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

Coding information into all infinite subsets of a dense set

Matthew Harrison-Trainor Lu Liu  and  Patrick Lutz
Abstract.

Suppose you have an uncomputable set XX and you want to find a set AA, all of whose infinite subsets compute XX. There are several ways to do this, but all of them seem to produce a set AA which is fairly sparse. We show that this is necessary in the following technical sense: if XX is uncomputable and AA is a set of positive lower density then AA has an infinite subset which does not compute XX. We also prove an analogous result for PA degree: if XX is uncomputable and AA is a set of positive lower density then AA has an infinite subset which is not of PA degree. We will show that these theorems are sharp in certain senses and also prove a quantitative version formulated in terms of Kolmogorov complexity. Our results use a modified version of Mathias forcing and build on work by Seetapun, Liu, and others on the reverse math of Ramsey’s theorem for pairs.

This work was supported by the National Science Foundation under Grant DMS-2153823 and Grant DMS-2203072.

1. Introduction

Suppose you have an uncomputable set XX and would like to find an infinite set AA\subseteq\mathbb{N} such that all infinite subsets of AA compute XX. Here’s one way to do this, due to Dekker and Myhill [3]: identify \mathbb{N} with 2<ω2^{<\omega} and let AA be the set of all finite initial segments of XX.

This is not the only way to encode an uncomputable set XX into all infinite subsets of AA. For example, if XX is hyperarithmetic then it can be computed from any sufficiently fast growing function ([9], Theorem 6.8). If we make sure AA has sufficiently large gaps between its elements then for any infinite subset BB of AA, the function which enumerates the elements of BB grows fast enough to compute XX and hence BB itself computes XX.

Note that both of these methods produce fairly sparse subsets of \mathbb{N}. If we use Dekker and Myhill’s method then the set AA will have just nn elements less than 2n2^{n}. If we use the second method (in the case where XX is hyperarithmetic) then the set AA will be even sparser—the gaps between successive elements of AA grow faster than any computable function.

It seems reasonable to informally conjecture that this is a necessary feature of such coding methods; in other words, to conjecture that if XX is uncomputable and every infinite subset of AA computes XX then AA must be sparse. We can turn this informal conjecture into a formal one by picking a precise definition of “sparse.”

The main theorem of this paper states that the conjecture holds if we define “sparse” to mean “lower density zero.” Recall that the lower density of a set of natural numbers AA\subseteq\mathbb{N} is

ρ¯(A)=lim infn|A[n]|n+1\operatorname{\underline{\rho}}(A)=\liminf_{n\to\infty}\frac{|A\cap[n]|}{n+1}

where [n][n] denotes the set {0,1,,n}\{0,1,\ldots,n\}. The main theorem of this paper is as follows.

Theorem 1.1.

For any uncomputable set XX and any set AA\subseteq\mathbb{N} of positive lower density, there is some infinite subset of AA which does not compute XX.

Our proof of this theorem relies on a theorem implicit in the work of Seetapun [15] and first proved explicitly by Dzhafarov and Jockusch [6].111This theorem is sometimes known by the name “strong cone avoidance for 𝖱𝖳21\mathsf{RT}^{1}_{2},” which originates from its connection to the reverse math of Ramsey’s theorem. We decided not to use that name here since we are not concerned with reverse math in this paper. For a proof, see [8], Theorem 6.63.

Theorem 1.2 (Seetapun’s theorem).

For any uncomputable set XX and any set AA\subseteq\mathbb{N}, either AA or A¯\overline{A} must contain some infinite subset which does not compute XX.

Our proof is partially inspired by proofs of results related to Seetapun’s theorem by Cholak, Jockusch and Slaman [2], Dzhafarov and Jockusch [6] and Monin and Patey [12]. More specifically, those results are proved using variations on Mathias forcing and use the low basis theorem (or the cone avoiding basis theorem) to show that certain sets of conditions are dense. Our proof also uses a variation on Mathias forcing and the cone avoiding basis theorem, but, in addition, uses Seetapun’s theorem in a manner similar to the cone avoiding basis theorem. We will explain our strategy more carefully in Section 3.

It is natural to ask whether Theorem 1.1 holds for stronger notions of sparsity. In Section 4, we will show that our theorem is sharp in the sense that it fails to hold for several such notions.

We will also prove two other results which show that it is difficult to encode information into all infinite subsets of a dense set. These two results are stated in terms of having PA degree and in terms of Kolmogorov complexity, respectively.

Avoiding PA degree

Theorem 1.1 can be rephrased in terms of the property of cone avoidance. Say that a set 𝒞2ω\mathcal{C}\subseteq 2^{\omega} avoids cones if for every nontrivial cone of Turing degrees, there is some element of 𝒞\mathcal{C} which is not in that cone. Then Theorem 1.1 can be restated as: for every set AA\subseteq\mathbb{N} of positive lower density, the set of infinite subsets of AA avoids cones.

Cone avoidance can be seen as a kind of computability-theoretic weakness. Another standard weakness notion is PA avoidance. Say that a set 𝒞2ω\mathcal{C}\subseteq 2^{\omega} avoids PA degree if there is some element of 𝒞\mathcal{C} which is not of PA degree. Though cone avoidance and PA avoidance are not equivalent, they do occur together relatively frequently.

One example of this comes from research on 𝖱𝖳22\mathsf{RT}^{2}_{2}, a statement of Ramsey theory which has been thoroughly studied in the field of Reverse Mathematics. Seetapun showed that over 𝖱𝖢𝖠0\mathsf{RCA}_{0}, 𝖱𝖳22\mathsf{RT}^{2}_{2} does not imply 𝖠𝖢𝖠0\mathsf{ACA}_{0}. A core part of the proof is Seetapun’s theorem above, which can be read as stating that for any set AA\subseteq\mathbb{N}, the set of infinite subsets of AA and A¯\overline{A} avoids cones. Later, Liu showed that 𝖱𝖳22\mathsf{RT}^{2}_{2} also does not imply 𝖶𝖪𝖫0\mathsf{WKL}_{0} and the heart of his proof was a theorem—analogous to Seetapun’s theorem—stating that for every set AA\subseteq\mathbb{N}, the set of infinite subsets of AA and A¯\overline{A} avoids PA degree.

Theorem 1.3 (Liu’s Theorem, [10]).

For any set AA\subseteq\mathbb{N}, either AA or A¯\overline{A} must contain an infinite set that is not of PA degree.

It seems natural to ask whether our main theorem can be modified to yield PA avoidance rather than cone avoidance, in the same way that Seetapun’s theorem can be modified to yield Liu’s theorem. We show that this is indeed the case.

Theorem 1.4.

For any set AA\subseteq\mathbb{N} of positive lower density, there is an infinite subset of AA which is not of PA degree.

Just as our proof of Theorem 1.1 uses Seetapun’s theorem, our proof of the theorem above uses Liu’s theorem (along with several techniques first developed to prove Liu’s theorem). We will also show that a certain natural-sounding common generalization of both Theorem 1.1 and Theorem 1.4 is false.

Kolmogorov complexity

We will also investigate a quantitative version of Theorem 1.1. Stated loosely, that theorem says that it is impossible to encode an infinite amount of information into all infinite subsets of a set of positive lower density. On the other hand, it is obvious that some finite information may be so encoded. For example, it is easy to encode one bit of information into all infinite subsets of a set of lower density 1/21/2 (by using the parity of the elements of the set) and, more generally, nn bits of information into all infinite subsets of a set of lower density 1/2n1/2^{n}. But just how much information can be encoded?

We can formulate a precise version of this question using Kolmogorov complexity. To do so, it is convenient to introduce the following definition.

Definition 1.5.

For a string σ2<ω\sigma\in 2^{<\omega} and a family of sets 𝒫()\mathcal{F}\subseteq\operatorname{\mathcal{P}}(\mathbb{N}), define

C(σ)=maxXCX(σ).C(\sigma\mid\mathcal{F})=\max_{X\in\mathcal{F}}C^{X}(\sigma).

Informally, C(σ)C(\sigma\mid\mathcal{F}) should be thought of as the Kolmogorov complexity of σ\sigma relative to an arbitrary element of \mathcal{F}. Recall that for any set AA\subseteq\mathbb{N}, [A]ω[A]^{\omega} denotes the family of all infinite subsets of AA. The quantity

C(σ)C(σ[A]ω)C(\sigma)-C(\sigma\mid[A]^{\omega})

can be thought of as the number of bits of information about σ\sigma that are encoded into all infinite subsets of AA. Thus a somewhat more formal version of our question above is: if σ\sigma is any string and AA\subseteq\mathbb{N} is a set of lower density at least δ>0\delta>0, then how large can C(σ)C(σ[A]ω)C(\sigma)-C(\sigma\mid[A]^{\omega}) be?

Based on our example above, it is perhaps natural to guess that this difference should not be much larger than log(1/δ)\log(1/\delta)—in other words, that it should not be possible to encode more than about log(1/δ)\log(1/\delta) bits of information about σ\sigma into all infinite subsets of a set of lower density δ\delta. Surprisingly, this is not the case.

Proposition 1.6.

For any string σ\sigma and δ(0,1]\delta\in(0,1], there is some set AA\subseteq\mathbb{N} of lower density at least δ\delta such that

C(σ[A]ω)max(0,C0(σ)log(1/δ))+O(loglog(1/δ)).C(\sigma\mid[A]^{\omega})\leq\max(0,C^{0^{\prime}}(\sigma)-\log(1/\delta))+O(\log\log(1/\delta)).

In other words, in addition to lowering the complexity of σ\sigma by log(1/δ)\log(1/\delta), we can also lower it to the 00^{\prime} complexity of σ\sigma. In fact, this upper bound on C(σ[A]ω)C(\sigma\mid[A]^{\omega}) is also optimal (up to a small error term).

Theorem 1.7.

For any string σ\sigma and set AA\subseteq\mathbb{N} of lower density at least δ(0,1]\delta\in(0,1],

C(σ[A]ω)C0(σ)log(1/δ)O(loglog(1/δ))C(\sigma\mid[A]^{\omega})\geq C^{0^{\prime}}(\sigma)-\log(1/\delta)-O(\log\log(1/\delta))

where the constant hidden by the O()O(\cdot) notation does not depend on σ\sigma or AA.

We will prove these two results in Section 6 and comment on some further questions around how much information can be encoded into all infinite subsets of a dense set.

Acknowledgements

We thank Andrew Marks for posing the question that motivated this paper and for several helpful conversations and Damir Dzhafarov for a useful conversation on the topic of Weihrauch reducibility.

2. Preliminaries

2.1. Notation

We will use the following notation for finite sets of natural numbers. For natural numbers n<mn<m,

[n]\displaystyle[n] ={0,1,,n}\displaystyle=\{0,1,\ldots,n\}
[n,m)\displaystyle[n,m) ={n,n+1,,m1}\displaystyle=\{n,n+1,\ldots,m-1\}
(n,m)\displaystyle(n,m) ={n+1,n+2,,m1}.\displaystyle=\{n+1,n+2,\ldots,m-1\}.

We will also use the following notation related to infinite sets of natural numbers. For a set AA\subseteq\mathbb{N},

A(n)\displaystyle A(n) = the nth bit of A, i.e. 0 if nA and 1 if nA\displaystyle=\text{ the $n^{\text{th}}$ bit of $A$, i.e.\ $0$ if $n\notin A$ and $1$ if $n\in A$}
A¯\displaystyle\overline{A} = the complement of A\displaystyle=\text{ the complement of $A$}
[A]ω\displaystyle[A]^{\omega} = the set of infinite subsets of A.\displaystyle=\text{ the set of infinite subsets of $A$}.

By a Turing functional we mean a program Φ\Phi with oracle access which has inputs in \mathbb{N} and outputs in {0,1}\{0,1\}. We will use the following notation related to Turing functionals. For a Turing functional Φ\Phi, oracle AA\subseteq\mathbb{N} and number nn\in\mathbb{N},

Φ(A,n)\displaystyle\Phi(A,n) = the output of Φ with oracle A on input n\displaystyle=\text{ the output of $\Phi$ with oracle $A$ on input $n$}
Φ(A)\displaystyle\Phi(A) = the partial function {0,1} given by nΦ(A,n)\displaystyle=\text{ the partial function $\mathbb{N}\to\{0,1\}$ given by $n\mapsto\Phi(A,n)$}
Φ(A,n)b\displaystyle\Phi(A,n)\neq b  means that either Φ(A,n) diverges or Φ(A,n)b.\displaystyle\;\text{ means that either $\Phi(A,n)$ diverges or $\Phi(A,n)\!\downarrow\;\neq b$}.

We will sometimes want to consider a finite set ss\subseteq\mathbb{N} as an initial segment of an oracle. We will use Φ(s,n)\Phi(s,n) to mean the output of Φ\Phi on input nn when run for at most max(s)\max(s) steps, using ss as an oracle and automatically diverging if there is any query to the oracle about a number larger than max(s)\max(s).

Similarly, we will sometimes want to consider the Kolmogorov complexity of a string σ\sigma relative to a finite set ss\subseteq\mathbb{N}, which we will denote by Cs(σ)C^{s}(\sigma). More precisely, Cs(σ)C^{s}(\sigma) will denote the length of the shortest program which outputs σ\sigma when using an oracle for ss and which runs for at most max(s)\max(s) steps, never makes an oracle query about a number larger than max(s)\max(s) and has length at most max(s)\max(s). Note that it is possible that no such program exists due to the limitations on length and running time. In this case we define Cs(σ)=C^{s}(\sigma)=\infty. Also note that Cs(σ)C^{s}(\sigma) is computable.

2.2. Density of sets of natural numbers

If AA\subseteq\mathbb{N} is a set of natural numbers then its lower density, denoted ρ¯\operatorname{\underline{\rho}}, and upper density, denoted ρ¯\operatorname{\overline{\rho}}, are defined by

ρ¯(A)\displaystyle\operatorname{\underline{\rho}}(A) =lim infn|A[n]|n+1\displaystyle=\liminf_{n\to\infty}\frac{|A\cap[n]|}{n+1}
ρ¯(A)\displaystyle\operatorname{\overline{\rho}}(A) =lim supn|A[n]|n+1.\displaystyle=\limsup_{n\to\infty}\frac{|A\cap[n]|}{n+1}.

If ρ¯(A)>0\operatorname{\underline{\rho}}(A)>0 then we say AA has positive lower density. Similarly, if ρ¯(A)>0\operatorname{\overline{\rho}}(A)>0 then AA has positive upper density.

In order to work with dense subsets of \mathbb{N}, it will be helpful to introduce some auxiliary terminology. To motivate this terminology, note that if the upper density of a set AA is strictly greater than δ\delta then there are infinitely many nn such that

|A[n]|δ(n+1).|A\cap[n]|\geq\delta\cdot(n+1).

It is useful to be able to speak about the collection of all such nn. To that end, we introduce the following terminology. For A,DA,D\subseteq\mathbb{N} and δ>0\delta>0:

  • AA is δ\delta-dense at nn if |A[n]|δ(n+1)|A\cap[n]|\geq\delta\cdot(n+1).

  • AA is δ\delta-dense if it is δ\delta-dense at every nn\in\mathbb{N} and dense if it is δ\delta-dense for some δ>0\delta>0.

  • AA is δ\delta-dense along DD if it is δ\delta-dense at every nDn\in D and dense along DD if it is δ\delta-dense along DD for some δ>0\delta>0.

The reason these notions are useful is that they act as lower complexity versions of the properties of having positive upper or lower density.

For example, a set AA has positive lower density if and only if {0}A\{0\}\cup A is δ\delta-dense for some δ>0\delta>0 (though note that δ\delta may need to be much lower than ρ¯(A)\operatorname{\underline{\rho}}(A)), but the property of having positive lower density is Σ20\Sigma^{0}_{2}, while the property of being δ\delta-dense is Π10\Pi^{0}_{1}.

Similarly, a set AA has positive upper density if and only if it is δ\delta-dense along DD for some δ>0\delta>0 and infinite set DD\subseteq\mathbb{N} (and this time, δ\delta can be arbitrarily close to ρ¯(A)\operatorname{\overline{\rho}}(A)), but the property of having positive upper density is Σ30\Sigma^{0}_{3} while the property of being δ\delta-dense along DD is again Π10\Pi^{0}_{1}.

2.3. Mathias forcing

Mathias forcing is a useful tool for constructing an infinite subset of a set while ensuring that the subset being constructed satisfies various properties. We will briefly review the basics of Mathias forcing; for a more complete introduction, see [8], Section 6.5.

A condition for Mathias forcing is a pair (s,A)(s,A) consisting of a finite set ss\subseteq\mathbb{N}, called the stem, and an infinite set AA\subseteq\mathbb{N}, called the reservoir, such that max(s)<min(A)\max(s)<\min(A). Often, the reservoir AA is required to come from some restricted class of sets, such as a Turing ideal or co-cone.

A Mathias condition, (s,A)(s,A), is extended by a condition, (s,A)(s^{\prime},A^{\prime}), written (s,A)(s,A)(s,A)\geq(s^{\prime},A^{\prime}), if all of the following hold.

  • sss\subseteq s^{\prime}

  • AAA\supseteq A^{\prime}

  • and for all nssn\in s^{\prime}\setminus s, nAn\in A.

In other words, (s,A)(s^{\prime},A^{\prime}) is formed from (s,A)(s,A) by choosing finitely many elements of the reservoir to add to the stem and by removing some elements (possibly infinitely many) from the reservoir.

A Mathias condition, (s,A)(s,A), should be thought of as partially specifying a subset GG of \mathbb{N} as follows: the stem ss consists of numbers that have already been put into GG and the reservoir AA consists of numbers that may be put into GG at some later stage. Thus any number which is not in sAs\cup A is definitely not in GG.

This can be made precise as follows. Any filter for Mathias forcing can be used to define a subset of \mathbb{N} by taking the union of all the stems in the filter. We will denote this subset by GG and, in a slight abuse of terminology, will often conflate it with the filter itself (i.e. we will refer to GG as the generic).

It is straightforward to check that if the filter is sufficiently generic then GG is infinite. Also, say that a set BB is compatible with a condition (s,A)(s,A) if sBsAs\subseteq B\subseteq s\cup A. If (s,A)(s,A) is any element of the filter defining GG then it is straightforward to check that GG is compatible with (s,A)(s,A).

Thus one way (though not the only way) to show that every sufficiently generic GG satisfies some property PP is to show that the following set of conditions

{(s,A)all infinite sets C compatible with (s,A) satisfy P}\{(s,A)\mid\text{all infinite sets $C$ compatible with $(s,A)$ satisfy $P$}\}

is dense in the Mathias forcing partial order.

2.4. Seetapun’s theorem

In the introduction, we mentioned that our proof of Theorem 1.1 uses Seetapun’s theorem.

See 1.2

We will also need to use a couple corollaries of this theorem.

Corollary 2.1.

Suppose AA\subseteq\mathbb{N} is infinite and does not compute XX. Then for any BAB\subseteq A there is some infinite subset of either BB or ABA\setminus B which does not compute XX.

Proof.

Let π:A\pi\colon A\to\mathbb{N} be a bijection which is computable from AA. The idea is to consider Seetapun’s theorem applied to π(B)\pi(B).

Note that since π\pi is a bijection, π(B)=π(AB)\mathbb{N}\setminus\pi(B)=\pi(A\setminus B). Thus by Seetapun’s theorem relativized to AA, either π(B)\pi(B) or π(AB)\pi(A\setminus B) has an infinite subset CC such that CAC\oplus A does not compute XX. Since π\pi is a bijection, π1(C)\pi^{-1}(C) is an infinite subset of either BB or ABA\setminus B. And since AA computes π\pi, CAC\oplus A computes π1(C)\pi^{-1}(C) and hence π1(C)\pi^{-1}(C) does not compute XX. ∎

Corollary 2.2.

Suppose AA\subseteq\mathbb{N} is infinite and does not compute XX. Then for any finite partition B1,,BnB_{1},\ldots,B_{n} of AA, at least one of the BiB_{i}’s has an infinite subset which does not compute XX.

Proof.

The idea of the proof is to use induction on nn and to reduce the inductive case to the previous corollary.

The base case, n=1n=1 is trivial since in this case B1=AB_{1}=A and thus B1B_{1} itself does not compute XX.

Now assume for induction that the statement holds for all AA, XX and partitions of length at nn. Fix AA and XX as in the statement of the corollary and suppose B1,,Bn+1B_{1},\ldots,B_{n+1} is a partition of AA. By applying Corollary 2.1 to the sets inBi\bigcup_{i\leq n}B_{i} and Bn+1B_{n+1}, we obtain an infinite set CC which is a subset of either inBi\bigcup_{i\leq n}B_{i} or Bn+1B_{n+1} such that CC does not compute XX.

If CBn+1C\subseteq B_{n+1} then we are done. If CinBiC\subseteq\bigcup_{i\leq n}B_{i} then we can apply the inductive assumption to the partition B1C,,BnCB_{1}\cap C,\ldots,B_{n}\cap C of CC to obtain an infinite set DD which does not compute XX and which is a subset of BiCB_{i}\cap C (and thus of BiB_{i} as well) for some ini\leq n. ∎

3. Proof of the main theorem

For the rest of this section, suppose that XX is uncomputable and AA\subseteq\mathbb{N} has positive lower density. We want to construct an infinite subset of AA which does not compute XX. We will construct this set using Mathias forcing. The basic strategy is as follows: let GG be a Mathias generic which is compatible with the condition (,A)(\varnothing,A) (i.e. the generic filter used to define GG contains the condition (,A)(\varnothing,A)). Then we have

  • Since GG is compatible with (,A)(\varnothing,A), GAG\subseteq A.

  • Since GG is sufficiently generic, GG is infinite.

If we could also show that any sufficiently generic GG does not compute XX then we would be done.

However, it is not hard to see that a generic for plain Mathias forcing—with no restrictions on the possible reservoirs—does not have this property. For example, suppose BB is an infinite set such that all infinite subsets of BB compute XX. Then any Mathias generic GG which is compatible with (,B)(\varnothing,B) will compute XX. To solve this problem, we will impose a restriction on the reservoirs of the Mathias conditions which guarantees that every sufficiently generic GG does not compute XX.

3.1. A natural idea that doesn’t work

Perhaps the most obvious restriction to put on the reservoirs is to require them to have positive lower density. Here’s why this seems natural. The problem with plain Mathias forcing is that there are conditions (s,B)(s,B) such that every set compatible with (s,B)(s,B) computes XX. However, if we believe the statement we are trying to prove—that every set of positive lower density contains an infinite subset which does not compute XX—then this same problem cannot occur when BB is required to have positive lower density.

However, this approach does not work. The problem, briefly stated is that even though every set of positive lower density contains an infinite subset which does not compute XX (as we will eventually show), there is a set of positive lower density, all of whose subsets of positive lower density uniformly compute XX. Thus if we only use reservoirs that have positive lower density, then given a condition (s,B)(s,B) and a Turing functional Φ\Phi, there is no obvious way to find a condition (s,B)(s,B)(s^{\prime},B^{\prime})\leq(s,B) forcing that the generic GG does not compute XX via Φ\Phi (since it could be that for any such (s,B)(s^{\prime},B^{\prime}), sBs^{\prime}\cup B^{\prime} itself computes XX via Φ\Phi).

In fact, it is even possible to show that if we only use reservoirs that have positive lower density then it is possible that GG is forced to compute XX. This can be shown using the following proposition, which we will prove in Section 4.3.

{restatable*}

propositiondensesubsetsofdense For any set XX, there is some set AA\subseteq\mathbb{N} of positive lower density such that all subsets of AA of positive lower density compute XX uniformly.

Proposition 3.1.

Let AA and XX be as in the above proposition. If GG is generic for Mathias forcing with reservoirs of positive lower density and GG is compatible with (,A)(\varnothing,A) then GG computes XX.

Proof.

Let Φ\Phi be a Turing functional witnessing the fact that all subsets of AA of positive lower density compute XX uniformly (in other words, such that for all BAB\subseteq A of positive lower density, Φ(B)=X\Phi(B)=X). For each nn, we will show that the set of conditions

𝒟n={(s,B)Φ(s,n)=X(n)}\mathcal{D}_{n}=\{(s,B)\mid\Phi(s,n)\!\downarrow\;=X(n)\}

is dense below (,A)(\varnothing,A). This shows that for each nn, Φ(G,n)=X(n)\Phi(G,n)\!\downarrow\;=X(n) and hence that Φ(G)=X\Phi(G)=X.

So fix nn for which we will show that 𝒟n\mathcal{D}_{n} is dense below (,A)(\varnothing,A). Let (s,B)(s,B) be an arbitrary condition extending (,A)(\varnothing,A). We need to show that there is some condition (s,B)(s^{\prime},B^{\prime}) extending (s,B)(s,B) such that Φ(s,n)=X(n)\Phi(s^{\prime},n)\!\downarrow\;=X(n).

Since sBs\cup B is a subset of AA of positive lower density, we must have Φ(sB,n)=X(n)\Phi(s\cup B,n)\!\downarrow\;=X(n). Let ss^{\prime} be an initial segment of sBs\cup B which is long enough to witness this and set B=BsB^{\prime}=B\setminus s^{\prime}. Then (s,B)(s,B)(s^{\prime},B^{\prime})\leq(s,B) and Φ(s,n)=X(n)\Phi(s^{\prime},n)\!\downarrow\;=X(n) as desired. ∎

3.2. Density Mathias forcing

We have seen that if we want to use Mathias forcing to construct an infinite subset of AA which does not compute XX then neither allowing all infinite sets as reservoirs nor requiring the reservoirs to have positive lower density works. The problem with the former is that there are too many reservoirs available, in particular there are reservoirs whose infinite subsets all compute XX. The problem with the latter is that there are too few reservoirs available, in particular, for a fixed reservoir BB and Turing functional Φ\Phi it may not be possible to find a reservoir BBB^{\prime}\subseteq B witnessing that not all infinite subsets of BB compute XX via Φ\Phi. Thus we want a requirement that lies in-between these two extremes.

We will now describe such a requirement. In essence, we will allow a set BB to be a reservoir if it has positive upper density and the fact that it has positive upper density is witnessed by a set that does not compute XX. We will refer to Mathias forcing with reservoirs satisfying this condition as density Mathias forcing.

Definition 3.2.

A condition for density Mathias forcing is a Mathias condition (s,B)(s,B) for which there is some set DD such that

  1. (1)

    BB is dense along DD

  2. (2)

    and DD does not compute XX.

Note that (,A)(\varnothing,A) itself is a density Mathias condition: AA is dense along [min(A)]\mathbb{N}\setminus[\min(A)] and since XX is uncomputable, [min(A)]\mathbb{N}\setminus[\min(A)] does not compute XX.

3.3. The proof

We will now show that if GG is sufficiently generic for density Mathias forcing then GG does not compute XX. We will begin with a technical lemma which we will use in the proof. Then we will prove the main lemma, which shows that for a single program Φ\Phi there is a dense set of conditions which guarantee that Φ(G)X\Phi(G)\neq X.

Lemma 3.3.

If (s,B)(s,B) is a density Mathias condition and B1,,BkB_{1},\ldots,B_{k} is a finite partition of BB then for some iki\leq k, (s,Bi)(s,B_{i}) is a density Mathias condition.

Proof.

Since (s,B)(s,B) is a density Mathias condition there is some δ>0\delta>0 and some infinite set DD such that BB is δ\delta-dense along DD and DD does not compute XX. Now define subsets D1,,DkD_{1},\ldots,D_{k} of DD by

Di={nDi is least such that Bi is δ/k-dense at n}.D_{i}=\{n\in D\mid i\text{ is least such that $B_{i}$ is $\delta/k$-dense at $n$}\}.

We claim that D1,,DkD_{1},\ldots,D_{k} partition DD. To see why, let nDn\in D. Thus B[n]δ(n+1)B\cap[n]\geq\delta\cdot(n+1). Since B1,,BkB_{1},\ldots,B_{k} partition BB, at least one of the sets

B1[n],B2[n],,Bk[n]B_{1}\cap[n],B_{2}\cap[n],\ldots,B_{k}\cap[n]

must have size at least δ(n+1)/k\delta\cdot(n+1)/k. Thus there is some ii such that BiB_{i} is δ/k\delta/k-dense at nn.

Since DD does not compute XX and D1,,DkD_{1},\ldots,D_{k} partition DD, we can apply Corollary 2.2 to Seetapun’s theorem to get an infinite set EE which is contained in some DiD_{i} and which does not compute XX. It is straightforward to check that BiB_{i} is δ/k\delta/k-dense along EE and thus (s,Bi)(s,B_{i}) is a density Mathias condition. ∎

Lemma 3.4.

For any Turing functional Φ\Phi and density Mathias condition (s,B)(s,B), there is some density Mathias condition (s,B)(s,B)(s^{\prime},B^{\prime})\leq(s,B) such that for any set CC compatible with (s,B)(s^{\prime},B^{\prime}), Φ(C)X\Phi(C)\neq X.

Proof.

Since (s,B)(s,B) is a density Mathias condition, we can fix some δ>0\delta>0 and infinite set DD such that BB is δ\delta-dense along DD and DD does not compute XX. There are two cases to consider.

  1. (1)

    It is possible to make Φ\Phi wrong on some input: there is some nn\in\mathbb{N} and finite set tBt\subseteq B such that Φ(st,n)X(n)\Phi(s\cup t,n)\!\downarrow\;\neq X(n).

  2. (2)

    It is impossible to make Φ\Phi wrong on any input: for every nn\in\mathbb{N} and finite set tBt\subseteq B, Φ(st,n)\Phi(s\cup t,n) either diverges or is equal to X(n)X(n).

The first case is easy: we can simply take s=sts^{\prime}=s\cup t and B=B[max(t)]B^{\prime}=B\setminus[\max(t)]. Thus we may assume we are in the second case.

Since we are in the second case, it follows that for every CC compatible with (s,B)(s,B), Φ(C)\Phi(C) never disagrees with XX—though it may diverge on some inputs. Our goal is to find (s,B)(s^{\prime},B^{\prime}) extending (s,B)(s,B) such that for every CC compatible with (s,B)(s^{\prime},B^{\prime}), Φ(C)\Phi(C) does diverge on at least one input.

Intuition. Here’s the basic idea of the proof. By using ideas from proofs of Seetapun’s theorem, it is not too hard to find a set B0B_{0} and number n0n_{0} such that for all CC compatible with (s,B0)(s,B_{0}), Φ(C,n0)X(n0)\Phi(C,n_{0})\neq X(n_{0}). It is also not too hard to ensure B0B_{0} is δ\delta-dense along DD. We would like to take B=BB0B^{\prime}=B\cap B_{0}. The reason is that any CC compatible with (s,BB0)(s,B\cap B_{0}) is compatible with both (s,B)(s,B) and (s,B0)(s,B_{0}) and thus for such a CC, Φ(C,n0)\Phi(C,n_{0}) is neither different from X(n0)X(n_{0}) (because CC is compatible with (s,B)(s,B)) nor equal to X(n0)X(n_{0}) (because CC is compatible with (s,B0)(s,B_{0})) and thus Φ(C,n0)\Phi(C,n_{0}) must diverge.

The problem with this idea is that BB0B\cap B_{0} might not be very dense—in fact, it might even be empty. So instead, we will iterate: we will find another set B1B_{1} and number n1n_{1} which have the same properties as B0B_{0} and n0n_{0} but such that B1B_{1} is disjoint from B0B_{0}. If neither BB0B\cap B_{0} nor BB1B\cap B_{1} work then we will keep going and find B2B_{2} disjoint from both, and so on.

The fact that sets are disjoint and all fairly dense will ensure that this process cannot go on forever: it is not possible to have more than 1/δ1/\delta disjoint sets which are all δ\delta-dense. Thus we will eventually find some BiB_{i} which works. In practice, carrying out this idea involves a lot of additional technical details.

General strategy. We will find disjoint sets B0,,BkB_{0},\ldots,B_{k} and numbers n0,,nkn_{0},\ldots,n_{k} such that

  1. (1)

    for each iki\leq k and all CC compatible with (s,Bi)(s,B_{i}), Φ(C,ni)X(ni)\Phi(C,n_{i})\neq X(n_{i}) (i.e. it may diverge or it may converge to some output that is not equal to X(ni)X(n_{i}))

  2. (2)

    and (s,B(B0Bk))(s,B\cap(B_{0}\cup\ldots\cup B_{k})) is a density Mathias condition.

Before explaining how to find the BiB_{i}’s, let’s explain why this is enough to finish the proof. Since (s,B(B0Bk))(s,B\cap(B_{0}\cup\ldots\cup B_{k})) is a density Mathias condition whose reservoir is partitioned by BB0,,BBkB\cap B_{0},\ldots,B\cap B_{k}, Lemma 3.3 implies that there is some ii such that (s,BBi)(s,B\cap B_{i}) is a density Mathias condition.

We now claim that we can take s=ss^{\prime}=s and B=BBiB^{\prime}=B\cap B_{i}. To see why, note that if CC is compatible with (s,BBi)(s,B\cap B_{i}) then it is compatible with both (s,B)(s,B) and (s,Bi)(s,B_{i}). Thus Φ(C,ni)\Phi(C,n_{i}) can neither disagree with X(ni)X(n_{i}) (since CC is compatible with (s,B)(s,B)) nor agree with X(ni)X(n_{i}) (since CC is compatible with (s,Bi)(s,B_{i})) and therefore Φ(C,ni)\Phi(C,n_{i}) must diverge.

Strategy to construct B𝟎,,Bk\bm{B_{0},\ldots,B_{k}}. We will build the sequence of BiB_{i}’s inductively. To make the induction work, we will require the sets BiB_{i} to satisfy some additional properties. In particular, we will construct a sequence of sets B0,B1,B2,B_{0},B_{1},B_{2},\ldots along with sets DD0D1D2D\supseteq D_{0}\supseteq D_{1}\supseteq D_{2}\supseteq\ldots and numbers n0,n1,n2,n_{0},n_{1},n_{2},\ldots such that for each ii,

  1. (1)

    DiD_{i} is infinite.

  2. (2)

    BiB_{i} is δ/2\delta/2-dense along DiD_{i}.

  3. (3)

    B0B1BiDiB_{0}\oplus B_{1}\oplus\ldots\oplus B_{i}\oplus D_{i} does not compute XX.

  4. (4)

    BiB_{i} is disjoint from B0Bi1B_{0}\cup\ldots\cup B_{i-1}

  5. (5)

    For each CC compatible with (s,Bi)(s,B_{i}), Φ(C,ni)X(ni)\Phi(C,n_{i})\neq X(n_{i}).

We will show by induction that if we have built sequences B0,,BkB_{0},\ldots,B_{k}, D0,,DkD_{0},\ldots,D_{k} and n0,,nkn_{0},\ldots,n_{k} satisfying these requirements then either (s,B(B0Bk))(s,B\cap(B_{0}\cup\ldots\cup B_{k})) is a density Mathias condition (and thus we are finished building the sequence) or we can extend the sequence—i.e. find Bk+1,Dk+1B_{k+1},D_{k+1} and nk+1n_{k+1} satisfying the requirements. To finish, we will show that any sequence satisfying the requirements cannot be infinite.

Extending the sequence: picking Dk+𝟏\bm{D_{k+1}}. Suppose that we have built sequences B0,,BkB_{0},\ldots,B_{k}, D0,,DkD_{0},\ldots,D_{k} and n0,,nkn_{0},\ldots,n_{k} satisfying the five requirements listed above. If we knew that B(B0Bk)B\cap(B_{0}\cup\ldots\cup B_{k}) was dense along DkD_{k} then we could stop: in that case (s,B(B0Bk))(s,B\cap(B_{0}\cup\ldots\cup B_{k})) would be a density Mathias condition. More generally, if we knew that B(B0Bk)B\cap(B_{0}\cup\ldots\cup B_{k}) was dense along any infinite subset of DkD_{k} which does not compute XX then we could stop. So let’s assume that’s not the case and show we can extend the sequence B0,,BkB_{0},\ldots,B_{k}.

Let EE be the subset of DkD_{k} defined by

E={nDkB(B0Bk) is δ/2-dense at n}.E=\{n\in D_{k}\mid B\cap(B_{0}\cup\ldots\cup B_{k})\text{ is $\delta/2$-dense at }n\}.

By (the relativized form of) Corollary 2.2 to Seetapun’s theorem, there is an infinite subset Dk+1D_{k+1} of either EE or DkED_{k}\setminus E such that B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1} does not compute XX. However, Dk+1D_{k+1} cannot be a subset of EE since then B(B0Bk)B\cap(B_{0}\cup\ldots\cup B_{k}) would be dense along Dk+1D_{k+1}, which contradicts our assumption above.

Thus Dk+1DkED_{k+1}\subseteq D_{k}\setminus E. This implies B(B0Bk)B\setminus(B_{0}\cup\ldots\cup B_{k}) is δ/2\delta/2-dense along Dk+1D_{k+1}. To see why, consider any nDk+1n\in D_{k+1}. Since Dk+1DkDD_{k+1}\subseteq D_{k}\subseteq D and BB is δ\delta-dense along DD, BB is δ\delta-dense at nn. Since nEn\notin E, B(B0Bk)B\cap(B_{0}\cup\ldots\cup B_{k}) is not δ/2\delta/2-dense at nn. Therefore B(B0Bk)B\setminus(B_{0}\cup\ldots\cup B_{k}) must be δ/2\delta/2-dense at nn.

Extending the sequence: picking Bk+𝟏\bm{B_{k+1}}. Observe that B(B0Bk)B\setminus(B_{0}\cup\ldots\cup B_{k}) has the following properties.

  1. (1)

    As we just showed, it is δ/2\delta/2-dense along Dk+1D_{k+1}.

  2. (2)

    It is disjoint from B0BkB_{0}\cup\ldots\cup B_{k}.

  3. (3)

    Since it is a subset of BB, for every finite subset tt and every nn\in\mathbb{N}, Φ(st,n)\Phi(s\cup t,n) either diverges or is equal to X(n)X(n).

These properties look almost like what we want from Bk+1B_{k+1} but in the last item above we would like to have “not equal to X(n)X(n)” rather than “equal to X(n)X(n).” In other words, we want to find a set which is very similar to B(B0Bk)B\setminus(B_{0}\cup\ldots\cup B_{k}) but differs in one key respect. We will argue that if we cannot find such a set then XX is computable from B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}.

For each nn\in\mathbb{N} and b{0,1}b\in\{0,1\}, define a set 𝒞n,b𝒫(N)\mathcal{C}_{n,b}\subseteq\operatorname{\mathcal{P}}(N) by

𝒞n,b={Y\displaystyle\mathcal{C}_{n,b}=\{Y\subseteq\mathbb{N}\mid\; Y is disjoint from B0Bk\displaystyle Y\text{ is disjoint from }B_{0}\cup\ldots\cup B_{k}
and Y is δ/2-dense along Dk+1\displaystyle\text{and }Y\text{ is $\delta/2$-dense along }D_{k+1}
and for all finite tY, either Φ(st,n) or Φ(st,n)=b}.\displaystyle\text{and for all finite $t\subseteq Y$, either $\Phi(s\cup t,n)\!\uparrow$ or $\Phi(s\cup t,n)\!\downarrow\;=b$}\}.

Note that each 𝒞n,b\mathcal{C}_{n,b} is a Π10\Pi^{0}_{1} class relative to B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}. Thus for each (n,b)(n,b), there is some B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}-computable binary tree Tn,bT_{n,b} whose paths are exactly the elements of 𝒞n,b\mathcal{C}_{n,b}. Furthermore, it is easy to see that Tn,bT_{n,b} can be computed uniformly in (n,b)(n,b).

We would now like to show that for some nn and bX(n)b\neq X(n), 𝒞n,b\mathcal{C}_{n,b} is nonempty. Suppose not. Then for each nn\in\mathbb{N} and b{0,1}b\in\{0,1\}, we have the following

  1. (1)

    If X(n)=bX(n)=b then 𝒞n,b\mathcal{C}_{n,b} is nonempty (as witnessed by B(B0Bk)B\setminus(B_{0}\cup\ldots\cup B_{k})) and thus Tn,bT_{n,b} is infinite.

  2. (2)

    If X(n)bX(n)\neq b then 𝒞n,b\mathcal{C}_{n,b} is empty and so, by König’s lemma, Tn,bT_{n,b} is finite.

Therefore to compute X(n)X(n) using B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}, we simply need to check which of Tn,0T_{n,0} and Tn,1T_{n,1} is finite.

Thus there is some nn and bX(n)b\neq X(n) such that 𝒞n,b\mathcal{C}_{n,b} is nonempty. By the cone avoiding basis theorem, relativized to B0BkDk+1B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}, we can find some Bk+1𝒞n,bB_{k+1}\in\mathcal{C}_{n,b} such that

(B0BkDk+1)Bk+1TX.\left(B_{0}\oplus\ldots\oplus B_{k}\oplus D_{k+1}\right)\oplus B_{k+1}\ngeq_{T}X.

It is straightforward to check that this Bk+1B_{k+1} satisfies all the necessary requirements.

What about the base case? Since we are forming the sequences B0,B1,B_{0},B_{1},\ldots and D0,D1,D_{0},D_{1},\ldots inductively, it would seem that we need to explain the base case: how to pick B0B_{0}, D0D_{0} and n0n_{0}. However, the argument we gave in the inductive case to pick Bk+1B_{k+1} and Dk+1D_{k+1} works just as well for the base case, with the exception that instead of having to pick D0D_{0} carefully we can just take D0=DD_{0}=D. Also in this case, we should interpret the union B0BkB_{0}\cup\ldots\cup B_{k} as the empty set.

The sequence cannot be infinite. We have shown that if we have formed sets B0,,BkB_{0},\ldots,B_{k} and D0,,DkD_{0},\ldots,D_{k} satisfying the five conditions listed above then either (s,B(B0Bk))(s,B\cap(B_{0}\cup\ldots\cup B_{k})) is a density Mathias condition (in which case we are done) or we can find Bk+1B_{k+1} and Dk+1D_{k+1} extending the sequence. But how do we know the first option ever holds? In other words, how do we know we cannot just extend the sequence indefinitely? We will now show that no sequence satisfying the five requirements can be longer than 2/δ2/\delta.

Suppose for contradiction that B0,,BkB_{0},\ldots,B_{k} and D0,,DkD_{0},\ldots,D_{k} do satisfy the five requirements and k>2/δk>2/\delta. Let nn be any element of DkD_{k}. Note that for each iki\leq k, DkDiD_{k}\subseteq D_{i} and hence BiB_{i} is δ/2\delta/2-dense at nn. In other words,

|Bi[n]|δ2(n+1).|B_{i}\cap[n]|\geq\frac{\delta}{2}\cdot(n+1).

However, since the BiB_{i}’s are all disjoint, this gives us k>2/δk>2/\delta disjoint subsets of [n][n], all of size at least δ/2(n+1)\delta/2\cdot(n+1), which is impossible. ∎

We can now prove Theorem 1.1.

See 1.1

Proof.

Let GG be a generic for density Mathias forcing which is compatible with the condition (,A)(\varnothing,A). We claim that GG is an infinite subset of AA which does not compute XX. Since GG is generic, it is infinite and since GG is compatible with (,A)(\varnothing,A), GAG\subseteq A. Now let Φ\Phi be any Turing functional. By Lemma 3.4, the following set of density Mathias conditions is dense

{(s,B)for all C compatible with (s,B)Φ(C)X}\{(s,B)\mid\text{for all $C$ compatible with $(s,B)$, $\Phi(C)\neq X$}\}

and so Φ(G)X\Phi(G)\neq X. Since this holds for all Φ\Phi, GG does not compute XX. ∎

3.4. An open question

Monin and Patey have proved the following variation on Seetapun’s theorem, in which computability is replaced by hyperarithmetic reducibility [14].

Theorem 3.5 (Monin and Patey).

For any non-hyperarithmetic set XX and any set AA\subseteq\mathbb{N}, there is an infinite subset BB of either AA or A¯\overline{A} such that XX is not hyperarithmetic in BB.

It seems natural to ask whether the analogous variation on Theorem 1.1 is true. Answering this question seems to be beyond the techniques we used to prove our theorem.

Question 3.6.

Suppose XX is not hyperarithmetic and AA\subseteq\mathbb{N} has positive lower density. Must AA have an infinite subset BB such that XX is not hyperarithmetic in BB?

4. The main theorem is sharp

Our goal in proving Theorem 1.1 was to formalize the intuition that if every infinite subset of a set AA can compute an uncomputable set XX then AA must be sparse. In our theorem, we interpreted sparse to mean lower density zero. However, this is not the only reasonable interpretation of what it means for a set of natural numbers to be sparse. In this section, we will consider a couple other notions of sparsity and give counterexamples showing that for each, our main theorem becomes false.

In other words, for various notions of sparsity, we will prove that there is an uncomputable XX and a set AA\subseteq\mathbb{N} which is not sparse such that all infinite subsets of AA compute XX. In nearly all cases, we will be able to strengthen the counterexample in two ways.

First, instead of just finding a single XX that can be encoded into a non-sparse set in this way, we will show that any set XX can be so encoded—i.e. for every XX there is some non-sparse AA\subseteq\mathbb{N} such that all infinite subsets of AA compute XX.

Second, instead of every subset of AA simply computing XX, we can ensure that they all compute XX uniformly—i.e. there is a single Turing functional Φ\Phi such that for all infinite subsets BB of AA, Φ(B)=X\Phi(B)=X.

4.1. Sets of positive upper density

Our first counterexample concerns replacing positive lower density with positive upper density. In this case, we can even find a counterexample where AA has upper density one.

Proposition 4.1.

For any set XX, there is some set AA\subseteq\mathbb{N} such that ρ¯(A)=1\operatorname{\overline{\rho}}(A)=1 and all infinite subsets of AA compute XX uniformly.

Proof.

Let A~\widetilde{A} be any infinite set, all of whose infinite subsets compute XX uniformly (for example, A~\widetilde{A} could be the set produced by the Dekker-Myhill method explained in the introduction). The idea is to build AA by copying A~\widetilde{A}, but add in a lot of redundancy so that AA is occasionally very dense.

To that end, pick a computable sequence of numbers n0<n1<n2<n_{0}<n_{1}<n_{2}<\ldots which grows fast enough that ni+1>inin_{i+1}>i\cdot n_{i} and define

A=iA~[ni,ni+1).A=\bigcup_{i\in\widetilde{A}}[n_{i},n_{i+1}).

In other words, for each iA~i\in\widetilde{A}, AA contains the entire interval [ni,ni+1)[n_{i},n_{i+1}).

First observe that AA has upper density 11. Indeed, for each iA~i\in\widetilde{A}, AA is i1i\frac{i-1}{i}-dense at ni+1n_{i+1}.

Second, observe that all infinite subsets of AA compute XX uniformly: given an infinite subset BAB\subseteq A, we can uniformly compute the following infinite subset B~\widetilde{B} of A~\widetilde{A},

B~={in[ni,ni+1)A},\widetilde{B}=\{i\mid\exists n\in[n_{i},n_{i+1})\cap A\},

and then use B~\widetilde{B} to compute XX. ∎

In spite of this counterexample, our main theorem can be strengthened to hold for certain sets of positive upper density.

Proposition 4.2.

For any uncomputable XX and set AA\subseteq\mathbb{N} such that AA is dense along some infinite set DD which does not compute XX, there is some infinite subset of AA which does not compute XX.

Proof.

Note that (,A)(\varnothing,A) is a density Mathias condition, as witnessed by DD. Thus there is a generic GG for density Mathias forcing which is compatible with (,A)(\varnothing,A). Exactly as in the proof of Theorem 1.1, GG is an infinite subset of AA which does not compute XX. ∎

4.2. Sets whose density goes to zero slowly

Our next counterexample concerns sets of lower density zero where the density approaches 0 very slowly. Define the density function of a set AA to be the function

dA(n)=|A[n]|n+1.d_{A}(n)=\frac{|A\cap[n]|}{n+1}.

Note that if a set AA has positive lower density then dA(n)d_{A}(n) is bounded away from 0. By contrast, most of the methods we have seen so far of coding a set XX into all infinite subsets of a set produce sets whose density functions converge to 0 very rapidly. For example, the Dekker-Myhill method we mentioned in the introduction produces a set AA such that dA(n)d_{A}(n) is approximately log(n)n\frac{\log(n)}{n}. In the previous subsection, we saw a method for which this is not quite true—in particular, we saw a method which produces a set AA of upper density one, which implies that dA(n)d_{A}(n) is close to 11 infinitely often. However, even for this method, dA(n)d_{A}(n) is also infinitely often smaller than log(n)n\frac{\log(n)}{n}.

Based on these examples, one might guess that if all infinite subsets of a set AA compute an uncomputable set XX then dA(n)d_{A}(n) cannot be lower bounded by any monotone function that goes to zero much slower than log(n)n\frac{\log(n)}{n}—in other words that there must be infinitely many places where the density of AA is exponentially small. However, this is false: we will show that it is possible to encode any set XX into all infinite subsets of a set AA whose density function goes to zero arbitrarily slowly.

Definition 4.3.

For any function f:[0,1]f\colon\mathbb{N}\to[0,1], a set AA\subseteq\mathbb{N} is ff-dense if for all nn,

dA(n)f(n).d_{A}(n)\geq f(n).
Proposition 4.4.

Suppose f:[0,1]f\colon\mathbb{N}\to[0,1] is a function such that

limnf(n)=0.\lim_{n\to\infty}f(n)=0.

Then for any set XX there is some AA such that AA is ff-dense and all infinite subsets of AA compute XX uniformly.

Proof.

The idea is to modify the Dekker-Myhill coding method explained in the introduction, but slow it down so that it produces an ff-dense set. In Dekker and Myhill’s scheme, each element of AA encodes one more bit of XX than the previous element. In our modified version, we will repeatedly encode the same number of bits of XX until ff becomes small enough to allow us to encode more. Roughly speaking, we will only start using elements of AA to encode the first nn bits of XX once ff drops below 1/2n1/2^{n}.

One comment before we go into the details of the construction: it might seem necessary to require ff to be computable but there is a trick that allows us to avoid requiring that. Essentially, we can use elements of AA to encode not just some bits of XX but also to encode how many bits are encoded.

To define AA formally, first pick a sequence 0<n1<n2<0<n_{1}<n_{2}<\ldots which grows fast enough that for all ii:

  1. (1)

    for all mnim\geq n_{i}, f(m)1/(522i+1)f(m)\leq 1/(5\cdot 2^{2i+1})

  2. (2)

    and ni+1nin_{i+1}-n_{i} is divisible by 22i+12^{2i+1}.

The idea is that in the interval [ni,ni+1)[n_{i},n_{i+1}), each element of AA will encode the first ii bits of XX. Note that the sequence n0,n1,n2,n_{0},n_{1},n_{2},\ldots is not required to be computable.

Next, for each ii, let σi\sigma_{i} denote the length 2i+12i+1 binary string consisting of the first ii bits of XX, followed by a 11, followed by a ii zeros—i.e.

σi=(Xi)1000i times.\sigma_{i}=(X\operatorname{\upharpoonright}i)^{\smallfrown}1^{\smallfrown}\underbrace{00\ldots 0}_{i\text{ times}}.

Now define AA as follows.

  1. (1)

    First put [0,n1)[0,n_{1}) into AA.

  2. (2)

    Next, for each i1i\geq 1 and m[ni,ni+1)m\in[n_{i},n_{i+1}), put mm into AA if the binary expansion of mm ends with the string σi\sigma_{i}.

We will now show that AA has the properties desired.

Claim.

AA is ff-dense.

Proof.

Consider a single interval [ni,ni+1)[n_{i},n_{i+1}). Note that this interval is composed of some number of disjoint intervals of length 22i+12^{2i+1} and that each one of these smaller intervals contains exactly one element of AA. Also note that AA contains all numbers less than n1n_{1}. From these two observations, it is easy to see that for each ii, AA is 1/22i+11/2^{2i+1}-dense at ni+1n_{i+1}. In other words, AA is sufficiently dense at the endpoints of each interval [ni,ni+1)[n_{i},n_{i+1}). It remains to check that AA is also sufficiently dense in the interior of these intervals.

Note that any number m(ni,ni+1)m\in(n_{i},n_{i+1}) can be written as m=ni+k22i+1+lm=n_{i}+k\cdot 2^{2i+1}+l for some kk and l<22i+1l<2^{2i+1}. Our observations above imply that AA is 1/22i+11/2^{2i+1}-dense at ni+k22i+1n_{i}+k\cdot 2^{2i+1}. We now want to show that f(m)f(m) and ll are small enough that AA is still sufficiently dense at mm.

To see why, note that by assumption, ni>22i1n_{i}>2^{2i-1} and thus l<4(ni+k22i+1)l<4\cdot(n_{i}+k\cdot 2^{2i+1}). Thus, since AA is 1/22i+11/2^{2i+1}-dense at ni+k22i+1n_{i}+k\cdot 2^{2i+1}, it is also 1/(522i+1)1/(5\cdot 2^{2i+1})-dense at mm. By our choice of nin_{i}, this implies that AA is f(m)f(m)-dense at mm. ∎

Claim.

Every infinite subset of AA computes XX uniformly.

Proof.

Suppose BB is an infinite subset of AA. For any element mBm\in B such that mn1m\geq n_{1}, the binary expansion of mm must end with some number of zeros. Suppose it ends with kk zeros. Then the binary expansion of mm must end with a string σ\sigma of length kk, followed by a one, followed by kk zeros. By construction of AA, this string σ\sigma is guaranteed to be an initial segment of XX. So to compute any bit X(i)X(i) of XX, we simply need to find an element mn1m\geq n_{1} of BB whose binary expansion ends with a long enough string of zeros. And such an element is guaranteed to exist by our construction of AA and the fact that BB is infinite. ∎

This concludes the proof. ∎

4.3. Dense subsets of dense sets

Our final counterexample is of a somewhat different nature than the first two. In our previous counterexamples, we considered possible strengthenings of our main theorem given by relaxing the density requirement on AA. We will now consider possible strengthenings given by restricting which subsets of AA we are allowed to use.

The main theorem of this paper shows that it is impossible to encode an uncomputable set XX into all infinite subsets of a set AA of positive lower density. But what if we only want to encode XX into all subsets of AA of positive lower density? Or into all subsets of positive upper density? We will show that such encoding is possible in both cases, but that it can only be done uniformly in the first case.

\densesubsetsofdense
Proof.

Suppose we only had to consider 1/21/2-dense subsets of AA. For each nn, any such subset must contain at least one element of AA between nn and 2n2n. Thus we could code one bit of XX into the parity of the elements of AA between nn and 2n2n and be sure that any 1/21/2-dense subset of AA will be able to recover this bit. However, we want our procedure to work for subsets of AA of any constant density greater than zero, not just 1/21/2-dense subsets. So we will encode each bit of XX infinitely many times, each time making a more conservative assumption about the density of the subset of AA doing the decoding.

More formally, pick a computable bijection π:×+\pi\colon\mathbb{N}\to\mathbb{N}\times\mathbb{Q}^{+} and a computable sequence 0=n0<n1<n2<0=n_{0}<n_{1}<n_{2}<\ldots such that for all ii,

  1. (1)

    if π(i)=(j,δ)\pi(i)=(j,\delta) then δni+1>ni\delta\cdot n_{i+1}>n_{i}

  2. (2)

    and nin_{i} is even.

The idea is that in the interval [ni,ni+1)[n_{i},n_{i+1}), elements of AA will code the jthj^{\text{th}} bit of XX in such a way that any δ\delta-dense subset of AA will be able to decode it.

Now define AA as follows. For each ii\in\mathbb{N} and m[ni,ni+1)m\in[n_{i},n_{i+1}), let (j,δ)=π(i)(j,\delta)=\pi(i) and put mm into AA if

{m is even and X(j)=0or m is odd and X(j)=1.\begin{cases}m\text{ is even and }X(j)=0\\ \text{or }m\text{ is odd and }X(j)=1.\end{cases}

In other words, in the interval [ni,ni+1)[n_{i},n_{i+1}), the jthj^{\text{th}} bit of XX is encoded into the parity of the elements of AA.

First observe that AA has lower density 1/21/2: for each mm, either 2m2m or 2m+12m+1 is in AA (this is why we wanted to make sure each nin_{i} is even).

Second, observe that all subsets of AA of positive lower density can compute XX uniformly. Given a subset BAB\subseteq A of positive lower density and a jj\in\mathbb{N}, we can compute X(j)X(j) as follows: search for any ii such that π(i)=(j,δ)\pi(i)=(j,\delta) for some δ\delta and [ni,ni+1)B[n_{i},n_{i+1})\cap B is nonempty. Then from the parity of any element of [ni,ni+1)B[n_{i},n_{i+1})\cap B we can recover X(j)X(j). The fact that such an ii must exist follows from our choice of nin_{i}’s and the fact that BB has positive lower density—if BB is δ\delta-dense and π(i)=(j,δ)\pi(i)=(j,\delta) then BB must have at least one element in the interval [ni,ni+1)[n_{i},n_{i+1}) because otherwise BB cannot be δ\delta-dense at ni+1n_{i+1}. ∎

For the case of subsets of positive upper density, the counterexample can be constructed using a result of Bienvenu, Day and Hölzl [1].

Theorem 4.5 (Bienvenu, Day and Hölzl).

For any set XX, there is a set AA such that for all partial functions f:{0,1}f\colon\mathbb{N}\to\{0,1\}, if the domain of ff has positive upper density and for all ndom(f)n\in\operatorname{dom}(f), f(n)=A(n)f(n)=A(n) then ff computes XX.

Corollary 4.6.

For any set XX, there is a set AA of positive lower density such that all subsets of AA of positive upper density compute XX.

Proof.

Let AA be the set from Bienvenu, Day and Hölzl’s theorem. Let A~={2nnA}{2n+1nA}\widetilde{A}=\{2n\mid n\in A\}\cup\{2n+1\mid n\notin A\}. Note that A~\widetilde{A} has lower density 1/21/2 and that from each element of A~\widetilde{A}, we can recover one bit of the characteristic function of AA.

Thus any subset BB of A~\widetilde{A} can compute a partial function fB:{0,1}f_{B}\colon\mathbb{N}\to\{0,1\} which agrees with AA everywhere it is defined. Furthermore, the density of dom(fB)\operatorname{dom}(f_{B}) at any nn is about twice the density of BB at 2n+22n+2. Therefore if BB has positive upper density then so does dom(fB)\operatorname{dom}(f_{B}). So for any such BB, our choice of AA implies that fBf_{B}, and hence BB itself, computes XX. ∎

As we noted above, it is impossible to strengthen this corollary to make all subsets of AA of positive upper density compute XX uniformly.

Proposition 4.7.

For any uncomputable set XX, set AA of positive lower density and Turing functional Φ\Phi, there is some subset BAB\subseteq A of positive upper density such that Φ(B)X\Phi(B)\neq X.

Proof.

This follows directly from Lemma 3.4. In particular, using that lemma we can find a density Mathias condition (s,B)(s,B) extending (,A)(\varnothing,A) such that for every set CC compatible with (s,B)(s,B), Φ(C)X\Phi(C)\neq X. In particular, if C=sBC=s\cup B then Φ(C)X\Phi(C)\neq X. Since (s,B)(s,B) is a density Mathias condition, CC has positive upper density and so we are done. ∎

5. Avoiding PA degree

In this section we prove the theorem promised in the introduction which modifies our main theorem to avoid PA degree rather than cones of Turing degrees.

See 1.4

In order to prove this theorem, it will be convenient to use the following alternative definition of PA degree.

Definition 5.1.

Given a partial function f:{0,1}f\colon\mathbb{N}\to\{0,1\}, a completion of ff is a total function g:{0,1}g\colon\mathbb{N}\to\{0,1\} such that for every nn in the domain of ff, f(n)=g(n)f(n)=g(n).

Definition 5.2.

A set AA is said to be of PA degree if for every partial computable function f:{0,1}f\colon\mathbb{N}\to\{0,1\}, AA computes a completion of ff.

We can thus obtain Theorem 1.4 as a corollary of the following theorem.

Theorem 5.3.

Suppose f:{0,1}f\colon\mathbb{N}\to\{0,1\} is a computable partial function with no computable completion. Then every set AA\subseteq\mathbb{N} of positive lower density contains an infinite subset which does not compute a completion of ff.

We show in Theorem 5.11 that the assumption that ff is computable is neccesary.

Our proof of this theorem follows the same general strategy as the proof of our main theorem, with a few notable changes. First, we replace Seetapun’s theorem with Liu’s theorem (as discussed in the introduction) and also make some more-or-less standard changes to the proof which are relevant for avoiding PA degrees (these changes are mostly taken from the work of Liu [10] and of Monin and Patey [13]). Second, and more interestingly, we can no longer rely on the cone avoiding basis theorem (the point is that there is no such thing as a “PA degree avoiding basis theorem” for somewhat obvious reasons). Thus we are forced to replace the argument using the cone avoiding basis theorem with a somewhat more elaborate argument. Third, we need an extra combinatorial fact about intersections of dense sets. This fact is not hard to prove, but it is not as simple as the combinatorics which appeared in the proof of our main theorem.

5.1. Liu’s theorem and Liu’s lemma

To prove the theorem above, we will use two results due to Liu. First, Liu’s theorem, which, as we discussed in the introduction, is the analogue of Seetapun’s theorem for this setting (and will play an analogous role in the proof). We stated a version of this theorem as Theorem 1.3, but we will use the following more general version.

Theorem 5.4 (Liu’s Theorem, [10]).

Suppose f:{0,1}f\colon\mathbb{N}\to\{0,1\} is a computable partial function with no computable completion. Then for every set AA\subseteq\mathbb{N}, there is an infinite set BB which is a subset of either AA or the complement of AA such that BB does not compute a completion of ff.

As with Seetapun’s theorem, it is easy to use this theorem to prove the following corollary.

Corollary 5.5.

Suppose f:{0,1}f\colon\mathbb{N}\to\{0,1\} is a computable partial function and AA\subseteq\mathbb{N} is an infinite set which does not compute a completion of ff. Then for every finite partition B1,,BnB_{1},\ldots,B_{n} of AA, at least one of the BiB_{i}’s has an infinite subset which does not compute a completion of ff.

In the course of proving Theorem 1.3, Liu implicitly used the following lemma (see Lemma 6.6 of [10]). It was also used in a slightly different form by Monin and Patey in [13] (see Lemmas 2.13 and 3.12).

Definition 5.6.

A valuation is a finite partial function p:{0,1}p\colon\mathbb{N}\to\{0,1\} represented as a lookup table (rather than, say, a code for a Turing machine which computes pp).

Definition 5.7.

Given a valuation pp and a partial function f:{0,1}f\colon\mathbb{N}\to\{0,1\}, pp is ff-correct if pfp\subseteq f—in other words if for all nn in the domain of pp, nn is also in the domain of ff and p(n)=f(n)p(n)=f(n).

Lemma 5.8 (Liu [10]).

If WW is a c.e. set of valuations and ff is a computable partial function with no total completion then either WW contains some ff-correct valuation or for every kk, there are at least kk many incompatible valuations outside of WW.

Proof.

Suppose WW does not contain any ff-correct valuations and fix a number kk. For any finite set ss\subseteq\mathbb{N} and valuation pp, say that pp is ff-correct mod ss if pp is ff-correct when we ignore the elements of ss—in other words, for all ndom(p)sn\in\operatorname{dom}(p)\setminus s, nn is in the domain of ff and p(n)=f(n)p(n)=f(n). We will show that there are numbers n1,,nkn_{1},\ldots,n_{k} such that no element of WW is ff-correct mod {n1,,nk}\{n_{1},\ldots,n_{k}\}. In particular, this implies that WW does not contain any element with domain contained in {n1,,nk}\{n_{1},\ldots,n_{k}\}. Since it is easy to see that there are at least kk such incompatible functions (in fact, at least 2k2^{k}), this is sufficient to prove the lemma.

We will construct n1,,nkn_{1},\ldots,n_{k} by induction. Suppose we have constructed n1,,nin_{1},\ldots,n_{i} and want to find ni+1n_{i+1}. In other words, we know WW contains no element which is ff-correct mod {n1,,ni}\{n_{1},\ldots,n_{i}\} and we want to find some n{n1,,ni}n\notin\{n_{1},\ldots,n_{i}\} such that no element is ff-correct mod {n1,,ni,n}\{n_{1},\ldots,n_{i},n\}.

Suppose for contradiction that no such nn exists. Then for every n{n1,,ni}n\notin\{n_{1},\ldots,n_{i}\}, WW contains some element pp which is ff-correct mod {n1,,ni,n}\{n_{1},\ldots,n_{i},n\}. At the same time, pp cannot be ff-correct mod {n1,,ni}\{n_{1},\ldots,n_{i}\}. By unrolling the definitions, it is possible to see that this implies that nn is in the domain of pp and either not in the domain of ff or p(n)f(n)p(n)\neq f(n). The key point is that in either case, 1p(n)1-p(n) does not disagree with f(n)f(n).

But this gives us a method to compute a completion gg of ff: For each n{n1,,ni}n\in\{n_{1},\ldots,n_{i}\}, hard-code some appropriate value for g(n)g(n). For each n{n1,,ni}n\notin\{n_{1},\ldots,n_{i}\}, search for some pWp\in W which is ff-correct mod {n1,,ni,n}\{n_{1},\ldots,n_{i},n\} (note that there is a computable enumeration of ff-correct valuations). For the first such pp that is found, set g(n)=1p(n)g(n)=1-p(n). Since ff has no computable completions, this gives us the desired contradiction. ∎

5.2. A combinatorial lemma

As we referred to above, we will need the following combinatorial lemma, which says that when you have enough dense sets, two of them must have fairly dense intersection. The proof is a standard counting argument, which we phrase as a probabilistic proof based on estimating the variance of a certain random variable.

Lemma 5.9.

For every δ>0\delta>0, there is some kk with the following property: For any n>0n>0 and any kk subsets A1,,Ak[0,n)A_{1},\ldots,A_{k}\subseteq[0,n), all of which have size at least δn\delta n, there is some pair of indices iji\neq j such that AiAjnδ2/2\frac{A_{i}\cap A_{j}}{n}\geq\delta^{2}/2.

Proof.

We will begin by thinking of kk as a variable and fixing a number nn and sets A1,,Ak[0,n)A_{1},\ldots,A_{k}\subseteq[0,n). We will then find a lower bound on the maximum of AiAjn\frac{A_{i}\cap A_{j}}{n} over all iji\neq j. At the end, we will check that if kk is large enough then this lower bound is at least δ2/2\delta^{2}/2 (and does not depend on nn).

So: fix nn and sets A1,,AkA_{1},\ldots,A_{k}. Consider choosing xx from [0,n)[0,n) uniformly at random and counting the number of sets AiA_{i} which contain xx. Let XX be the random variable denoting this count. For each ii, let 1Ai1_{A_{i}} be the random variable indicating whether xx is in AiA_{i} or not (1Ai=11_{A_{i}}=1 if xAix\in A_{i} and 0 otherwise). Note that we have

X=ik1Ai.X=\sum_{i\leq k}1_{A_{i}}.

The key to the proof is to estimate the variance of XX. First, by a standard calculation we have

Var[X]\displaystyle\operatorname{Var}[X] =E[X2]E[X]2\displaystyle=\operatorname{E}[X^{2}]-\operatorname{E}[X]^{2}
=E[(ik1Ai)2]E[ik1Ai]2\displaystyle=\operatorname{E}\Bigl{[}\bigl{(}\sum_{i\leq k}1_{A_{i}}\bigr{)}^{2}\Bigr{]}-E\Bigl{[}\sum_{i\leq k}1_{A_{i}}\Bigr{]}^{2}
=i,jkE[1Ai1Aj](ikE[1Ai])2\displaystyle=\sum_{i,j\leq k}\operatorname{E}\left[1_{A_{i}}1_{A_{j}}\right]-\bigl{(}\sum_{i\leq k}\operatorname{E}[1_{A_{i}}]\bigr{)}^{2}
=i,j|AiAj|n(i|Ai|n)2\displaystyle=\sum_{i,j}\frac{|A_{i}\cap A_{j}|}{n}-\bigl{(}\sum_{i}\frac{|A_{i}|}{n}\bigr{)}^{2}
=ij|AiAj|n+i|Ai|n(i|Ai|n)2.\displaystyle=\sum_{i\neq j}\frac{|A_{i}\cap A_{j}|}{n}+\sum_{i}\frac{|A_{i}|}{n}-\bigl{(}\sum_{i}\frac{|A_{i}|}{n}\bigr{)}^{2}.

Using the fact that Var[X]0\operatorname{Var}[X]\geq 0 and letting y=i|Ai|ny=\sum_{i}\frac{|A_{i}|}{n}, we have

ij|AiAj|ny2y.\sum_{i\neq j}\frac{|A_{i}\cap A_{j}|}{n}\geq y^{2}-y.

Since each AiA_{i} has size at least δn\delta n, yδky\geq\delta k. Let’s assume k>1/δk>1/\delta and thus y>1y>1. Since y2yy^{2}-y is monotonic in yy for y>1y>1, we have

ij|AiAj|ny2y(δk)2δk.\sum_{i\neq j}\frac{|A_{i}\cap A_{j}|}{n}\geq y^{2}-y\geq(\delta k)^{2}-\delta k.

We can now calculate a lower bound ε\varepsilon on the maximum value of |AiAj|n\frac{|A_{i}\cap A_{j}|}{n} over all iji\neq j. For we have

ij|AiAj|nεk(k1)εk2\sum_{i\neq j}\frac{|A_{i}\cap A_{j}|}{n}\leq\varepsilon k(k-1)\leq\varepsilon k^{2}

and thus

εk2(δk)2δk.\varepsilon k^{2}\geq(\delta k)^{2}-\delta k.

Simple algebraic manipulation then yields εδ2δ/k\varepsilon\geq\delta^{2}-\delta/k.

We are now in position to finish the proof. Recall that we wanted kk large enough that εδ2/2\varepsilon\geq\delta^{2}/2. The lower bound above shows that it’s enough to take k2/δk\geq 2/\delta. Since this value does not depend on nn, we are done. ∎

5.3. The proof

For the rest of this section, fix a computable partial function ff with no computable completion. Define a modified version of density Mathias forcing where conditions are Mathias conditions (s,A)(s,A) such that there is some infinite set DD which does not compute a completion of ff and so that AA is dense along DD. It suffices to prove the following key lemma; the rest of the proof can be completed as in the proof of Theorem 1.1.

Lemma 5.10.

For any condition (s,A)(s,A) and Turing functional Φ\Phi, there is some condition (s,A)(s,A)(s^{\prime},A^{\prime})\leq(s,A) such that for all sets BB compatible with (s,A)(s^{\prime},A^{\prime}), Φ(B)\Phi(B) is not a completion of ff.

Proof.

Fix a set DD and number δ>0\delta>0 such that AA is δ\delta-dense along DD and DD does not compute a completion of ff.

Given a set BB and a valuation pp, say that pp is consistent with BB if for all finite sets tBt\subseteq B, Φ(st)\Phi(s\cup t) and pp are compatible (i.e. for every nn either at least one of Φ(st,n)\Phi(s\cup t,n) and p(n)p(n) is not defined or they are both defined and are equal). Say that a sequence of sets B1,,BlB_{1},\ldots,B_{l} is lawful if it satisfies the following criteria:

  1. (1)

    The BiB_{i}’s are all disjoint.

  2. (2)

    For each BiB_{i}, there are incompatible valuations pp and qq such that BiB_{i} is consistent with both pp and qq. Note that this implies that for all CC compatible with (s,Bi)(s,B_{i}), Φ(C)\Phi(C) must be compatible with both pp and qq and hence cannot be total.

  3. (3)

    There is an infinite set EE such that each BiB_{i} is δ2/8\delta^{2}/8-dense along EE, A(B1Bl)A\setminus(B_{1}\cup\ldots\cup B_{l}) is δ/2\delta/2-dense along EE and EE does not compute a completion of ff.222For avoiding a cone in our main theorem, we were also able to have that B1BlEB_{1}\oplus\cdots\oplus B_{l}\oplus E avoids the cone. This was because of the cone avoiding basis theorem. Here, we are not able to ask that B1BlEB_{1}\oplus\cdots\oplus B_{l}\oplus E does not compute a completion of ff. Before, we used B1,,BlB_{1},\ldots,B_{l} in the definition of 𝒞n,b\mathcal{C}_{n,b}, so that it was a Π10\Pi^{0}_{1} class relative to B1BkEB_{1}\oplus\cdots\oplus B_{k}\oplus E. Now, when defining the analogous class, it must be a Π10\Pi^{0}_{1} class relative only to EE alone. One consequence of this is that while previously we constructed our sequences B1,,BlB_{1},\ldots,B_{l} by extension, here, the sequence of length l+1l+1 will not necessarily extend the sequence of length ll.

We will prove that if there is a lawful sequence of length ll then either there is a condition (s,A)(s^{\prime},A^{\prime}) with the property we want or there is a lawful sequence of length l+1l+1. Also, since the BiB_{i} are disjoint and δ2/8\delta^{2}/8-dense along EE, there is no lawful sequence of length strictly greater than 8/δ28/\delta^{2}, which suffices to finish the proof.

Finding lawful sequences. Suppose that B1,,BlB_{1},\ldots,B_{l} is a lawful sequence, as witnessed by EE and (p1,q1),(p2,q2),,(pl,ql)(p_{1},q_{1}),(p_{2},q_{2}),\ldots,(p_{l},q_{l}). We want to either find (s,A)(s,A)(s^{\prime},A^{\prime})\leq(s,A) satisfying the conclusion of the lemma or a lawful sequence of length l+1l+1.333A slightly subtle point here is that we allow l=0l=0—that is, we allow the starting sequence to be empty. As in the proof of our main theorem, it is possible to check that the whole proof still works even in this case.

For every valuation rr, define 𝒞r\mathcal{C}_{r} to be the set of tuples (X1,,Xl,Y)(X_{1},\ldots,X_{l},Y) such that

  1. (1)

    X1,,Xl,YX_{1},\ldots,X_{l},Y are all disjoint.

  2. (2)

    Each XiX_{i} is δ2/8\delta^{2}/8-dense along EE.

  3. (3)

    YY is δ/2\delta/2-dense along EE.

  4. (4)

    For each ii, XiX_{i} is consistent with both pip_{i} and qiq_{i}.

  5. (5)

    YY is consistent with rr.

Note that the sets 𝒞r\mathcal{C}_{r} are uniformly Π10(E)\Pi^{0}_{1}(E). Let WW be the set of valuations rr such that 𝒞r\mathcal{C}_{r} is empty. Note that WW is c.e. relative to EE. Since EE does not compute a completion of ff, we can apply Liu’s Lemma relative to EE to get that one of the following holds.

  1. (1)

    There is some rWr\in W which is ff-correct.

  2. (2)

    For every kk, there are at least kk incompatible valuations outside of WW.

Case 1: an ff-correct valuation in WW. Suppose that some rWr\in W is ff-correct. Thus 𝒞r\mathcal{C}_{r} is empty. In particular, (B1,,Bl,A(B1Bl))(B_{1},\ldots,B_{l},A\setminus(B_{1}\cup\ldots\cup B_{l})) is not in 𝒞r\mathcal{C}_{r}. By our choice of B1,,BlB_{1},\ldots,B_{l} and EE, this can only be because rr is not consistent with A(B1Bl)A\setminus(B_{1}\cup\ldots\cup B_{l}). In other words, there is some finite set tAt\subset A and some nn such that Φ(st,n)\Phi(s\cup t,n) and r(n)r(n) are both defined and not equal. Moreover, since rr is ff-correct, this implies that Φ(st,n)\Phi(s\cup t,n) and f(n)f(n) are both defined and not equal. In this case we are done because we can set (s,A)=(st,A[max(st)])(s^{\prime},A^{\prime})=(s\cup t,A\setminus[\max(s\cup t)]).

Case 2: many incompatible valuations outside WW. Suppose that there are kk many incompatible valuations r1,,rkr_{1},\ldots,r_{k} outside WW, where kk is chosen as in the combinatorial lemma (but with δ/2\delta/2 rather than δ\delta). Thus we can find sequences

(B11,,\displaystyle(B_{1}^{1},\ldots, Bl1,C1)𝒞r1\displaystyle B_{l}^{1},C_{1})\in\mathcal{C}_{r_{1}}
(B12,,\displaystyle(B_{1}^{2},\ldots, Bl2,C2)𝒞r2\displaystyle B_{l}^{2},C_{2})\in\mathcal{C}_{r_{2}}
\displaystyle\vdots
(B1k,,\displaystyle(B_{1}^{k},\ldots, Blk,Ck)𝒞rk.\displaystyle B_{l}^{k},C_{k})\in\mathcal{C}_{r_{k}}.

Since each CiC_{i} is δ/2\delta/2-dense along EE, it follows from the combinatorial lemma (and our choice of kk) that for each nEn\in E, there are iji\neq j such that CiCjC_{i}\cap C_{j} is δ2/8\delta^{2}/8-dense at nn. We can now apply the same trick we used in the proof of our main theorem (but using Liu’s theorem instead of Seetapun’s theorem) to get an infinite set E1EE_{1}\subseteq E and a pair iji\neq j such that CiCjC_{i}\cap C_{j} is δ2/8\delta^{2}/8-dense along E1E_{1} and E1E_{1} does not compute a completion of ff.

Now let B1,B2,,BlB_{1}^{\prime},B_{2}^{\prime},\ldots,B_{l}^{\prime} denote B1i,B2i,,BliB_{1}^{i},B_{2}^{i},\ldots,B_{l}^{i} and let Bl+1B_{l+1}^{\prime} denote CiCjC_{i}\cap C_{j}. We have the following facts about the sequence B1,,Bl+1B_{1}^{\prime},\ldots,B_{l+1}^{\prime}:

  1. (1)

    They are all disjoint. This follows from the facts that B1i,,BliB_{1}^{i},\ldots,B_{l}^{i} are all disjoint from each other and from CiC_{i} and that Bl+1CiB_{l+1}^{\prime}\subseteq C_{i}.

  2. (2)

    For each BhB_{h}^{\prime}, there are incompatible valuations pp and qq such that BhB_{h}^{\prime} is consistent with both pp and qq. For B1,,BlB_{1}^{\prime},\ldots,B_{l}^{\prime} we can just take php_{h} and qhq_{h}. For Bl+1B_{l+1}^{\prime} we can take rir_{i} and rjr_{j}.

  3. (3)

    E1E_{1} is infinite, does not compute a completion of ff and each BhB_{h}^{\prime} is δ2/8\delta^{2}/8-dense along E1E_{1}. For B1,,BlB_{1}^{\prime},\ldots,B_{l}^{\prime}, this last part is because E1EE_{1}\subseteq E. For Bl+1B_{l+1}^{\prime}, this follows from our choice of E1E_{1}.

The point is that B1,,Bl+1B_{1}^{\prime},\ldots,B_{l+1}^{\prime} is very close to a lawful sequence of length l+1l+1. All that’s missing is that A(B1Bl+1)A\setminus(B_{1}^{\prime}\cup\ldots\cup B_{l+1}^{\prime}) be δ/2\delta/2-dense along E1E_{1}. However, this is easily fixed.

Again using the same trick as in the proof of our main theorem, we can get an infinite set E2E1E_{2}\subseteq E_{1} which does not compute a completion of ff and such that either

  • A(B1Bl+1)A\cap(B_{1}^{\prime}\cup\ldots\cup B_{l+1}^{\prime}) is δ/2\delta/2-dense along E2E_{2}

  • or A(B1Bl+1)A\setminus(B_{1}^{\prime}\cup\ldots\cup B_{l+1}^{\prime}) is δ/2\delta/2-dense along E2E_{2}.

In the first case, we can use the trick once more to find some infinite E3E2E_{3}\subseteq E_{2} which does not compute a completion of ff and some hl+1h\leq l+1 such that ABhA\cap B_{h}^{\prime} is dense along E3E_{3} and so we can finish by taking (s,A)=(s,ABh)(s^{\prime},A^{\prime})=(s,A\cap B_{h}^{\prime}). In the latter case, we have found a lawful sequence of length l+1l+1. ∎

5.4. A false generalization

There is a common generalization of both Theorem 1.1 and Theorem 1.4 which at one point the authors thought might be true. Namely, that given any partial function f:{0,1}f\colon\mathbb{N}\to\{0,1\} with no computable completion, and any set AA\subseteq\mathbb{N} of positive lower density, there is an infinite subset of AA that does not compute any completion of ff. (Theorem 1.1 is the case when ff is already total, and Theorem 1.4 is the case when ff is computable.) However, this is false.

Theorem 5.11.

There is a partial function f:{0,1}f\colon\mathbb{N}\to\{0,1\} and a set AA\subseteq\mathbb{N} of positive lower density such that any infinite subset BAB\subseteq A computes a completion of ff.

Proof.

We will define a partial function ff while simulateneously defining x0,x1,x2,x_{0},x_{1},x_{2},\ldots. We will make use of a computable function π:\pi\colon\mathbb{N}\to\mathbb{N} such that π1(n)\pi^{-1}(n) is infinite for each nn. For each xix_{i}, we will have π(xi)=i\pi(x_{i})=i. The idea is that while the construction will be non-computable, the function π(x)\pi(x) will allow us to computably recover from xx the function Φπ(x)\Phi_{\pi(x)} that xx would have been used to diagonalize against if xx was used to diagonalize at all.

Begin with x0=0x_{0}=0. Suppose that we have defined xix_{i}. We have two cases:

  1. (1)

    If Φi(xi)\Phi_{i}(x_{i})\downarrow, let sis_{i} be the number of stages it takes to converge, and define f(xi)=1Φi(xi)f(x_{i})=1-\Phi_{i}(x_{i}). Let xi+1>six_{i+1}>s_{i} be such that π(xi+1)=i+1\pi(x_{i+1})=i+1.

  2. (2)

    Otherwise, if Φi(xi)\Phi_{i}(x_{i})\uparrow, leave f(xi)f(x_{i}) undefined, and let xi+1>xix_{i+1}>x_{i} be such that π(xi+1)=i+1\pi(x_{i+1})=i+1.

For any xx outside of {x0,x1,}\{x_{0},x_{1},\ldots\}, f(x)f(x) is undefined.

Now define A2<ωA\subseteq 2^{<\omega} as follows. Put σA\sigma\in A if whenever Φi(xi)\Phi_{i}(x_{i})\downarrow but Φi,|σ|(xi)\Phi_{i,|\sigma|}(x_{i})\uparrow, σ(xi)=f(xi)\sigma(x_{i})=f(x_{i}). We claim that AA has lower density at least 1/41/4. For this, it is important that in identifying 2<ω2^{<\omega} with \mathbb{N}, we list out the finite binary strings in order of increasing length. Thus it suffices to show that for each nn, there is some x<nx<n and t{0,1}t\in\{0,1\} such that A2n{σ2nσ(x)=t}A\cap 2^{n}\supseteq\{\sigma\in 2^{n}\mid\sigma(x)=t\}. If not, there would be i<ji<j with xi<xj<nx_{i}<x_{j}<n such that

  1. (1)

    Φi(xi)\Phi_{i}(x_{i})\downarrow but Φi,n(xi)\Phi_{i,n}(x_{i})\uparrow and

  2. (2)

    Φj(xj)\Phi_{j}(x_{j})\downarrow but Φj,n(xj)\Phi_{j,n}(x_{j})\uparrow.

From (1), it follows that n<sin<s_{i}, but by construction xj>six_{j}>s_{i}. This contradicts the choice of xj<nx_{j}<n.

Now suppose that BB is an infinite subset of AA. We will compute from BB a completion gg of ff. For each xx, look for some σB\sigma\in B with σ(x)\sigma(x)\downarrow. Check whether Φπ(x),|σ|(x)\Phi_{\pi(x),|\sigma|}(x)\downarrow. If so, define g(x)=1Φπ(x),|σ|(x)g(x)=1-\Phi_{\pi(x),|\sigma|}(x). Otherwise, define g(x)=σ(x)g(x)=\sigma(x). It is easy to see that gg is total.

To see that gg is a completion of ff, it suffices to check that for all ii, g(xi)=f(xi)g(x_{i})=f(x_{i}). Let σB\sigma\in B be the string used to define g(xi)g(x_{i}). If Φi,|σ|(xi)\Phi_{i,|\sigma|}(x_{i})\downarrow, then we defined both g(xi)g(x_{i}) and f(xi)f(x_{i}) to be 1Φi(xi)1-\Phi_{i}(x_{i}). Otherwise, if Φi,|σ|(xi)\Phi_{i,|\sigma|}(x_{i})\uparrow, then we defined g(xi)=σ(xi)g(x_{i})=\sigma(x_{i}), and if f(xi)f(x_{i}) is defined then Φi(xi)\Phi_{i}(x_{i})\downarrow and so, as σA\sigma\in A, σ(xi)=f(xi)\sigma(x_{i})=f(x_{i}). ∎

6. Preserving Kolmogorov complexity

Recall from the introduction that for any string σ\sigma and any collection of sets 𝒫()\mathcal{F}\subseteq\operatorname{\mathcal{P}}(\mathbb{N}), we define

C(σ)=maxXCX(σ).C(\sigma\mid\mathcal{F})=\max_{X\in\mathcal{F}}C^{X}(\sigma).

Also recall that C(σ)C(σ[A]ω)C(\sigma)-C(\sigma\mid[A]^{\omega}) can be thought of as the number of bits of information about σ\sigma coded into all infinite subsets of AA. The goal of this section is to investigate this quantity in the case where AA is a set of positive lower density.

We noted in the introduction that it is possible to encode about log(1/δ)\log(1/\delta) bits of information into all infinite subsets of a set of lower density δ\delta. For example, if δ=1/2n\delta=1/2^{n} then for any binary string σ\sigma of length nn, we can encode σ\sigma into all infinite subsets of a set AA of lower density 1/2n1/2^{n} by letting AA consist of all numbers mm such that the last (i.e. least significant) bits of the binary expansion of mm match σ\sigma.

However, this is not the only way to encode information into all infinite subsets of a dense set. Roughly speaking, for any fixed number NN, it is possible to encode “an arbitrary integer greater than NN.” We will now make this precise.

Definition 6.1.

For any string σ2<ω\sigma\in 2^{<\omega} and number NN\in\mathbb{N}, define

C(σN)=maxnNC(σn).C(\sigma\mid\;\geq N)=\max_{n\geq N}C(\sigma\mid n).
Proposition 6.2.

For any string σ\sigma and number NN, there is some set AA of lower density 1 such that

C(σ[A]ω)C(σN)+O(1)C(\sigma\mid[A]^{\omega})\leq C(\sigma\mid\;\geq N)+O(1)

where the constant hidden by the O(1)O(1) does not depend on σ\sigma or NN.

Proof.

Let A={nnN}A=\{n\mid n\geq N\}. If BB is an infinite subset of AA then all elements of BB are greater than or equal to NN. In other words, from any infinite subset of AA we can uniformly extract a number greater than or equal to NN. Thus C(σ[A]ω)C(\sigma\mid[A]^{\omega}) is at most C(σN)+O(1)C(\sigma\mid\;\geq N)+O(1), where the extra constant represents the complexity of the extraction procedure. ∎

It may not be immediately apparent exactly how much information can be gained from knowing “an arbitrary integer greater than NN” (i.e. how large C(σ)C(σN)C(\sigma)-C(\sigma\mid\;\geq N) can be). A result of Vereshchagin [16] gives a precise characterization.

Theorem 6.3 (Vereshchagin).

For any string σ\sigma,

C0(σ)=minNC(σN)±O(1).C^{0^{\prime}}(\sigma)=\min_{N}C(\sigma\mid\;\geq N)\pm O(1).

Together with our proposition above, this shows that for any string σ\sigma, it is possible for all infinite subsets of a dense set AA to lower the complexity of σ\sigma to its 00^{\prime} complexity. This can be combined with the first coding method we mentioned to give a result claimed in the introduction.

See 1.6

Proof.

Let k=C0(σ)k=C^{0^{\prime}}(\sigma) and let ρ\rho be a string witnessing this fact—i.e. a string such that |ρ|=k|\rho|=k and for the universal oracle machine UU, U0(ρ)=σU^{0^{\prime}}(\rho)=\sigma. It is a standard fact about Kolmogorov complexity that there is some NN such that for all nNn\geq N, C(σρ,n)=O(1)C(\sigma\mid\rho,n)=O(1) (this fact is more or less the easy direction of Vereshchagin’s theorem mentioned above).

Next, let mm\in\mathbb{N} be such that 2m1<δ2m2^{-m-1}<\delta\leq 2^{-m} and define AA by

A={nN\displaystyle A=\{n\geq N\mid the binary representation of nn mod 2m2^{m}
is equal to the first m bits of ρ}.\displaystyle\qquad\qquad\text{is equal to the first $m$ bits of $\rho$}\}.

If the length of ρ\rho is less than mm then first extend ρ\rho to length mm by adding a 11 followed by a run of 0s to the end.

Note that the lower density of AA is 2m2^{-m}, which is greater than or equal to δ\delta. Furthermore, from any infinite subset BAB\subseteq A, we can uniformly extract a number nNn\geq N. And if we know mm then we can also uniformly extract the first mm bits of ρ\rho. Since C(σρ,n)=O(1)C(\sigma\mid\rho,n)=O(1), to reconstruct σ\sigma from this information, we only need to know the last kmk-m bits of ρ\rho. In other words, if we are given an infinite subset BAB\subseteq A then to reconstruct σ\sigma we just need to know mm and the last kmk-m bits of ρ\rho. Thus we have

C(σ[A]ω)(km)+O(C(m))km+O(log(m)).C(\sigma\mid[A]^{\omega})\leq(k-m)+O(C(m))\leq k-m+O(\log(m)).

Since log(1/δ)mlog(1/δ)1\log(1/\delta)\geq m\geq\log(1/\delta)-1, this gives us

C(σ[A]ω)klog(1/δ)+O(loglog(1/δ))C(\sigma\mid[A]^{\omega})\leq k-\log(1/\delta)+O(\log\log(1/\delta))

as desired. ∎

In this section, we will show that this upper bound is essentially optimal. In particular, we will prove the following theorem.

See 1.7

It is also possible to prove a corresponding result for Seetapun’s theorem. In particular, for any set AA, define

𝖱𝖳21(A)=[A]ω[A¯]ω.\mathsf{RT}^{1}_{2}(A)=[A]^{\omega}\cup[\overline{A}]^{\omega}.

The quantity C(σ)C(σ𝖱𝖳21(A))C(\sigma)-C(\sigma\mid\mathsf{RT}^{1}_{2}(A)) can be viewed as the number of bits of information about σ\sigma that can be encoded into all infinite subsets of AA and A¯\overline{A}. Just as in the case of subsets of a dense set, it is always possible to lower the complexity of σ\sigma to its 00^{\prime} complexity.

Proposition 6.4.

For any string σ\sigma, there is some set AA\subseteq\mathbb{N} such that

C(σ𝖱𝖳21(A))C0(σ)+O(1).C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\leq C^{0^{\prime}}(\sigma)+O(1).
Proof.

Once again, by Vereshchagin’s theorem it is enough to prove that for any NN there is some set AA such that C(σ𝖱𝖳21(A))C(σN)+O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\leq C(\sigma\mid\;\geq N)+O(1). And for this, we can simply take A={nnN}A=\{n\mid n\geq N\}. ∎

Just like in the case of dense sets, it is possible to show that this is optimal (and unlike in the case of dense sets, the error term here is constant).

Theorem 6.5.

For any string σ\sigma and set AA\subseteq\mathbb{N},

C(σ𝖱𝖳21(A))C0(σ)O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C^{0^{\prime}}(\sigma)-O(1)

where the constant hidden by the O(1)O(1) does not depend on σ\sigma or AA.

The remainder of this section is focused on proving Theorems 1.7 and 6.5. As a warm-up, and to demonstrate our proof strategy, we will first prove a much easier version of Theorem 6.5. We will then prove the full theorem and, finally, adapt the techniques from the proof to prove Theorem 1.7.

6.1. Kolmogorov complexity and Seetapun’s theorem: easy version

We will now prove a version of Theorem 6.5 where the 00^{\prime} oracle is replaced by an oracle for a complete Σ21\Sigma^{1}_{2} set. In particular, let WW denote a complete Σ21\Sigma^{1}_{2} set and fix a string σ\sigma and set AA\subseteq\mathbb{N}. We will prove that

C(σ𝖱𝖳21(A))CW(σ)O(1).C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C^{W}(\sigma)-O(1).

This may appear a bit absurd—WW is ridiculously powerful—but it helps demonstrate the basic strategy that we will use in our proofs of Theorems 6.5 and 1.7.

Here’s the core idea of the proof. Suppose kk is a number such that C(σ𝖱𝖳21(A))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(A))<k. We would like to show that CW(σ)k+O(1)C^{W}(\sigma)\leq k+O(1). To do this, we will identify some property of σ\sigma which is shared by at most 2k2^{k} other strings and which can be recognized using an oracle for WW. One obvious special property of σ\sigma is that there is a set XX such that C(σ𝖱𝖳21(X))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(X))<k. It is not hard to show that at most 2k2^{k} strings have this property and that WW is powerful enough to check which strings have it (in fact, we chose WW to make this part obvious).

We will now give a more formal version of this argument. We will describe a program EE with the following properties.

  • For each kk, EW(k)E^{W}(k) enumerates a set of at most 2k2^{k} strings.

  • For any set AA, string σ\sigma and number kk such that C(σ𝖱𝖳21(A))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(A))<k, σ\sigma is in the set enumerated by EW(k)E^{W}(k).

By general properties of Kolmogorov complexity, for any number kk and string τ\tau enumerated by EW(k)E^{W}(k), CW(τ)k+O(1)C^{W}(\tau)\leq k+O(1). In particular, if C(σ𝖱𝖳21(A))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(A))<k then CW(σ)k+O(1)C^{W}(\sigma)\leq k+O(1).

We will now describe EE. First, say that a string τ\tau is good-for-kk if there is a set XX such that C(τ𝖱𝖳21(X))<kC(\tau\mid\mathsf{RT}^{1}_{2}(X))<k. We claim that both of the following hold.

  1. (1)

    It is computable in WW to check if a string τ\tau is good-for-kk (i.e. the set of pairs (τ,k)(\tau,k) such that τ\tau is good-for-kk is computable relative to WW).

  2. (2)

    There are at most 2k2^{k} strings which are good-for-kk.

To see why the first claim holds, simply note that the statement “τ\tau is good-for-kk” is equivalent to the statement “there is some XX such that for every YY, if YXY\subseteq X or YX¯Y\subseteq\overline{X}, then CY(τ)<kC^{Y}(\tau)<k,” which is a Σ21\Sigma^{1}_{2} formula of τ\tau and kk.

To see why the second claim holds, suppose that τ1,,τl\tau_{1},\ldots,\tau_{l} are distinct good-for-kk strings, as witnessed by X1,,XlX_{1},\ldots,X_{l}. Note that the collection of all Boolean combinations of the sets X1,,XlX_{1},\ldots,X_{l} forms a finite partition of \mathbb{N}. Thus one of these Boolean combinations is infinite. Let BB be this Boolean combination and note that for each ili\leq l, BB is either a subset of XiX_{i} or of Xi¯\overline{X_{i}} and, either way, is an element of 𝖱𝖳21(Xi)\mathsf{RT}^{1}_{2}(X_{i}). Therefore for each ili\leq l, CB(τi)<kC^{B}(\tau_{i})<k, which is impossible unless l2kl\leq 2^{k}.

The program EW(k)E^{W}(k) works as follows. Given the input kk, it goes through all strings in some fixed order and uses WW to check whether each one is good-for-kk. Each time it finds a string which is good-for-kk, it enumerates it. As we showed above, EW(k)E^{W}(k) will never enumerate more than 2k2^{k} strings. Also, for any set AA, string σ\sigma and number kk such that C(σ𝖱𝖳21(A))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(A))<k, it is obvious that σ\sigma is good-for-kk and thus will be enumerated by EW(k)E^{W}(k). It follows that for such a σ\sigma and kk, CW(σ)k+O(1)C^{W}(\sigma)\leq k+O(1).

6.2. Kolmogorov complexity and Seetapun’s theorem: hard version

We will now prove Theorem 6.5, which states that for any string σ\sigma and set AA\subseteq\mathbb{N}, C(σ𝖱𝖳21(A))C0(σ)O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C^{0^{\prime}}(\sigma)-O(1). Our strategy is the same as in the previous subsection—i.e. let’s assume that C(σ𝖱𝖳21(A))<kC(\sigma\mid\mathsf{RT}^{1}_{2}(A))<k and identify a property of σ\sigma which is shared by at most 2k2^{k} strings and which can be recognized using an oracle for 00^{\prime}.

In identifying such a property, there is a natural trade-off between how easy the property is to describe and the computational power required to check if a string has the property. In the previous subsection, we just needed a property which can be recognized using a complete Σ21\Sigma^{1}_{2} set and this allowed us to use a very straightforward property of σ\sigma. Now, however, we want a property that can be recognized by 00^{\prime}, which forces us to use a property that has a more intricate description. The key definition is the following.

Definition 6.6.

A finite set of strings FF is kk-safe if there is some finite444We could take n=2|F|n=2^{|F|} but this is not needed. partition X1,,XnX_{1},\ldots,X_{n} of \mathbb{N} and some number mm such that for all ini\leq n and all finite subsets sXis\subseteq X_{i},

|s|m|{τCs(τ)<k}F|2k.|s|\geq m\implies|\{\tau\mid C^{s}(\tau)<k\}\cup F|\leq 2^{k}.

The idea is that FF is kk-safe as witnessed by a partition X1,,XnX_{1},\ldots,X_{n} if for any XiX_{i} and any infinite subset BB of XiX_{i}, we may safely assume that BB will assign all strings in FF complexity less than kk. More specifically, it may not actually be the case that BB assigns a complexity less than kk to each string in FF, but if we assume that it does then we will never see a contradiction of the form “BB assigns complexity less than kk to too many strings.”

Essentially the key property of σ\sigma is that it is a member of every maximal kk-safe set. More precisely, we will show that any maximal kk-safe set has size at most 2k2^{k} and contains σ\sigma and that there is some maximal kk-safe set which can be enumerated by 00^{\prime}, uniformly in kk. The key facts are

  1. (1)

    Every kk-safe set has size at most 2k2^{k}.

  2. (2)

    If FF is kk-safe then so is F{σ}F\cup\{\sigma\}.

  3. (3)

    The collection of kk-safe sets is c.e. relative to 00^{\prime} (uniformly in kk).

Before proving these three facts, let’s see how they can be used to finish the proof. Consider the following procedure for using 00^{\prime} to enumerate a maximal kk-safe set.

  1. 1.

    Set F=F=\varnothing

  2. 2.

    While true:

  3. 3.

    Search for a string τ\tau such that F{τ}F\cup\{\tau\} is kk-safe

  4. 4.

    If such a τ\tau is found, enumerate τ\tau and set F=F{τ}F=F\cup\{\tau\}

It is easy to show by induction that at every step of the above algorithm, FF is kk-safe. Since no kk-safe set has size more than 2k2^{k}, the above algorithm can add at most 2k2^{k} strings to FF. Thus after some point, no new strings will be added to FF. We claim that at this point, σ\sigma must be an element of FF. If not, then since F{σ}F\cup\{\sigma\} is kk-safe, eventually the algorithm will discover this fact and add σ\sigma to FF, contradicting our assumption that no new elements are added to FF. All this implies that the algorithm must eventually enumerate σ\sigma and also that the algorithm will enumerate at most 2k2^{k} strings.

To finish, observe that the above procedure was uniform in kk. In other words, there is a single program EE such that for each ll, E0(l)E^{0^{\prime}}(l) carries out the above algorithm with k=lk=l. As in the previous subsection, for every number ll and string τ\tau enumerated by E0(l)E^{0^{\prime}}(l), C0(τ)l+O(1)C^{0^{\prime}}(\tau)\leq l+O(1). Since σ\sigma is enumerated by E0(k)E^{0^{\prime}}(k), this shows that C0(σ)k+O(1)C^{0^{\prime}}(\sigma)\leq k+O(1), as desired.

We will now prove each of the three facts above.

Lemma 6.7.

Every kk-safe set has size at most 2k2^{k}.

Proof.

Suppose FF is a kk-safe set. Let X1,,XnX_{1},\ldots,X_{n} be a partition witnessing that FF is kk-safe. Since X1,,XnX_{1},\ldots,X_{n} is a finite partition of \mathbb{N}, some XiX_{i} must be infinite. It is straightforward to check that since XiX_{i} is infinite, we must have

|{τCXi(τ)<k}F|2k|\{\tau\mid C^{X_{i}}(\tau)<k\}\cup F|\leq 2^{k}

which implies that |F|2k|F|\leq 2^{k}. ∎

Lemma 6.8.

If FF is a kk-safe set then so is F{σ}F\cup\{\sigma\}.

Proof.

Suppose that FF is kk-safe as witnessed by X1,,XnX_{1},\ldots,X_{n} and mm. Recall that the string σ\sigma and set AA have the property that for all infinite subsets BB of AA or A¯\overline{A}, CB(σ)<kC^{B}(\sigma)<k. The main idea of this proof is that the fact that F{σ}F\cup\{\sigma\} is kk-safe can be witnessed by the partition

X1A,X1A¯,,XnA,XnA¯.X_{1}\cap A,X_{1}\cap\overline{A},\ldots,X_{n}\cap A,X_{n}\cap\overline{A}.

More precisely, let mm^{\prime} be a number larger than both the size of any finite piece of this partition and mm. We will show that F{σ}F\cup\{\sigma\} is kk-safe as witnessed by mm^{\prime} and X1A,X1A¯,,XnA,XnA¯X_{1}\cap A,X_{1}\cap\overline{A},\ldots,X_{n}\cap A,X_{n}\cap\overline{A}.

To prove this claim, suppose ss is a finite subset of some set in this partition and |s|m|s|\geq m^{\prime}. Without loss of generality, let’s assume that sX1As\subseteq X_{1}\cap A. By our choice of mm^{\prime} and the assumption that |s|m|s|\geq m^{\prime}, X1AX_{1}\cap A must be infinite. Thus we can find some subset BX1AB\subseteq X_{1}\cap A which agrees with ss below max(s)\max(s) and which is infinite. It is straightforward to check the following facts

  1. (1)

    Since BB is an infinite subset of X1X_{1}, |{τCB(τ)<k}F|2k|\{\tau\mid C^{B}(\tau)<k\}\cup F|\leq 2^{k}.

  2. (2)

    Since ss is an initial segment of BB, {τCs(τ)<k}{τCB(τ)<k}\{\tau\mid C^{s}(\tau)<k\}\subseteq\{\tau\mid C^{B}(\tau)<k\}.

  3. (3)

    Since BB is an infinite subset of AA, CB(σ)<kC^{B}(\sigma)<k and thus σ{τCB(τ)<k}\sigma\in\{\tau\mid C^{B}(\tau)<k\}.

Putting these together, we have

{τCs(τ)<k}F{σ}\displaystyle\{\tau\mid C^{s}(\tau)<k\}\cup F\cup\{\sigma\} {τCB(τ)<k}F{σ}\displaystyle\subseteq\{\tau\mid C^{B}(\tau)<k\}\cup F\cup\{\sigma\}
={τCB(τ)<k}F\displaystyle=\{\tau\mid C^{B}(\tau)<k\}\cup F

and thus

|{τCs(τ)<k}F{σ}||{τCB(τ)<k}F|2k|\{\tau\mid C^{s}(\tau)<k\}\cup F\cup\{\sigma\}|\leq|\{\tau\mid C^{B}(\tau)<k\}\cup F|\leq 2^{k}

as desired. ∎

Lemma 6.9.

The collection of kk-safe sets is c.e. relative to 00^{\prime}, uniformly in kk.

Proof.

Given a number kk, a finite set of strings FF, a number mm and sets X1,,XnX_{1},\ldots,X_{n}\subseteq\mathbb{N}, the statement that mm and X1,,XnX_{1},\ldots,X_{n} witness that FF is kk-safe is uniformly Π10\Pi^{0}_{1}. In other words, there is some computable tree T(k,F,m,n)T(k,F,m,n) such that X1XnX_{1}\oplus\ldots\oplus X_{n} is a path through T(k,F,m,n)T(k,F,m,n) if and only if mm and X1,,XnX_{1},\ldots,X_{n} witness that FF is kk-safe and, furthermore, T(k,F,m,n)T(k,F,m,n) is uniformly computable in kk, FF, mm and nn.

By König’s lemma, T(k,F,m,n)T(k,F,m,n) has an infinite path if and only if T(k,F,m,n)T(k,F,m,n) is itself infinite. Thus to check if FF is kk-safe, it suffices to check if there are mm and nn such that T(k,F,m,n)T(k,F,m,n) is infinite. This is a Σ20\Sigma^{0}_{2} property (uniformly in kk and FF) and thus c.e. relative to 00^{\prime}. ∎

6.3. Kolmogorov complexity and dense sets

We will now prove Theorem 1.7. Actually, we will instead prove the following apparently weaker version and then derive Theorem 1.7 as a corollary.

Theorem 6.10.

For any string σ\sigma and set AA\subseteq\mathbb{N} which is δ\delta-dense for δ(0,1]\delta\in(0,1],

C(σ[A]ω)C0(σ)log(1/δ)O(loglog(1/δ))C(\sigma\mid[A]^{\omega})\geq C^{0^{\prime}}(\sigma)-\log(1/\delta)-O(\log\log(1/\delta))

where the constant hidden by the O()O(\cdot) notation does not depend on σ\sigma or AA.

Note that this theorem differs from Theorem 1.7 in that the set AA is required to be δ\delta-dense rather than to have lower density δ\delta (essentially AA is required to be dense everywhere rather than just dense at all sufficiently large points).

We will begin by showing how to use this theorem to prove Theorem 1.7.

Proof of Theorem 1.7 using Theorem 6.10.

The basic point is that for any string σ\sigma and set AA of lower density δ\delta, there is a set A~\widetilde{A} which is δ/4\delta/4-dense such that

C(σ[A~]ω)C(σ[A]ω)+O(1)C(\sigma\mid[\widetilde{A}]^{\omega})\leq C(\sigma\mid[A]^{\omega})+O(1)

and thus Theorem 6.10 implies that

C(σ[A]ω)\displaystyle C(\sigma\mid[A]^{\omega}) C(σ[A~]ω)O(1)\displaystyle\geq C(\sigma\mid[\widetilde{A}]^{\omega})-O(1)
C0(σ)log(4/δ)O(loglog(4/δ))\displaystyle\geq C^{0^{\prime}}(\sigma)-\log(4/\delta)-O(\log\log(4/\delta))
=C0(σ)log(1/δ)O(loglog(1/δ)).\displaystyle=C^{0^{\prime}}(\sigma)-\log(1/\delta)-O(\log\log(1/\delta)).

The set A~\widetilde{A} is simple to describe. Let NN be large enough that for all nNn\geq N, AA has density at least δ/2\delta/2 at nn. Then define

A~={nn is even and n2N}{2n+1nA}.\widetilde{A}=\{n\mid n\text{ is even and }n\leq 2N\}\cup\{2n+1\mid n\in A\}.

In other words, A~\widetilde{A} contains a copy of AA on the odd numbers, together with enough even numbers to make sure A~\widetilde{A} is relatively dense everywhere.

To finish, note that any infinite subset of A~\widetilde{A} must contain an infinite number of odd numbers and thus can uniformly compute an infinite subset of AA and so we have

C(σ[A~]ω)C(σ[A]ω)+O(1)C(\sigma\mid[\widetilde{A}]^{\omega})\leq C(\sigma\mid[A]^{\omega})+O(1)

as desired. ∎

We will now prove Theorem 6.10. The proof closely follows the proof from the previous subsection, but with a few additional difficulties. We will begin by modifying the definition of a kk-safe set.

Definition 6.11.

For any kk\in\mathbb{N} and rational δ(0,1]\delta\in(0,1], a finite set of strings F={τ1,,τn}F=\{\tau_{1},\ldots,\tau_{n}\} is kk-safe at density δ\delta if there is a number mm and sets X1,,XnX_{1},\ldots,X_{n}\subseteq\mathbb{N} of density at least δ\delta such that for all i1<i2<<ili_{1}<i_{2}<\ldots<i_{l} and all finite subsets sXi1Xils\subseteq X_{i_{1}}\cap\ldots\cap X_{i_{l}},

|s|m|{τCs(τ)<k}{τi1,,τil}|2k.|s|\geq m\implies|\{\tau\mid C^{s}(\tau)<k\}\cup\{\tau_{i_{1}},\ldots,\tau_{i_{l}}\}|\leq 2^{k}.

As in the previous subsection, we have the following claims.

  1. (1)

    If FF is kk-safe at density δ\delta then |F|2k/δ|F|\leq 2^{k}/\delta.

  2. (2)

    Suppose we have a set AA and string σ\sigma such that AA is δ\delta dense and C(σ[A]ω)<kC(\sigma\mid[A]^{\omega})<k. If FF is kk-safe at density δ\delta then so is F{σ}F\cup\{\sigma\}.

  3. (3)

    The collection of sets which are kk-safe at density δ\delta is c.e. relative to 00^{\prime}, uniformly in kk and δ\delta.

Before proving these claims, let’s see how to use them to finish the proof. In a similar fashion to the previous subsection, we can find a program EE such that for any kk and nn, E0(k,n)E^{0^{\prime}}(k,n) enumerates a maximal set which is kk-safe at density 2n2^{-n}. Furthermore, E0(k,n)E^{0^{\prime}}(k,n) enumerates at most 2k2n2^{k}\cdot 2^{n} strings. By standard facts about Kolmogorov complexity, this implies that for any τ\tau which is enumerated by E0(k,n)E^{0^{\prime}}(k,n), C0(τn)k+n+O(1)C^{0^{\prime}}(\tau\mid n)\leq k+n+O(1). Thus for such a τ\tau we have

C0(τ)C0(τn)+O(C0(n))k+n+O(C0(n))C^{0^{\prime}}(\tau)\leq C^{0^{\prime}}(\tau\mid n)+O(C^{0^{\prime}}(n))\leq k+n+O(C^{0^{\prime}}(n))

and so

C0(τ)k+n+O(log(n)).C^{0^{\prime}}(\tau)\leq k+n+O(\log(n)).

Now fix a string σ\sigma and a set AA of density at least δ(0,1/2]\delta\in(0,1/2]. Let k=C(σ[A]ω)+1k=C(\sigma\mid[A]^{\omega})+1 and nn be such that 2n<δ2n+12^{-n}<\delta\leq 2^{-n+1}. Then by the facts above, we have that σ\sigma is enumerated by E0(k,n)E^{0^{\prime}}(k,n) and thus

C0(σ)k+n+O(log(n))=k+log(1/δ)+O(loglog(1/δ)).C^{0^{\prime}}(\sigma)\leq k+n+O(\log(n))=k+\log(1/\delta)+O(\log\log(1/\delta)).

Note that in the calculation above we used that n1log(1/δ)<nn-1\leq\log(1/\delta)<n.

We will now prove the claims above.

Lemma 6.12.

Suppose FF is kk-safe at density δ\delta. Then |F|2k/δ|F|\leq 2^{k}/\delta.

Proof.

Suppose F={τ1,,τn}F=\{\tau_{1},\ldots,\tau_{n}\} and that the sets X1,,XnX_{1},\ldots,X_{n} witness that FF is kk-safe at density δ\delta.

We first claim that there is some collection of at least δn\delta n many XiX_{i}’s whose intersection is infinite. To see why, suppose for contradiction that each such collection is finite. Then there is some number NN large enough that every xNx\geq N is in fewer than δn\delta n of the XiX_{i}’s. Fix some NN^{\prime} much larger than NN and consider a random point xx from the interval [N,N)[N,N^{\prime}). We will derive a contradiction by calculating the expected number of XiX_{i}’s which contain xx.

First consider a single fixed XiX_{i}. Since XiX_{i} is δ\delta dense at NN^{\prime}, XiX_{i} has at least δN\delta N^{\prime} elements less than NN^{\prime} and thus has at least δNN\delta N^{\prime}-N elements in the interval [N,N)[N,N^{\prime}). So the probability that xx is in XiX_{i} is at least

δNNN=δNN.\frac{\delta N^{\prime}-N}{N^{\prime}}=\delta-\frac{N}{N^{\prime}}.

By linearity of expectation, this implies that the expected number of XiX_{i}’s which contain xx is at least

δnNnN.\delta n-\frac{Nn}{N^{\prime}}.

Thus there is some xx in the interval [N,N)[N,N^{\prime}) which is contained in at least δnNnN\delta n-\frac{Nn}{N^{\prime}} many XiX_{i}’s. Since the number of XiX_{i}’s containing xx must be an integer, this xx must actually be contained in

δnNnN\left\lceil\delta n-\frac{Nn}{N^{\prime}}\right\rceil

many XiX_{i}’s. If NN^{\prime} is large enough then this is simply equal to δn\lceil\delta n\rceil. In other words, xx is contained in at least δn\delta n many XiX_{i}’s, contradicting our assumption.

We have now shown that there is some collection of at least δn\delta n many XiX_{i}’s whose intersection is infinite. Suppose for convenience that these XiX_{i}’s are simply X1,,XlX_{1},\ldots,X_{l} where l=δnl=\lceil\delta n\rceil and let B=X1XlB=X_{1}\cap\ldots\cap X_{l}. Note that by definition of “kk-safe at density δ\delta,” we must have

|{τCB(τ)<k}{τ1,,τl}|2k,|\{\tau\mid C^{B}(\tau)<k\}\cup\{\tau_{1},\ldots,\tau_{l}\}|\leq 2^{k},

which is impossible unless l2kl\leq 2^{k}. To summarize, we have δn2k\delta n\leq 2^{k}, which implies that n2k/δn\leq 2^{k}/\delta, as desired. ∎

Lemma 6.13.

Suppose AA is δ\delta dense, C(σ[A]ω)<kC(\sigma\mid[A]^{\omega})<k and FF is kk-safe at density δ\delta. Then F{σ}F\cup\{\sigma\} is also kk-safe at density δ\delta.

Proof.

The proof is essentially the same as the proof of Lemma 6.8. Namely, let F={τ1,,τn}F=\{\tau_{1},\ldots,\tau_{n}\} and fix sets X1,,XnX_{1},\ldots,X_{n} and a number mm witnessing that FF is kk-safe at density δ\delta. Let m>mm^{\prime}>m be large enough that for any intersection of finitely many of X1,,Xn,AX_{1},\ldots,X_{n},A, if that intersection happens to be finite, then it has at most mm^{\prime} elements. By copying the proof of Lemma 6.8, one can show that F{σ}F\cup\{\sigma\} is kk-safe at density δ\delta, as witnessed by X1,,Xn,AX_{1},\ldots,X_{n},A and mm^{\prime}. ∎

Lemma 6.14.

The collection of sets which are kk-safe at density δ\delta is c.e. relative to 00^{\prime}, uniformly in kk and δ\delta.

Proof.

The proof is essentially the same as the proof of Lemma 6.9. The key point is that the statement “{τ1,,τn}\{\tau_{1},\ldots,\tau_{n}\} is kk-safe at density δ\delta, as witnessed by X1,,XnX_{1},\ldots,X_{n} and mm” is Π10\Pi^{0}_{1}. ∎

6.4. Open questions

The presence of the 00^{\prime} oracle in Theorems 6.5 and 1.7 is somewhat unsatisfying. In one sense it is necessary: Propositions 6.4 and 1.6 show that it cannot be removed. However, there is another sense in which it may be possible to strengthen the two theorems.

For the sake of concreteness, we will focus on a possible strengthening of Theorem 6.5. In that theorem, we showed that for any set AA and string σ\sigma, C(σ𝖱𝖳21(A))C0(σ)O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C^{0^{\prime}}(\sigma)-O(1). By Vereshchagin’s theorem, this is equivalent to showing that for each string σ\sigma, there is some number NN such that C(σ𝖱𝖳21(A))C(σN)O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C(\sigma\mid\;\geq N)-O(1). However, this allows the possibility that even for a fixed set AA, there is a different such NN for each σ\sigma. It seems natural to wonder whether this is necessary, or whether, for each set AA, there is a single NN which works for all σ\sigma.

Conjecture 6.15.

For any set AA\subseteq\mathbb{N}, there is a number NN such that for any string σ\sigma,

C(σ𝖱𝖳21(A))C(σN)O(1).C(\sigma\mid\mathsf{RT}^{1}_{2}(A))\geq C(\sigma\mid\;\geq N)-O(1).

Note that the set AA we constructed to show that the bound in Theorem 6.5 is tight is not a counterexample to this conjecture: for that set AA there is a single number NN such that for all strings σ\sigma, C(σ𝖱𝖳21(A))=C(σN)±O(1)C(\sigma\mid\mathsf{RT}^{1}_{2}(A))=C(\sigma\mid\;\geq N)\pm O(1).

One could also ask for a similar strengthening of Theorem 1.7.

Conjecture 6.16.

For any set AA\subseteq\mathbb{N} of lower density δ\delta, there is a number NN and a string τ\tau of length at most log(1/δ)\log(1/\delta) such that for any string σ\sigma,

C(σ[A]ω)C(σN,τ)O(loglog(1/δ)).C(\sigma\mid[A]^{\omega})\geq C(\sigma\mid\;\geq N,\tau)-O(\log\log(1/\delta)).

There is another conjecture, also related to Theorem 1.1 and Kolmogorov complexity, but quite distinct from the conjectures above, that we would like to mention here. Namely, it might be possible to strengthen Theorem 1.1 to say that given AA of positive lower density, and XX uncomputable, not only is there a subset of AA that does not compute XX, but there is a subset of AA which does not help to compress initial segments of XX.

Conjecture 6.17.

Let XX be any set. For any set AA\subseteq\mathbb{N} of lower density δ\delta there is an infinite subset BAB\subseteq A such that for infinitely many nn,

CB(Xn)C(Xn)O(1).C^{B}(X\operatorname{\upharpoonright}n)\geq C(X\operatorname{\upharpoonright}n)-O(1).

Here we take inspiration from a theorem due to Chaitin (see [4], Theorem 3.4.4) that XX is computable if and only if C(Xn)C(n)+O(1)C(X\operatorname{\upharpoonright}n)\leq C(n)+O(1) for all nn. If XX is uncomputable, there is a function f:f\colon\mathbb{N}\to\mathbb{N} such that lim supnf(n)=\limsup_{n\to\infty}f(n)=\infty and for all nn, C(Xn)C(n)+f(n)O(1)C(X\operatorname{\upharpoonright}n)\geq C(n)+f(n)-O(1). If BAB\subseteq A is an infinite set which preserves the complexity of infinitely many initial segments of XX, then for infinitely many nn,

CB(Xn)C(Xn)O(1)C(n)+f(n)O(1)CB(n)+f(n)O(1),C^{B}(X\operatorname{\upharpoonright}n)\geq C(X\operatorname{\upharpoonright}n)-O(1)\geq C(n)+f(n)-O(1)\geq C^{B}(n)+f(n)-O(1),

and hence XX is not computable relative to BB, which is exactly the conclusion of Theorem 1.1.

Note that in this last conjecture we must only ask that CB(Xn)C(Xn)O(1)C^{B}(X\operatorname{\upharpoonright}n)\geq C(X\operatorname{\upharpoonright}n)-O(1) for infinitely many nn, rather than for all nn, as whenever BAB\subseteq A is uncomputable it will given shorter descriptions for infinitely many nn, and thus BB can further compress even some initial segments of the empty set X=X=\varnothing.

7. The relationship between the main theorem and Seetapun’s theorem

It seems clear that Theorem 1.1 and Seetapun’s theorem are closely related. Both give limitations on coding information into all infinite subsets of a set and we used Seetapun’s theorem in our proof of Theorem 1.1. However, the two theorems are even more closely related than this suggests. In particular, Theorem 1.1 directly implies Seetapun’s theorem.

To see why, let XX be uncomputable and AA\subseteq\mathbb{N} be arbitrary. Consider the following set BB\subseteq\mathbb{N} of lower density 1/21/2:

B={2nnA}{2n+1nA}.B=\{2n\mid n\in A\}\cup\{2n+1\mid n\notin A\}.

Given any subset CBC\subseteq B, we can read off a subset CevenC_{\text{even}} of AA by looking at the even elements of CC and a subset CoddC_{\text{odd}} of A¯\overline{A} by looking at the odd elements of CC. Moreover, if CC is infinite then at least one of CevenC_{\text{even}} and CoddC_{\text{odd}} must be infinite as well. By Theorem 1.1, BB must have some infinite subset CC that does not compute XX. But then either CevenC_{\text{even}} is an infinite subset of AA which does not compute XX or CoddC_{\text{odd}} is an infinite subset of A¯\overline{A} which does not compute XX.

The argument above does not constitute a new proof of Seetapun’s theorem because we used Seetapun’s theorem in our proof of Theorem 1.1. However, it might make one wonder whether there is a similar direct proof of Theorem 1.1 from Seetapun’s theorem (in which case, the somewhat complicated proof of Theorem 1.1 that we gave in Section 3 would be pointless). In this section, we will show that this is not the case. However, we first have to make precise what sort of proof we are trying to rule out.

7.1. Strong omniscient computable reducibility

The proof of Seetapun’s theorem from Theorem 1.1 that we gave above can be seen as an example of a strong computable reduction, a notion closely related to Weihrauch reducibility and reverse math which was first introduced by Dzhafarov [5].

Definition 7.1.

Given two partial functions P,Q:2ω𝒫(2ω)P,Q\colon 2^{\omega}\to\operatorname{\mathcal{P}}(2^{\omega}), PP is strongly computably reducible to QQ, written PscQP\leq_{sc}Q, if for any xdom(P)x\in\operatorname{dom}(P),

  1. (1)

    there is some x~Tx\widetilde{x}\leq_{T}x such that x~dom(Q)\widetilde{x}\in\operatorname{dom}(Q)

  2. (2)

    and for any y~Q(x~)\widetilde{y}\in Q(\widetilde{x}), there is some yTy~y\leq_{T}\widetilde{y} such that yP(x)y\in P(x).

Intuitively, a partial function P:2ω𝒫(2ω)P\colon 2^{\omega}\to\operatorname{\mathcal{P}}(2^{\omega}) can be seen as a problem: elements xx of the domain of PP are instances of the problem and elements yy of P(x)P(x) are solutions to the instance xx. Under this interpretation, a problem PP is strongly computably reducible to a problem QQ if any instance xx of the problem PP can compute an instance x~\widetilde{x} of QQ such that any solution to x~\widetilde{x} can compute a solution to the original problem xx.

As we said, the proof of Seetapun’s theorem from Theorem 1.1 that we gave above can be seen as an example of a strong computable reduction. First, we can define problems corresponding to Seetapun’s theorem and Theorem 1.1:

  • 𝖱𝖳21\mathsf{RT}^{1}_{2} is the problem in which an instance is a set AA\subseteq\mathbb{N} and a solution to that instance is an infinite subset of either AA or A¯\overline{A}.

  • 𝖲𝖣\mathsf{SD} (which stands for “subset of a dense set”) is the problem in which an instance is a set AA\subseteq\mathbb{N} of positive lower density and a solution is an infinite subset of AA.

Next, we can translate the main idea of the proof into a strong computable reduction of 𝖱𝖳21\mathsf{RT}^{1}_{2} to 𝖲𝖣\mathsf{SD}:

  1. (1)

    Given an instance AA of 𝖱𝖳21\mathsf{RT}^{1}_{2}, we can compute the instance B={2nnA}{2n+1nA}B=\{2n\mid n\in A\}\cup\{2n+1\mid n\notin A\} of 𝖲𝖣\mathsf{SD}.

  2. (2)

    Given a solution CC to the instance BB (i.e. an infinite subset CBC\subseteq B), we can compute a solution to the instance AA of 𝖱𝖳21\mathsf{RT}^{1}_{2} (namely, either CevenC_{\text{even}} or CoddC_{\text{odd}}).

Furthermore, it is easy to check that if PP and QQ are problems such that PscQP\leq_{sc}Q and XX is a set such that there is some instance PP, all of whose solutions compute XX then there is some instance of QQ all of whose solutions compute XX. Thus the fact that 𝖱𝖳21sc𝖲𝖣\mathsf{RT}^{1}_{2}\leq_{sc}\mathsf{SD} plus Theorem 1.1 implies Seetapun’s theorem.

So to show that there is no proof of this sort of Theorem 1.1 from Seetapun’s theorem, we can show that there is no strong computable reduction of 𝖲𝖣\mathsf{SD} to 𝖱𝖳21\mathsf{RT}^{1}_{2}. In fact, we will prove a somewhat stronger statement: instead of strong computable reductions of 𝖲𝖣\mathsf{SD} to 𝖱𝖳21\mathsf{RT}^{1}_{2}, we will consider strong omniscient computable reductions.

Strong omniscient computable reducibility is a natural weakening of strong computable reducibility, introduced by Monin and Patey [11], in which the transformation of instances is not required to be computable (though the transformation of solutions still is).

Definition 7.2.

Given two partial functions P,Q:2ω𝒫(2ω)P,Q\colon 2^{\omega}\to\operatorname{\mathcal{P}}(2^{\omega}), PP is strongly omnisciently computably reducible to QQ, written PsocQP\leq_{soc}Q, if for any xdom(P)x\in\operatorname{dom}(P),

  1. (1)

    there is some x~\widetilde{x} such that x~dom(Q)\widetilde{x}\in\operatorname{dom}(Q)

  2. (2)

    and for any y~Q(x~)\widetilde{y}\in Q(\widetilde{x}), there is some yTy~y\leq_{T}\widetilde{y} such that yP(x)y\in P(x).

Note that the only difference from strong computable reducibility is that here, x~\widetilde{x} is not required to be computable from xx.

It is clear that PscQP\leq_{sc}Q implies PsocQP\leq_{soc}Q, but not always vice-versa, and hence showing that 𝖲𝖣soc𝖱𝖳21\mathsf{SD}\nleq_{soc}\mathsf{RT}^{1}_{2} is a stronger result than showing that 𝖲𝖣sc𝖱𝖳21\mathsf{SD}\nleq_{sc}\mathsf{RT}^{1}_{2}.

7.2. Theorem 1.1 is not reducible to Seetapun’s theorem

We will now prove that 𝖲𝖣soc𝖱𝖳21\mathsf{SD}\nleq_{soc}\mathsf{RT}^{1}_{2}. In our proof, we will construct infinite sets AA and GG such that AA has lower density at least 1/21/2 and no infinite subset of GG computes any infinite subset of AA.

To see why constructing such an AA and GG is sufficient to prove 𝖲𝖣soc𝖱𝖳21\mathsf{SD}\nleq_{soc}\mathsf{RT}^{1}_{2}, suppose for contradiction that such an AA and GG exist and that 𝖲𝖣soc𝖱𝖳21\mathsf{SD}\leq_{soc}\mathsf{RT}^{1}_{2}. Viewing AA as an instance of 𝖲𝖣\mathsf{SD}, we get a set BB (i.e. an instance of 𝖱𝖳21\mathsf{RT}^{1}_{2}) such that every infinite subset of BB and B¯\overline{B} (i.e. every solution to 𝖱𝖳21(B)\mathsf{RT}^{1}_{2}(B)) computes an infinite subset of AA (i.e. a solution to 𝖲𝖣(A)\mathsf{SD}(A)). However, since GG is infinite, at least one of GBG\cap B and GB¯G\cap\overline{B} is also infinite and does not compute any infinite subset of AA.

In our construction of AA and GG, we will use a corollary of the Galvin-Prikry theorem [7].

Theorem 7.3 (Galvin-Prikry theorem).

For any Borel set 𝒳𝒫()\mathcal{X}\subseteq\operatorname{\mathcal{P}}(\mathbb{N}), there is some infinite set BB\subseteq\mathbb{N} such that either all infinite subsets of BB are in 𝒳\mathcal{X} or no infinite subsets of BB are in 𝒳\mathcal{X}.

Corollary 7.4.

For any finite set kk, Borel coloring c:𝒫()kc\colon\operatorname{\mathcal{P}}(\mathbb{N})\to k and infinite set BB, there is some infinite set BBB^{\prime}\subseteq B such that all infinite subsets of BB^{\prime} have the same color.

Before proving the existence of AA and GG, we will try to motivate the construction and explain how the Galvin-Prikry theorem will be used. Our goal is to construct AA and GG such that for all Turing functionals Φ\Phi and all infinite subsets CGC\subseteq G, Φ(C)\Phi(C) is not an infinite subset of AA. Suppose that instead of having to handle all Turing functionals, we just had to handle a single Turing functional, Φ\Phi. By the Galvin-Prikry theorem, there is some infinite set B0B_{0} such that either

  1. (1)

    for all infinite subsets CB0C\subseteq B_{0}, Φ(C,0)=1\Phi(C,0)\!\downarrow\;=1

  2. (2)

    or for all infinite subsets CB0C\subseteq B_{0}, Φ(C,0)1\Phi(C,0)\neq 1 (i.e. Φ(C,0)\Phi(C,0) either diverges or converges to something besides 11).

In the first case, we are done: we can take G=B0G=B_{0} and make sure 0A0\notin A (e.g. take A={0}A=\mathbb{N}\setminus\{0\}). The second case is more difficult. Taking G=B0G=B_{0} is not enough because it does not give us much control over what is computed by infinite subsets of GG—it just ensures that if CC is an infinite subset of GG then 0 is not an element of the set Φ(C)\Phi(C).

However, requiring GG to be a subset of B0B_{0} does seem like progress towards our goal: if we could similarly ensure that for each nn\in\mathbb{N} and infinite set CGC\subseteq G, nn is not an element of Φ(C)\Phi(C), then we would be done (because this would imply that for each such CC, Φ(C)\Phi(C) is either not total or computes the empty set).

This suggests the following strategy: form a sequence of sets

B0B1B2B_{0}\supseteq B_{1}\supseteq B_{2}\supseteq\ldots

where for each nn, BnB_{n} is chosen using the Galvin-Prikry theorem so that either

  1. (1)

    for all infinite subsets CBnC\subseteq B_{n}, Φ(C,n)=1\Phi(C,n)\!\downarrow\;=1

  2. (2)

    or for all infinite subsets CB0C\subseteq B_{0}, Φ(C,n)1\Phi(C,n)\neq 1.

If the first case holds for some nn, we can take G=BnG=B_{n} and forbid nn from being included in AA. If the second case holds for every nn, then we would like to take G=nBnG=\bigcap_{n\in\mathbb{N}}B_{n} since, as we noted above, this ensures that for each infinite CGC\subseteq G, Φ(C)\Phi(C) is either not total or computes the empty set. However, there is one problem with this: nBn\bigcap_{n\in\mathbb{N}}B_{n} could be finite, or even empty.

To deal with this problem, we will construct GG using a sequence of Mathias conditions, at each step restricting the reservoir in a manner similar to what we have just described, while also adding some elements to the stem to ensure GG is infinite. We will now give the details of the proof. Because of the necessity of using Mathias conditions, of dealing with all Turing functionals (rather than just one), and of ensuring that AA has high density, our proof is somewhat more elaborate than the sketch we have just given. However, the key ideas are the same.

Theorem 7.5.

𝖲𝖣soc𝖱𝖳21\mathsf{SD}\nleq_{soc}\mathsf{RT}^{1}_{2}.

Proof.

As explained above, we will show how to construct infinite sets AA and GG such that AA has lower density at least 1/21/2 and no infinite subset of GG computes an infinite subset of AA. To construct GG, we will build a sequence of Mathias conditions

(,)=(s0,B0)(s1,B1)(s2,B2)(\varnothing,\mathbb{N})=(s_{0},B_{0})\geq(s_{1},B_{1})\geq(s_{2},B_{2})\geq\ldots

and take G=nsnG=\bigcup_{n}s_{n}. Along the way, we will forbid certain numbers from being included in AA and then take AA to be the set of all non-forbidden numbers. For convenience, we will ensure that each sns_{n} has size exactly nn.

For each ee, we must satisfy the following requirement:

Requirement e\bm{e}: for every infinite CGC\subseteq G, either there is some mm such that Φe(C,m)=1\Phi_{e}(C,m)\!\downarrow\;=1 and mAm\notin A or for all but finitely many mm, Φe(C,m)1\Phi_{e}(C,m)\neq 1.

To satisfy requirement ee, we will take some action at each step n>en>e of the construction. Roughly speaking, on step nn we will try to satisfy the requirement for all mm in the interval [2e+n+3,2e+n+4)[2^{e+n+3},2^{e+n+4}).

More precisely, to satisfy requirement ee, we will ensure that for each step n>en>e and each tsnt\subseteq s_{n}, either

  1. (1)

    there is some m[2e+n+3,2e+n+4)m\in[2^{e+n+3},2^{e+n+4}) such that for all infinite sets CC compatible with (t,Bn)(t,B_{n}), Φ(C,m)=1\Phi(C,m)\!\downarrow\;=1

  2. (2)

    or for all m[2e+n+3,2e+n+4)m\in[2^{e+n+3},2^{e+n+4}) and infinite sets CC compatible with (t,Bn)(t,B_{n}), Φ(C,m)1\Phi(C,m)\neq 1.

If the first case holds, we will also pick one such mm to forbid from AA.

How to pick sn+𝟏\bm{s_{n+1}} and Bn+𝟏\bm{B_{n+1}}. For each nn, form sn+1s_{n+1} by picking an arbitrary element of BnB_{n} to add to sns_{n}. Then pick Bn+1B_{n+1} as follows.

For each e<n+1e<n+1, tsn+1t\subseteq s_{n+1} and m[2e+n+1+3,2e+n+1+4)m\in[2^{e+n+1+3},2^{e+n+1+4}), define a coloring ce,t,m:𝒫(){0,1}c_{e,t,m}\colon\operatorname{\mathcal{P}}(\mathbb{N})\to\{0,1\} by

ce,t,m(B)={1if Φe(tB,m)=10otherwise.c_{e,t,m}(B)=\begin{cases}1&\text{if }\Phi_{e}(t\cup B,m)\!\downarrow\;=1\\ 0&\text{otherwise}.\end{cases}

Next, define a coloring cc on 𝒫()\operatorname{\mathcal{P}}(\mathbb{N}) by setting c(B)c(B) to be the sequence

c(B)=ce,t,m(B)e<n+1,tsn+1,m[2e+n+1+3,2e+n+1+4).c(B)=\langle c_{e,t,m}(B)\mid e<n+1,t\subseteq s_{n+1},m\in[2^{e+n+1+3},2^{e+n+1+4})\rangle.

Note that cc has finite range and can easily be seen to be Borel. Thus by the corollary to the Galvin-Prikry theorem, we can choose Bn+1B_{n+1} to be an infinite subset of BnB_{n} such that cc assigns the same value to all infinite subsets of Bn+1B_{n+1}. It is straightforward to check that Bn+1B_{n+1} has the properties desired.

Finally, recall that we must forbid some elements from AA. For each e<n+1e<n+1 and tsn+1t\subseteq s_{n+1}, if there is any mm in the interval [2e+n+1+3,2e+n+1+4)[2^{e+n+1+3},2^{e+n+1+4}) such that ce,t,mc_{e,t,m} has constant value 11 on Bn+1B_{n+1} then pick the least such mm and forbid it from AA. Also recall that at the end of the construction, we take AA to consist of all numbers not forbidden from AA.

This concludes the construction of GG and AA. All that remains now is to check that they have the necessary properties.

The set G\bm{G} satisfies all requirements. Fix any ee and we will show that requirement ee is satisfied. Suppose CC is an infinite subset of GG. We will show that either Φ(C)\Phi(C) is not a subset of AA or it is not infinite.

For each nn, define tn=Csnt_{n}=C\cap s_{n} and note that since GsnBnG\subseteq s_{n}\cup B_{n}, CC is compatible with (tn,Bn)(t_{n},B_{n}). We know that for each n>en>e, one of two possibilities holds:

  1. (1)

    there is some m[2e+n+3,2e+n+4)m\in[2^{e+n+3},2^{e+n+4}) such that mm is forbidden from AA and for all infinite sets DD compatible with (tn,Bn)(t_{n},B_{n}), Φ(D,m)=1\Phi(D,m)\!\downarrow\;=1

  2. (2)

    or for every m[2e+n+3,2e+n+4)m\in[2^{e+n+3},2^{e+n+4}) and every infinite set DD compatible with (tn,Bn)(t_{n},B_{n}), Φ(D,m)1\Phi(D,m)\neq 1.

On the one hand, if the first possibility holds for any nn then there is some mm such that Φ(C,m)=1\Phi(C,m)\!\downarrow\;=1 and mm is forbidden from AA, which implies that Φ(C)\Phi(C) is not a subset of AA.

On the other hand, if the second possibility always holds then for every n>en>e and every mm in the interval [2e+n+3,2e+n+4)[2^{e+n+3},2^{e+n+4}), Φ(C,m)1\Phi(C,m)\neq 1. Since the union of these intervals consists of all numbers greater than or equal to 22e+42^{2e+4}, this shows that either Φ(C)\Phi(C) is not total or it only contains numbers less than 22e+42^{2e+4} and thus is not infinite.

The set A\bm{A} has high density. For each ee, let AeA_{e} denote the set of numbers forbidden from AA on behalf of requirement ee. Note that A¯=eAe\overline{A}=\bigcup_{e}A_{e}. We will show that each AeA_{e} never has density greater than 1/2e+21/2^{e+2} and thus that their union never has density more than

e12e+2=12.\sum_{e\in\mathbb{N}}\frac{1}{2^{e+2}}=\frac{1}{2}.

This implies that A¯\overline{A} never has density more than 1/21/2 and thus that AA is 1/21/2-dense. Note, by the way, that when we say that each AeA_{e} never has density greater than 1/2e+21/2^{e+2}, we mean that for each nn, AeA_{e} is at most 1/2e+21/2^{e+2}-dense at nn (in the sense of Section 2.2); it is not good enough for it to just have upper density at most 1/2e+21/2^{e+2}.

So fix ee and we will argue that AeA_{e} never has density more than 1/2e+21/2^{e+2}. First consider a single interval of the form [2e+n+3,2e+n+4)[2^{e+n+3},2^{e+n+4}) for n>en>e. The only time numbers from this interval will be added to AeA_{e} is on step nn of the construction and on that step, at most one number will be added to AeA_{e} per subset of sns_{n}. Since sns_{n} has size exactly nn, this means at most 2n2^{n} such numbers will be added. Thus we have

|Ae[2e+n+3,2e+n+4)|2n=12e+3|[2e+n+3,2e+n+4)|.|A_{e}\cap[2^{e+n+3},2^{e+n+4})|\leq 2^{n}=\frac{1}{2^{e+3}}\cdot|[2^{e+n+3},2^{e+n+4})|.

In other words, in the interval [2e+n+3,2e+n+4)[2^{e+n+3},2^{e+n+4}), AeA_{e} has density at most 1/2e+31/2^{e+3}.

From this fact it can easily be shown that at each power of 22, AeA_{e} has density at most 1/2e+31/2^{e+3}, i.e. for each m>0m>0,

|Ae[2m1]|2m2e+3.|A_{e}\cap[2^{m}-1]|\leq\frac{2^{m}}{2^{e+3}}.

It remains to check that the density of AeA_{e} is high enough in-between powers of 22. But this follows easily from what we have already established. If 2mk<2m+12^{m}\leq k<2^{m+1} then we have

|Ae[k]|\displaystyle|A_{e}\cap[k]| |Ae[2m+11]|\displaystyle\leq|A_{e}\cap[2^{m+1}-1]|
2m+12e+3=2m2e+2k+12e+2\displaystyle\leq\frac{2^{m+1}}{2^{e+3}}=\frac{2^{m}}{2^{e+2}}\leq\frac{k+1}{2^{e+2}}

and thus AeA_{e} is at most 1/2e+21/2^{e+2}-dense at kk, as promised. ∎

The proof of Theorem 7.5 actually yields a somewhat stronger result. Let 𝖱𝖳<1\mathsf{RT}^{1}_{<\infty} denote the problem in which an instance is a finite partition A1,,AnA_{1},\ldots,A_{n} of \mathbb{N} and a solution is an infinite set BB which is a subset of some AiA_{i}. Roughly speaking, 𝖱𝖳<1\mathsf{RT}^{1}_{<\infty} is the problem corresponding to Corollary 2.2 of Seetapun’s theorem and thus the corollary below shows that there is no direct proof of our main theorem from Corollary 2.2 in the same sense that the theorem above showed there is no direct proof of our main theorem from Seetapun’s theorem.

Corollary 7.6.

𝖲𝖣soc𝖱𝖳<1\mathsf{SD}\nleq_{soc}\mathsf{RT}^{1}_{<\infty}.

Proof.

Let AA and GG be as in the proof of Theorem 7.5 and suppose for contradiction that 𝖲𝖣soc𝖱𝖳<1\mathsf{SD}\leq_{soc}\mathsf{RT}^{1}_{<\infty}. Then, thinking of AA as an instance of 𝖲𝖣\mathsf{SD}, we get a partition B1,,BnB_{1},\ldots,B_{n} of \mathbb{N} such that for any BiB_{i} and any infinite subset CBiC\subseteq B_{i}, CC computes an infinite subset of AA. However, since GG is infinite, there must be some BiB_{i} such that GBiG\cap B_{i} is infinite. Since GBiG\cap B_{i} is an infinite subset of GG, it does not compute any infinite subset of AA, which gives a contradiction. ∎

References

  • [1] Laurent Bienvenu, Adam R Day, and Rupert Hölzl. From bi-immunity to absolute undecidability. The Journal of Symbolic Logic, 78(4):1218–1228, 2009.
  • [2] Peter A Cholak, Carl G Jockusch, and Theodore A Slaman. On the strength of ramsey’s theorem for pairs. The Journal of Symbolic Logic, 66(1):1–55, 2001.
  • [3] JCE Dekker and John Myhill. Retraceable sets. Canadian journal of mathematics, 10:357–373, 1958.
  • [4] Rodney G Downey and Denis R Hirschfeldt. Algorithmic randomness and complexity. Springer Science & Business Media, 2010.
  • [5] Damir D. Dzhafarov. Cohesive avoidance and strong reductions. Proc. Amer. Math. Soc., 143(2):869–876, 2015.
  • [6] Damir D Dzhafarov and Carl G Jockusch. Ramsey’s theorem and cone avoidance. The Journal of Symbolic Logic, 74(2):557–578, 2009.
  • [7] Fred Galvin and Karel Prikry. Borel sets and ramsey’s theorem. The Journal of Symbolic Logic, 38(2):193–198, 1973.
  • [8] Denis R Hirschfeldt. Slicing the truth: On the computable and reverse mathematics of combinatorial principles. World Scientific, 2015.
  • [9] Carl G Jockusch. Uniformly introreducible sets. The Journal of Symbolic Logic, 33(4):521–536, 1969.
  • [10] Jiayi Liu. 𝖱𝖳22\mathsf{RT}_{2}^{2} does not imply 𝖶𝖪𝖫0\mathsf{WKL}_{0}. J. Symb. Log., 77(2):609–620, 2012.
  • [11] Benoit Monin and Ludovic Patey. Π10\Pi^{0}_{1}-encodability and omniscient reductions. Notre Dame Journal of Formal Logic, 60(1):1–12, 2019.
  • [12] Benoit Monin and Ludovic Patey. Pigeons do not jump high. Advances in Mathematics, 352:1066–1095, 2019.
  • [13] Benoit Monin and Ludovic Patey. 𝖲𝖱𝖳22\mathsf{SRT}_{2}^{2} does not imply 𝖱𝖳22\mathsf{RT}_{2}^{2} in ω\omega-models. Adv. Math., 389:32, 2021. Id/No 107903.
  • [14] Benoit Monin and Ludovic Patey. The weakness of the pigeonhole principle under hyperarithmetical reductions. J. Math. Log., 21(3):41, 2021. Id/No 2150013.
  • [15] David Seetapun and Theodore A Slaman. On the strength of ramsey’s theorem. Notre Dame Journal of Formal Logic, 36(4):570–582, 1995.
  • [16] Nikolai K. Vereshchagin. Kolmogorov complexity conditional to large integers. Theoretical Computer Science, 271(1-2):59–67, 2002. Kolmogorov complexity.