Halin’s Infinite Ray Theorems: Complexity and Reverse Mathematics: Version E††thanks: All the authors were partially supported by NSF Grant DMS-1161175.
Abstract
Halin [1965] proved that if a graph has many pairwise disjoint rays for each then it has infinitely many pairwise disjoint rays. We analyze the complexity of this and other similar results in terms of computable and proof theoretic complexity. The statement of Halin’s theorem and the construction proving it seem very much like standard versions of compactness arguments such as König’s Lemma. Those results, while not computable, are relatively simple. They only use arithmetic procedures or, equivalently, finitely many iterations of the Turing jump. We show that several Halin type theorems are much more complicated. They are among the theorems of hyperarithmetic analysis. Such theorems imply the ability to iterate the Turing jump along any computable well ordering. Several important logical principles in this class have been extensively studied beginning with work of Kreisel, H. Friedman, Steel and others in the 1960s and 1970s. Until now, only one purely mathematical example was known. Our work provides many more and so answers Question 30 of Montalbán’s Open Questions in Reverse Mathematics [2011]. Some of these theorems including ones in Halin [1965] are also shown to have unusual proof theoretic strength as well.
Mathematics Subject Classification 2020: Primary 05C63, 03D55, 03B30; Secondary 03D80, 03F35, 05C38, 05C69, 05C70
1 Introduction
In this paper we analyze the complexity of several results in infinite graph theory. These theorems are said to be ones of Halin type or, more generally, of ubiquity theory. The classical example is a theorem of Halin [11]: If a countable graph contains, for each , a sequence of disjoint rays (a ray is a sequence of distinct vertices such that there is an edge between each and ) then it contains an infinite such sequence of rays. (Note: As will be described in Definition 3.2, when we talk about disjoint rays we always mean pairwise disjoint.) Halin actually deals with arbitrary graphs and formulates the result differently. The uncountable cases, however, are essentially just counting arguments. We deal only with countable structures but discuss his formulation in §6. This standard formulation of his theorem seems like a typical compactness theorem going from arbitrarily large finite collections of objects to an infinite collection. The archetypical example here is König’s Lemma: If a finitely branching tree has paths of length for every then it has a branch, i.e. an infinite path. In outline, a modern proof of Halin’s theorem for countable graphs (due to Andreae, see [5, Theorem 8.2.5(i)]) seems much like that of König’s Lemma (and many others in infinite graph theory). The construction of the desired sequence of rays proceeds by a recursion through the natural numbers in which each step is a simple procedure. While the procedure is much more delicate than for König’s Lemma, it is basically of the same complexity. It uses Menger’s theorem for finite graphs at each step but this represents a computable procedure (for finite graphs). The other parts of the step depend on the same type of information as in König’s Lemma. They ask, for example, if various sets (computable in the given graph) are nonempty or infinite. Nonetheless, we prove that the complexity of this construction and theorem are much higher than that for König’s Lemma or other applications of compactness. The concepts from graph theory, computability theory and proof theory/reverse mathematics that we need for our analysis are discussed in §3. Basic references for terminology, background and standard results not explicitly stated or otherwise attributed are Diestel [5] for graph theory; Rogers [22] and Sacks [24] for computability theory; and for reverse mathematics Simpson [28] with an approach which is primarily in terms of formal systems and Hirschfeldt [13] with one primarily emphasizing computability.
We follow two well established procedures for measuring the complexity of constructions and theorems. The first is basically computability theoretic. It has its formal beginnings in the 1950s but has much earlier roots in constructive or computable mathematics reaching back to antiquity. (See Ershov et al. [6] for history and surveys of the approach in several areas of combinatorics, algebra and analysis.) The measuring rod here is relative computability. We say a set of natural numbers is (Turing) computable from a set , , if there is an algorithm (say on a Turing machine or any other reasonable model of general computation) that, when given access to all membership facts about (an oracle for ) computes membership in . The standard hierarchies of complexity here are based on iterations of the Turing jump. This operator takes to , the halting problem relativized to , i.e. the set of programs with oracle for , , such that halts on input . For example, if the tree of König’s Lemma is computable in then there is a branch computable in the double jump of .
The second approach is proof theoretic. It measures the complexity of a theorem by the logical strength of the axioms needed to prove it. This approach also has a long history but the formal subject, now called reverse mathematics, starts with H. Friedman’s work in the 1970s (e.g. [7, 8]). One compares axiomatic systems and by saying that is stronger than , ( proves ) if one can prove every sentence from the axioms of . Of course, we know what it means for to be provable in . The goal here is to characterize to the extent possible the axioms needed to prove a given mathematical theorem . To this end, one begins with a weak base theory. Then one wants to find a system such that not only does but also (with the weak base theory) proves all the axioms of . Hence the name reverse mathematics as we seek to prove the “axioms” of from the theorem . Typically, the systems here are formalized in arithmetic with quantification over sets as well as numbers. The standard base theory () corresponds to the axioms needed to do computable constructions. Stronger systems are then usually generated by adding comprehension axioms which assert the existence of specific families of sets. For example, a very important system is . It is equivalent in the sense of reverse mathematics just described to König’s Lemma. Formally, it asserts that every subset of the natural numbers defined by a formula that quantifies only over numbers (and not sets) exists. This is also equivalent to asserting that for every set , the set exists.
The early decades of reverse mathematics were marked by a large variety of results characterizing a wide array of theorems and constructions as being one of five specific levels of complexity including and . Each of these systems (Simpson’s “big five”) have corresponding specific recursion theoretic construction principles. In more recent decades, there has been a proliferation of results placing theorems and constructions outside the big five. Sometimes these are inserted linearly and sometimes with incomparabilities. They are now collectively often called the “zoo” of reverse mathematics. (See https://rmzoo.math.uconn.edu/diagrams/ for pictures.)
Theorems and constructions in combinatorics in general, and graph theory in particular, have been a rich source of such denizens of this zoo. Almost all of them have fallen below (König’s Lemma) and so have the objects they seek constructible computably in finitely many iterations of the Turing jump. Ramsey theory, in particular, has provided a very large class of constructions and theorems of distinct complexity. One example of the infinite version of a classical theorem of finite graph theory that is computationally and reverse mathematically strictly stronger than is König’s Duality Theorem () for countable graphs. (Every bipartite graph has a matching and a cover consisting of one vertex from each edge of the matching.) The proofs of this theorem for infinite graphs (Podewski and Steffens [21] for countable and Aharoni [1] for arbitrary ones) are not just technically difficult but explicitly used both transfinite recursions and well orderings of all subsets of the given graph. These techniques lie far beyond . Aharoni, Magidor and Shore [2] proved that this theorem is of great computational strength in that there are computable graphs for which the required matching and cover compute all the iterations of the Turing jump through all computable well-orderings. They also showed that it was strong reverse mathematically as it implied , the standard system above used to deal with such transfinite recursions. Some of the lemmas used in each of the then known proofs were shown to be equivalent to the next and final of the big five systems, - and of corresponding computational strength. Simpson [27] later provided a new proof of the theorem using logical methods that avoided these lemmas and showed that the theorem itself is equivalent to and so strictly weaker than the lemmas both computationally and in terms of reverse mathematics.
The situation for the theorems of Halin type that we study here is quite different. The standard proofs do not seem to use such strong methods. Nonetheless, as we mentioned above, the theorems are much stronger than with some versions not even provable in . We prove that these theorems occupy a few houses in the area of the reverse mathematics zoo devoted to what are called theorems (or theories) of hyperarithmetic analysis, THAs (Definition 3.13). Computationally, for each computable well ordering , there is a computable instance of any THA which has all of its required objects Turing above , the th iteration of the Turing jump. On the other hand, they are computationally and proof theoretically much weaker than and so . The point here is that there is a single computable graph such that the matching and cover required by lies above for all the computable well-orderings , while for each computable instance of a THA there is a computable well-ordering such that computes the desired object. In our cases, the instances are graphs with arbitrarily many disjoint rays and the desired object is an infinite sequence of disjoint rays. (The general usage of terms like instances and solutions of a theorem or principle is described at the end of §3.)
Beginning with work of Kreisel [15], H. Friedman [9], Steel [29] and others in the 1960s and 1970s and continuing into the last decade (by Montalbán [16, 17], Neeman [19, 20] and others), several axiomatic systems and logical theorems were found to be THAs and proven to lie in a number of distinct classes in terms of proof theoretic complexity. Until now, however, there has been only one mathematical but not logical example, i.e. one not mentioning classes of first order formulas or their syntactic complexity. This was a result (INDEC) about indecomposability of linear orderings in Jullien’s thesis [14] (see Rosenstein [23, Lemma 10.3]). It was shown to be a THA by Montalbán [16].
The natural quest then became to find out if there are any other THAs in the standard mathematical literature. The issue was raised explicitly in Montalbán’s “Open Questions in Reverse Mathematics” [18, Q30]. As our answer, we provide many examples. Most of them are provable in a well known system above gotten by adding on a weak form of the axiom of choice ().
Several of the basic Halin type theorems (the defined after Definition 3.4) have versions (the of Definition 6.1) like those appearing in the original papers that show that there are always families of disjoint rays of maximal cardinality which are of the same computational strength as the basic versions (Proposition 6.3 and Corollary 6.4). On the other hand, the are strictly stronger proof theoretically than the because they imply more induction than is available in (Theorem 6.8 and Corollary 6.9). Two of the variations we consider are as yet open problems of graph theory ([3] and Bowler, personal communication). We show that if we restrict the class of graphs to directed forests the principles are not only provable but reverse mathematically equivalent to . Note that as [28, IX.4.7], these theorems are not provable even in or from (Corollary 6.14). We do not know of other mathematical but nonlogical theorems of this strength. Other versions that require maximal sets of rays (Definition 6.17) are much stronger and, in fact, equivalent to - (Theorem 6.18).
2 Outline of Paper
Section 3 discusses basic concepts and background information. The first subsection (3.1) provides what we need from graph theory. Almost all the definitions are standard. At times we give slight variations that are equivalent to the standard ones but make dealing with the computability and proof theoretic analysis easier. We also state the theorems of Halin and some variants that are the main targets of our analysis.
The second subsection (3.2), assumes an intuitive view of computability of functions such as having an algorithm given by a program in any standard computer language (possibly with access to an “oracle” providing information about a given set or function). It then gives the standard notions and theorems that can be found in basic texts on computability theory needed to follow our analysis of the computational complexity of the graph theoretic theorems we study. In particular, it notes the Turing jump operator and its iterations along countable well orderings. These are our primary computational measuring rods. The final subsection (3.3) provides the syntax and semantics for the formal systems of arithmetic that are used to measure proof theoretic complexity. It also describes the standard basic axiomatic systems and their connections to the computational measures of the previous subsection. It includes the formal definition of the class of theorems which includes most of our graph theoretic examples, the THAs, Theorems (or Theories) of Hyperarithmetic Analysis. These are defined in terms of the transfinite iterations of the Turing jump and the hyperarithmetic sets of the previous subsection. In addition it defines a weak version of the axiom of choice that is an early well known example of such theories and plays a crucial role in our analysis.
Section 4 provides the proof that Halin’s original theorem (Definition 3.4) is computationally very complicated. For example, given any iteration of the Turing jump, there is a computable graph satisfying the hypotheses of such that any instance of its conclusion computes . Indeed, is a THA. At times, theorems or lemmas are stated in terms of the formal systems of §3.3, but the proofs rely only on the computational notions of §3.2.
Section 5 studies several variations of Halin’s where we consider directed as well as undirected graphs, edge rather than vertex disjointness for the rays and double as well as single rays. (See Definitions 3.1 and 3.2 and the discussion after Definition 3.4.) We provide reductions over between many of the pairs of the eight possible variants. The proofs of these reductions proceed purely combinatorially by providing one computational process that takes an instance of some , i.e. a graph satisfying its hypotheses, and produces a graph satisfying the hypotheses of another and another computable process that takes any solution to the instance, i.e. any sequence of rays satisfying the conclusion of , and produces a solution to the original instance of . (See Propositions 5.3, 5.5 and 5.7 and the associated Lemmas. An additional reduction using a stronger base theory is given in the next section (Theorem 6.15).)
We then show that five of the eight possible variants of are THAs (Theorem 5.1). As mentioned in §1, of the remaining three, two are still open problems in graph theory. We do, however, have an analysis of their restrictions to special classes of graphs in Theorem 5.16 and §6.1. The last of the variations, , has been proven more recently by Bowler, Carmesin, Pott [3] using more sophisticated methods than the other results. We have some lower bounds (Theorem 5.9) but we have yet to fully analyze the complexity of their construction.
In the next section (§6) we study some variations of that ask for different types of maximality for the solutions. The first sort actually follow the original formulation of in Halin [11]: In any graph there is a set of disjoint rays of maximum cardinality. For uncountable graphs this amounts to a basic counting argument on uncountable cardinals as all rays are countable. When restricted to countable graphs these variations, , are easily seen to be equivalent to our more modern formulation by induction. Technically, the induction used is for formulas () which is not available in . More specifically we show (Proposition 6.3) that and are equivalent (over ). As the definition of THAs only depends on standard models where full induction holds, if is a THA then so is .
We then prove that these maximal cardinality variants are strictly stronger proof theoretically than the basic (when they are known to be provable in ). This is done by showing (Theorem 6.8 and the Remark that follows it) that the relevant all imply weaker versions of that are analogous to the restrictions of embodied in weak (or unique)- and finite- (Definitions 7.1 and 7.2). In all the cases, it is enough induction to prove (with the apparatus of the basic ) the consistency of and so by Gödel’s second incompleteness theorem they cannot be proved in (Corollary 6.9).
As for proving full induction from an we are in much the same situation mentioned above for and . In particular, and for directed forests each proves as well as (Theorems 6.12 and 6.13) and so are equivalent to . As before, this shows that they are strictly stronger than (Corollary 6.14). Indeed, as mentioned at the end of §1, they are not even provable in . We do not know of any other mathematical theorem with this level of reverse mathematical strength.
The second variation of maximality, , studied in §6.2 is also mentioned in the original Halin paper [11]. It asks for a set of disjoint rays which is maximal in the sense of set containment. Of course, this follows immediately from Zorn’s Lemma for all graphs. For countable graphs we provide a reverse mathematical analysis, showing that each of the is equivalent to - (Theorem 6.18).
In §7, we discuss the reverse mathematical relationships between the THAs associated with variations of Halin’s theorem and previously studied THAs as well as one new logical THA (finite- of Definition 7.1). Basically, all the (and so ) imply H. Friedman’s (Definition 7.4) by Theorem 7.7 and finite- (Theorem 7.3). On the other hand, none of them are implied by it (Theorem 7.10) or by - (Definition 7.8 and Theorem 7.9). does, however, imply finite- which is not implied by weak (unique)- (Goh [10]). Figure 4 summarizes many of the known relations with references.
In the penultimate section (§8) we study the only use of in each of our proofs of . It consists of which says we can go from the hypothesis that there are arbitrarily many disjoint rays to a sequence in which each is a sequence of many disjoint rays. We analyze the strength of the and the weakenings of which each take the existence of such a sequence as its hypothesis in place of there being arbitrarily many disjoint rays. For example, for all the which are consequences of and so are THAs, is equivalent to over (Corollary 8.5) and so all of them are also THAs. For the same choices of , proves over . While a natural strengthening of does imply and indeed is equivalent to it (Theorem 8.9), we do not know if itself implies . All we can prove is that it is not a consequence of (Theorem 8.10).
In the last section (§9), we mention some open problems.
3 Basic Notions and Background
We begin with basic notions and terminology from graph theory. At times we use formalizations that are clearly equivalent to more standard ones but are easier to work with computationally or proof-theoretically. The following two subsections supply background and basic information about the standard computational and logical/proof theoretic notions that we use here to measure the complexity of the graph theorems and constructions that we analyze in the rest of this paper. Note that we denote the set of natural numbers by when we may be thinking of them in a model of arithmetic as in §3.3 and by when we emphasize that we specifically want the standard natural numbers.
3.1 Graph Theoretic Notions
Definition 3.1.
A graph is a pair consisting of a set (of vertices) and a set of unordered pairs with from (called edges). These structures are also called undirected graphs (or here U-graphs). A structure of the form as above is a directed graph (or here D-graph) if consists of ordered pairs of vertices with . To handle both cases simultaneously, we often use to stand for undirected (U) or directed (D). We then use to stand for the appropriate kind of edge, i.e. or .
An -subgraph of the -graph is an -graph such that and . It is an induced -subgraph if .
Definition 3.2.
An -ray in H is a pair consisting of an -subgraph of and an isomorphism from with edges for to . (Note that this implies that the range of is the set .) We say that the ray begins at . We also describe this situation by saying that contains the -ray . We sometimes abuse notation by saying that the sequence of vertices is an -ray in . A tail of an -ray is a final segment of said -ray. Similarly we consider double X-rays where the isomorphism is from the integers with edges for . A tail of a double -ray is a final segment of said -ray, or an initial segment of said -ray considered in reverse order.
We use -ray to stand for either a (single) ray (S) or double ray (D) and so we have, in general, --rays or just -rays if the type of graph (U or D) is already established. For brevity, when we describe rays we will often only list their vertices in order instead of defining and explicitly. However the reader should be aware that we always have and in the background.
contains many --rays for if there is a sequence such that each is a --ray in (with ). contains many disjoint (or vertex-disjoint) --rays if the are pairwise disjoint. contains many edge-disjoint --rays if the are pairwise disjoint. We often use to stand for either vertex (V) or edge (E) as in the following definitions.
An -graph contains arbitrarily many Y-disjoint Z-X-rays if it contains many such rays for every .
An -graph contains infinitely many Y-disjoint Z-X-rays if there is an -subgraph of and a sequence such that each is a --ray in (with ) such that the or , respectively for , are pairwise disjoint and and .
Definition 3.3.
An -path in an -graph is defined similarly to single rays except that the domain of is a proper initial segment of instead of itself. Thus they are finite sequences of distinct vertices with edges between successive vertices in the sequence. If is a path, we say it is a path of length between and . Our notation for truncating and combining paths , and is as follows: , , and we use concatenation in the natural way, e.g., if the union of , and is a path, we denote it by . We treat rays as we do paths in this notation, as long as it makes sense, writing, for example, for the ray which is gotten by starting at an element of ; is the path which is the initial segment of ending in and we use concatenation as for paths as well.
The starting point of the work in this paper is a theorem of Halin [11] that we call the infinite ray theorem as expressed in [5, Theorem 8.2.5(i)].
Definition 3.4 (Halin’s Theorem).
, the infinite ray theorem, is the principle that every graph which contains arbitrarily many disjoint rays contains infinitely many disjoint rays.
The versions of Halin’s theorem which we consider in this paper allow for to be an undirected or a directed graph and for the disjointness requirement to be vertex or edge. We also allow the rays to be single or double. The corresponding versions of Halin’s Theorem are labeled as for appropriate values of and to indicate whether the graphs are undirected or directed (U or D); whether the disjointness refers to the vertices or edges (V or E) and whether the rays are single or double (S or D), respectively, in the obvious way. We often state a theorem for several or all and then in the proof use “graph”, “edge” and “disjoint” unmodified with the intention that the proof can be read for any of the cases. This is convenient for minimizing repetition in some of our arguments.
We will also consider restrictions of these theorems to specific families of graphs. We need a few more notions to define them.
Definition 3.5.
A tree is a graph with a designated element called its root such that for each vertex there is a unique path from to . A branch on (or in) is a ray that begins at its root. We denote the set of its branches by and say that is well-founded if and otherwise it is ill-founded. A forest is an effective disjoint union of trees, or more formally, a graph with a designated set (of vertices called roots) such that for each vertex there is a unique such that there is a path from to and, moreover, there is only one such path. In general, the effectiveness we assume when we take effective disjoint unions of graphs means that we can effectively (i.e. computably) uniquely identify each vertex in the union with the original vertex (and the graph to which it belongs) which it represents in the disjoint union.
A directed tree is a directed graph such that its underlying graph where is a tree. A directed forest is a directed graph whose underlying graph is a forest.
Definition 3.6.
An -graph is locally finite if, for each , the set of neighbors of is finite. A locally finite -tree is also called finitely branching. (Note this does not mean there are finitely many branches in the tree.)
Of course, there are many well known equivalent definitions of trees and associated notions. We have given one possible set of graph-theoretic definitions. In the case of undirected graphs our definition is equivalent to all the standard ones. Readers are welcome to think in terms of their favorite definition. Note, however, we are restricting ourselves to what would (in set theory) be called countable trees with all nodes of finite rank. Thus, we typically think of trees as subtrees of , i.e. the sets of finite strings of numbers (as vertices) with an edge between and if and only if they differ by one being an extension of the other by one element, e.g. .
3.2 Computability Hierarchies
While we may cite results about uncountable graphs, all sets and structures actually studied in this paper will be countable. Thus for purposes of defining their complexity, we can think of all of them as being subsets of, or relations or functions on, .
We do not give a formal stand alone definition of computability for sets or functions but assume an at least intuitive grasp of some model of computation such as by a Turing or Register machine that has unbounded memory and is allowed to run for unboundedly many steps. (We do provide in §3.3 a definition via definability in arithmetic that is equivalent to the formal versions of machine model definitions.) Thus we say a function is computable if there is a program for one of these machines that computes as output when given input . A set is computable if its characteristic function is computable. Note that as the alphabets or our languages are finite, there are only countably many programs and as our formation rules are effective, we have a computable list of the programs and hence one, , of the partial functions they compute. (They are only partial as, of course, some programs fail to halt on some inputs.)
Fundamental to measuring the relative computational complexity of sets or functions is the notion of machines with oracles and Turing reduction. Given a set or function we consider machines augmented by the ability to produce or if it has already produced . We say that such a machine is one with an oracle for or . We then say that is computable from (or Turing reducible to) if there is a machine with oracle which computes via some reduction . We write this as . We say is of the same (Turing) degree as , , if and . We use all the same terminology and notations for functions.
The first level beyond the computable in our basic hierarchy of computable complexity is given by the halting problem converges that is is set of such that the computation of the th machine on input eventually halts. We then define an operator on sets converges that is is the set of such that the computation of the th machine with oracle , on input eventually halts. (It is easy to see that .) The crucial fact here is the undecidability of the halting problem (for every oracle), i.e. for every , is strictly above in terms of Turing computability. The other basic fact that we need about is that it is computably enumerable, i.e. there is a computable function whose range is . If we say that is enumerated in, or enters, at (stage) . If we view as defined by using the empty oracle , the procedure that takes us from the halting problem to the Turing jump by replacing as oracle by is an instance of a general procedure called relativization. It takes any computable function or proof about computable functions or degrees (i.e. ones with oracle ) to the same function, or proof about functions, computable in (or degrees above that of ). Almost always this procedure trivially transforms correct proofs with oracle to ones with arbitrary oracle . Typically, this transformation keeps the same programs doing the required work with any oracle. For example, is computably enumerable in (or relative to ), i.e. there is a function whose range is and this can be taken to be the same such that enumerates . We also use to denote the set of numbers enumerated in (or that have entered) by stage . This phenomena of the procedure or result not depending on the particular oracle or depending in a fixed computable way on some other parameters is described as its being uniform in the oracle or other parameters. We describe an important example of uniformity in Remark 3.9.
We can now generate a hierarchy of computational complexity by iterating the jump operator beginning with any set : ; . While the finite iterations of the jump capture most construction techniques and theorems in graph theory (and most other areas of classical countable/separable mathematics), we will be interested in ones that go beyond such techniques and proofs. The basic idea is that we continue the hierarchy by iteration into the transfinite while still tying the iteration to computable procedures.
Definition 3.7.
We represent well-orderings or ordinals as well-ordered relations on . Typically such ordinal notations are endowed with various additional structure such as identifying , successor and limit ordinals and specifying cofinal -sequences for the limit ordinals. If we have a representation of then restricting the well-ordering to numbers in its domain provides representations of each ordinal . We generally simply work with ordinals and omit concerns about translating standard relations and procedures to the representation. An ordinal is recursive (in a set ) if it has a recursive (in ) representation. For a set and ordinal (notation) computable from , we define the transfinite iterations of the Turing jump of by transfinite induction on : ; and for a limit ordinal , (or as the effective disjoint sum over the in the specified cofinal sequence in ).
Definition 3.8.
, the collection of all sets hyperarithmetic in consists of those sets recursive in some for an ordinal recursive in . We say that is hyperarithmetic in or hyperarithmetically reducible to , if .
These sets too, will be characterized by a definability class in arithmetic in §3.3. For now we just note that they clearly go far beyond the sets computable from the finite iterations of the jump.
The computational strength of our graph theoretic theorems such as is measured by this hierarchy as we will show that, for every set and every set hyperarithmetic in , there is a graph computable from which satisfies the hypotheses of but for which any collection of rays satisfying its conclusion computes . On the other hand, placing an upper bound on the strength of requires analyzing its proof and the principles used in it. The relevant one is a form of the axiom of choice. We define it in the next subsection along with a general class of such principles, the theorems/theories of hyperarithmetic analysis which are, computationally, the primary objects of our analysis in this paper.
We note one important well known basic fact relating the jumps of to trees computable from . We will need it for our proofs that and its variants are computationally complex enough to compute all the sets hyperarithmetic in any given set (as the instances of the graphs range over graphs computable from ).
Remark 3.9.
For any set and any ordinal computable from , there is a sequence computable from of trees (necessarily) computable from such that each tree has exactly one branch and is of the same complexity as , i.e. . The procedure for computing this sequence is uniform in and the index for the program computing the well ordering from , i.e. there is one computable function that when given an oracle for , an index for (i.e. the such that is the well ordering ) and a in the ordering, computes the whole sequence and the indices for the reductions between and . (See, e.g. [25, Theorem 2.3]). We may also easily assure that the are effectively disjoint so that their union is a forest.
Some versions of the variations on (see §6.2) that call for types of maximality for the infinite set of disjoint rays are stronger both computationally and proof theoretically than the described above. Their computational strength is captured by a kind of jump operator that goes beyond all the hyperarithmetic ones. It captures the ability to tell if a computable ordering is a well-ordering.
Definition 3.10.
The hyperjump of , , is the set is (the characteristic function of) a subtree of which is well-founded.
This operator also corresponds to a syntactically defined level of comprehension as we note in §3.3.
3.3 Logical and Axiomatic Hierarchies
The basic notions from logic that we need here are those of languages, structures and axiomatic systems and proofs. As we will deal only with countable sets and structures, we can assume that we are dealing just with the natural numbers with a way to define and use sets and functions on them. Thus, at the beginning, we have in mind the natural numbers along with the usual apparatus of the language of (first order) arithmetic, say and along with the syntax of standard first order logic (the Boolean connectives and ; the variables such as and ranging over the numbers with the usual quantifiers and as well as the standard equality relation ). A structure for this language is a set along with elements for and , binary functions for and and a binary relation for . We also need a way of talking about subsets of (or functions on) the numbers. We follow the standard practice in reverse mathematics of using sets and defining functions in terms of their graphs. So we expand our language by adding on new classes of (second order) variables such as and and the associated quantifiers and along with a new relation symbol between numbers and sets.
A structure for this language is one of the form where its restriction is a structure for first order arithmetic and is a specified nonempty collection of subsets of disjoint from , the set of “numbers” of , over which the second order quantifiers and variables of our language range. It is called the second order part of . The usual membership symbol always denotes the standard membership relation between elements of and subsets of that are in and the language only allows atomic formulas using which are of the form for a term of the first order language and is a second order variable. So a sentence is true in , , if first order quantification is interpreted as ranging over , second order quantification ranges over and the relations and functions of the language are as described. This specifies the semantics for second order arithmetic. Note that, following [28], we do not take equality for sets to be a primitive relation on this structure. The notation for it is viewed as being defined by .
Proof theoretic notions deal with all possible structures for the language and axiom systems to specify what we need in any particular argument. For most of our purposes and all of the computational ones, one can restrict attention to standard models of arithmetic, i.e. ones with and some with the usual interpretations of the functions and relations. We generally abbreviate these structures as with , or simply , as all the functions and relations are then fixed.
We view the syntax as one for a two sorted first order logic. So the (first order) variables range over the first sort () and the second order ones over the second sort (). We assume any standard proof theoretic system with the caveat that is interpreted as true equality with the equality axioms included only for . For it is a relation defined as above. This generates the provability notion used above to define our notion of logical strength and equivalences of theories (sets of sentences often called axioms) as above. We now define the standard weak base theory used to define the logical strength of mathematical theorems as described above. We then define a few other common systems that will be used later. The formal details can be found in [28].
Each axiomatic subsystem of second order arithmetic that we consider contains the standard basic axioms for , , and (which say that is a discrete ordered semiring) and an Induction Axiom:
() | . |
Typically axiom systems for second order arithmetic are defined by adding various types of set existence axioms although at times additional induction axioms are used as well. In order to define them we need to specify various standard syntactic classes of formulas determined by quantifier complexity. As usual, we add to our language bounded quantifiers and for first order (i.e. arithmetic) terms defined in the standard way. We typically denote formulas by capital Greek letters except that the indexed and refer, as above, to our fixed enumeration of the Turing machines and associated partial functions.
Definition 3.11.
The and formulas of second order arithmetic are just the ones with only bounded quantifiers but we allow parameters for elements of either or when working with a structure . Proceeding inductively, a formula is () if it is of the form () where is (). We assume some computable coding of all these formulas (viewed as strings of symbols from our language) by natural numbers. We say is arithmetic if it is or for some . It is () if it is of the form () where is arithmetic. (One can continue to define and in the natural way but we will not need to consider such formulas here.) We say a set is in one of these classes relative to (i.e. with as a parameter) if there is a formula such that . If is both in and in it is called in .
We mention a few additional standard connections between the syntactic complexity of the definition of a set and ’s properties in terms of computability and graph theoretic notions. They can all be found in [22].
Proposition 3.12.
The sets are in . A set is computable in if and only if it is in . More generally, it is computable in if and only if it is in . It is hyperarithmetic in if and only if it is in . There is a computable function such that if is in via the formula with code then for every , is (the characteristic function of) a tree such that has a branch.
The first system for analyzing the proof theoretic strength of theorems and theories in reverse mathematics is just strong enough to prove the existence of the computable sets and so supplies us with all the usual computable functions such as pairing or more generally those coding finite sequences as numbers. In particular, it provides the predicates defining the (codes of) the partial computable functions and the relations saying the computation halts in many steps with output . Thus we have the basic tools to define and discuss Turing reducibility and the Turing jump. It is our weak base theory and is assumed to be included in every system we consider.
() Recursive Comprehension Axioms: In addition to the ones mentioned above, its axioms include the schemes of recursive (generally called ) comprehension and induction:
(-) | for all |
---|---|
formulas and formulas in which is not free. | |
() | for all formulas . |
Note that these formulas may have free set or number variables. As usual, the existence assertion of the axiom is taken to mean that for each instantiation of the free variables (by numbers or sets, as appropriate, called parameters) there is an as described. We take this for granted as well as the restriction that the is not free in the rest of the formula in all of the set existence axioms of any of our systems.
suffices to define and manipulate the basic notions of computability listed at the beginning of Section 3.2. The standard models of are just those whose second order part is closed under Turing reduction and disjoint union (). As suggested above what are now often called the computable in sets which are, as mentioned above, the in sets, were originally called the sets recursive in . Hence the terminology in RCA0.
Any axiom system we consider from now on will be assumed to include . If we have some axiom scheme or principle we typically denote the system formed by adding it to by . We next move up to the arithmetic comprehension axiom and its system.
() for every arithmetic formula .
As mentioned above the are defined by a formula with as a parameter. So one can show that this system is equivalent (over RCA0) to the totality of the Turing jump operator, i.e. for every , exists. Its standard models are those of RCA0 whose second order part is also closed under Turing jump. It is also equivalent (in the sense of reverse mathematics) to König’s Lemma, which asserts that every finitely branching tree with paths of arbitrarily long length has a branch.
In general, we say one system of axioms is logically or reverse mathematically reducible to another over one if for every sentence . Note that and/or may be a single sentence or theorem. We say that and are equivalent over if each is reducible to the other. If no system is specified we assume that RCA0 is intended.
As we will not deal with it, we have omitted the formal definition of the usual system WKL0 which falls strictly between RCA0 and ACA0. It is characterized by the restriction of König’s Lemma to trees that are subsets of , the tree of finite binary strings under extension.
The next system of the five basic ones after ACA0 is ATR0. Its defining axiom says that arithmetic comprehension can be iterated along any countable well-order and so implies the existence of the sets hyperarithmetic in for each but is computationally stronger than this assumption. As usual the formal definition can be found in [28].
Instead, we formally describe the computationally defined class of theorems/theories that are the main focus of this paper and include several variations of . The definition is semantic, not axiomatic and involves only standard models. (Indeed by Van Wesep [30, 2.2.2], there can be no axiomatic characterization of this class in second order arithmetic.)
Definition 3.13.
A sentence (theory) is a theorem (theory) of hyperarithmetic analysis (THA) if
-
1.
For every , and
-
2.
For every , if and then .
It is worth pointing out some of the relations between THAs and . THAs are defined by only using standard models and iterations of the jump over true well orderings. talks about all models of and asserts the existence of iterates of the jump over all orderings that appear well-founded in the model. Thus for standard models it implies the second clause of the definition of THAs (Definition 3.13). On the other hand, there is a recursive linear order with no hyperarithmetic infinite descending sequence and so it seems well-founded in but it has a well-founded part longer than every recursive ordinal. Thus iterating the jump along this ordering would yield a set strictly Turing above every hyperarithmetic set. In particular, is not a model of which therefore is not a THA. (See e.g. [28, V.2.6].)
The last of the standard axiomatic systems, -CA0, is characterized by the comprehension axiom for formulas:
(-CA) for every formula .
Remark 3.14.
The hyperjump, , is clearly a set with parameter . In fact, every set with parameter is reducible to . Indeed, there is a computable function such that for every index for a formula with parameter and every , [22, Corollary 16.XX(b)]. Thus - corresponds to closure under the hyperjump. We will see it appear as equivalent to a version of where we ask for a maximal set of disjoint rays in Theorem 6.18.
For this paper, the most important other existence axiom is a restricted form of the axiom of choice.
(-AC) where is arithmetic and is the column of .
A more common but clearly equivalent version of this axiom allows to be . A variant commonly called weak--AC (introduced in Definition 7.1 as unique--AC) requires the corresponding to be arithmetic. To make these and other choice axioms uniform we have adopted the format required elsewhere and equivalent here to be used for all the variations. The system -AC0 is well known to be a THA (essentially in Kreisel [15]). Thus it is strictly stronger than . On the other hand, it is strictly weaker than . (It is known that -AC0 [28, V.8.3] but the converse fails as -AC is true in HYP while , as we have pointed out, is not.) This choice axiom plays a crucial role in our analysis because we provide the upper bound on the strength of most of our theorems by showing that they follow from -AC0. This provides the computational upper bound for being a THA as any consequence of a THA must also satisfy Definition 3.13(1). Thus the bulk of our proofs for the computational complexity of the theorems we study consist of showing that they imply Definition 3.13(2), i.e. closure under “hyperarithmetic in”.
Over the past fifty years, several other logical axioms have been shown to be THA. We will discuss some of them in §7. However, as we discussed in §1, only one somewhat obscure purely mathematical theorem was previously known to be a THA. We provide several more in this paper (Theorem 5.1, Corollary 6.4 and Theorem 6.13). We also introduce a new logical axiom, finite- (Definition 7.2) which is a THA as well.
For those interested in the proof theory and so nonstandard models, we also at times explicitly consider the induction axiom at the same level.
() for every formula .
This axiom does not imply the existence of any infinite sets and is, of course, true in every standard model. Thus the readers interested only in the computational complexity of the Halin type theorems can safely ignore these considerations.
It is in the nature of reverse mathematics that sentences and sets of sentences of second order arithmetic are often viewed in several different ways. In different contexts they may be seen as mathematical or logical principles, axioms, axiom schemes, theories, theorems or the like. We point out what may be a less familiar terminology that is currently popular. What might be seen as a typical axiom or theorem asserting that for every of some sort there is a with some relation to , i.e. a sentence of the form may be called a principle. With this terminology come the notions of an instance of the principle, i.e. an satisfying and a solution for , i.e. a such that holds.
4 and Hyperarithmetic Analysis
We devote this section to the proof of
Theorem 4.1.
is a theorem of hyperarithmetic analysis.
In this section we consider only vertex-disjoint single rays in undirected graphs, as in the statement of . The proof of Theorem 4.1 will be split into two parts. The first part verifies that satisfies the second clause of Definition 3.13.
Theorem 4.2.
Every standard model of is closed under hyperarithmetic reduction.
Proof.
Fix a standard model of . First, we show that contains . By relativizing the proof, it follows that is closed under Turing jump.
For each , consider the tree consisting of all strings of the form such that some number below is enumerated into at stage , and either or . Observe that has a unique computable branch , where is the smallest number such that .
Consider the disjoint union . Observe that satisfies the premise of (in ), because each has a computable branch. Apply to to obtain a sequence of disjoint rays in . Each is contained in some . We can, uniformly in , extend or truncate to the unique branch of . Hence computes a sequence of infinitely many distinct branches , which in turn computes longer and longer initial segments of . This proves that contains .
Next we show that, for each computable limit ordinal , if contains for every then contains . (Again the desired result follows by relativization.) By Proposition 3.9, there is a computable sequence of trees such that each tree has exactly one branch with these reductions computed uniformly. Fix an increasing computable sequence which is cofinal in and consider the disjoint union . Observe that satisfies the premise of (in ): for each , computes the branches for . Apply to to obtain a sequence of disjoint rays in . As before, computes a sequence of infinitely many distinct branches , and hence a sequence of infinitely many distinct . Each uniformly computes for , so we conclude that computes as desired. ∎
It follows that is not provable in , despite the apparent similarity between and a compactness result. (Indeed, is not even provable in - (Theorem 7.9).)
Next, we present essentially the proof of attributed to Andreae (see [5, Theorem 8.2.5 and bottom of pg. 275]) and then analyze it with an eye to the axioms which can be used to formalize it. We then use this analysis to complete the proof of Theorem 4.1.
The key combinatorial lemma implicit in Andreae’s proof is:
Lemma 4.3.
Given disjoint rays and disjoint rays there are disjoint rays such that for each , and start at the same vertex.
Before proving Lemma 4.3, let us use it to prove .
Proof of assuming Lemma 4.3.
Given a graph which has arbitrarily many disjoint rays, we build by recursion on sequences of disjoint rays with initial segments of length such that is followed by one more vertex for . The required infinite sequence of disjoint rays will then be given by .
Suppose that we have and . By assumption, let be a sequence of disjoint rays. Discard all rays which contain a vertex of some . There are at most many of them, so by discarding and renumbering if necessary we are left with .
For each , let denote the first vertex on after . Apply Lemma 4.3 to and . We obtain disjoint rays such that for each , begins at vertex . Now let for and . These are disjoint by construction. This completes the inductive step of the construction of the and so provides the required witnesses for . ∎
It remains to prove Lemma 4.3. The key ingredient is Menger’s theorem for finite graphs. If and are disjoint sets of vertices in a graph, we say that is an - path if starts with some vertex in and ends with some vertex in . A set of vertices separates and if any - path contains at least one vertex in .
Theorem 4.4 (Menger, see [5, Theorem 3.3.1]).
Let be a finite graph. If and are disjoint sets of vertices in , then the minimum size of a set of vertices which separate and is equal to the maximum size of a set of disjoint - paths.
We now present the proof of Lemma 4.3.
Proof of Lemma 4.3.
Suppose we are given disjoint rays and disjoint rays . First, define the set
Then we perform the following recursive procedure. At each step, check if there is some such that has not been defined and intersects at most many rays which have not been discarded. If there is no such , we end the procedure. Otherwise, find the least such and do the following:
-
1.
discard all rays which intersect ;
-
2.
define .
After the procedure is complete, let be the set of for which has not been defined. Let be the set of rays which have not been discarded. Let . We observe that , because
Next, for each , let be the first vertex on such that meets exactly many rays in . (Each exists by construction of .)
Observe that the finite set meets at most many rays in . Since , we may pick some ray in which does not meet . We define to be said ray. Then, discard all rays in which do not meet .
Finally, we use Menger’s theorem to define for each . For each , let denote the first vertex of . For each such that remains in , let be the first vertex on such that and are disjoint. Then consider the following finite sets of vertices:
We want to apply Menger’s theorem to . Towards that end, we claim that cannot be separated from in by fewer than vertices.
Suppose that and . Since and is disjoint, there is some such that does not meet . Next, since meets many disjoint rays in , there is some such that and meets , but does not meet . Let be any vertex in both and . Then is a path in from to which does not meet . This proves our claim.
By Menger’s theorem, there are many disjoint - paths in . Then, for each , define by starting from , then following the - path given by Menger’s theorem to some , and finally following .
We have constructed a collection of rays. It is straightforward to check that they are disjoint, and that for each , and start at the same vertex. ∎
We now analyze these proofs from a reverse mathematical perspective to show that follows from the THA - and so also satisfies the first clause of Definition 3.13.
Theorem 4.5.
i) and - each implies .
ii) - implies . Hence for every , satisfies .
Proof.
i) The proof of the first part of Theorem 4.2 essentially shows that implies . One point to note is that for each , proves there is some such that (using e.g. [28, II.3.9]). Now note that the same argument also proves the fact (essentially in [15]) that - implies as it directly supplies the sequence of branches in .
ii) The proof of presented above is easily seen to be one in , except for two points. First, each step of the induction assumed we had available a sequence of disjoint rays in our graph. However, all we know is that for each , there is some collection of disjoint rays of size . To access this information for each in a recursive construction, we require that there is a single sequence such that the th entry of the sequence is a collection of disjoint rays of size . Such a sequence can be obtained using the axiom of choice. In this case, since the predicate “there exists many disjoint rays” is , such a sequence can be obtained using -. Therefore we may assume we have a sequence with each a sequence of disjoint rays and the graph consisting of all the vertices and edges occurring in any . We begin our construction with .
The second point is the iterated application of Lemma 4.3. The proof of that Lemma can be done in . (In particular, one can check that Menger’s theorem is provable in by following the first proof for it given in [5, Theorem 3.3.1].) However, here we need a bit more information as generally we cannot, in , carry out recursive constructions that increase complexity (even by some fixed number of jumps) at each step. We can only carry out recursions where each step is done computably in some set of the model. So it suffices to show that the constructions for all the instances of Lemma 4.3 needed in its iterations in the proof can be done computably in .
The crucial observation is that, by an induction starting with , we may take each to be of the form where is a finite path in and is a tail of one of the . To see this simply note that each ray constructed in the Lemma starts with an initial segment of some ; it continues with a finite path in and ends with a tail of some . Thus computably in we can at every stage find one of these descriptions that provides the disjoint rays needed for the next stage of the recursion. Thus the sequences and are computable from . The sequence is then also computable from and computing the formal presentation of infinitely many disjoint rays then takes only one more jump. This completes the analysis of our proof of in -. ∎
We will establish some implications and nonimplications between and other THAs in §7.
5 Variants of and Hyperarithmetic Analysis
In this section, we show that at least five of the eight principles are THAs:
Theorem 5.1.
All single-ray variants of (i.e., ) and are theorems of hyperarithmetic analysis.
Of the other three variants, and are open problems of graph theory ([3] and Bowler, personal communication). We do, however, have interesting results about these principles when restricted to directed forests (Theorem 6.13, Corollary 6.14). The other one, , was proved by Bowler, Carmesin, Pott [3] using structural results about ends. We hope to analyze its strength in future work.
The proof of Theorem 5.1 consists of several variations of the proof of Theorem 4.1. One of which () requires some additional ideas.
In order to minimize repetition, we establish some implications between some variants of over . The proofs of each of these reductions follow the same basic plan. To deduce from we provide computable maps , and which, provably in , take -graphs to -graphs , -disjoint -rays or sets of -disjoint -rays in to -disjoint -rays or sets of -disjoint -rays in , and -disjoint -rays or sets of -disjoint -rays in to -disjoint -rays or sets of -disjoint -rays in , respectively. These functions are designed to take witnesses of the hypothesis of in to witnesses of the hypothesis of in and witnesses to the conclusion of in to witnesses to the conclusion of in . Clearly it suffices to provide such computable maps to establish the desired reduction in . (Those familiar with Weihrauch reducibility will recognize that these arguments establish Weihrauch reductions between certain problems corresponding to variants of .)
Unless otherwise noted all definitions and proofs in this section are in .
Lemma 5.2.
Given an undirected graph , we can uniformly compute a directed graph and mappings between Z-rays in and Z-rays in which preserve Y-disjointness.
Proof.
We define a computable map from undirected graphs to directed graphs as follows. The set of vertices of consists of the vertices of , together with two new vertices and for each edge in . The set of edges of consists of five edges , , , , for each edge in .
Next we define a computable map : given a ray in , maps it to the ray in . Conversely, we define a computable map from rays in into rays in as follows. Observe that exactly one of the first three vertices in is a vertex in , because the only outgoing edges from a vertex lead to or to , and the only outgoing edge from a vertex leads to . We take this vertex (say ) to be the first vertex of . Every outgoing edge from leads to some . Combining the above observations, we deduce that the tail has the form . Then maps to the ray in .
Similarly, given a double ray in , maps it to the double ray in . We can show that every double ray in has this form by considering the incoming edges to each vertex in . Therefore we can define a computable map from double rays in to double rays in by .
It is straightforward to check that , , and preserve -disjointness. ∎
Therefore we have
Proposition 5.3.
The directed variants of imply their corresponding undirected variants, i.e., implies for each value of and .
Lemma 5.4.
Given a directed graph , we can uniformly compute a directed graph and mappings between -rays in and -rays in which satisfy the following properties: if two -rays in are vertex-disjoint, then the corresponding -rays in are edge-disjoint, and if two -rays in are edge-disjoint, then the corresponding -rays in are vertex-disjoint.
Proof.
We define a computable map from directed graphs to directed graphs as follows. The set of vertices of is , where and stand for incoming and outgoing respectively. The set of edges of consists of for each , and for each .
Next we define a computable map : given a ray in , maps it to the ray in . Conversely, we define a computable map from rays in to rays in as follows. Given , the ray visits the vertex in whenever appears in . (For example, we map to and we map to .)
Similarly, maps a given double ray in to the double ray
in . Every double ray in has this form, so we may define .
It is straightforward to check that the above mappings have the desired properties. ∎
Therefore we have
Proposition 5.5.
The directed edge-disjoint variants of imply their corresponding directed vertex-disjoint variants, i.e., implies for each value of .
Lemma 5.6.
Given a directed graph , we can uniformly compute a directed graph and mappings between sets of -disjoint rays in and sets of -disjoint double rays in which preserve cardinality.
Proof.
We define a computable map from directed graphs to directed graphs containing as follows. For each vertex of we add new vertices for each and edges and for all .
Next we define a computable map from sets of -disjoint rays in to sets of -disjoint double rays in as follows. Given a set of -disjoint rays in , we first ensure that each ray begins at a different vertex, by replacing it with a tail if necessary. (This is only relevant if the rays are edge-disjoint rather than vertex-disjoint.) Then for each ray , we consider the double ray in . This yields a set of -disjoint double rays in of the same cardinality.
Finally we define a computable map from sets of -disjoint double rays in to sets of -disjoint rays in . Given a double ray in , we search for the least such that is a vertex in . (If none of these vertices were in then as there are edges between them they would all have to be of the form for a single in . The only edges between these make them into a copy of the reverse order on . This order cannot have any subsequence of order type .) Now we map the given double ray to the ray in . It is straight forward to check that this map induces a cardinality-preserving map from sets of -disjoint double rays in to sets of -disjoint rays in . ∎
Therefore we have
Proposition 5.7.
The directed double ray variants of imply their corresponding directed single ray variants, i.e., implies for each value of .
Figure 1 summarizes the known implications between variants of over . We will show that implies over (Theorem 6.15).
Remark 5.8.
Bowler, Carmesin, Pott [3, pg. 2 l. 3–7] describe an implication from to which appears to use . It turns out that the graph-theoretic principle used to carry out the implication does not imply even over (and is hence much weaker than ), but when combined with , yields a THA. It and several other principles with the same property (almost theorems/theories of hyperarithmetic analysis) are analyzed in Shore [26].
We return to the goal of proving Theorem 5.1.
Theorem 5.9.
For each choice of XYZ, implies . Furthermore, every standard model of and is closed under hyperarithmetic reduction.
Proof.
By Proposition 5.3, it suffices to prove the desired result for the undirected variants of . Theorems 4.2 and 4.5(i) together assert the desired result for . We describe how to modify the proofs of Theorems 4.2 and 4.5(i) to prove the desired result for the other variants of .
Observe that in the aforementioned proofs, we only applied to forests such that each of the constituent trees has a unique branch. In such graphs, none of the constituent trees can contain two rays which are edge-disjoint. Hence the aforementioned proofs establish the desired result for as well.
In order to prove the desired result for and , we modify the aforementioned proofs as follows. For each tree, we relabel as needed to then add one computable branch of new vertices (other than the root of the tree) such that the new branches are also disjoint. The resulting trees satisfy the following properties:
-
•
Each tree contains some double ray which is Turing equivalent to the branch on the original tree.
-
•
No two double rays in the tree can be vertex-disjoint or edge-disjoint.
-
•
Given any double ray in the tree, we can uniformly compute the branch on the original tree.
It is straightforward to check that the modified proofs establish the desired result for and . ∎
Henceforth we will not explicitly mention uses of whenever we are assuming any or -.
Next, we show that and are provable in - (Theorems 5.10, 5.15). It then follows from Propositions 5.3 and 5.5 that , and are also provable in , completing the proof of Theorem 5.1.
Theorem 5.10.
- implies .
Proof.
The mathematical result is due to [11] and is Exercise 42 of Chapter 8 in [5]. Our proof is very similar to that of Theorem 4.5(ii) which follows [5, Theorem 8.2.5(i)] except we need to grow our family in “two directions”.
Our plan is to construct, by recursion on , sequences and such that the are disjoint double rays with subpaths of length such that, for , is extended by a new vertex at each end. Our required sequence of disjoint double rays is then .
As we want to reuse the proof of Lemma 4.3, we want to decompose double rays into single rays. To that end we introduce some notation. If and are double rays we let and while and are and , respectively, but with order reversed. We use or to stand for one of or . For single rays we use to denote the reverse sequence of vertices. So, for example, if is an edge in a double ray , we have .
We note the changes needed in the proof of Theorem 4.5. Now our sequence given by - consists of double rays and is the corresponding graph. Our construction of the desired and again proceeds recursively in and at the end uses one more jump to get the required sequence of double rays. We begin our recursion by setting and as any subpath of length . Here we note that by construction every will be of the form where is a finite path in while and are each or for some . Again, finding each and in turn is recursive in once we prove in that it exists given and .
Thus it suffices to prove an analog of Lemma 4.3 in : Given and and for we can construct and as required. To simplify the notation we omit the superscripts and . For , let be the first vertex in before and be the first vertex in after . We will arrange for to be preceded by and followed by which will then also be .
We first discard from any sharing a vertex with any . As there are such vertices we have at least many remaining. We relabel these as for and choose an edge in each . We now essentially follow the proof of Lemma 4.3 but for the sets of single rays and . Our goal for each , and , is to find a suitable replacement (beginning with the same vertex while maintaining the required disjointness) so that we can assemble the from them for . We also want an disjoint from all the double rays with which will be .
We follow the procedure described in the proof of Lemma 4.3 (using for as has elements) to define single ray replacements for the from which we will construct the . When in that procedure we would keep the old ray we do so here as well and let . When the procedure is completed we let , and we have not defined and let . Clearly we have discarded at most many more of the rays in .
Now continue the construction for the and the first vertices on such that meets exactly many rays in . The union of these finite paths meets at most many rays in and so at most many double rays , . Thus there must be at least one such that it does not meet and no was discarded during the process. We let one such be and take any subpath of length as .
Now discard all single rays in not meeting . For each remaining in let be the first vertex on such that is disjoint from . As in the proof of Lemma 4.3, we apply Menger’s theorem to and where
This produces a set of disjoint paths in from each , , to a in . We can now define for as the single ray beginning with and then continuing with after .
We can now define the required double rays for as and check that these have all the desired properties. ∎
As for , instead of following the proof of Theorem 4.5, we will reduce to the problem of finding an infinite sequence of vertex-disjoint rays in a certain locally finite graph (see [3, pg. 2 l. 3–7]). To carry out this reduction, we define the line graph:
Definition 5.11.
The line graph of an -graph is the -graph whose vertices are the edges of and whose edges are the , where and are edges in .
Lemma 5.12.
Let be an X-graph. There is a computable mapping from rays in to rays in such that if two rays in are edge-disjoint, then their images are vertex-disjoint.
Proof.
Map to . ∎
However, vertex-disjoint rays in do not always yield edge-disjoint rays in . An extreme counterexample is what is called the (undirected) star graph which consists of a single vertex with infinitely many neighbors: It does not contain any rays yet its line graph is isomorphic to the complete graph on which contains infinitely many vertex-disjoint rays. Nonetheless, if is locally finite, then vertex-disjoint rays in do correspond to edge-disjoint rays in :
Lemma 5.13 ().
Let be a locally finite X-graph. There is a mapping from rays in to rays in such that if two rays in are vertex-disjoint, then their images are edge-disjoint rays in .
Proof.
Given a ray in , we construct a ray in by recursion. Say that . Start by defining to be . Having defined , we define as follows. Let be the largest such that is an endpoint of . Such exists because is locally finite and is a ray. We can find by . Then define to be the endpoint of other than . This completes the recursion. Note that the are strictly increasing: as is not an endpoint of by the maximality of . The next edge then includes and so . Also note that if the graph is directed the must be of the form and the last occurrence of any in an must be as its first element. By construction is infinite and contains no repeated vertices because of the maximality requirement, hence it is a ray. Observe that every edge in is a vertex in , so the above mapping maps vertex-disjoint rays in to edge-disjoint rays in . ∎
It remains to show that we can restrict our attention to locally finite graphs. We accomplish this with the help of -. Given a directed graph with arbitrarily many edge-disjoint rays, we can use - to choose a family where, for each , the rays are edge-disjoint. From this family, we may construct an appropriate locally finite subgraph of :
Lemma 5.14.
Suppose that is an X-graph and there is some family of rays in such that for each , the rays are edge-disjoint. Then there is some locally finite -subgraph of and some family of rays in such that for each , the rays are edge-disjoint.
Proof.
Define the vertices of to be the vertices of , say . We specify the set of edges of by providing a recursive construction of sets of edges putting in a set of edges at each step. We guarantee that each is a union of finitely many finite sets of edge-disjoint rays in and that after stage no edge with a vertex for as an endpoint is ever put into .
Begin at stage by putting all the edges in into . Proceeding recursively at stage we have and consider the edge-disjoint rays , . For each , say . Each for appears at most once in as is a ray. For each , since we have access to the set of vertices of , we can decide whether contains and, if so, find the index such that . Call it . If there is no such , set . Define to be the tail of after . We put all the edges in , , into . Let .
It remains to show that is locally finite. Consider any vertex . No edge containing as an endpoint is put in after stage . On the other hand, is the union of finitely many finite sets of edge-disjoint rays (all of which have been computed uniformly). Each set of edge-disjoint rays in this union has appearing at most once in each of its rays. Thus at most two edges containing appear in each of the finitely many rays in this set. Therefore there are only finitely many edges containing in each of the finite sets of edge-disjoint rays making up . All in all, only finitely many edges in contain . ∎
We are ready to prove
Theorem 5.15.
- implies for each value of X.
Proof.
Given an -graph with arbitrarily many edge-disjoint rays, we can use - to choose a family such that for each , the rays , , are edge-disjoint. By Lemma 5.14, there is a locally finite subgraph of and a family such that for each , the , , are edge-disjoint rays in . By Lemma 5.12, there is a family such that for each , the , , are vertex-disjoint rays in . By the second part of the proof of Theorem 4.5(ii) (which can be carried out in ), has infinitely many vertex-disjoint rays. Finally by Lemma 5.13, has infinitely many edge-disjoint rays. Hence has infinitely many edge-disjoint rays. ∎
Finally, we give a proof of for directed forests using (recall that remains open). We will see that and for directed forests are equivalent over but note that - does not imply . (See Theorem 6.13 and the comment following it).
Theorem 5.16.
implies for directed forests.
We first prove two lemmas.
Lemma 5.17 ().
Let be a directed forest and let , be directed double rays in . Suppose and have an edge in common. Then there are vertices and such and , have no vertices in common outside of those in . Note that we allow for the possibility that and/or in the sense that for any double ray . We call the intersection of and .
Proof.
Suppose , provide a counterexample. As they have an edge in common, they lie in the same directed tree in and can be viewed as (undirected) double rays in (the underlying graph for ). As they form a counterexample to the lemma, there must be either a first (i.e. earliest in the double ray ) such that or a last such that but and have a vertex in common outside the common interval. The situations are symmetric and we consider the second. The immediate successors and of in and , respectively, must be different by our choice of . Consider now the location of in . If it is after then the paths from to in and to in (both considered now as undirected graphs within ) are different as the immediate successor of in is while in it is either or a vertex in . Thus there are two different paths in from to contradicting ’s being a tree. If, on the other hand is before in , it must be before and a similar argument provides different paths from to in and . ∎
Lemma 5.18 ().
There is a computable function such that given any sequence of DED rays with subpaths of length in a directed tree and sequence of DED rays in , we can construct a sequence of DED rays with subpaths of length in such that extends at each end for . Indeed, we may take .
Proof.
First we remove all the that contain an edge in any at cost
of at most many . Consider any remaining in the second
sequence. By Lemma 5.17, its intersections with
the are intervals of edges in which are disjoint as
the are. By our first thinning of the list, none of the
intersect any of the so each must lie entirely
above or entirely below. We associate to each a label
consisting of the set ; the elements
of in the order in which the (for ) appear
in (in terms of the ordering of ) along with a or
depending on which side of it falls in . We write
for the starting vertex of and for the ending one. As
above, we allow the values for these endpoints if the intervals
are infinite. Now there are, of course, at most finitely many such labels. In
particular, there are at most such labels. Thus if we have
many left at least two of them, say and
have the same label say with set .
Claim:
.
For the sake of a contradiction, assume we have
in with preceding in the ordering of in the label.
Say is the ray such that is before in . Note that and are edge disjoint as and are. We consider two cases: (1) is before in and (2)
is before in . We now produce, for each case, two
vertices with two distinct sequences (i) and (ii) of adjacent edges in
connecting these two vertices. These sequences are illustrated in Figures
2 and
3. Note that by our assumptions on the
orderings of the intervals (for and ) as
displayed, all of the starting or ending points of that appear in
our sequences are vertices in one of the rays (i.e. none are ):
-
(1i)
Start at in and go to then in go to .
-
(1ii)
Start at in and go to then go in to .
-
(2i)
Start at in and go to then in go to then in go to .
-
(2ii)
Start at in and go to .
To see that the two sequences of vertices are different, note for (1) that (1i) contains an edge in but (1ii) does not. For (2) note that (2i) contains an edge in but (2ii) does not. We now, in each case, view the associated two distinct sequences of vertices with the same endpoints in the underlying (undirected) tree . The only way one can have such sequences in a tree is for one of the sequences to contain some vertices in order. However, any three successive vertices in any of these sequences lie within one of the or or both and so cannot have two instances of the same vertex. The crucial point is that each is in both and and has at least two vertices. Any transition along the sequence between an and an (in either order) goes through and so any three consecutive vertices are all contained in one or one (or both).
Knowing now that is or , we complete the proof of the Lemma. If , then both and are disjoint from all the and so we may add on either one of them as with an arbitrary subpath of length while keeping and extending the appropriately for all . Otherwise say . Let be the one of or such that is closer to . (Remember that they are both on the same side of this interval in by our fixing the label.) Now replace the tail of starting with and going away from by the tail of starting with and going in the same direction. Let this ray be . Note that it is disjoint from all the , as it contains only edges that are in or neither of which share any edges with such . It is also disjoint from where is the one of , which is not since all the edges of are either in or in outside of by our choice of as closer to . As is also disjoint from all the for by our fixing the label, we may define and appropriately for , and and choosing of length arbitrarily so as to get the sequence required in the Lemma. ∎
Lemma 5.18 provides the inductive step for the following proof:
Proof of Theorem 5.16.
Assume we are given a directed forest with arbitrarily many DED rays. By we may take a sequence such that, for each , is a sequence of many DED rays in . If there are infinitely many of the trees making up each of which contains some then we are done. So we may assume that all of them are in one directed tree . We now wish to define by recursion such that, for each , is a sequence of many DED rays with subpaths of length such that for each , extends at each end so that the form an infinite sequence of DED rays in as required. Lemma 5.18 provides precisely the required inductive step for the construction since we have the required sequences of DED rays at each step of the construction. Once again we just have to note that Lemma 5.18 provides witnesses for the double rays that are composed of finite paths in and final segments in one direction or the other of some of the and so they can be found recursively in . ∎
6 Variations on Maximality
In this section, we consider variants of whose solutions are required to be maximal in terms of cardinality or maximal in terms of set inclusion.
6.1 Maximum Cardinality Variants of
Definition 6.1.
Let be the statement that every -graph has a set of -disjoint -rays of maximum cardinality, or more formally, the statement that for every -graph :
-
•
there is no -ray in , or
-
•
there is some and some such that is a sequence of -disjoint -rays in , and there is no such that is a sequence of -disjoint -rays in , or
-
•
there is some such that is a sequence of -disjoint -rays in .
When we talk about the cardinality of a (possibly empty) finite sequence we mean the number (which may be ). Of course a sequence is said to have infinite cardinality.
was proved by Halin [11], who also proved the corresponding statement for uncountable graphs.
Remark 6.2.
The notation in Definition 6.1 is inspired by the well known version of which, for every , asserts the existence of , the th jump of , for every :
This asserts (in addition to ) particular instances of . So too (in addition to ) do the as we are about to see.
Proposition 6.3.
For each choice of XYZ, implies over and implies over . Therefore and are equivalent over . In particular, they have the same standard models.
Proof.
The first implication holds because if an -graph has arbitrarily many -disjoint -rays, then any sequence of -disjoint -rays in the graph of maximum cardinality must be infinite. To prove the second implication, let be an -graph and be the formula which says that there is a sequence of length of -disjoint -rays in . If , then by , holds and so by there is a sequence of -disjoint -rays in as required. On the other hand, if there is an such that holds but fails, then this witnesses . ∎
Corollary 6.4.
and are theorems of hyperarithmetic analysis.
It follows from Proposition 6.3 and Theorem 5.9 that implies , so we will not explicitly mention uses of whenever we are assuming any .
Proposition 6.5.
implies , implies and implies .
Next, we show that proves sufficient induction in order to transcend -. This implies that is strictly stronger than for certain choices of (Corollary 6.10). The connection between and graphs is obtained by viewing the set of solutions of an arithmetic predicate as the set of (projections of) branches on a subtree of . In detail:
Lemma 6.6 ([28, V.5.4]).
If is an arithmetic formula, proves that there is a tree such that
and |
In fact, as the proof in [28, V.5.4] shows, the required functions are what are called the minimal Skolem functions and are arithmetically defined uniformly in and the formula .
The following easy corollary will be useful.
Lemma 6.7.
If is an arithmetic formula, proves that there is a sequence of subtrees of such that for each ,
and |
Proof.
Say that holds if and only if holds, where is such that . Apply Lemma 6.6 to the arithmetic formula to obtain a tree . For each , define to be the set of all such that . It is straightforward to check that satisfies the desired properties. ∎
Theorem 6.8.
proves .
Before proving the above theorem, we derive some corollaries:
Corollary 6.9.
proves the consistency of . Therefore it is not provable in .
Proof.
Corollary 6.10.
is strictly stronger than for the following choices of XYZ: XYS and UVD.
Proof.
We now prove Theorem 6.8:
Proof that implies .
By Proposition 6.5, it suffices to prove the desired result for . To prove from , begin by using Lemma 6.7 to define for each a sequence of trees such that for each and there is at most one such that and
We want to show that each is ill-founded. Note that if and is ill-founded, then so is . Therefore it suffices to show that for cofinally many , is ill-founded.
Apply to the disjoint union to obtain a collection of -disjoint rays of maximum cardinality. We prove that is infinite. Suppose not. Then there is some maximum such that contains a ray in . A ray in can be computably truncated or extended to a branch on , so is ill-founded. Hence is ill-founded as well (by ). But then there is a collection of -disjoint rays in which has cardinality greater than that of , contradiction.
We have proved that is infinite. Next we prove that each has at most one branch. That would imply that each contains at most one ray in , so contains rays in cofinally many , as desired.
If has two distinct branches and , then by the “at most one” condition in the definition of the . Consider the least such that . Such exists by . Note that because and for . But then and , contradiction.
This proves that implies . In order to prove that implies , we modify the above proof by adding to each a computable branch consisting of new vertices to form a tree . Apply to to obtain a collection of -disjoint double rays of maximum cardinality. Following the above proof, we may prove that is infinite and each contains at most one double ray in . So contains double rays in cofinally many , as desired. ∎
Remark 6.11.
The same proof shows that implies the following induction scheme: Suppose is a sequence of trees such that
-
1.
has a unique branch;
-
2.
for all , the number of branches on is the same as the number of branches on .
Then for all , there is a sequence such that for each , is the unique branch on . It also shows that implies (i.e. closure under the -jump) and much more. Indeed, similar ideas prove in Theorem 7.3 that implies unique- (Definition 7.1). also implies a similar induction scheme analogous to finite- (Definition 7.2).
We can prove that even fragments of give more induction than the specific instances derived in Theorem 6.8.
Theorem 6.12.
(even for directed forests) implies over .
Proof.
Suppose is a formula such that and hold. Let be a sequence of subtrees of such that is ill-founded if and only if holds. Let be a sequence of subtrees of such that for each , consists of all sequences where for each , is (a code for) a string in . We order these sequences by component-wise extension. It is clear that is ill-founded if and only if holds.
For each , orient each edge in towards its root and add a computable -ray of new vertices which starts at its root. This forms a directed tree . If contains a double ray, is ill-founded and holds. Furthermore, no two disjoint double rays can lie in the same .
Let be the directed forest . By , there is a sequence of disjoint double rays in of maximum cardinality, so the sequence may be for for some or . Since holds, is nonempty. If is finite, let be maximal such that contains some . Then holds, so holds as well. It follows that contains some double ray, which we can then add to to obtain a larger sequence of disjoint double rays in for the desired contradiction. Therefore is infinite. Since each contains at most one , infinitely many contain some . Therefore holds for all . ∎
In fact, we have the following equivalences:
Theorem 6.13.
The following are equivalent (over ):
-
1.
;
-
2.
for directed forests;
-
3.
for directed forests;
-
4.
for directed forests;
-
5.
for directed forests.
Proof.
(1) (2) follows from Theorem 5.16. (2) (3) follows from the proof of Proposition 6.3. (3) (4) follows from the observation that the mapping of graphs defined in Lemma 5.4 sends a directed forest to a directed forest. (4) (5) follows from Theorem 6.12 and the proof of Proposition 6.3.
To prove (5) (1), suppose is an arithmetic formula such that . By Lemma 6.7, there is a sequence of subtrees of such that
and |
By assumption on , each is ill-founded. We use to construct a sequence of directed trees as we did in the proof of Theorem 6.12 to construct from the .
By , the directed forest contains arbitrarily many disjoint double rays. Therefore contains infinitely many disjoint double rays , by . Note that any double ray in any must contain the computable ray we added, so any two double rays in the same must intersect. This implies that each belongs to some distinct . Therefore for every , there is some and some such that is a double ray in . When we remove the added computable ray from we are left with a branch in which is of the form where consists of witnesses for . ∎
Since (, even) does not prove [28, IX.4.7], it follows that
Corollary 6.14.
(even for directed forests) is not provable in , and strictly implies over .
Next, we show that implies over (see Figure 1).
Theorem 6.15.
implies over . Therefore (1) implies over ; (2) if any standard model of satisfies , then it satisfies as well.
Proof.
Let be a graph which contains arbitrarily many disjoint single rays. By , there is a sequence of disjoint double rays in of maximum cardinality. If this sequence is infinite, then there are infinitely many disjoint single rays in as desired. Otherwise, suppose that is a sequence of disjoint double rays in of maximum cardinality . Let be the subgraph of consisting of the union of all . Let be the induced subgraph of consisting of all vertices which do not lie in . Note that does not contain any double ray, otherwise would contain many disjoint double rays. Next, we expand to the graph , defined below.
Decompose into its connected components (there may only be finitely many). Any two single rays in the same must intersect, because if and are disjoint single rays in the same , then we can construct a double ray in by connecting them (start with a path between and of minimum length, then connect it to the tails of and which begin at the endpoints of the path).
For each , define by adding a computable ray of new vertices to , which begins at the -least vertex in . Define to be the disjoint union .
By , there is a sequence of disjoint double rays in of maximum cardinality.
Case 1. If this sequence is infinite, then contains infinitely many disjoint single rays because each double ray in the sequence has a tail which lies in . In this case we are done.
Case 2. Otherwise, does not contain arbitrarily many disjoint double rays. Since any two single rays in the same must intersect, we can transform any collection of disjoint single rays in into a collection of disjoint double rays in of equal cardinality by connecting each single ray to the -least vertex in its connected component and then following the computable ray we added. It follows that does not contain arbitrarily many disjoint single rays. Fix such that does not contain many disjoint single rays.
Towards a contradiction, we construct a collection of -many disjoint double rays in as follows. Fix a collection of many disjoint single rays in . First, at most of these single rays lie in . In fact at most of these single rays can have finite intersection with , because given a collection of disjoint single rays each of which has finite intersection with , we can obtain a collection of disjoint single rays in of the same cardinality by replacing each ray with an appropriate tail. Second, by reasoning analogous to the above, at most of these single rays can have finite intersection with . Therefore, there are at least many disjoint single rays in each of which have infinite intersection with both and .
Next, choose an edge in each and split into two single rays and . By the pigeonhole principle, there is some single ray of the form or , and at least many disjoint single rays in , each of which have infinite intersection with both and . Call these rays . Discard all the other rays in . Below we describe how to connect pairs of single rays using segments of to form a collection of -many disjoint double rays in .
Let denote the vertices of . Since each single ray has infinite intersection with , by the pigeonhole principle, there is a pair of disjoint rays and such that for each tail of , there is a vertex in and a vertex in such that no intersects between these two vertices. (Formally, we justify this by defining the following coloring recursively. Start from the first vertex in which is also in some . Search for the next vertex on which intersects some , . Then we color with the unordered pair . Then we search for the next vertex on which intersects some , and color with , and so on. Some color must appear infinitely often.) Then we commit to connecting and (but we do not do so just yet). Applying the pigeonhole principle again, there is a pair of disjoint rays and (with all distinct) such that for each tail of , there is a vertex in and a first vertex in (after in ) such that no , except perhaps , or , intersects between these two vertices. We may eliminate any elements of by changing (if necessary) to the last element of in . Again we commit to connecting and . Repeat this process until we have obtained pairs of single rays. That is, when we have and for an , we find and with and distinct from all previous and such that for each tail of there is a vertex and a first after in such that no , except perhaps or for , intersects between these two vertices. This process stops when we define and .
Finally, we connect these pairs of single rays in the opposite order in which we defined them: Start by picking some and some . Then we define a double ray by following until , then following until , and finally following , i.e., . Having defined , define , where and are chosen as follows: Consider a tail of such that the union of is disjoint from (1) ; (2) for each ; (3) for each . By choice of and , there are vertices and such that none of or intersect .
It is straightforward to check that each of is disjoint from . This process yields disjoint double rays in , contradicting the maximality of . ∎
Using some of the ideas in the previous proof, we can prove
Theorem 6.16.
for forests implies for forests over . Therefore for forests implies for forests over .
Proof.
Let be a forest. If happens to have arbitrarily many disjoint double rays, then by , has infinitely many disjoint double rays. Therefore there is an infinite sequence of disjoint single rays in . Such a sequence has maximum cardinality, so we are done in this case.
Suppose does not have arbitrarily many disjoint double rays. By for forests, there is a sequence of disjoint double rays in of maximum cardinality. Following the proof of Theorem 6.15, define the forests , , and . There, we proved that no two single rays in the same connected component of can be disjoint.
By for forests, there is a sequence of disjoint double rays in of maximum cardinality. If this sequence is infinite, then there is an infinite sequence of disjoint single rays in because each double ray in the sequence has a tail which lies in . This is a sequence of disjoint single rays of maximum cardinality in , so we are done in this case.
Otherwise, suppose is a disjoint sequence of double rays in of maximum cardinality. Consider the following disjoint sequence of single rays in . First, for each , consider the single ray formed by intersecting and the double ray . Second, for each , we can split the double ray into a pair of disjoint single rays in . This yields a finite sequence of disjoint single rays in .
We claim that is a sequence of disjoint single rays in of maximum cardinality. Suppose there is a larger sequence of disjoint single rays in . Since is a forest, any two single rays in which share infinitely many edges or vertices must share a tail. Therefore there is a single ray in this larger sequence which only shares finitely many edges and vertices with each . Then some tail of , say , is vertex-disjoint from each . In particular, is vertex-disjoint from each , i.e. lies in . Extend to a double ray in by first connecting to the -least vertex in its connected component , then following the computable ray which we added. The resulting double ray is disjoint from every , because no can lie in the same as (for is vertex-disjoint from by construction). This contradicts the maximality of . ∎
6.2 Maximal Variants of
Instead of sets of disjoint rays of maximum cardinality, we could consider sets of disjoint rays which are maximal with respect to set inclusion. For uncountable graphs, Halin [11] observed that any uncountable maximal set of disjoint rays is in fact of maximum cardinality (because rays are countable). This suggests another variant of , which we call maximal :
Definition 6.17.
Let be the statement that every -graph has a (possibly finite) sequence of -disjoint -rays which is maximal, i.e., for any -ray in , there is some such that and are not -disjoint.
immediately follows from Zorn’s Lemma. It is straightforward to show that implies - (see the proof of Theorem 6.18 below), hence is much stronger than or even . We show below that is equivalent to -. This situation is reminiscent of König’s duality theorem for countable graphs. Aharoni, Magidor, Shore [2] proved that the theorem implies and that - suffices to prove the required existence of a König cover. Simpson [27] later proved that actually suffices. The covers produced in [2], and indeed in all then known proofs of this duality theorem actually had various maximality properties. Aharoni, Magidor, Shore proved that the existence of covers with any of a variety of maximality properties actually implies -.
Theorem 6.18.
- is equivalent to .
Proof that implies -.
We first prove that implies by adapting the proof of Theorem 5.9: If we apply instead of to any of the forests constructed in that proof, we obtain a sequence containing a -ray in each tree which constitutes the forest. This is more than sufficient for carrying out the remainder of the proof of Theorem 5.9.
To prove that implies -, suppose we are given a set . Consider the disjoint union of all -computable trees (this exists, by ). Any maximal sequence of -disjoint rays in this forest must contain a ray in each ill-founded -computable tree. Hence its jump computes the hyperjump . This shows that implies -. To prove that the other imply -, it suffices to exhibit a computable procedure which takes trees to -graphs such that is ill-founded if and only if contains a -ray. For , it suffices to modify each tree by adding a computable branch which is not already on the tree (as we did in the proof of Theorem 5.9). For , it suffices to orient each of the graphs we constructed above in the obvious way. ∎
Proof that - implies .
First, we give a mathematical proof for that is a direct construction not relying on Zorn’s Lemma or the like. We will then explain how to modify it to apply to the other cases and then how to get it to work in -.
Suppose we are given an -graph whose vertices are elements of . We build a sequence of disjoint -rays in by recursion. If there are none we are done. Otherwise start with as any -ray in . Suppose at stage we have constructed disjoint -rays for some such that for each , begins at . If there is a -ray beginning at which is disjoint from the for , choose one as , if not move on to stage . This construction produces a (possibly finite) sequence of disjoint -rays in . We show that this sequence is maximal. If is a -ray which is disjoint from every , then go to stage of the construction, where is the first vertex of . If we did insert some during stage , then would not be disjoint from . Hence we did not insert any -ray during stage . But is a -ray that begins at and is disjoint from , contradiction.
To prove , we modify the above construction as follows. At stage , we search for a -ray which is disjoint from the previous rays and has as its first edge. The rest of the proof proceeds as above.
The only real obstacle in formalizing the above proofs even in is being able to find out at stage if there is an as requested and, if so, choosing one. The question is and so in - we can answer it and then perhaps use some construction or choice principle to produce it. As we only get a yes answer some of the time, - does not seem sufficient. Also later choices depend on previous ones. A computability argument using the Gandy basis theorem and its uniformities works but requires more background development. A choice principle that returns an element if there is one satisfying a property (but may act arbitrarily otherwise) and that can be iterated in a recursion is strong - which consists of the scheme
for any formula . It is known that strong - and - are equivalent [28, VII.6.9]. This clearly has the right flavor and the only issue is defining the required with parameter . This is slightly fussy but not problematic. We provide the details: To define , first recursively define a finite sequence . If have been defined, define to be the least number (if any) above and below such that is a -ray in which is disjoint from . It is clear that there is an arithmetic formula with parameter which defines from and . Next, we say that holds if is a -ray in which begins with , and is disjoint from .
Apply strong - for the formula to obtain some set . By -comprehension with parameter , we may inductively define a (possibly finite) sequence , just as we did in the definition of . Clearly is a sequence of disjoint -rays in . We claim that it is maximal.
Suppose that is a -ray in which is disjoint from every . Suppose that begins with vertex . Then is disjoint from , where is the largest below . It follows that holds. So holds, i.e., is a Z-ray which begins with and is disjoint from . By definition of , that means that . But then and are not disjoint, contradiction.
A slicker proof suggested by the referee requires perhaps more background in the metamathematics of Reverse Mathematics. It uses a “countable coded -model” with the given graph as an element. By -, for any set there is a set such that and is a -model, i.e. any formula with parameters from among the is true if and only if it is true in [28, VII.2.10]. Now one carries out the mathematical proof above but whenever one asks if there is a ray with some property one asks if there is one such among the . This converts the questions to ones of fixed arithmetic complexity in . Then, if the answer is yes, finding an appropriate is also arithmetic in with fixed complexity. This converts the entire construction to one arithmetic in (and so certainly in -). Consider now the claim that the sequence of rays given by this construction over is actually maximal. The existence of a counter-example with first element gives the same contradiction as before. The point is that, as is a -model, the answer to the question asked at stage of the construction of whether there is a ray in with first vertex disjoint from the finite sequence is the same as the question as to whether there exists one at all. The only other point to note is that even though the sequence of rays is constructed outside of each finite initial segment is in by an external induction and the fact that is obviously a model of . The argument with the adjustments for is then the same. ∎
7 Relationships Between and Other Theories of Hyperarithmetic Analysis
In this section, we establish implications and nonimplications between variants of and THAs other than . One such standard theory is as follows:
Definition 7.1.
The theory unique- consists of and the principle
for each arithmetic formula .
The above theory is typically known as weak- (e.g., [28, VIII.4.12]). We deviate from this terminology to introduce a new choice principle where the requirement for unique solutions is replaced by one for finitely many solutions.
Definition 7.2.
The theory finite- consists of and the principle
for each arithmetic formula . Formally, “ nonzero finitely many ” means that there is a nonempty sequence such that for each , holds if and only if for some .
Similarly to -, each of these two choice principles are equivalent to ones where is allowed to be of the form or nonzero finitely many , respectively. However, unlike - neither of these two principle is equivalent to the version where is allowed to be . Not only would those versions fail to capture the idea that we are dealing with unique or finitely many witnesses and paths through trees but they should be stronger than the stated principles. It is easy to see, for example, that even the unique version with implies - (Definition 7.8) which is stronger than unique-- by Van Wesep [30].
Since the THA implies finite- which in turn implies unique- whose models are closed under hyperarithmetic reducibility by Proposition 3.9, it follows that finite- is a THA (as is unique-). Goh [10] shows that finite- is strictly stronger than unique-. We were led to study this version of choice by realizing that a variant of our original proof that implies unique- worked for the finite version.
Theorem 7.3.
implies finite- over . (It follows that implies finite- over , but this is superseded by Theorem 7.7 below.)
Proof.
We first prove that for forests implies finite-. By Lemma 6.7, it suffices to prove that for any sequence of subtrees of such that each has finitely many branches, a sequence exists with each . As in the proof of Theorem 6.12, we construct a sequence of trees such that for each , the branches on are precisely those of the form where is a branch on for .
By (for forests) there is a sequence of -disjoint rays in of maximum cardinality. We claim that is infinite. If not, let be least such that there is no in . Then we can increase the cardinality of by adding any ray from while maintaining disjointness by our choice of . The point here is that if for any then they are both in as the trees are disjoint and there are no edges between them. Therefore has a ray in infinitely many . Thus we may construct the desired sequence recursively by searching at stage for an in for some and take to be the branch in which shares a tail with the th coordinate of .
Another theory of hyperarithmetic analysis which follows from is arithmetic Bolzano-Weierstrass ():
Definition 7.4.
The theory consists of and the following principle: If is an arithmetic predicate on , either there is a finite sequence which contains every such that holds or there is an such that every one of its neighborhoods has two such that holds. Such an is called an accumulation point of the class .
Friedman [8] introduced and asserted that it follows from (with unrestricted induction). Conidis [4] proved Friedman’s assertion and established relationships between and most then known theories of hyperarithmetic analysis. Goh [10] shows that implies finite-. We do not know if is strictly stronger than finite-.
The following two lemmas will be useful in deriving from . The first lemma describes a connection between sets of solutions of arithmetic predicates and disjoint rays in trees.
Lemma 7.5 ().
Suppose is an arithmetic predicate. Then there is a tree such that if there is a sequence of distinct solutions of , then there is a sequence of Y-disjoint single rays in of the same cardinality, and vice versa.
Proof.
By Lemma 6.6, there is a tree such that
and |
If is a sequence of distinct solutions of , then, as the required are arithmetic uniformly in the , there is a sequence of distinct branches on of the same cardinality.
By taking an appropriate tail of each branch, we obtain a sequence of vertex-disjoint (hence edge-disjoint) single rays in of the same cardinality with each one being a tail of : As no two distinct branches in a tree can have infinitely many vertices in common, simply take to be the tail of starting after all vertices it has in common with any , .
Conversely, suppose there is a sequence of -disjoint single rays in . For each , we define a branch on which corresponds to it as follows. Let be the vertex in which is closest to the root of . Then we can extend to the root to obtain a branch on . We claim that is a sequence of distinct solutions of . For each , since and are -disjoint, they cannot share a tail. So and must be distinct. Since for each , there is at most one such that is a branch on , it follows that as desired. ∎
The second lemma is essentially the well-known fact that the Bolzano-Weierstrass theorem is provable in :
Lemma 7.6 ([28, III.2.7]).
proves that if is a sequence of distinct elements of , then there is some which is an accumulation point of .
Theorem 7.7.
implies over . Therefore implies over .
Proof.
By Proposition 6.5, it suffices to show that the undirected variants of imply .
Suppose is an arithmetic predicate on such that no finite sequence contains every such that holds. By Lemma 7.5, there is a tree such that for any sequence of distinct solutions of , there is a sequence of -disjoint single rays in of the same cardinality, and vice versa.
By , or by and Theorem 6.16, there is a sequence of -disjoint single rays in of maximum cardinality. This yields a sequence of distinct solutions of of the same cardinality.
If this sequence is finite, then there is a solution of not in the sequence by our assumption. Hence there is a sequence of distinct solutions of of larger cardinality, which yields a sequence of -disjoint single rays in of larger cardinality for the desired contradiction.
Thus there is an infinite sequence of distinct solutions of . By Lemma 7.6, there is an accumulation point of , which is of course an accumulation point of , as desired. ∎
We now turn our attention to nonimplications. One prominent theory of hyperarithmetic analysis is the scheme of -comprehension (studied by Kreisel [15]):
Definition 7.8.
The theory - consists of and the principle
for all formulas and .
Theorem 7.9.
-.
Proof.
Theorem 7.10.
.
Proof.
By Propositions 5.3 and 6.3, it suffices to show that . Van Wesep [30, I.1] constructed a standard model which satisfies unique-- but not -. Conidis [4, Theorem 4.1], using the approach of [19], showed that satisfies . We show below that does not satisfy .
In order to define , van Wesep constructed a tree and branches of such that (1) contains and infinitely many (distinct) (see [30, pg. 13 l. 1–11]); (2) does not contain any infinite sequence of distinct branches of (see [30, pg. 12 l. 7–9] and Steel [29, Lemma 7].) Then is an instance of in which has no solution in . This shows that does not satisfy for trees. (The reader who wants to follow the details of the proofs in [4] and [30, I.1] should look at the presentation of the basic methods in [19].)
Since is a standard model, it satisfies full induction. By Theorem 6.16, it follows that does not satisfy for forests. ∎
Figure 4 illustrates some of our results. In order to simplify the diagram, we have omitted all variants of except .
8 Isolating the Use of - in Proving
We isolate the use of - in our proofs of and (Theorems 4.5, 5.10, 5.15) by identifying the following principles:
Definition 8.1.
Let be the assertion that if is an -graph with arbitrarily many -disjoint -rays, then there is a sequence of sets such that for each , is a set of -disjoint -rays in .
Let be the assertion that if is an -graph and there is a sequence of sets such that for each , is a set of -disjoint -rays in , then has infinitely many -disjoint -rays.
stands for Strongly Collecting Rays. stands for Weak Infinite Ray Theorem.
It is clear that - implies and implies . The only use of - in our proofs of and is to prove and respectively:
Theorem 8.2.
proves and .
Proof.
For , see the proof of Theorem 4.5(ii). In particular, note that the hypothesis of is exactly the instance of - needed in that proof (and the one referenced there). From then on, the argument proceeds in to give the conclusion of and so of as desired. Similarly, for , see the proof of Theorem 5.10 and for , see the proof of Theorem 5.15. The desired result for then follows from Lemma 5.4. ∎
Next we will use the above result to show that and are equivalent over to and respectively. First, observe that implies for each choice of . Second, Lemmas 5.2, 5.4 and 5.6 imply
Proposition 8.3.
implies , implies and implies .
Proposition 8.4.
implies .
Proof.
By Proposition 8.3, it suffices to establish the desired result for the undirected variants of . The proofs are almost identical to those of Theorems 4.5(i) and 5.9. There, we applied to forests , where each contains a -ray, and no two -rays in can be -disjoint. Any infinite sequence of -disjoint -rays in must contain a -ray in cofinally many graphs . Therefore from such a sequence we can uniformly compute -rays in cofinally many graphs , which establishes by the construction of . If we assume instead of , we only have access to a sequence such that for each , is a set of -disjoint -rays in . From such a sequence we can still uniformly compute -rays in cofinally many graphs , because for any , must contain a -ray in some , . ∎
Corollary 8.5.
and are equivalent over for the following choices of XYZ: XYS and UVD.
Proposition 8.6.
implies , implies and implies .
Recall that and are provable in (Theorem 8.2). and are open, because implies , and and are open (see comments after Theorem 5.1). We do not have an upper bound on the proof-theoretic strength of (an upper bound on would yield an upper bound on , which we do not currently have).
We do not know if any is equivalent to . In an effort to clarify the situation, we define an apparent strengthening of and show that it implies :
Definition 8.7.
Let nonuniform- be the assertion that if is an -graph and there is a sequence of -rays in such that for each , there are such that are -disjoint, then has infinitely many -disjoint -rays.
Every instance of is also an instance of nonuniform-, so nonuniform- implies . Conversely, we have
Proposition 8.8.
implies nonuniform-.
Proof.
Suppose is an instance of nonuniform-, i.e., is an -graph and is a sequence of -rays in such that for each , there are such that are -disjoint. Then can find such uniformly in . Therefore by , is an instance of . By , has infinitely many -disjoint -rays as desired. ∎
Theorem 8.9.
Nonuniform- implies over . It follows that nonuniform- and nonuniform- are both equivalent to over .
Proof.
Next, we show that nonuniform- implies . By Lemma 5.2, it suffices to consider the undirected versions of nonuniform-. First, we prove that nonuniform- implies by constructing a computable instance of nonuniform- such that every nonuniform- solution computes . (The desired result follows by relativization.) We use a variation of the graph used in the analogous result in Theorem 4.2.
Construction of : and some number below is enumerated into at stage , and either or . and . is clearly computable.
Verification: It is clear that there is exactly one ray in beginning with for and the sequence is also computable. Note that if then this ray is . Otherwise, it is . Next observe that for each , and is and contained in and so is a set by bounded comprehension [28, II.3.9]. Thus the finite function taking (the least number in ) to the last stage at which an is enumerated in is also (coded by) a finite set. So we have, for each , a sequence of V-disjoint rays of length as required for the hypothesis of nonuniform-.
Suppose then that is the sequence of rays in a solution for nonuniform-. We wish to compute from this solution. As the are E-disjoint at most one of them contains the edge . So by eliminating that one, we can assume none of the contain . If any of the remaining rays contain some edge of the form for then (as it does not contain ) it must contain for every . Thus there can be at most one such ray among the remaining and so we can discard it and assume there are no such rays in our list. No remaining ray can have as its first vertex as if it did its second vertex would have to be of the form with . Any continuation of this sequence would have to follow the with descending from and so would have to terminate at and not be a ray. Thus all the remaining are of the form for some with for . So the remaining witness the conclusion of nonuniform- as desired.
As we can replace the first vertex of by the sequence beginning with and ending with its second vertex, we know that . Since the sequence and so that of the is infinite, given any we can find an and so compute as as required.
To show that nonuniform- implies , define as above. Consider the graph gotten by adding on for each new vertices for and edges and for . The witnesses for the hypothesis of nonuniform- in supply ones for nonuniform- by tacking on the before in reverse order. The witnesses for the conclusion of nonuniform- can be converted into ones for the conclusion of nonuniform- in by removing the new vertices. So once again we can compute . ∎
We are unable to show that implies , but we can prove the following:
Theorem 8.10.
is not provable in .
Proof.
By Proposition 8.6, it suffices to consider the undirected variants of . For all these variants it suffices to construct a computable graph on and a computable sequence such that (1) for each , the for are pairwise vertex-disjoint double rays in and (2) there is no computable sequence of edge-disjoint single rays in . It is clear that the constructed for this is a counterexample to each in the standard model of with second order part the recursive sets. Of course, as this model is standard, is not provable in , plus induction for all formulas.
The computable construction will be a finite injury priority argument. At the end of stage of our construction, for each , we will have defined a path with lengths strictly increasing with which is intended to be a segment of the double ray . We think of these paths as having domain a segment of containing . Its endpoints are and . The intention is that the will be the desired double rays such that, for each , the for will be vertex-disjoint. We will also have put all numbers less than in as vertices in at least one of these . In future stages, we will not add any edges between vertices which are currently in any for . Thus will be a computable graph given by the union of the double rays . We let be the graph defined so far, i.e. We let be its subgraph defined as the union of the for and similarly for and other interval notations.
We say that the disjointness condition, d.c., holds at stage if for every and distinct and , and are vertex-disjoint. Otherwise we say we have violated the d.c. Clearly, if we never violate the d.c. the (for fixed ) are pairwise vertex-disjoint. We arrange the construction so that we obviously never violate the d.c.
So it suffices to also meet the following requirements:
The requirements are listed in order of priority. During our construction, if all else fails, we will attempt to satisfy each at some stage by merging certain rays and using vertices and which are endpoints of and , respectively. We do this by adding the least new number (the merge point of this merger) as a vertex of as well as edges and which are appended to each with or , respectively, as an endpoint. We also ensure that and henceforth agree after the vertex as they grow in the corresponding directions.
Without loss of generality, and to simplify notation later, we make the assumption that if is convergent for any , and then so is for every .
Construction. At stage of the construction, we are given a finite graph consisting of, for each , finite vertex-disjoint paths as described above. We let be the final stage before at which was initialized. For notational convenience when and are specified we simply write for .
First, we act for the requirement of highest priority with which requires attention as described below. All requirements are initialized and unsatisfied at stage 0 and are initialized and declared to be unsatisfied whenever we act for a with .
We say that requires attention at stage if is not satisfied and there are such that , , is a path in disjoint from which can be extended in only one way to a maximal path in and this extension eventually reaches an which is an endpoint of some for and similarly for and for some such that and are not both endpoints of the same . We also require that the merger using and would not violate the d.c. We then let etc. be the associated witnesses for the least such computation. In this case, the actions for is to perform the merger using and as defined above and declare to be satisfied.
Finally, for each , in turn, which is an endpoint of any the paths as now defined we extend those paths by taking the least new number which we append after in each of these paths (and so add as a new edge). For each , in turn, we also take the next least new numbers and let the be these numbers in order for . This defines the for and completes stage of the construction. As promised we let .
Verification. It is clear that, for each , is a double ray and that is a computable graph consisting of the union of these rays. It is also clear that by construction we never violate the d.c. and so for each the for are pairwise vertex disjoint as required. Thus we only need to prove that we meet each .
We now state a series of facts about each of which follows immediately (or by simple inductions) from the construction and previous facts on the list.
Lemma 8.11.
Every vertex which is a merge point has exactly three neighbors and they are the and used in the merger and the added on after in the final part of the action at the merger stage. In addition, no endpoint of any is a merge point and every vertex which is not a merge point has exactly two neighbors.
If and so for some , then for any , and if so .
If then not both and are merge points. If neither are merge points then . If one, say , is a merge point for a merger at some stage necessarily less than using some and with the edge added on at the end of stage , then the other () is , or ; ; .
Any ray in which begins in remains in until it reaches an endpoint of some .
Each requirement acts at most once after the last time it is initialized. So, by induction, each requirement acts and is initialized only finitely often. Thus there is an infinite sequence such that at each we act for some and we never act for any with afterwards. The observation to make here is that if there were a last stage at which we act for any then there would be no mergers of any and for . In this case all the , for would be disjoint. It is then easy to see from the observations above that at some stage after we would act for a with such that and for .
If is one of the just defined, and then . The point to notice here is that by construction no merger at a stage can use any as every which can act after is initialized at by the choice of the . Thus any path in which starts with a never enters .
Suppose now that is total and defines a sequence of edge-disjoint single rays in . By Lemma 8.11, we may choose a for some after which no for ever acts again.
If is not satisfied at the end of stage , we argue that we act for it later to get a contradiction. Once a ray begins in , it must stay there by Lemma 8.11 and once beyond all the (finitely many) merge points in (none can be put in after stage and no ray has repeated vertices) it remains in some for with which it shares a tail (as each vertex which is not a merge point has exactly two neighbors). So by the edge-disjointness of the there is an such that is not in . It starts in with for some and so shares a tail with some with . Similarly there is such that begins in some with and shares a tail with some with . So eventually we have a stage such that and define paths in which go beyond the points by which and share tails with and , respectively, and after which neither tail contains a merge point and so these paths have unique extensions to paths in (determined by the appropriate tails of and ) ending with the endpoints of and , respectively. Finally, note that the merger of and at would not violate d.c. as by Lemma 8.11 can share vertices only with with and can share vertices only with with . Thus at stage we would act to satisfy for the desired contradiction.
So is satisfied at and was satisfied at some . With the notations as at , and define initial segments of and which can each be extended in only one way to maximal paths in eventually reaching the vertices and , respectively, as described at . The merger performed at puts the merge point in both and and so in and (as the successor of and , respectively, as by Lemma 8.11 is the only neighbor in of other than its predecessor in and similarly for and ). We then add in to at stage . By Lemma 8.11, the only neighbors of are , and so any continuation of and after must produce a shared edge, as can continue only with either which is an edge of or with and can continue only with which is an edge of or . This yields the final contradiction. ∎
9 Open Questions
In addition to the variations of the Halin type theorems investigated here that remain open problems of graph theory ( and ) the most intriguing computational and reverse mathematical questions are about either separating the variants or providing additional reductions or equivalences among the and -. Clearly the most important issue is deciding if any (or even all) the which are known to be provable in - are actually equivalent to it. We extend this problem to include the and .
Question 9.1.
Can one show that any of the which are provable in ( and ) do not imply over or even over ? An intermediate result might be that (for one of these versions) does not imply over .
Should any of these be strictly weaker than -, the question would then be to determine the relations among the and analogously the .
Question 9.2.
Can any additional arrows be added to Figure 1 over or ? (This includes the question of whether .)
As we noted in Remark 5.8 there is an apparent additional reduction in Bowler, Carmesin, Pott [3, pg. 2 l. 3–7]. They use an intermediate reduction to locally finite graphs in the sense of relying on the fact that if a graph has arbitrarily many disjoint rays it has a locally finite subgraph with arbitrarily many disjoint rays. This is the principle to which that Remark refers. It plus is a THA but over it does not imply and is provably very weak (in the sense of being highly conservative over ). Shore [26] proves these results and further analyzes this and many similar principles some related to the and others to an array of classical logical principles.
Any reductions in as requested in the Question above would, of course, provide the analogous ones for the . However, it is possible that other implications can be proven for the :
Question 9.3.
Can any implications of the form be proven in other than the ones known to hold for the versions?
Probably more challenging is the problem of separating the principles.
Question 9.4.
Can one prove any nonimplication over or over for any pair of the ?
Of course, any such separation for the of Question 9.1 would answer a case of that question by proving that at least one of these principles is strictly weaker than -. In addition, a separation by standard models or even ones over for the would give nonimplication for the corresponding but it might be that nonstandard models could be used to separate one pair of versions but not the other.
The next natural question looks below in Figure 4.
Question 9.5.
Can one prove that finite- does not imply over or ?
The weaker versions, , of the , prompt a question about .
Question 9.6.
Do any of the (especially the ones provable from ) imply ? An easier question might be whether they imply ?
References
- [1] Ron Aharoni. König’s duality theorem for infinite bipartite graphs. J. London Math. Soc. (2), 29(1):1–12, 1984.
- [2] Ron Aharoni, Menachem Magidor, and Richard A. Shore. On the strength of König’s duality theorem for infinite bipartite graphs. J. Combin. Theory Ser. B, 54(2):257–290, 1992.
- [3] Nathan Bowler, Johannes Carmesin, and Julian Pott. Edge-disjoint double rays in infinite graphs: a Halin type result. J. Combin. Theory Ser. B, 111:1–16, 2015.
- [4] Chris J. Conidis. Comparing theorems of hyperarithmetic analysis with the arithmetic Bolzano-Weierstrass theorem. Trans. Amer. Math. Soc., 364(9):4465–4494, 2012.
- [5] Reinhard Diestel. Graph theory, volume 173 of Graduate Texts in Mathematics. Springer, Berlin, fifth edition, 2017.
- [6] Yu. L. Ershov, S. S. Goncharov, A. Nerode, J. B. Remmel, and V. W. Marek, editors. Handbook of recursive mathematics. Vol. 2, volume 139 of Studies in Logic and the Foundations of Mathematics. North-Holland, Amsterdam, 1998. Recursive algebra, analysis and combinatorics.
- [7] Harvey Friedman. Algorithmic procedures, generalized Turing algorithms, and elementary recursion theory. In Logic Colloquium ’69 (Proc. Summer School and Colloq., Manchester, 1969), pages 361–389, 1971.
- [8] Harvey Friedman. Some systems of second order arithmetic and their use. Proceedings of the International Congress of Mathematicians (Vancouver, B. C., 1974), Vol. 1, pages 235–242, 1975.
- [9] Harvey M. Friedman. Subsystems of set theory and analysis. 1967. Thesis (Ph. D.)–Massachusetts Institute of Technology, Dept. of Mathematics.
- [10] Jun Le Goh. The strength of an axiom of finite choice for branches in trees. To appear in the Journal of Symbolic Logic.
- [11] Rudolf Halin. Über die Maximalzahl fremder unendlicher Wege in Graphen. Math. Nachr., 30:63–85, 1965.
- [12] Rudolf Halin. Die Maximalzahl fremder zweiseitig unendlicher Wege in Graphen. Mathematische Nachrichten, 44(1‐6):119–127, 1970.
- [13] Denis R. Hirschfeldt. Slicing the truth, volume 28 of Lecture Notes Series. Institute for Mathematical Sciences. National University of Singapore. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2015. On the computable and reverse mathematics of combinatorial principles, Edited and with a foreword by Chitat Chong, Qi Feng, Theodore A. Slaman, W. Hugh Woodin and Yue Yang.
- [14] Pierre Jullien. Contribution à l’étude des types d’ordres dispersés. 1968. Thesis (Ph.D.)–Marseille.
- [15] Georg Kreisel. The axiom of choice and the class of hyperarithmetic functions. Nederl. Akad. Wetensch. Proc. Ser. A 65 = Indag. Math., 24:307–319, 1962.
- [16] Antonio Montalbán. Indecomposable linear orderings and hyperarithmetic analysis. J. Math. Log., 6(1):89–120, 2006.
- [17] Antonio Montalbán. On the -separation principle. MLQ Math. Log. Q., 54(6):563–578, 2008.
- [18] Antonio Montalbán. Open questions in reverse mathematics. Bull. Symbolic Logic, 17(3):431–454, 2011.
- [19] Itay Neeman. The strength of Jullien’s indecomposability theorem. J. Math. Log., 8(1):93–119, 2008.
- [20] Itay Neeman. Necessary use of induction in a reversal. J. Symbolic Logic, 76(2):561–574, 2011.
- [21] Klaus-Peter Podewski and Karsten Steffens. Injective choice functions for countable families. J. Combinatorial Theory Ser. B, 21(1):40–46, 1976.
- [22] Hartley Rogers, Jr. Theory of recursive functions and effective computability. MIT Press, Cambridge, MA, second edition, 1987.
- [23] Joseph G. Rosenstein. Linear orderings, volume 98 of Pure and Applied Mathematics. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York-London, 1982.
- [24] Gerald E. Sacks. Higher recursion theory. Perspectives in Mathematical Logic. Springer-Verlag, Berlin, 1990.
- [25] Richard A. Shore. On the strength of Fraïssé’s conjecture. In Logical methods (Ithaca, NY, 1992), volume 12 of Progr. Comput. Sci. Appl. Logic, pages 782–813. Birkhäuser Boston, Boston, MA, 1993.
- [26] Richard A. Shore. Almost theorems of hyperarithmetic analysis. J. Symbolic Logic, 88(2):664–696, 2023.
- [27] Stephen G. Simpson. On the strength of König’s duality theorem for countable bipartite graphs. J. Symbolic Logic, 59(1):113–123, 1994.
- [28] Stephen G. Simpson. Subsystems of second order arithmetic. Perspectives in Logic. Cambridge University Press, Cambridge; Association for Symbolic Logic, Poughkeepsie, NY, second edition, 2009.
- [29] John R. Steel. Forcing with tagged trees. Ann. Math. Logic, 15(1):55–74, 1978.
- [30] Robert Alan Van Wesep. Subsystems of Second-order Arithmetic, and Descriptive Set Theory under the Axiom of Determinateness. ProQuest LLC, Ann Arbor, MI, 1977. Thesis (Ph.D.)–University of California, Berkeley.