\subseteq\mathbb{N}, define a tree T​(n,A)={Οƒ:βˆ€y<|Οƒ|​R​(n,σ​[y],A​[y])}T(n,A)=\{\sigma:\forall y<|\sigma|R(n,\sigma[y],A[y])\}. For each f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} and AβŠ†β„•A\subseteq\mathbb{N}, define F(f,A)={n:βˆƒp(f+pΒ majorizesΒ T(n,A))F(f,A)=\{n:\exists p(f_{+p}\text{ majorizes }T(n,A)).

Definition 6.11.

Define Ξ¦βˆ—β€‹(f,A)\Phi_{\ast}(f,A) by

βˆ€n<f​(0)​(f+2​ majorizes ​T​(n,A)β†’f+1​ majorizes ​T​(n,A)).\forall n<f(0)(f_{+2}\text{ majorizes }T(n,A)\to f_{+1}\text{ majorizes }T(n,A)).
Lemma 6.12.

(Essentially [13] VI.6.1) 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that for any AβŠ†β„•A\subseteq\mathbb{N} and h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}}, if βˆ€g​(Ξ¦βˆ—β€‹(h∘g,A))\forall g(\Phi_{\ast}(h\circ g,A)), then F​(h,A)=HJ⁑(A)F(h,A)=\operatorname{HJ}(A).

Proof.

We first remark that for each nn and AA, T​(n,A)T(n,A) has a path if and only if n∈HJ⁑(A)n\in\operatorname{HJ}(A).

Take h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}} such that βˆ€g​(Ξ¦βˆ—β€‹(h∘g,A))\forall g(\Phi_{\ast}(h\circ g,A)). We show that F​(h,A)βŠ†HJ⁑(A)F(h,A)\subseteq\operatorname{HJ}(A). If n∈F​(h,A)n\in F(h,A) then f+pf_{+p} majorizes T​(n,A)T(n,A) for some pp, and hence [T​(n,A)]β‰ βˆ…[T(n,A)]\neq\varnothing. Thus, n∈HJ⁑(A)n\in\operatorname{HJ}(A).

Conversely, assume n∈HJ⁑(A)n\in\operatorname{HJ}(A) and show that n∈F​(h,A)n\in F(h,A). Now, T​(n,A)T(n,A) has a path. We claim that h+(n+2)h_{+(n+2)} majorizes T​(n,A)T(n,A). Let g0∈[T​(n,A)]g_{0}\in[T(n,A)] and define g​(n)=g0​(0)+β‹―+g0​(n)g(n)=g_{0}(0)+\cdots+g_{0}(n). Then g∈[β„•]β„•g\in[\mathbb{N}]^{\mathbb{N}} and it majorizes T​(n,A)T(n,A). Assume h+(n+2)h_{+(n+2)} does not majorize T​(n,A)T(n,A). Take k>0k>0 such that βˆ€Ο„βˆˆT(n,A)(|Ο„|=kβ†’βˆƒm<k(Ο„(m)>h+(n+2)(m))\forall\tau\in T(n,A)(|\tau|=k\to\exists m<k(\tau(m)>h_{+(n+2)}(m)) by the previous lemma. Define ff by

f​(x)={h+(n+1)​(x)​ if ​x<k+1h+(n+k+2)​(g​(x))​ otherwise.\displaystyle f(x)=\begin{cases}h_{+(n+1)}(x)\text{ if }x<k+1\\ h_{+(n+k+2)}(g(x))\text{ otherwise.}\end{cases}
Claim 6.12.1.

Now we have

  1. 1.

    f+(k+1)f_{+(k+1)} majorizes T​(n,A)T(n,A) but

  2. 2.

    f+1f_{+1} does not majorize T​(n,A)T(n,A).

Proof of Claim.

(1.) For any xβˆˆβ„•x\in\mathbb{N}, f+(k+1)​(x)=f​(x+k+1)=h+(n+1)​(g​(x))f_{+(k+1)}(x)=f(x+k+1)=h_{+(n+1)}(g(x)). Since h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}}, we have h+(n+1)​(g​(x))β‰₯g​(x)β‰₯g0​(x)h_{+(n+1)}(g(x))\geq g(x)\geq g_{0}(x). Hence g0​(x)g_{0}(x) ensures that f+(k+1)f_{+(k+1)} majorizes T​(n,A)T(n,A).

(2.) For a contradiction, assume f+1f_{+1} majorizes T​(n,A)T(n,A). Then there exists Ο„k∈T​(n,A)\tau_{k}\in T(n,A) such that

|Ο„k|=kβˆ§βˆ€m<k​(Ο„k​(m)≀f+1​(m)).|\tau_{k}|=k\land\forall m<k(\tau_{k}(m)\leq f_{+1}(m)).

Now, for any m<km<k,

f+1​(m)\displaystyle f_{+1}(m) =f​(m+1)\displaystyle=f(m+1)
=h+(n+1)​(m+1)\displaystyle=h_{+(n+1)}(m+1)
=h​(m+n+2)=h+(n+2)​(m).\displaystyle=h(m+n+2)=h_{+(n+2)}(m).

Therefore, βˆ€m<k​(Ο„k​(m)≀h+(n+2)​(m))\forall m<k(\tau_{k}(m)\leq h_{+(n+2)}(m)). However, this contradicts the definition of kk. ∎

For each k′≀kk^{\prime}\leq k, define gkβ€²βˆˆ[β„•]β„•g_{k^{\prime}}\in[\mathbb{N}]^{\mathbb{N}} such that

gk′​(0)\displaystyle g_{k^{\prime}}(0) =0,\displaystyle=0,
gk′​(1)\displaystyle g_{k^{\prime}}(1) =1,\displaystyle=1,
gk′​(x+2)\displaystyle g_{k^{\prime}}(x+2) ={x+kβ€²+n+2​ if ​x+kβ€²+1<k+1g​(x+kβ€²+1)+(n+k+2)​ otherwise.\displaystyle=\begin{cases}x+k^{\prime}+n+2\text{ if }x+k^{\prime}+1<k+1\\ g(x+k^{\prime}+1)+(n+k+2)\text{ otherwise}.\end{cases}

Then, for any k′≀kk^{\prime}\leq k, gkβ€²βˆˆ[β„•]β„•g_{k^{\prime}}\in[\mathbb{N}]^{\mathbb{N}} and f+(kβ€²+1)=(h∘gkβ€²)+2f_{+(k^{\prime}+1)}=(h\circ g_{k^{\prime}})_{+2}. Indeed, for any xx,

f+(kβ€²+1)​(x)\displaystyle f_{+(k^{\prime}+1)}(x) =f​(x+kβ€²+1)\displaystyle=f(x+k^{\prime}+1)
={h​(x+kβ€²+n+2)​ if ​x+kβ€²+1<k+1h​(g​(x+kβ€²+1)+(n+k+2))​ otherwise\displaystyle=\begin{cases}h(x+k^{\prime}+n+2)\text{ if }x+k^{\prime}+1<k+1\\ h(g(x+k^{\prime}+1)+(n+k+2))\text{ otherwise}\end{cases}

and

(h∘gkβ€²)+2​(x)\displaystyle(h\circ g_{k^{\prime}})_{+2}(x) =h​(gk′​(x+2))\displaystyle=h(g_{k^{\prime}}(x+2))
={h​(x+kβ€²+n+2)​ if ​x+kβ€²+1<k+1h​(g​(x+kβ€²+1)+(n+k+2))​ otherwise.\displaystyle=\begin{cases}h(x+k^{\prime}+n+2)\text{ if }x+k^{\prime}+1<k+1\\ h(g(x+k^{\prime}+1)+(n+k+2))\text{ otherwise}.\end{cases}

Now we have

  1. 1.

    f+(k+1)f_{+(k+1)} majorizes T​(n,A)T(n,A),

  2. 2.

    for any k′≀kk^{\prime}\leq k, if f+(kβ€²+1)f_{+(k^{\prime}+1)} majorizes T​(n,A)T(n,A) then f+kβ€²f_{+k^{\prime}} majorizes T​(n,A)T(n,A).

The second condition is proved as follows. We see n<n+1≀h​(n+1)=f​(0)<fkβ€²+1​(0)=(h∘gkβ€²)+2​(0)n<n+1\leq h(n+1)=f(0)<f_{k^{\prime}+1}(0)=(h\circ g_{k^{\prime}})_{+2}(0). Therefore, by the assumption on hh, if fkβ€²+1=(h∘gkβ€²)+2f_{k^{\prime}+1}=(h\circ g_{k^{\prime}})_{+2} majorizes T​(n,A)T(n,A), then (h∘gkβ€²)+1=fkβ€²(h\circ g_{k^{\prime}})_{+1}=f_{k^{\prime}} also majorizes T​(n,A)T(n,A).

By these conditions and arithmetical induction, we have f+1f_{+1} majorizes T​(n,A)T(n,A). However, this is a contradiction. ∎

Definition 6.13.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) For f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}} and AA, define Fn​(f,A)F^{n}(f,A) by

F0​(f,A)\displaystyle F^{0}(f,A) =A,\displaystyle=A,
Fn+1​(f,A)\displaystyle F^{n+1}(f,A) =F​(f,Fn​(f,A)).\displaystyle=F(f,F^{n}(f,A)).
Definition 6.14.

Define an arithmetical formula Ξ¦n​(f,A)\Phi_{n}(f,A) by

Ξ¦n​(f,A)≑⋀i<nΞ¦βˆ—β€‹(f,Fi​(f,A)).\displaystyle\Phi_{n}(f,A)\equiv\bigwedge_{i<n}\Phi_{\ast}(f,F^{i}(f,A)).
Lemma 6.15.

Let kβˆˆΟ‰k\in\omega. Then 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that for any h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}} and AA, if RPΞ¦k​(h,A)\mathrm{RP}_{\Phi_{k}}(h,A) holds, then then (βˆ€g​Φk​(h∘g,A))(\forall g\Phi_{k}(h\circ g,A)) holds.

Proof.

Take hh and AA as above. For the sake of contradiction, assume (βˆ€g​¬Φk​(h∘g,A))(\forall g\lnot\Phi_{k}(h\circ g,A)).

For each nn define gng_{n} by gn​(0)=0,gn​(x+1)=n+x+1g_{n}(0)=0,g_{n}(x+1)=n+x+1. Then, since ¬Φ​(h∘g,A)\lnot\Phi(h\circ g,A) holds, there exists m<h​(g​(0))=h​(0)m<h(g(0))=h(0) such that

⋁i<k(h∘gn)+2​ majorizes ​T​(m,Fi​((h∘gn)+2,A))​, but ​(h∘gn)+1​ does not.\displaystyle\bigvee_{i<k}(h\circ g_{n})_{+2}\text{ majorizes }T(m,F^{i}((h\circ g_{n})_{+2},A))\text{, but }(h\circ g_{n})_{+1}\text{ does not. }

We note that (h∘gn)+1​(x)=(h∘gn)​(x+1)=h​(gn​(x+1))=h​(n+x+1)=h+(n+1)​(x)(h\circ g_{n})_{+1}(x)=(h\circ g_{n})(x+1)=h(g_{n}(x+1))=h(n+x+1)=h_{+(n+1)}(x) and (h∘gn)+2​(x)=h+(n+2)​(x)(h\circ g_{n})_{+2}(x)=h_{+(n+2)}(x). Hence, for each nn, there exists m<h​(0)m<h(0) such that

⋁i<kh+(n+2)​ majorizes ​T​(m,Fi​(h+(n+2),A))​ but ​h+(n+1)​ does not.\displaystyle\bigvee_{i<k}h_{+(n+2)}\text{ majorizes }T(m,F^{i}(h_{+(n+2)},A))\text{ but }h_{+(n+1)}\text{ does not. }
Claim 6.15.1.

For each i<ki<k and nn, Fi​(h+(n+2),A)=Fi​(h,A)F^{i}(h_{+(n+2)},A)=F^{i}(h,A).

Proof of Claim.

We will show by (meta-)induction on ii.

The base step is immediate. By definition, F0​(h+(n+2),A)=A=F0​(h,A)F^{0}(h_{+(n+2)},A)=A=F^{0}(h,A).

For the induction step, assume Fi​(h+(n+2),A)=Fi​(h,A)F^{i}(h_{+(n+2)},A)=F^{i}(h,A). Then

Fi+1​(h+(n+2),A)\displaystyle F^{i+1}(h_{+(n+2)},A) =F​(h+(n+2),Fi​(h+(n+2),A))\displaystyle=F(h_{+(n+2)},F^{i}(h_{+(n+2)},A))
=F​(h+(n+2),Fi​(h,A))\displaystyle=F(h_{+(n+2)},F^{i}(h,A))
={x:βˆƒp​((h+(n+2))+p​ majorizes ​T​(x,Fi​(h,A)))}\displaystyle=\{x:\exists p((h_{+(n+2)})_{+p}\text{ majorizes }T(x,F^{i}(h,A)))\}
={x:βˆƒp​(h+p​ majorizes ​T​(x,Fi​(h,A)))}=F​(h,Fi​(h,A)).\displaystyle=\{x:\exists p(h_{+p}\text{ majorizes }T(x,F^{i}(h,A)))\}=F(h,F^{i}(h,A)).

This completes the proof. ∎

Now, for each nn, there exists m<h​(0)m<h(0) such that

⋁i<kh+(n+2)​ majorizes ​T​(m,Fi​(h,A))​ but ​h+(n+1)​ does not.\displaystyle\bigvee_{i<k}h_{+(n+2)}\text{ majorizes }T(m,F^{i}(h,A))\text{ but }h_{+(n+1)}\text{ does not. }

Put θ​(i,n)≑h+(n+2)​ majorizes ​T​(m,Fi​(h,A))​ but ​h+(n+1)​ does notΒ \theta(i,n)\equiv h_{+(n+2)}\text{ majorizes }T(m,F^{i}(h,A))\text{ but }h_{+(n+1)}\text{ does not }.

For each nn, define m​(n)m(n) be the smallest m<h​(0)m<h(0) satisfying the above condition.

Claim 6.15.2.

For each mm, there are at most kk-many nn such that m​(n)=mm(n)=m.

Proof of Claim.

Fix an mm. For the sake of contradiction, let n0,…,nkn_{0},\ldots,n_{k} be such that n0<β‹―<nkn_{0}<\cdots<n_{k} and m​(nj)=mm(n_{j})=m for all j<kj<k. Then, for each j<kj<k, there exists an ij<ki_{j}<k such that θ​(ij,nj)\theta(i_{j},n_{j}) holds.

We claim that if jβ‰ jβ€²j\neq j^{\prime} then ijβ‰ ijβ€²i_{j}\neq i_{j^{\prime}}. Let j<j′≀kj<j^{\prime}\leq k. If ij=ijβ€²i_{j}=i_{j^{\prime}}, then

h+(nj+2)​ majorizes ​T​(m,Fij​(h,A))​ but ​h+(nj+1)​ does not ∧\displaystyle h_{+(n_{j}+2)}\text{ majorizes }T(m,F^{i_{j}}(h,A))\text{ but }h_{+(n_{j}+1)}\text{ does not }\ \land
h+(njβ€²+2)​ majorizes ​T​(m,Fij​(h,A))​ but ​h+(njβ€²+1)​ does notΒ .\displaystyle h_{+(n_{j^{\prime}}+2)}\text{ majorizes }T(m,F^{i_{j}}(h,A))\text{ but }h_{+(n_{j^{\prime}}+1)}\text{ does not }.

However, since nj<njβ€²n_{j}<n_{j^{\prime}} and h+(nj+2)h_{+(n_{j}+2)} majorizes T​(m,Fij​(h,A))T(m,F^{i_{j}}(h,A)), h+(njβ€²+1)h_{+(n_{j^{\prime}}+1)} should majorize T​(m,Fij​(h,A))T(m,F^{i_{j}}(h,A)). This is a contradiction.

Now {ij:j≀k}\{i_{j}:j\leq k\} should be a k+1k+1-element subset of {0,…,kβˆ’1}\{0,\ldots,k-1\}. This is a contradiction. ∎

Now, m​(n)m(n) is a total function such that {m​(n):nβˆˆβ„•}\{m(n):n\in\mathbb{N}\} is bounded by h​(0)h(0). However, for any y∈{m​(n):nβˆˆβ„•}y\in\{m(n):n\in\mathbb{N}\}, there are at most kk-many nn such that m​(n)=ym(n)=y. This is a contradiction. ∎

Lemma 6.16.

(𝖠𝖒𝖠0\mathsf{ACA}_{0}) For any h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}} and AA, if βˆ€g​Φn​(h∘g,A)\forall g\Phi_{n}(h\circ g,A) then Fi​(h,A)=HJi⁑(A)F^{i}(h,A)=\operatorname{HJ}^{i}(A) for any i≀ni\leq n.

Proof.

Take hh and AA as above. Then, for the formula Ξ¦βˆ—\Phi_{\ast} in Definition 6.11, βˆ€g​(Ξ¦βˆ—β€‹(h∘g,A))\forall g(\Phi_{\ast}(h\circ g,A)) holds. Hence F​(h,A)=HJ⁑(A)F(h,A)=\operatorname{HJ}(A). Thus, βˆ€g​(Ξ¦βˆ—β€‹(h∘g,HJ⁑(A)))\forall g(\Phi_{\ast}(h\circ g,\operatorname{HJ}(A))) also holds, and hence F​(h,F​(h,A))=F​(h,HJ⁑(A))=HJ2⁑(A)F(h,F(h,A))=F(h,\operatorname{HJ}(A))=\operatorname{HJ}^{2}(A). Iterating this argument, we have Fi​(h,A)=HJi⁑(A)F^{i}(h,A)=\operatorname{HJ}^{i}(A) for any i≀ni\leq n. ∎

Theorem 6.17.

Over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}^{+}_{0}, {rel⁑(Ξ£n0​𝖱𝖺𝗆):nβˆˆΟ‰}\{\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}):n\in\omega\} and {Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} prove the same sentences. Therefore, any Ξ 21\Pi^{1}_{2} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is provable from 𝖠𝖒𝖠0++rel⁑(Ξ£n0​𝖱𝖺𝗆)\operatorname{\mathsf{ACA}}^{+}_{0}+\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}) for some nβˆˆΟ‰n\in\omega.

Proof.

As we have proved in Theorem 6.6, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) implies rel⁑(Ξ£n0​𝖱𝖺𝗆)\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}) for each nn.

We show that Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) is provable from {rel⁑(Ξ£n0​𝖱𝖺𝗆):nβˆˆΟ‰}\{\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}):n\in\omega\}. Assume 𝖠𝖒𝖠0++{rel⁑Σn0​𝖱𝖺𝗆:nβˆˆΟ‰}\operatorname{\mathsf{ACA}}^{+}_{0}+\{\operatorname{rel}\Sigma^{0}_{n}\mathsf{Ram}:n\in\omega\}. Take an arbitrary nβˆˆΟ‰n\in\omega. Let XX be a set. We will show that there exists a coded Ο‰\omega-model β„³\mathcal{M} such that Xβˆˆβ„³X\in\mathcal{M} and β„³βŠ§βˆƒY​(Y=HJn⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{n}(X)).

Now Ξ¦n​(f,X)\Phi_{n}(f,X) is an arithmetical formula for f∈[β„•]β„•f\in[\mathbb{N}]^{\mathbb{N}}. Therefore, there exists an h∈[β„•]β„•h\in[\mathbb{N}]^{\mathbb{N}} such that pRPΞ¦n​(h,X)\mathrm{pRP}_{\Phi_{n}}(h,X) holds. By 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, take β„³={Y:Y≀TahβŠ•X}\mathcal{M}=\{Y:Y\leq_{\mathrm{T}}^{\mathrm{a}}h\oplus X\}. Then β„³βŠ§π– π–’π– 0+(βˆ€g​Φn​(h∘g,X))∨(βˆ€g​¬Φn​(h∘g,X))\mathcal{M}\models\mathsf{ACA}_{0}+(\forall g\Phi_{n}(h\circ g,X))\lor(\forall g\lnot\Phi_{n}(h\circ g,X)). Thus, by Lemma 6.15, β„³βŠ§π– π–’π– 0+βˆ€g​Φn​(h∘g,X)\mathcal{M}\models\mathsf{ACA}_{0}+\forall g\Phi_{n}(h\circ g,X). Therefore, by Lemma 6.16, β„³βŠ§βˆƒY​(Y=HJn⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{n}(X)). ∎

In this section, we have proved that

  • β€’

    for any nβˆˆΟ‰n\in\omega, Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) implies rel⁑(Ξ£n0​𝖱𝖺𝗆)\operatorname{rel}(\Sigma^{0}_{n}\mathsf{Ram}), and

  • β€’

    for any nβˆˆΟ‰n\in\omega, there is some mβˆˆΟ‰m\in\omega such that rel⁑(Ξ£m0​𝖱𝖺𝗆)\operatorname{rel}(\Sigma^{0}_{m}\mathsf{Ram}) implies Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n)

over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}. Therefore, it is natural to ask

Question 6.18.

Is Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) provable from rel⁑(Ξ£n+10​𝖱𝖺𝗆)\operatorname{rel}(\Sigma^{0}_{n+1}\mathsf{Ram}) over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+} ?

A precise analysis of relationship between hyperjumps and Ramsey’s theorem is studied in [5] in the context of Weihrauch degrees. It is expected a natural formalization of this work would give a positive answer for this question.

7 Pseudo determinacy and Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}

In this section, we introduce a Ξ 21\Pi^{1}_{2} variant of determinacy which we call pseudo-determinacy. Then we consider the hierarchy of pseudo-determinacy for (Ξ£10)k(\Sigma^{0}_{1})_{k} games. First of all, we introduce the notion of difference hierarchy and the determinacy.

Definition 7.1 (Difference hierarchy).

Let kβˆˆΟ‰k\in\omega. We define the class (Ξ£10)k(\Sigma^{0}_{1})_{k} of formulas as follows:

  • β€’

    a formula is (Ξ£10)1(\Sigma^{0}_{1})_{1} if it is Ξ£10\Sigma^{0}_{1},

  • β€’

    a formula Ο†\varphi is (Ξ£10)k+1(\Sigma^{0}_{1})_{k+1} if there are a Ξ£10\Sigma^{0}_{1} formula ψ\psi and a (Ξ£10)k(\Sigma^{0}_{1})_{k} formula ΞΈ\theta such that Ο†β‰‘Οˆβˆ§Β¬ΞΈ\varphi\equiv\psi\land\lnot\theta.

We call the hierarchy formed by {(Ξ£10)k:kβˆˆΟ‰}\{(\Sigma^{0}_{1})_{k}:k\in\omega\} the difference hierarchy of Ξ£10\Sigma^{0}_{1} formulas.

Definition 7.2.

We call a function S:β„•<β„•β†’β„•S:\mathbb{N}^{<\mathbb{N}}\to\mathbb{N} a strategy777Usually, a strategy is a function defined over ⋃nβ„•2​n\bigcup_{n}\mathbb{N}^{2n} or ⋃nβ„•2​n+1\bigcup_{n}\mathbb{N}^{2n+1}.. Let SS and Sβ€²S^{\prime} be strategies. We say a function fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}} is the play along (S,Sβ€²)(S,S^{\prime}) (write f=SβŠ—Sβ€²f=S\otimes S^{\prime}) if

βˆ€n​(f​(2​n)=S​(f​[2​n])∧f​(2​n+1)=S′​(f​[2​n+1])).\forall n(f(2n)=S(f[2n])\land f(2n+1)=S^{\prime}(f[2n+1])).

Let φ​(Xβ†’,f)\varphi(\vec{X},f) be a formula for fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}}. Let Xβ†’\vec{X} be sets. We say φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is determined if

βˆƒS0β€‹βˆ€S1​φ​(Xβ†’,S0βŠ—S1)βˆ¨βˆƒS1β€‹βˆ€S0​¬φ​(Xβ†’,S0βŠ—S1).\displaystyle\exists S_{0}\forall S_{1}\varphi(\vec{X},S_{0}\otimes S_{1})\lor\exists S_{1}\forall S_{0}\lnot\varphi(\vec{X},S_{0}\otimes S_{1}).

Especially, if a strategy S0S_{0} satisfies

βˆ€S1​φ​(Xβ†’,S0βŠ—S1),\displaystyle\forall S_{1}\varphi(\vec{X},S_{0}\otimes S_{1}),

then we say φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is determined via a strategy S0S_{0} for player 0. If a strategy S1S_{1} satisfies

βˆ€S0​¬φ​(Xβ†’,S0βŠ—S1),\displaystyle\forall S_{0}\lnot\varphi(\vec{X},S_{0}\otimes S_{1}),

then we say φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is determined via a strategy S1S_{1} for player 11.

We also say Ο†\varphi is determined if for any Xβ†’\vec{X}, φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is determined. Let Ξ“\Gamma be a class of formulas for fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}}. We say Ξ“\Gamma is determined (write Ξ“βˆ’π–£π–Ύπ—\Gamma\mathchar 45\relax\mathsf{Det}) if any Ο†βˆˆΞ“\varphi\in\Gamma is determined.

It is well-known that over 𝖱𝖒𝖠0\mathsf{RCA}_{0},

  • β€’

    𝖠𝖳𝖱0\mathsf{ATR}_{0} is equivalent to Ξ£10βˆ’π–£π–Ύπ—\Sigma^{0}_{1}\mathchar 45\relax\mathsf{Det}, and

  • β€’

    Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is equivalent to (Ξ£10)kβˆ’π–£π–Ύπ—(\Sigma^{0}_{1})_{k}\mathchar 45\relax\mathsf{Det} for k>1k>1.

In this section, we consider Ξ 21\Pi^{1}_{2} variants of the second clause.

Definition 7.3 (pseudo-determinacy).

Let φ​(Xβ†’,f)\varphi(\vec{X},f) be a formula for fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}}. For given sets Xβ†’\vec{X}, we say φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is pseudo-determined if

βˆƒS0βˆ€S1≀TaS0βŠ•Xβ†’Ο†(Xβ†’,S0βŠ—S1)βˆ¨βˆƒS1βˆ€S0≀TaS1βŠ•X→¬φ(Xβ†’,S0βŠ—S1)).\displaystyle\exists S_{0}\forall S_{1}\leq_{\mathrm{T}}^{a}S_{0}\oplus\vec{X}\varphi(\vec{X},S_{0}\otimes S_{1})\lor\exists S_{1}\forall S_{0}\leq_{\mathrm{T}}^{a}S_{1}\oplus\vec{X}\lnot\varphi(\vec{X},S_{0}\otimes S_{1})).

If φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is pseudo-determined via SS, then we say SS is a pseudo-winning strategy for the game defined by Ο†\varphi and Xβ†’\vec{X}. If Xβ†’\vec{X} is clear from the context, we just say SS is a pseudo-winning strategy for the game defined by Ο†\varphi. If φ​(Xβ†’,βˆ™)\varphi(\vec{X},\bullet) is pseudo-determined for any Xβ†’\vec{X}, then we say Ο†\varphi is pseudo-determined.

Let Ξ“\Gamma be a class of formulas for fβˆˆβ„•β„•f\in\mathbb{N}^{\mathbb{N}}. We say Ξ“\Gamma is pseudo-determined if any Ο†βˆˆΞ“\varphi\in\Gamma is pseudo-determined. Let rel⁑(Ξ“βˆ’π–£π–Ύπ—)\operatorname{rel}(\Gamma\mathchar 45\relax\mathsf{Det}) denote the assertion that every game defined by a formula in Ξ“\Gamma is pseudo-determined.

We note that the statement [Ο†\varphi is pseudo-determined] is of the form rel⁑(ψ)\operatorname{rel}(\psi) for some ψ\psi in the sense of Section 5. Moreover, each rel⁑((Ξ£10)nβˆ’π–£π–Ύπ—)\operatorname{rel}((\Sigma^{0}_{1})_{n}\mathchar 45\relax\mathsf{Det}) is axiomatizable by a Ξ 21\Pi^{1}_{2} sentence of the form rel⁑(ψ)\operatorname{rel}(\psi).

Remark 7.4.

For any (Ξ£10)k(\Sigma^{0}_{1})_{k} formula φ​(X1,…,Xn,f)\varphi(X_{1},\ldots,X_{n},f), there is a (Ξ£10)k(\Sigma^{0}_{1})_{k} formula Οˆβ€‹(X,f)\psi(X,f) such that 𝖱𝖒𝖠0βŠ’βˆ€X1,…,Xn,f(Ο†(X1,…,Xn,f)β†”Οˆ(X1βŠ•β‹―βŠ•Xn,f)\mathsf{RCA}_{0}\vdash\forall X_{1},\ldots,X_{n},f(\varphi(X_{1},\ldots,X_{n},f)\leftrightarrow\psi(X_{1}\oplus\cdots\oplus X_{n},f). Therefore, when considering rel(Ξ£10)kβˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{k}\mathchar 45\relax\mathsf{Det}, we may assume that any (Ξ£10)k(\Sigma^{0}_{1})_{k} formula has exactly one variable XX and one function variable ff.

Let φ​(X,f)\varphi(X,f) be a (Ξ£10)k(\Sigma^{0}_{1})_{k} formula. We may assume that for any XX, all strategies for the game φ​(X,βˆ™)\varphi(X,\bullet) computes XX. Indeed, if SS is a strategy for player 0, then SS is enough to be defined at sequences with even length. Thus, we may assume that X​(n)=S​(12​n+1)X(n)=S(1^{2n+1}). Similarly, if SS is a strategy for player 11, then we may assume that X​(n)=S​(12​n)X(n)=S(1^{2n}). Therefore, φ​(X,f)\varphi(X,f) is pseudo-determined if

βˆ€X​(βˆƒS0β€‹βˆ€S1≀TaS0​φ​(X,S0βŠ—S1)βˆ¨βˆƒS1β€‹βˆ€S0≀TaS1​¬φ​(X,S0βŠ—S1)).\displaystyle\forall X(\exists S_{0}\forall S_{1}\leq_{\mathrm{T}}^{a}S_{0}\varphi(X,S_{0}\otimes S_{1})\lor\exists S_{1}\forall S_{0}\leq_{\mathrm{T}}^{a}S_{1}\lnot\varphi(X,S_{0}\otimes S_{1})).

We note that for any kβˆˆΟ‰,k>1k\in\omega,k>1, rel(Ξ£10)kβˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{k}\mathchar 45\relax\mathsf{Det} is a Ξ 21\Pi^{1}_{2} statement provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}. Thus, each rel(Ξ£10)kβˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{k}\mathchar 45\relax\mathsf{Det} is provable from some Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n) over 𝖠𝖒𝖠0\mathsf{ACA}_{0}. We show that Ξ²01​𝖱π–₯𝖭⁑(k;𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k;\mathsf{ATR}_{0}) is enough to prove rel(Ξ£10)kβˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{k}\mathchar 45\relax\mathsf{Det}.

To prove (Ξ£10)2(\Sigma^{0}_{1})_{2} determinacy from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0}, Tanaka proved the following lemma.

Lemma 7.5.

Let φ​(X,f)\varphi(X,f) be a (Ξ£10)2(\Sigma^{0}_{1})_{2} formula with exactly displayed free set variables. Assume 𝖠𝖳𝖱0\mathsf{ATR}_{0}. For any XX, if HJ⁑(X)\operatorname{HJ}(X) exists, then φ​(X,βˆ™)\varphi(X,\bullet) is determined.

Proof.

See [17]. ∎

We extend this lemma to (Ξ£10)k(\Sigma^{0}_{1})_{k} formulas.

Lemma 7.6.

Let kβˆˆΟ‰k\in\omega and φ​(X,f)\varphi(X,f) be a (Ξ£10)k+1(\Sigma^{0}_{1})_{k+1} formula with exactly displayed set variables. Over 𝖠𝖳𝖱0\mathsf{ATR}_{0}, for any set XX, if HJk⁑(X)\operatorname{HJ}^{k}(X) exists, then φ​(X,βˆ™)\varphi(X,\bullet) is determined.

Proof.

We will show by induction on kβˆˆΟ‰k\in\omega. The case k=0k=0 is trivial because 𝖠𝖳𝖱0\mathsf{ATR}_{0} proves Ξ£10\Sigma^{0}_{1} determinacy.

Let kβˆˆΟ‰k\in\omega. Assume that 𝖠𝖳𝖱0\mathsf{ATR}_{0} proves that for any XX, if HJk⁑(X)\operatorname{HJ}^{k}(X) exists then any (Ξ£10,X)k+1(\Sigma^{0,X}_{1})_{k+1} game is determined.

We reason in 𝖠𝖳𝖱0\mathsf{ATR}_{0}. Let XX be a set such that HJk+1⁑(X)\operatorname{HJ}^{k+1}(X) exists. Let φ​(A,f)\varphi(A,f) be a (Ξ£10)k+2(\Sigma^{0}_{1})_{k+2} formula. Then, there is a Ξ£10\Sigma^{0}_{1} formula Οˆβ€‹(A,f)\psi(A,f) and a (Ξ£10)k+1(\Sigma^{0}_{1})_{k+1} formula θ​(A,f)\theta(A,f) such that φ​(A,f)β‰‘Οˆβ€‹(A,f)βˆ§Β¬ΞΈβ€‹(A,f)\varphi(A,f)\equiv\psi(A,f)\land\lnot\theta(A,f). Moreover, Οˆβ€‹(A,f)\psi(A,f) is written as βˆƒn​η​(A,f​[n])\exists n\eta(A,f[n]) for a Ξ£00\Sigma^{0}_{0} formula Ξ·\eta. Take a coded Ξ²\beta model β„³\mathcal{M} such that Xβˆˆβ„³βŠ§βˆƒY​(Y=HJk⁑(X))X\in\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{k}(X)). Define WW by

W={Οƒβˆˆβ„•<β„•:η​(X,Οƒ)βˆ§β„³βŠ§player 0 wins for ​¬θσ​(X,βˆ™)}.\displaystyle W=\{\sigma\in\mathbb{N}^{<\mathbb{N}}:\eta(X,\sigma)\land\mathcal{M}\models\text{player 0 wins for }\lnot\theta^{\sigma}(X,\bullet)\}.

Here, ¬θσ​(X,f)\lnot\theta^{\sigma}(X,f) is the formula ¬θ​(X,Οƒβˆ—f)\lnot\theta(X,\sigma\ast f) and Οƒβˆ—f\sigma\ast f denotes the concatenation of Οƒ\sigma and ff. Let φ′​(f)β‰‘βˆƒn​(f​[n]∈W)\varphi^{\prime}(f)\equiv\exists n(f[n]\in W). Then, the game defined by Ο†β€²\varphi^{\prime} is determined.

Case 1: player 0 wins the game Ο†β€²\varphi^{\prime} via S0S_{0}. Now, for any ΟƒβˆˆW\sigma\in W, β„³βŠ§βˆƒS0Οƒβ€‹βˆ€S1σ​¬θ​(X,Οƒβˆ—S0ΟƒβŠ—S1Οƒ)\mathcal{M}\models\exists S_{0}^{\sigma}\forall S_{1}^{\sigma}\lnot\theta(X,\sigma\ast S_{0}^{\sigma}\otimes S_{1}^{\sigma}). Take a sequence ⟨S0ΟƒβŸ©ΟƒβˆˆW\langle S_{0}^{\sigma}\rangle_{\sigma\in W}. Then, each S0ΟƒS_{0}^{\sigma} satisfies βˆ€S1σ​¬θ​(X,Οƒβˆ—S0ΟƒβŠ—S1Οƒ)\forall S_{1}^{\sigma}\lnot\theta(X,\sigma\ast S_{0}^{\sigma}\otimes S_{1}^{\sigma}) because β„³\mathcal{M} is a Ξ²\beta-model. Consider a strategy for player 0 such that

  • β€’

    while the play is not in WW, the strategy mimics S0S_{0},

  • β€’

    once the play Οƒ\sigma is in WW, the strategy mimics S0ΟƒS_{0}^{\sigma}.

Then this strategy is a winning strategy for player 0.

Case 2: player 1 wins the game Ο†β€²\varphi^{\prime} via S1S_{1}. Now we have that for each play Οƒ\sigma consistent with S1S_{1}, if η​(X,Οƒ)\eta(X,\sigma) then

β„³βŠ§player 0 has no winning strategy for ​¬θσ​(X,βˆ™).\displaystyle\mathcal{M}\models\text{player 0 has no winning strategy for }\lnot\theta^{\sigma}(X,\bullet).

By induction hypothesis, β„³βŠ§ΞΈΟƒβ€‹(X,βˆ™)​ is determined\mathcal{M}\models\theta^{\sigma}(X,\bullet)\text{ is determined}. Hence for each Οƒ\sigma consistent with S1S_{1}, if η​(X,Οƒ)\eta(X,\sigma) then β„³βŠ§βˆƒS1Οƒβ€‹βˆ€S0σ​θ​(X,Οƒβˆ—S0ΟƒβŠ—S1Οƒ)\mathcal{M}\models\exists S_{1}^{\sigma}\forall S_{0}^{\sigma}\theta(X,\sigma\ast S_{0}^{\sigma}\otimes S_{1}^{\sigma}). Take a sequence ⟨S1Οƒ:σ​ is consistent with ​S1βˆ§Ξ·β€‹(X,Οƒ)⟩\langle S_{1}^{\sigma}:\sigma\text{ is consistent with }S_{1}\land\eta(X,\sigma)\rangle such that β„³βŠ§βˆ€S0σ​θ​(X,Οƒβˆ—S0ΟƒβŠ—S1Οƒ)\mathcal{M}\models\forall S_{0}^{\sigma}\theta(X,\sigma\ast S_{0}^{\sigma}\otimes S_{1}^{\sigma}). Then, since β„³\mathcal{M} is a coded Ξ²\beta-model, for each S0ΟƒS_{0}^{\sigma}, βˆ€S1σ​θ​(X,Οƒβˆ—S0ΟƒβŠ—S1Οƒ)\forall S_{1}^{\sigma}\theta(X,\sigma\ast S_{0}^{\sigma}\otimes S_{1}^{\sigma}) holds. Consider a strategy for player 1 such that

  • β€’

    while the play Οƒ\sigma does not satisfy η​(X,Οƒ)\eta(X,\sigma), the strategy mimics S1S_{1},

  • β€’

    once the play Οƒ\sigma satisfies η​(X,Οƒ)\eta(X,\sigma), the strategy mimics S1ΟƒS_{1}^{\sigma}.

Then this strategy is a winning strategy for player 1. ∎

Theorem 7.7.

For each kβˆˆΟ‰,k>0k\in\omega,k>0, Ξ²01​𝖱π–₯𝖭⁑(k,𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k,\mathsf{ATR}_{0}) proves rel(Ξ£10)k+1𝖣𝖾𝗍\operatorname{rel}(\Sigma^{0}_{1})_{k+1}\mathsf{Det}.

Proof.

Let φ​(A,f)\varphi(A,f) be a (Ξ£10)k+1(\Sigma^{0}_{1})_{k+1} formula.

Assume Ξ²01​𝖱π–₯𝖭⁑(k,𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k,\mathsf{ATR}_{0}). Let XX be a set. By Ξ²01​𝖱π–₯𝖭⁑(k,𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(k,\mathsf{ATR}_{0}) take a coded Ο‰\omega-model β„³\mathcal{M} such that Xβˆˆβ„³βŠ§π– π–³π–±0+βˆƒY​(Y=HJk⁑(X))X\in\mathcal{M}\models\mathsf{ATR}_{0}+\exists Y(Y=\operatorname{HJ}^{k}(X)). Then, by the previous lemma, β„³βŠ§Ο†β€‹(X,βˆ™)​ is determined\mathcal{M}\models\varphi(X,\bullet)\text{ is determined}. Without loss of generality, we may assume that β„³βŠ§player 0 wins for φ​(X,βˆ™)\mathcal{M}\models\text{player 0 wins for $\varphi(X,\bullet)$}. Let S0βˆˆβ„³S_{0}\in\mathcal{M} be a winning strategy of player 0. Then, for any S1≀TaS0S_{1}\leq_{\mathrm{T}}^{\mathrm{a}}S_{0}, S1βˆˆβ„³S_{1}\in\mathcal{M} and hence φ​(X,S0βŠ—S1)\varphi(X,S_{0}\otimes S_{1}). ∎

We next see the implication from pseudo-determinacy to Ξ²01​𝖱π–₯𝖭\beta^{1}_{0}\operatorname{\mathsf{RFN}}. First we see the implication from rel(Ξ£10)2βˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}.

Lemma 7.8.

There exists a (Ξ£10)2(\Sigma^{0}_{1})_{2} formula φ​(f,X,Y)\varphi(f,X,Y) such that 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves the following.

βˆ€Ξ±:WOβˆ€X(φ​(βˆ™,X,Ξ±)Β is determinedβ†’βˆƒY(Y=(HJ(X))(Ξ±)).\forall\alpha:\operatorname{WO}\forall X(\text{$\varphi(\bullet,X,\alpha)$ is determined}\to\exists Y(Y=(\operatorname{HJ}(X))^{(\alpha)}).
Proof.

See [17]. ∎

Theorem 7.9.

Over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, rel(Ξ£10)2βˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det} proves 𝖳𝖫𝖯𝖯\mathsf{TLPP}.

Proof.

By 4.23, it is enough to show that

βˆ€Ξ±:WOβˆ€Xβˆƒβ„³(\displaystyle\forall\alpha:\operatorname{WO}\forall X\exists\mathcal{M}( Ξ±,Xβˆˆβ„³,\displaystyle\alpha,X\in\mathcal{M},
β„³βŠ§π– π–’π– 0+βˆƒY​(Y=HJ⁑(X)),\displaystyle\mathcal{M}\models\mathsf{ACA}_{0}+\exists Y(Y=\operatorname{HJ}(X)),
β„³Β is closed underΒ Ξ±-jump).\displaystyle\text{$\mathcal{M}$ is closed under $\alpha$-jump}).

Let φ​(f,X,Y)\varphi(f,X,Y) be the formula in the Lemma 7.8. Take a well-ordering Ξ±\alpha and a set XX. By rel(Ξ£10)2βˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det} take a pseudo-winning strategy SS for the game φ​(βˆ™,X,Ξ±β‹…Ο‰)\varphi(\bullet,X,\alpha\cdot\omega). Let β„³\mathcal{M} be a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} containing X,Ξ±β‹…Ο‰,SX,\alpha\cdot\omega,S. Then β„³βŠ§βˆƒY​(Y=(HJ⁑(X))(Ξ±β‹…Ο‰))\mathcal{M}\models\exists Y(Y=(\operatorname{HJ}(X))^{(\alpha\cdot\omega)}).

We define a coded Ο‰\omega-model β„³β€²\mathcal{M}^{\prime} by

β„³β€²={Aβˆˆβ„³:β„³βŠ§βˆƒn​(A≀T(HJ⁑(X))(Ξ±β‹…n))}.\mathcal{M}^{\prime}=\{A\in\mathcal{M}:\mathcal{M}\models\exists n(A\leq_{\mathrm{T}}(\operatorname{HJ}(X))^{(\alpha\cdot n)})\}.

Then, β„³β€²\mathcal{M}^{\prime} is a coded Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} closed under Ξ±\alpha-jump. Moreover, since β„³β€²βˆˆβ„³\mathcal{M}^{\prime}\in\mathcal{M} and β„³βŠ§[HJ⁑(X)βˆˆβ„³β€²]\mathcal{M}\models[\operatorname{HJ}(X)\in\mathcal{M}^{\prime}], β„³β€²βŠ§βˆƒY​(Y=HJ⁑(X))\mathcal{M}^{\prime}\models\exists Y(Y=\operatorname{HJ}(X)) by Lemma 2.28. ∎

Corollary 7.10.

Ξ²01​𝖱π–₯𝖭⁑(1)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1) is not enough to prove rel(Ξ£10)2βˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}.

Proof.

It follows from Ξ²01𝖱π–₯𝖭(1)<𝖳𝖫𝖯𝖯≀rel(Ξ£10)2βˆ’π–£π–Ύπ—\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1)<\mathsf{TLPP}\leq\operatorname{rel}(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}. ∎

Remark 7.11.

In his paper [17], Tanaka pointed that 𝖠𝖒𝖠0\mathsf{ACA}_{0} is not strong enough to prove β€œΞ 11\Pi^{1}_{1} comprehension implies (Ξ£10)2βˆ’π–£π–Ύπ—(\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}” by using the axiomatic system lightface Ξ 11\Pi^{1}_{1} comprehension. The previous corollary is another expression of this fact.

At last, we see the general case. The point is the following lemma.

Lemma 7.12.

Let n>0n>0. Then there is a (Ξ£10)p​(n)(\Sigma^{0}_{1})_{p(n)} formula Gn​(X,f)G_{n}(X,f) such that 𝖠𝖒𝖠0\mathsf{ACA}_{0} proves that

  • β€’

    βˆ€X​(Gn​(X,βˆ™)​ is determinedβ†’HJn⁑(X)​ exists.)\forall X(G_{n}(X,\bullet)\text{ is determined}\to\operatorname{HJ}^{n}(X)\text{ exists.})

  • β€’

    Any winning strategy for GnG_{n} computes XX.

Here, p​(n)p(n) is a certain primitive recursive function.

Proof.

See the proof of Proposition 6 in [9]. ∎

Theorem 7.13.

Let n>0n>0. Then, over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}_{0}^{+}, rel(Ξ£10)p​(n)βˆ’π–£π–Ύπ—\operatorname{rel}(\Sigma^{0}_{1})_{p(n)}\mathchar 45\relax\mathsf{Det} implies Ξ²01​𝖱π–₯𝖭⁑(n)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n).

Proof.

We reason in 𝖠𝖒𝖠0++rel(Ξ£10)p​(n)βˆ’π–£π–Ύπ—\operatorname{\mathsf{ACA}}_{0}^{+}+\operatorname{rel}(\Sigma^{0}_{1})_{p(n)}\mathchar 45\relax\mathsf{Det}. Let XX be a set. By rel(Ξ£10)p​(n)\operatorname{rel}(\Sigma^{0}_{1})_{p(n)}, take a pseudo-winning strategy SS for the game Gn​(X,βˆ™)G_{n}(X,\bullet). Let β„³\mathcal{M} be an Ο‰\omega-model of 𝖠𝖒𝖠0\mathsf{ACA}_{0} containing SS. Then β„³βŠ§[Gn​(X,βˆ™)​ is determined via ​S]\mathcal{M}\models[G_{n}(X,\bullet)\text{ is determined via }S], and hence β„³βŠ§βˆƒY​(Y=HJn⁑(X))\mathcal{M}\models\exists Y(Y=\operatorname{HJ}^{n}(X)). This completes the proof. ∎

Corollary 7.14.

Over 𝖠𝖒𝖠0+\operatorname{\mathsf{ACA}}^{+}_{0}, {rel⁑((Ξ£10)nβˆ’π–£π–Ύπ—):nβˆˆΟ‰}\{\operatorname{rel}((\Sigma^{0}_{1})_{n}\mathchar 45\relax\mathsf{Det}):n\in\omega\} and {Ξ²01​𝖱π–₯𝖭⁑(n):nβˆˆΟ‰}\{\beta^{1}_{0}\operatorname{\mathsf{RFN}}(n):n\in\omega\} prove the same sentences. Therefore, any Ξ 21\Pi^{1}_{2} sentence provable from Ξ 11βˆ’π–’π– 0\Pi^{1}_{1}\mathchar 45\relax\mathsf{CA}_{0} is provable from 𝖠𝖒𝖠0++rel((Ξ£10)nβˆ’π–£π–Ύπ—))\operatorname{\mathsf{ACA}}^{+}_{0}+\operatorname{rel}((\Sigma^{0}_{1})_{n}\mathchar 45\relax\mathsf{Det})) for some nβˆˆΟ‰n\in\omega.

Remark 7.15.

We have proved in Lemma 3.7, Theorem 7.7 and 7.9 that

  • β€’

    𝖳𝖫𝖯𝖯≀rel⁑((Ξ£10)2βˆ’π–£π–Ύπ—)≀β01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\mathsf{TLPP}\leq\operatorname{rel}((\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det})\leq\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}),

  • β€’

    𝖳𝖫𝖯𝖯<Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\mathsf{TLPP}<\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}).

Thus, at least one of 𝖳𝖫𝖯𝖯\mathsf{TLPP} or Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0}) is not equivalent to rel⁑((Ξ£10)2βˆ’π–£π–Ύπ—)\operatorname{rel}((\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}).

Question 7.16.

Can we separate rel⁑((Ξ£10)2βˆ’π–£π–Ύπ—)\operatorname{rel}((\Sigma^{0}_{1})_{2}\mathchar 45\relax\mathsf{Det}) and 𝖳𝖫𝖯𝖯\mathsf{TLPP} or Ξ²01​𝖱π–₯𝖭⁑(1;𝖠𝖳𝖱0)\beta^{1}_{0}\operatorname{\mathsf{RFN}}(1;\mathsf{ATR}_{0})?

References

  • [1] SamuelΒ R. Buss, editor. Handbook of Proof Theory, volume 137. Amsterdam, 1998.
  • [2] DamirΒ D. Dzhafarov and Carl Mummert. Reverse mathematicsβ€”problems, reductions, and proofs. Theory and Applications of Computability. Springer, Cham, [2022] Β©2022.
  • [3] D.Β FernΓ‘ndez-Duque, P.Β Shafer, H.Β Towsner, and K.Β Yokoyama. Metric fixed point theory and partial impredicativity. Philos. Trans. Roy. Soc. A, 381(2248):Paper No. 20220012, 20, 2023.
  • [4] Anton Freund. FraΓ―ssé’s conjecture, partial impredicativity and well-ordering principles, part i. arXiv preprint arXiv:2406.13485, 2024.
  • [5] Gian Marco and Albert Marcone. The Galvin-Prikry theorem in the weihrauch lattice. arXiv preprint arXiv:2410.06928.
  • [6] Alberto Marcone. On the logical strength of Nash-Williams’ theorem on transfinite sequences. In Logic: from foundations to applications: European logic colloquium, pages 327–351, 1996.
  • [7] Antonio MontalbΓ‘n. FraΓ―ssé’s conjecture in Ξ 11\Pi_{1}^{1}-comprehension. J. Math. Log., 17(2):1750006, 12, 2017.
  • [8] Antonio MontalbΓ‘n and RichardΒ A Shore. Conservativity of ultrafilters over subsystems of second order arithmetic. The Journal of Symbolic Logic, 83(2):740–765, 2018.
  • [9] Leonardo Pacheco and Keita Yokoyama. Determinacy and reflection principles in second-order arithmetic. arXiv preprint arXiv:2209.04082, 2022.
  • [10] Fedor Pakhomov and Giovanni SoldΓ . On Nash-Williams’ theorem regarding sequences with finite range. arXiv preprint arXiv:2405.13842, 2024.
  • [11] Pavel PudlΓ‘k. The lengths of proofs. In Buss [1], chapter VIII, pages 547–637.
  • [12] Paul Shafer. Menger’s theorem in Ξ 11βˆ’CA0\Pi^{1}_{1}\mathrm{-CA}_{0}. Arch. Math. Logic, 51(3-4):407–423, 2012.
  • [13] StephenΒ G. Simpson. Subsystems of second order arithmetic. Perspectives in Logic. Cambridge University Press, Cambridge; Association for Symbolic Logic, Poughkeepsie, NY, second edition, 2009.
  • [14] Yudai Suzuki. Relative leftmost path principles and omega-model reflections of transfinite inductions. arXiv preprint arXiv:2407.13504, 2024.
  • [15] Yudai Suzuki and Keita Yokoyama. Searching problems above arithmetical transfinite recursion. Annals of Pure and Applied Logic, page 103488, 2024.
  • [16] Kazuyuki Tanaka. The Galvin-Prikry theorem and set existence axioms. Annals of Pure and Applied Logic, 42(1):81–104, 1989.
  • [17] Kazuyuki Tanaka. Weak axioms of determinacy and subsystems of analysis I: Ξ”20\Delta^{0}_{2} games. Mathematical Logic Quarterly, 36(6):481–491, 1990.
  • [18] Henry Towsner. Partial impredicativity in reverse mathematics. J. Symbolic Logic, 78(2):459–488, 2013.