New jump operators on equivalence relations
Abstract
We introduce a new family of jump operators on Borel equivalence relations; specifically, for each countable group we introduce the -jump. We study the elementary properties of the -jumps and compare them with other previously studied jump operators. One of our main results is to establish that for many groups , the -jump is proper in the sense that for any Borel equivalence relation the -jump of is strictly higher than in the Borel reducibility hierarchy. On the other hand, there are examples of groups for which the -jump is not proper. To establish properness, we produce an analysis of Borel equivalence relations induced by continuous actions of the automorphism group of what we denote the full -tree, and relate these to iterates of the -jump. We also produce several new examples of equivalence relations that arise from applying the -jump to classically studied equivalence relations and derive generic ergodicity results related to these. We apply our results to show that the complexity of the isomorphism problem for countable scattered linear orders properly increases with the rank.
1. Introduction
The backdrop for our study is the Borel complexity theory of equivalence relations. Recall that if are equivalence relations on standard Borel spaces then is Borel reducible to , written , if there exists a Borel function such that
The notion of Borel reducibility gives rise to a preorder structure on equivalence relations. As with other complexity hierarchies, it is natural to study operations such as jumps.
Definition 1.1. We say that a mapping on Borel equivalence relations is a proper jump operator if it satisfies the following properties for Borel equivalence relations :
Monotonicity: implies ; | |||||
Properness: whenever has at least two equivalence classes. |
Note that the terms jump or jump operator may be used for a monotone mapping with in a context where strict properness is not relevant or has not been established. We may also apply a jump operator to analytic equivalence relations; in this case we do not expect or require properness. Indeed, for all of the jump operators discussed below, one can find analytic equivalence relations which are fixed points for the mapping up to Borel bireducibility.
One may also ask for some definability condition on a jump operator. While we do not require any particular such conditions, it is the case that all the jump operators discussed below are uniformly definable in the sense that if is a Borel set so that each section is an equivalence relation on , then the set given by iff is also Borel, where denotes the domain of .
Several jump operators have been studied extensively, including the Friedman–Stanley jump [13] and the Louveau jump [24], which we discuss below. There is also a jump operator on quasi-orders introduced by Rosendal [25]; see also subsequent work by Camerlo et al.[6].
Here, we introduce a new class of jump operators which are associated with countable groups.
Definition 1.2. Let be an equivalence relation on , and let be a countable group. The -jump of is the equivalence relation defined on by
We will use the term Bernoulli jump as a collective name for any member of the family of -jumps. Indeed, note that if then is the orbit equivalence relation induced by the classical Bernoulli shift action of , and if for a Polish space then is the orbit equivalence relation induced by the “generalized” Bernoulli action of .
We reserve the notation for the product of countably many copies of with index set . Thus is Borel isomorphic to , and is an equivalence relation of countable index over . Indeed, letting act on by the left shift , we have that are -equivalent iff there is with .
It is clear that the -jump operator is monotone for any . We will be concerned with whether and when the -jump is proper. Before addressing this question, we recall the situation with the Friedman–Stanley and Louveau jumps.
Definition 1.3. Let be a Borel equivalence relation on . The Friedman–Stanley jump of is the equivalence relation defined on by
Theorem 1.1 ([13]). The mapping is a proper jump operator.
Definition 1.4. Let be a Borel equivalence relation on and let be a free filter on . The Louveau jump of with respect to is the equivalence relation defined on by
Theorem 1.2 ([24]). For any free filter , the mapping is a proper jump operator.
The original proof of the Friedman–Stanley result used Friedman’s theorem on the non-existence of Borel diagonalizers (recall a Borel diagonalizer for is a homomorphism from to so that ). However, both the Friedman–Stanley result and the Louveau result can be proved using the concept of potential complexity of equivalence relations, which we briefly introduce. First, we will say that a Borel class is a pointclass consisting of Borel sets and closed under continuous preimages. For example, is a Borel class for any .
Definition 1.5. Let be an equivalence relation on the Polish space , and let be a Borel class. We say is potentially , written , if there is a topology on such that have the same Borel sets and such that is in with respect to .
We remark that is potentially if and only if is essentially , i.e. there is some in with .
Definition 1.6. We say that a family of Borel equivalence relations has cofinal potential complexity if for every Borel class there is such that .
Louveau established that if is an equivalence relation with at least two classes, then the family of iterates of the Louveau jump of with respect to a free filter has cofinal potential complexity. Since being potentially is equivalent to being essentially , a family of cofinal potential complexity cannot have a maximum element with respect to . Hence it follows that the Louveau jump with respect to a free filter is a proper jump operator.
Meanwhile it is known that the family of Borel equivalence relations induced by actions of has cofinal potential complexity. Moreover, Hjorth–Kechris–Louveau [19] established that if is an equivalence relation with at least two classes, then any Borel equivalence relation induced by an action of is Borel reducible to some iterate of the Friedman–Stanley jump of . Thus they arrived at a “potential complexity” proof that the Friedman–Stanley jump is a proper jump operator.
Returning to the Bernoulli jumps, we will establish the following theorem.
Theorem 1. Let be a countable group so that or for prime is a quotient of a subgroup of . Then the mapping is a proper jump operator.
In the proof, we will use a result of Solecki [28] which implies that if is one of the groups or for prime, then the family of Borel equivalence relations induced by actions of the group has cofinal potential complexity. Our main work will be to show that such equivalence relations are Borel reducible to iterates of the -jump. Before stating this result, we provide the following notation for the iterates of the -jump.
Definition 1.7. For an equivalence relation and a countable group we define the iterates of the -jump recursively by
The tower of is of particular interest and will figure in an application to the classification of scattered linear orders. We note that may be replaced by any nontrivial Polish space to produce equivalent iterates for . Also, in the definition of , the third rule may be used for successor ordinals as well.
Returning to the proof of Theorem 1, we will actually consider a more complicated group than . For a countable group , we will introduce the notion of a -tree, which is a tree in which the children of each node carry the structure of a subset of the group . We will see that -trees are closely tied to iterates of the -jump in a way analogous to the way regular trees are tied to iterates of the Friedman–Stanley jump. Namely, the iterated -jump will be Borel bireducible with the isomorphism relation on well-founded -trees of rank .
In particular, we will introduce the full -tree , which may be naturally identified with . Its automorphism group, , is a closed subgroup of , and the group is a closed subgroup of . We will establish the following theorem.
Theorem 2. Let be a countable group. Then for any Borel equivalence relation such that is induced by a Borel action of a closed subgroup of , there exists such that is Borel reducible to .
Since this result applies to the group , when it is combined with the result of Solecki showing such groups have non-Borel orbit equivalence relations, it is sufficient to prove Theorem 1.
The Bernoulli jumps are not always proper jump operators. In particular, we will establish the following theorem.
Theorem 3. Let be a countable group with no infinite sequence of strictly descending subgroups. Then the mapping is not a proper jump operator.
To establish this result, we directly calculate that for such , we have that is Borel bireducible with .
Theorems 1 and 3 leave open the question of precisely when the -jump is proper. Indeed, we shall see that there exist groups which do not meet the hypothesis of either result.
We will also study the structure of specific Borel equivalence relations with respect to the Bernoulli jumps and see that they provide new examples of equivalence relations whose complexity lies between and . We first establish:
Theorem 4. is generically -ergodic.
This result has subsequently been strengthened by Allison and Panagiotopoulos [2] to show that is generically ergodic with respect to any orbit equivalence relation of a TSI Polish group. Using this, we establish:
Theorem 5. We have the following:
. | |||||
. | |||||
and are -incomparable. |
Shani [26] has produced further non-reducibility results about Bernoulli jumps of countable Borel equivalence relations. For instance, he has shown that , and that and are -incomparable. He has also produced another example of an equivalence relation strictly between and which is -incomparable with and .
One application of the theory of Bernoulli jumps is to the classification of countable scattered linear orders. Recall that a linear order is scattered if it has no subordering isomorphic to . The class of scattered linear orders carries a derivative operation where one identifies points such that the interval is finite. This derivative operation may furthermore be used to define a rank function on the countable scattered orders with values in , as we will discuss in Sec. 9. We will establish the following theorem.
Theorem 6. The isomorphism equivalence relation on countable scattered linear orders of rank is Borel bireducible with .
This result, together with the fact that the -jump is proper, implies that the complexity of the classification of countable scattered linear orders increases strictly with the rank.
This paper is organized as follows. In Sec. 2, we establish some of the basic properties of the -jump. In Sec. 3, we compare the -jump with the Friedman–Stanley jump. In Sec. 4, we introduce -trees and relate iterates of the -jump to isomorphism of well-founded -trees. In Sec. 5, we study actions of and prove Theorem 2. In Sec. 6, we investigate the properness of the -jump and in particular prove Theorems 1 and 3. In Sec. 7, we revisit the proof from Sec. 5 and show that in some cases, we can achieve better lower bounds on the complexity of the iterated jumps. In Sec. 8, we investigate new equivalence relations arising from Bernoulli jumps, and compare them against well-known equivalence relations, establishing Theorems 4 and 5. In Sec. 9, we discuss the connection between the -jump and scattered linear orders, and in particular prove Theorem 6.
2. Properties of the -Jump for a Countable Group
In this section, we explore some of the basic properties of the -jump. We begin by recording the following three properties, whose proofs are immediate.
Proposition 2.1. For any countable group and equivalence relations and , we have
(a) | If is Borel (respectively, analytic), then is Borel (respectively, analytic). | ||||
(b) | . | ||||
(c) | If then . |
The next result strengthens Proposition 2.1(b).
Proposition 2.2. If is a Borel equivalence relation and is an infinite group, then .
Proof. If has just finitely many classes, the left-hand side is smooth. Otherwise for any , we have that is Borel bireducible with . Hence, it is sufficient to define a reduction from to .
Let be an infinite subset such that implies . (The set of such is of measure 1.) Define a function by
The -jump may be thought of as a kind of wreath product. Recall that if are groups then the full support wreath product is defined as follows. Let be the full support group product of many copies of , and let act on by the shift. Then is the semidirect product with respect to this action.
Proposition 2.3. If is the orbit equivalence relation of some (possibly uncountable) group acting on , and is any countable group, then is the orbit equivalence relation of acting on .
This result implies that the -jump preserves many properties of orbit equivalence relations. For the next statement, recall that a cli group is one which carries a complete, left-invariant metric.
Corollary 2.1. If is the orbit equivalence relation of a Polish group then is the orbit equivalence relation of a Polish group. The same statement holds if we replace “Polish group” with any of the following: solvable group, cli group, or closed subgroup of .
This follows for cli groups from the preservation of being a cli group under wreath products [14, Theorem 2.2.11]. Note, however, that being induced by a TSI group is not preserved (here a TSI group is one which carries a two-sided invariant metric). Indeed, it is shown in [2] that cannot be the orbit equivalence relation of a TSI group.
By contrast, the Louveau jump does not satisfy any of these preservation properties. For example, the Louveau jump of with respect to the Fréchet filter is , and it is well-known is not reducible to a Polish group action [15]. The Friedman–Stanley jump does satisfy the above preservation property with respect to Polish groups and subgroups of . However, it does not satisfy the above preservation property with respect to solvable groups, TSI groups, or cli groups. For example, the Friedman–Stanley jump of is and it is well-known is not induced by a cli group action [21].
Proposition 2.4. Let be an equivalence relation on . If is a subgroup or quotient of , then .
Proof. We adapt the argument from the case when is an equality relation (see [14, Proposition 7.3.4]). If is a quotient of , let be a surjective homomorphism and define by . Then whenever we have
Next, assume that . Fix an element , and let be defined by
Next, we can relate jumps for finite powers of a group to iterates of the jump for .
Lemma 2.1. For any countable group , we have . In particular, .
Proof. Beginning with the first statement, we show the case of , with larger being similar. Given let be the function from to so that is a code for all of the , viewed from the base point , i.e., . We now define a reduction from to by setting
If is equivalent to , witnessed by , then is equivalent to with the witness for the outer jump and for each coordinate of the inner jump. Conversely if is equivalent to , then there exists such that is equivalent to . It follows that witnesses that is equivalent to .
The second statement follows from Proposition 2.2. □
We can also absorb countable powers in certain -jumps.
Proposition 2.5. If is a Borel equivalence relation and and are infinite groups, then .
Proof. Arguing as in Proposition 2.2, we may assume has infinitely many classes, and find a reduction from to for some . Let . For , define by
In particular, if has a subgroup isomorphic to for some infinite group , then .
Another property of equivalence relations is that of being pinned, which is defined using forcing. We briefly recall the definition; we refer the reader to [21] or [30] for properties of pinned equivalence relations.
Definition 2.1. Let be an analytic equivalence relation on . A virtual -class is a pair where is a poset and is a -name so that , where and are the interpretations of using the left and right generics, respectively. A virtual -class is pinned if there is some from the ground model so that . We say that is pinned if every virtual -class is pinned.
Theorem 2.1. If is pinned then is pinned.
Proof. Suppose is pinned and let be a virtual -class. Then . Hence . We can therefore find a condition and a such that
Now, since is pinned we can find such that for all we have . It follows that . Finally, we claim this is in fact unconditionally forced. Indeed, if is any other condition then . Again using the transitivity of , we conclude that too. □
By contrast, the Friedman–Stanley jump does not preserve pinned-ness, as the relation is not pinned (see [21, 17.1.3]). Kanovei also shows in [21] that pinned-ness is preserved under Fubini products, so that the Louveau jump with the respect to the Fréchet filter does preserve pinned-ness.
Remark 2.1. We feel that Corollary 2.1 and Theorem 2.1 justify the following Proclamation: The -jump is a kinder, gentler jump operator than the Louveau jump or the Friedman–Stanley jump.
We may view the Friedman–Stanley jump and the -jumps as special cases of a more general construction. Let be an equivalence relation on , and let be a permutation group of a countable set . We define the jump on by
We may also ask how -jumps for different groups compare to one another.
Question 2.1. Given a fixed , how many distinct complexities can arise as For countable groups and , when is ?
Here, Shani [26] has characterized strong ergodicity between -jumps of countable Borel equivalence relations in terms of group-theoretic properties, showing in particular:
Theorem 2.2 ([26, Corollary 1.4]). Let be a generically ergodic countable Borel equivalence relation. Then:
| |||||
and are -incomparable. |
This contrasts sharply with the case of group actions, where actions of any two infinite countable abelian groups produce hyperfinite equivalence relations. It also shows that there is no “least” -jump, although the -jump is the most complicated. We do not know whether incomparability can be extended to iterated jumps, such as through some notion of rigidity. Note, for instance, that by Lemma 2.1 we have .
Question 2.2. Are there countable groups and so that and are -incomparable for all
We also introduce two restrictions of Bernoulli jumps.
Definition 2.2. Let be an equivalence relation on and a countable group. The free part of and the pairwise-inequivalent part of are given by
We immediately have . We will see below that in certain cases all three are bireducible, but this is not true in general. Indeed, this fails already for , as is the shift equivalence relation which is bireducible with the universal countable Borel equivalence relation by [11, Proposition 1.8], whereas is the free part of the shift, which is bireducible with the universal treeable countable Borel equivalence relation and hence strictly below (see, e.g. [20, Theorem 3.17 and Corollary 3.28]).
Question 2.3. For which and do we have or ?
3. Comparing -Jumps to Friedman–Stanley Jumps
We begin by recalling that the Friedman–Stanley tower for is defined analogously to the iterated -jump.
Definition 3.1. The equivalence relation for is defined recursively by
We compare the -jump to the Friedman–Stanley jump.
Definition 3.2. We say that is weakly absorbing if has perfectly many classes and .
Note that if has perfectly many classes and , then is weakly absorbing, so in particular this holds for , , and for all .
Lemma 3.1. For any with at least two classes, is weakly absorbing.
Proof. When has at least two classes, is above and hence has perfectly many classes. We define a reduction from to as follows. Fix -inequivalent points and . Given , let it map to the set , where
Lemma 3.2. If has perfectly many classes then .
Proof. Given a non-periodic -sequence of -representatives, we construct a new sequence as follows. For each we first define a real by iff . We then let be the least such that and , if such exists, and otherwise. Finally, we let . Thus, each coordinate has several pieces, and we say that two such coordinates and are equivalent if they are equivalent with respect to .
We claim that the entries are pairwise inequivalent. If this is not the case, we can find such that is equivalent to . Let . Note that . Then the blocks and are pointwise -equivalent. Next, since we have for , and since , we can conclude . Using this reasoning inductively, we can conclude that for all . The fact that can be used right-to-left to obtain the same for negative as well. In other words, is periodic with period . This contradicts our assumption, and completes the claim.
Thus the map is a reduction from the free part of to the pairwise inequivalent part of . Since has perfectly many classes we have , so . □
Theorem 3.1. If is weakly absorbing then .
Proof. Since has perfectly many classes, we may fix with so that . From the previous lemma we then have that there is a reduction from to . Let be the set of periodic elements, where consists of such that for all we have . Let be the least so that if , and if . We now define a reduction from to as follows; since is weakly absorbing this will be sufficient. If , we let map to . If , let map to , noting that the concatenation map provides a reduction from to .
Suppose first that . If then and , so that
Suppose conversely that . We must have either both or both . If then we must have , and there are and so that for , and hence for all , so . Suppose then . Then , so for each there is an with and . Because the ’s are pairwise inequivalent, we conclude that for all , so there is with for all . Thus, so . □
It turns out that is not reducible to in general. In particular the following is shown in [3]:
Theorem 3.2 (Shani). is not potentially . Hence, neither nor is reducible to .
Noting that , we thus have that .
Question 3.1. For which and is ?
A slight modification of the argument in 3.1 can be used to show:
Lemma 3.3. , i.e. is Borel reducible to the pairwise inequivalent part of .
Proof. Note that is hyperfinite, as is any equivalence relation of finite index over , and hence reducible to . Thus, using Lemma 3.2, we may fix pairwise -inequivalent and find a reduction from to , where . We extend to the periodic part of as follows. With the notation from above, for let
The same technique can be applied to the -jump of some other equivalence relations, such as and , but we do not know if this is true for other .
We can now compare the hierarchy with the hierarchy .
Theorem 3.3. For all , is Borel reducible to .
Proof. By induction on . The case of follows from Theorem 3.1 since and , and each iteration of the -jump preserves the property of being weakly absorbing so we may again apply Theorem 3.1. □
As noted earlier, -jumps are gentler than the Friedman–Stanley jump so the reverse of Theorem 3.3 is false.
Proposition 3.1. is not Borel reducible to for any .
Proof. is pinned because the identity relation is pinned, and -jumps and countable products of pinned relations are pinned. On the other hand, is not pinned and being pinned is preserved downward under . □
As , we get the following corollary.
Corollary 3.1. and .
Although none of the ’s are above , we will see below that they are unbounded in Borel complexity.
The result of Shani noted above shows that Theorem 3.3 does not hold for all countable groups , as we do not always have , e.g. for (since ). Proposition 7.1 will show that this also does not hold for . Potential complexity bounds do allow us to give a weaker comparison, which Proposition 7.1 will show to be the best possible in general.
We refer below to the equivalence relations , which are defined in [19, Sec. 6]. These may be viewed as restricted forms of the relations ; we do not need the specifics of their definitions, but note the key point that for all for which they are defined. We also recall that for a pointclass , the pointclass consists of all sets which are the difference of two sets in .
Lemma 3.4. for , and hence for .
Proof. Theorem 2 of [19] shows that for a Borel equivalence relation induced by an action of a closed subgroup of we have iff for not a limit, and iff for a limit. Then is still induced by an action of a closed subgroup of and is in for not a limit or in for a limit. From [19, Theorem 4.1] we then have that for not a limit, and hence by [19, Corollary 6.4] we have in both cases for . The case of is immediate since is countable and hence strictly below . □
From this we conclude the following proposition.
Proposition 3.2. For any countable group and we have , so for any we have .
Proof. Note that and since it is countable. We proceed by induction on . Successor steps follow immediately from the previous lemma. For a limit ordinal , if for all , then
From [19, Corollary 6.4] we then have the following corollary.
Corollary 3.2. For any countable group we have , , and:
for not a limit; | |||||
for a limit. |
4. -Trees
The Friedman–Stanley jump naturally corresponds to the group and its iterates correspond to isomorphism of well-founded trees. Namely, the equivalence relation is bireducible with the isomorphism relation on countable well-founded trees of rank at most . We will see that -jumps naturally correspond to certain group actions, and the iterates correspond to isomorphism of well-founded -trees, which are trees where the children of each node carry the structure of a subset of . We make this precise as follows.
Definition 4.1. Let be a countable group. The language consists of the binary relation together with binary relations . A -tree is an -structure which is a rooted tree with child relation and satisfies the additional -formulas:
, for each with | |||||
| |||||
| |||||
, for each | |||||
, for each . |
We say that a -tree is well-founded if it is well-founded as a tree, and we define rank in the usual way. Note that our definition ensures all -trees are countable. Next, we introduce the full -tree, . The automorphism group of will figure crucially in the following section.
Definition 4.2. Let be a countable group. The full -tree, denoted , is the non-empty -tree which additionally satisfies , and for each :
This produces a categorical theory and defines the full -tree uniquely up to isomorphism. For one model of , we take the universe to be and interpret each by .
Definition 4.3. We let denote the isomorphism relation on -trees. For we let denote the isomorphism relation on well-founded -trees of rank at most .
Since every -tree is isomorphic to a substructure of the full -tree , we have that is induced by an action of its automorphism group, . This group is isomorphic to the infinite wreath power , a cli group, and hence and any other -action is pinned.
We now can relate iterated -jumps to isomorphism of well-founded -trees. Nodes of rank 1 in a -tree carry more structure than in a regular tree, where there are only countably many isomorphism types; hence the indexing differs by 1 from the case of the and tree isomorphism. Also note that is , whereas is .
Proposition 4.1. For each , is Borel bireducible with .
Proof. For , is the shift of on , . Given an element we naturally associate a -tree of rank 2 consisting of a root with children , where we interpret . If , with , then the map given by and is an isomorphism from to . Conversely, let be an isomorphism from to , so . Fix and suppose for some . Then for all with we have , so ; hence . Thus , so . Hence witnesses . For the reverse reduction, given a rank 2 -tree , choose any non-root node and define by . Then , so the map witnesses .
Induction steps are similar. Given , with a reduction , we can send to the tree with children of the root so that the subtree for each to show . The reverse reduction is handled in an analogous manner, as are limit stages. □
Although we will see that in many instances the -jump is proper on the Borel equivalence relations, we will always have that is a fixed point of the -jump. This is analogous to the case of tree isomorphism, which is a fixed point of the Friedman–Stanley jump.
Proposition 4.2. For a countable group , is a fixed point of the -jump, i.e. .
Proof. Given with each coding a -tree , we map to the -tree with children of the root so that the subtree to show . □
We also note that none of the are of maximal complexity among -actions, since they are all pinned.
Proposition 4.3. For every countable group , , so is not Borel complete.
We will see in the following section that is of maximal complexity among -actions.
5. Reducing Actions of to Iterated -Jumps
In this section, we will establish that if is a countable group, then every Borel equivalence relation induced by a Borel action of a closed subgroup of the automorphism group of the full -tree, , is Borel reducible to some iterate of the -jump.
In the following section, we will use this, together with the fact that for certain groups the power has actions of cofinal essential complexity, to show that for such groups the -jump is a proper jump operator. As a preview, note that by [5, Theorem 2.3.5], if is a closed subgroup of then every Polish -space is reducible to a Polish -space; similarly, we may reduce a Borel -space to a Polish -space. Since is isomorphic to a closed subgroup of , if we know has actions of cofinal essential complexity, we will be able to conclude the iterates are properly increasing in complexity.
Recall from the previous section the full -tree may be identified with . We let be the restriction of to branches of length . Then is the wreath product of -many copies of , and is isomorphic the direct limit of the finite wreath powers of .
We are now ready to state the main theorem of this section.
Theorem 5.1. Let be a countable group and let be a Borel equivalence relation induced by a continuous action of a closed subgroup of . Let be an ordinal such that is .
(a) | If , , then . | ||||
(b) | If , a limit, then . | ||||
(c) | If , a limit, , then . |
The proof of this theorem is based on some concepts and techniques from [19], and we begin by recalling the relevant portions of this work, adapted slightly to our setting. Fix a countable group and a closed subgroup of . Let be a orbit equivalence relation given by a continuous action of on a Polish space .
The next definition concerns codes for subsets of . First, we fix for the remainder of this section an open basis for .
Definition 5.1. A -code is a pair where is a well-founded tree of rank so that if is not terminal then for all , and . We write . For , define the Borel set coded below , , by
Let be a -code for . For , let be the rank of in . For , let be its height in , i.e. its length as a finite sequence.
We now define a convenient basis for .
Definition 5.2. Let be an enumeration of a finite partial function from to given by , where for each we have equal to the length of these sequences in . We define .
The collection of all such is a countable basis for . Note that may be empty, and the same partial function will have multiple enumerations. Let consist of all such , and consist of just those with maximum height of elements of the domain at most .
Definition 5.3. Given we define the type of to be the sequence of heights of elements of the domain of . Let denote the set of types. For , let be the size of the domain of all of type , for , and the sum of heights of elements of the domain of (equivalently, heights of elements of the range). For types and , we write if the sequence of heights for extends that for . Observe that if and , and is the type of , then there is a type and of type with , , and .
We next introduce an action of on this basis.
Definition 5.4. Let be the action of on given by , i.e. the corresponding partial function satisfies . Then . We use the same notation for the action of on . For we write if and are in the same -orbit (equivalently, the same -orbit for ). If then and have the same type. Note, though, that acts on the enumerations, not the functions themselves, so two enumerations of the same partial function need not be -equivalent.
For and open subsets of , we define the double Vaught transform
Definition 5.5. For , with , and , define by
When this is essentially a real, and in general for it is a hereditarily countable set of rank . Note that can be recovered from . As in [19], we have the following translation property:
Lemma 5.1. For , .
Definition 5.6. Let be the topology of . For any and define the topology as the one generated by and the sets for and .
Then each is a Polish topology extending . For a set , let . We summarize the key properties of these topologies from [19, Secs. 2 and 3].
Proposition 5.1 (Hjorth–Kechris–Louveau). Let be an open basis for a topology on . With the definitions above, we have
for | |||||
The action of on is continuous for | |||||
If then . The latter condition implies | |||||
If and are then iff iff | |||||
If is for , then is in , so | |||||
The same holds with in place of when is for a limit, and with in place of when is for a limit and . |
In case is we have that is reducible to , so we begin with the case where is for . We show that every Polish -space is reducible to some iterate of the -jump of the identity relation, . We note the modifications for cases for other later.
Definition 5.7. Define the following hereditarily countable sets:
The idea is that codes the topology , and codes the orbit closure of in this topology. Hjorth–Kechris–Louveau establish the following in [19, Lemma 3.2]:
Theorem 5.2 (Hjorth–Kechris–Louveau). if and only if .
We will show that the equality of the sets and can be encoded into iterates of the -jump. We first inductively define functions and equivalence relations reducible to iterates of the -jump in order to encode . The bounds on the number of iterates follow from properties of the -jump established earlier; we give only brief comment on the calculations in proofs in this section, and will discuss these bounds further in Sec. 7. We assume for unless stated otherwise.
Lemma 5.2. For with , and of type , there is a Borel function and an equivalence relation , reducible to for , so that for and and of type , we have
(a) | If then | ||||
(b) | If then . |
Proof. (a) We use induction on . For , let and . If then , and if then .
For , define by
If , fix and with . Then for each we have
(b) If then for all and we have
so by inductive assumption we have . Thus
To see that is reducible to , note that for we have , and iterated jumps starting with are equivalent to those starting with for , as noted earlier, so we may safely consider to be for . Assume now that and the bound holds for all with , so . Then
Definition 5.8. For , let .
Lemma 5.3. For , there is a Borel function and an equivalence relation , reducible to , so that
(a) | If then | ||||
(b) | If then . |
Proof. (a) Let and , where for . Suppose , so there is with . Fix , so we have for all from the previous lemma. Thus for each we have
(b) Suppose , so for each we have . Thus there is so that for all we have
Lemma 5.4. There is a Borel function and an equivalence relation , reducible to , so that
(a) | If then | ||||
(b) | If then . |
Proof. Since can be recovered from we have if and only if for each . Letting and suffices. □
Definition 5.9. For , and , let
Lemma 5.5. There is a Borel function and an equivalence relation , reducible to , so that
(a) | If then | ||||
(b) | If then . |
Proof. For , , , and with let
We are now ready to conclude the proof of Theorem 5.1.
Proof of Theorem 5.1. Let be .
(a) | For , the function is a reduction of to as shown above, so . | ||||
(b) | For , a limit, we repeat the preceding argument using the topology in place of . | ||||
(c) | For , a limit and , we use the topology . |
Corollary 5.1. Let be a countable group and let be a Borel equivalence relation induced by a continuous action of . Then for some .
In particular, the above shows that a Polish -space is reducible to . We can improve this slightly in the case of .
Proposition 5.2. Let be a countable group and let be a Borel equivalence relation induced by a continuous action of a closed subgroup of . If is then .
Proof. Since is a closed subgroup of , is essentially countable by [17, Theorem 3.8], so let be a reduction of to . For each , there is some so that the set is nonmeager, and hence comeager in some . Define the relation on by
Suppose , so there is with . Then as in the proof of Lemma 5.3 there is so that for all we have , so that witnesses . Conversely, if , then there are some , , and with and . Then , so . □
Corollary 5.2. Let be a countable Borel equivalence relation. If for some then .
Proof. Let be a reduction of to . Then the -saturation of the range of is Borel by the Lusin–Novikov Theorem, and . Thus is potentially by [17, Theorem 3.8], so by [5, Theorem 5.1.5] we can refine the topology of so that becomes a Polish -space which is . Thus , and hence , is reducible to . □
Note that, e.g. when we already have that , since this is the shift equivalence relation which is bireducible with by [11, Proposition 1.8]. Thus, the above bound is not always optimal, but we do not know if this is true for other , such as . The following seems quite optimistic:
Question 5.1. If is a countable Borel equivalence relation reducible to some , is reducible to or even to
The same techniques will also show that isomorphism of -trees is maximal among -actions. We introduce a useful augmentation of -trees.
Definition 5.10. A labeled -tree is the full -tree with each node labeled with a real. We use to denote isomorphism of labeled -trees.
We may identify a labeled -tree with a function from to . We can see that isomorphism of -trees is bireducible with isomorphism of labeled -trees.
Lemma 5.6. .
Proof. To see that isomorphism of -trees is reducible to isomorphism of labeled -trees, we can identify a -tree with a subtree of , and assign the labeling where a node is labeled ‘1’ if and ‘0’ if .
For the reverse reduction, let be an injection so that for we have that and are not isomorphic via a shift; such an injection exists since the shift action of on has perfectly many equivalence classes. Let be a labeled -tree. For a node , let . Let be the subtree of given by
Theorem 5.3. Let . For any Polish -space we have .
Proof. It will suffice to show that . Fix an enumeration of , and let be the type for each . For with we let be such that and for . Given , we let be the labeled -tree defined as follows. For with for some we let , and let for other . We claim that the map is a reduction from to .
Suppose first that . Then for with we have
Conversely, suppose via . Let be the subtree of consisting of initial segments of those with (equivalently, ), so that is a Polish subspace of . Define the set of branches is an automorphism of . We claim that is comeager in . To see this, note that for we will have that exactly when each node in appears as with for some , and for any such any node in may be extended to another node in for which this is true. Now, since induces a homeomorphism from to , there are and with (i.e. for all ). Let be the induced automorphisms, so and . Since for all we must then have . □
The group corresponds to the -jump in an analogous way to that of the group with respect to the Friedman–Stanley jump, so it is natural to ask if it satisfies similar properties. For instance, Friedman showed:
Theorem 5.4 ([12, Theorem 1.5]). If is a Borel equivalence relation reducible to some -action, then is reducible to for some .
Allison has noted that the analogous result holds for -actions:
Corollary 5.3 (Allison). If is a Borel equivalence relation reducible to some -action, then is reducible to for some .
Proof. Since is a closed subgroup of , Friedman’s theorem implies that is reducible to some , and hence to some orbit equivalence relation of . By [1, Theorem 2.4], is then reducible to an action of with a potentially equivalence relation. By [1, Theorem 2.7], is then reducible to a continuous action of with a orbit equivalence relation, and hence to some by Theorem 5.1. □
Friedman and Stanley asked in [13] if every given by an -action which satisfies for all must be Borel complete; this remains open. We can similarly ask:
Question 5.2. If is given by an -action which satisfies for all , is reducible to
We note that there are Borel equivalence relations induced by actions of which are not reducible to one induced by an action of . This follows from the result of Allison and Panagiotopoulos that is generically ergodic with respect to any orbit equivalence relation of a TSI Polish group [2, Corollary 2.3] and the fact that is TSI for every countable .
We do not know if there is a canonical obstruction to reducibility to -actions, in the way that turbulence is an obstruction for -actions.
Question 5.3. Is there a dynamical characterization of when a Borel equivalence relation is reducible to an -action
We close this section by noting that Shani has observed that -jumps give “natural” examples of equivalence relations at intermediate levels of the Borel hierarchy, defined in [19, Sec. 6]. Namely, the -jump of (which is in the notation of [19]) is reducible to the relation defined there, and has potential complexity precisely . We do not know for which other equivalence relations and groups this holds, since the relations require pairwise-inequivalent successors at each node, and we do not know in general when is reducible to . We will see in Proposition 7.1 that this same precise complexity is obtained for .
6. Properness of the -Jump
In this section, we consider the question of when the -jump is a proper jump. To begin, we use our results of the previous section together with a result of Solecki to establish that the -jump is proper for a large class of countable groups, including the -jump.
Theorem 6.1. Let be a countable group such that or for a prime is a quotient of a subgroup of . Then the -jump is a proper jump operator.
Proof. Suppose towards a contradiction that is a Borel equivalence relation such that and . Then by induction on we have that for all , successor stages following from our assumption and limit stages from Proposition 2.2 and the fact that for a limit ordinal . Hence for all , so by Theorem 5.1, every Borel orbit equivalence relation induced by an action of is reducible to . It is a theorem of Solecki [28, Theorem 1] that if is one of the groups or for a prime then admits non-Borel orbit equivalence relations, and by [16] it follows that and hence induces Borel orbit equivalence relations of cofinal essential complexity. By our hypothesis and Proposition 2.4 it then follows that induces Borel orbit equivalence relations of cofinal essential complexity, contradicting that they are all reducible to . □
Although the -jump has no Borel fixed points for such , there are always analytic equivalence relations with . Of course, the universal analytic equivalence relation is an example. Moreover, by Proposition 4.2, the isomorphism relation on countable -trees is a fixed point too. Thus, we have the following corollary:
Corollary 6.1. Let be a countable group such that or for a prime is a quotient of a subgroup of . Then is not Borel.
For such , gives a new example of a non-Borel, non-Borel complete isomorphism relation. Other known examples include for instance isomorphism of abelian -groups [13], as well as several given in [29].
Friedman and Stanley’s proof that utilized a theorem of Friedman’s on the non-existence of Borel diagonalizers. We do not know if an analogous result holds for -jumps.
Definition 6.1. A Borel diagonalizer for is a Borel -invariant mapping such that for all and we have .
The following is the analogue of Friedman–Stanley’s application of diagonalizers to the jump:
Lemma 6.1. If then there is a Borel diagonalizer for .
Proof. Let be a Borel reduction from to . We may find so that for all and for all . Define by
If there is with , then , so (noting this is true for from its choice). Thus , so . But , so there would be with , a contradiction. Hence there is no with , so there is no with . But then we also have , and there is hence some with , again producing a contradiction. □
Question 6.1. Is it the case that doesn’t admit a Borel diagonalizer when the -jump is proper Note that this statement is at least as strong as Friedman’s theorem, since a diagonalizer for easily produces a diagonalizer for .
We now turn to the question of finding -jumps that are not proper. Recall that a group satisfies the descending chain condition if it does not have an infinite properly descending chain of subgroups. For example, the Prüfer -groups (also called quasi-cyclic groups) satisfy the descending chain condition. More generally if is quasi-finite, meaning it is infinite with no infinite proper subgroups, then satisfies the descending chain condition.
We use the descending chain condition to obtain the condition that a descending chain of cosets of subgroups has nonempty intersection. One may verify that for countable groups, the descending chain condition is equivalent to this latter condition.
Lemma 6.2. If satisfies the descending chain condition, then is Borel bireducible with for a limit and with at least two classes.
Proof. First, we show for any infinite . Let and . Define by ; then is a reduction from to .
For the other direction, we define a reduction from to , which is bireducible with when has at least two classes. Fix some in the domain of and let
Using Lemma 3.2, one can show that the above result also holds when , but we do not know whether it holds for other groups .
Theorem 6.2. If satisfies the descending chain condition, then . In particular, for such the -jump is not a proper jump operator.
Proof. From the previous lemma we have and so it suffices to define a Borel reduction from to . For this we let .
To see that is a homomorphism, we calculate:
Now, suppose conversely that . For each , let
We do not know if the statement of Theorem 6.2 is tight in the sense that is the least fixed-point. If is a group such that the -jump is improper, one may ask what is the least such that . In the following section, we will show that for all we have .
The following is a consequence of Theorem 6.2 together with the proofs of Theorem 6.1 and Corollary 5.1.
Corollary 6.2. If satisfies the descending chain condition, then the family of Borel orbit equivalence relations induced by continuous actions of the group is bounded with respect to .
For abelian groups , having non-Borel orbit equivalence relations is equivalent to the family of Borel orbit equivalence relations induced by continuous actions of being unbounded with respect to by [16, Theorem 5.11], but we do not know if this holds for . It suffices to ask about :
Question 6.2. If satisfies the descending chain condition, is Borel
From Lemma 6.1, we also have the following corollary.
Corollary 6.3. If satisfies the descending chain condition then there is a Borel diagonalizer for .
We conclude this section by exploring the gap between our results on proper and improper -jumps. In light of Theorem 6.1, it is natural to ask for which countable groups we have non-Borel orbit equivalence relations of or of . For abelian groups , the answer to this question is known in the case of -actions. We recall the following definition from [28].
Definition 6.2. Let be a group and a prime number. Then is said to be -compact if, for any descending chain of subgroups such that , we have . (Here denotes the projection.)
If satisfies the descending chain condition, then is -compact for all primes . Indeed, given as above, for each let . Then is a descending chain of cosets of subgroups of , so their intersection is non-empty, meaning .
An example of a group which is -compact for all , but does not satisfy the descending chain condition, is . Moreover is a group for which the hypotheses of both Theorems 6.1 and 6.2 are not satisfied.
It is natural to ask whether Theorem 6.2 can be generalized to all groups which are -compact for all . Indeed, it is shown in [28, Theorem 2] that if is -compact for all , then does not have a non-Borel orbit equivalence relation. It is further shown in [28] that for abelian groups , this is a complete characterization. The question of which non-abelian admit non-Borel orbit equivalence relations is open. Of course it may also be possible that some -jump is proper without this condition holding.
Question 6.3. Which countable groups give rise to proper jump operators If is -compact for all primes , is the -jump improper Does the group give rise to a proper jump operator
7. Bounds on Potential Complexities
The bounds in the statement of Theorem 5.1 are not always tight, that is, sometimes it is possible to reduce an equivalence relation to fewer iterates of the jump. As a consequence, the lower bounds on the potential complexity of that it gives are also not always tight. The bounds provided in Theorem 5.1 are derived from the definitions of the equivalence relations introduced in the proof, together with Proposition 2.2, Lemma 2.1, and the fact that for limit (see the proof of Lemma 6.2). In general this technique requires about -many iterates of the -jump to ensure the potential complexity has increased by one level in the Borel hierarchy.
In this section, we provide a more direct proof that the iterated jumps have cofinal potential complexity in the special case when , and produce sharp bounds on potential complexity of the iterates of the -jump. As a consequence we obtain an alternate proof that the -jump is proper. Recall that for a pointclass , the pointclass consists of all sets which are the difference of two sets in , and the pointclass is its dual, consisting of the complements of all such sets.
Proposition 7.1. For we have the following:
for not a limit; | |||||
for a limit. |
Proof. The upper bounds follow from Corollary 3.2. To establish the lower bounds, we use the properly increasing tower of equivalence relations defined by Hjorth–Kechris–Louveau in [19, §5] (where our corresponds to their for finite , and to for infinite ), and show that for each . We will let denote throughout this proof. To define , we adopt the following notation. For any we may regard as an element of , and we write for the th element of in . Also for any we let .
We let and . Given , we define if and only if:
(a) | for all , either or , and | ||||
(b) | for all but finitely many , . |
For a limit ordinal we fix an increasing sequence cofinal in and define if and only if:
(a) | for all , either or , and | ||||
(b) | for all but finitely many , . |
Hjorth–Kechris–Louveau show, in [19, Theorem 5.8] and its proof, that (with our indexing) for not a limit, and that for a limit. Hence the desired lower bounds will be established once we show that each .
For and , this is immediate. Next, suppose ; we will show that . We first claim that for each . Admitting this claim, since , we have for any by Proposition 2.5. Hence if then .
To establish the claim, let be given and define for and by
For a limit ordinal , suppose for all ; we show that . A direct modification of the proof of the previous claim gives that . Similarly, a direct modification of the proof of Proposition 2.5 gives that for . Since
We do not know the optimal complexity bounds in the case of groups other than . In particular, we do not know precise bounds on the complexities of the iterates of the -jump.
Question 7.1. What are the exact potential Borel complexities of the ’s
8. Generic -Ergodicity of -Jumps
In this section, we will show that -jumps of countable Borel equivalence relations produce new examples of equivalence relations intermediate between and . Furthermore, we will see that although some -jumps admit Borel fixed points, any such fixed point must be strictly above . The key notion in establishing these results is the following definition.
Definition 8.1. We say that is generically -ergodic if for every Baire measurable homomorphism from to there is a so that is comeager. We say is generically ergodic when it is generically -ergodic.
Note that if is generically -ergodic, and does not have a comeager equivalence class, then . Furthermore observe that each -class is meager. Also, when is an orbit equivalence relation, the following lemma shows that we may assume is Borel when verifying generic -ergodicity.
Lemma 8.1. Let be an orbit equivalence relation on and an equivalence relation on . If is a Baire measurable homomorphism from to , then there is a Borel homomorphism from to with for a comeager set of .
Proof. Let be a dense set so that is continuous on . Let be the Vaught transform, so that is an -invariant dense set (see [5, Lemma 5.1.7]). Let be given by
Our main result will be to show that is generically -ergodic for any countable group . Throughout this section we let be a countably infinite group and fix an increasing sequence of finite subsets with .
Definition 8.2. Let be a Borel -action inducing , and let act coordinate-wise via to induce the orbit equivalence relation . Then together with the shift action of on generates . For each and each , let , so that is a countable basis for .
The following lemma is derived from [18, Theorem 7.3]:
Lemma 8.2. Let be a Borel homomorphism. Then there is a Borel such that for all , and there are -invariant sets with such that for with and we have .
Proof. For each , is countable, so there is some so that is nonmeager, and hence comeager in some . Hence:
For , let be the least (in some fixed enumeration of ) so that . Note that if and then . We can now let be the unique so that . □
Now, is nonmeager for some , and since is induced by a continuous group action with dense orbits it must be comeager. Let be this . Let be a comeager set with continuous on . Let so that is comeager and -invariant.
Lemma 8.3. For all with we have .
Proof. Let with . We can then choose with and . Hence and , so, replacing and by and , we may assume .
Suppose , and let and be disjoint neighborhoods with and . We can then find neighborhoods and in with and , so that if then and if then . Since -orbits are dense, the set is non-empty and open, so it contains some with . But then by the previous lemma; however, then but , a contradiction. □
Lemma 8.4. For all , if there is with then .
Proof. We have , and since is -invariant we have . Since and is a homomorphism we have , and similarly for , so the result follows. □
Now, we are ready for the main result of this section:
Theorem 8.1. is generically -ergodic.
Proof. Let be a Borel homomorphism, and let , , and be as above. By the Kuratowski–Ulam theorem we have
When is generically -ergodic it is also generically -ergodic, so we immediately get:
Theorem 8.2. is generically -ergodic.
Corollary 8.1. is generically -ergodic.
Since is Borel reducible to , we have in particular that is properly below in complexity.
Allison and Panagiotopoulos have since strengthened the last result to show that is generically ergodic with respect to any orbit equivalence relation of a TSI Polish group [2, Corollary 2.3]. More recently, Allison has shown that an equivalence relation is generically ergodic with respect to any orbit equivalence relation with non-Archimedean abelian if and only if is generically -ergodic ([1, Theorem 6.6]), and is generically ergodic with respect to any orbit equivalence relation with non-Archimedean TSI if and only if is generically -ergodic ([1, Theorem 6.5]).
From the above, we can see that -jumps do not have fixed points at the first two levels.
Corollary 8.2. For any countable group we have .
Proof. For any countable , , and is the shift action of on , which is countable and generically ergodic. □
In Sec. 9, we consider which countable Borel equivalence relations are reducible to , and establish in Corollary 9.2 that , which immediately implies that . We may now summarize the situation to see that and are new examples of natural equivalence relations below which provide genuinely new levels in the Borel complexity hierarchy.
Theorem 8.3. We have the following :
| |||||
| |||||
and are -incomparable. |
Previously, the only known examples of equivalence relations between and were the examples constructed in [30], which are non-pinned. Shani has also produced a new example of an intermediate pinned equivalence relation in [26], denoted , which is also incomparable to and .
Theorem 8.4 ([26, Theorem 1.7]). is pinned and satisfies:
| |||||
and for any infinite countable group . |
Note that these results give several other equivalence relations strictly between and , such as and (where is the free group on 2 generators). Shani has shown, as consequences of [26, Corollary 5.6 and Proposition 5.14]:
Theorem 8.5 (Shani). The following hold:
| |||||
| |||||
and are -incomparable; | |||||
for any generically ergodic countable Borel equivalence relation . |
One can also ask if there are any equivalence relations strictly between and .
Question 8.1. If , does either or
From the Hjorth–Kechris dichotomy for , we can obtain a weak partial result.
Lemma 8.5. If then either or .
Proof. Let be a reduction of to . Set iff , so that is of countable index below and . Hence either or . In the first case, let be a reduction of to , and set iff . Then and is of countable index over , so that (see [8, Corollary 8.32]) and hence . In the second case, let be a reduction of to , and define the equivalence relation by iff . Then is a countable analytic equivalence relation with , so is essentially countable (see, e.g. [10, Proposition 7.1]); hence . □
Question 8.2. If , does either or
9. -Jumps and Scattered Linear Orders
In this section, we give an application of our results about the -jump to the classification of countable scattered linear orders. We also consider countable Borel equivalence relations which admit an assignment of scattered linear orders to each equivalence class, and establish some complexity bounds on them. Finally, we study the classification of countable complete linear orders, and give an application to classifying more general classes of countable models.
Definition 9.1. A linear order is said to be scattered if there does not exist an embedding (a one-to-one order-preserving map) from to .
The scattered linear orders admit a derivative or collapse operation as well as a rank function. To begin, define an equivalence relation on by if the interval between is finite. The -equivalence classes are convex, so we may form a quotient ordering . Next, we define equivalence relations for any ordinal as follows. If has been defined, we let iff , where denote the -equivalence classes of . If is a limit ordinal and have been defined for all , we let iff there exists such that .
Now, a linear order is scattered if and only if there exists such that is a single point (see [23, Exercise 34.18]). Thus, we may define the rank of a scattered linear order as the least such that is a single point. Let denote the class of countable scattered linear orders, denote the class of countable scattered linear orders of rank , and for the class of countable scattered linear orders of rank at most .
Proposition 9.1 ([23, Exercises 33.2, 34.18]). The set is a -complete set and the subsets and are Borel sets. Moreover, the function which maps a countable scattered linear order to its rank is a -rank on .
We will write to abbreviate . Furthermore, although is a non-Borel class of countable structures, we will write for the isomorphism relation on all countable scattered orderings.
Proposition 9.2. There is no absolutely reduction from to a Borel equivalence relation. On the other hand, no Borel complete equivalence relation is Borel reducible to .
Proof. For the first statement we recall from [16] (see remarks following Corollary 3.3) that there is no absolutely reduction from to a Borel equivalence relation, where is the isomorphism relation on countable well-orders. Moreover, we have that is absolutely -reducible to (see, e.g. [9, Theorem 3.5]). Finally compositions of absolutely reductions are again absolutely .
For the second statement, it will follow from the results below that is not reducible to , but we can give a more basic proof. Suppose there exists a Borel reduction from some Borel complete equivalence relation to . Then by the Boundedness Theorem, the range of is contained in some . Since Borel complete equivalence relations are not Borel, it follows that is not a Borel equivalence relation. But it is not difficult to see that is Borel, for instance see Proposition 9.1. This is a contradiction. □
The classification of scattered linear orders is closely related to the classification of -trees. We use a modification of the definition of -trees given in Sec. 4, which is slightly simpler and more natural in the context of orders.
Definition 9.2. A scattered order tree is a rooted tree together with, for each node , a linear ordering on the set of immediate successors of which is isomorphic to a subordering of .
If is a scattered order tree, we will use the symbol for the order relation on the immediate successors of any node of . We let denote the space of scattered order trees, and the space of well-founded scattered order trees of rank .
Lemma 9.1. The isomorphism relation on countable scattered linear orders of rank is Borel bireducible with the isomorphism relation on .
Proof. We first show is Borel reducible to the isomorphism relation on . Given a scattered linear order of rank we define a scattered order tree whose nodes consist of the equivalence classes for , with the ordering iff . It is not difficult to see that is a Borel reduction as desired.
To show that the isomorphism relation on is Borel reducible to , we proceed by induction. For the base case, it is clear that both the isomorphism relation on and are bireducible with .
For the inductive step, let and assume that for all there exists a Borel reduction from the isomorphism relation on to . Further assume that for all and all we have that does not contain any -class of size 1. (We may do so without loss of generality, by inserting an immediate successor to each point of .)
Given a scattered order tree of rank , let denote the order type of the children of the root of , so that is a suborder of . For each let denote the subtree of rooted at the th child of , and let be the rank of . Finally, let
Note that the notion of scattered order tree carries less information than our earlier notion of -tree, since for instance there are many scattered order trees of rank 2 up to isomorphism, while there are continuum many (complexity ) -trees of rank 2 up to isomorphism. However, after this level the two classifications do align in complexity, with the ranks adjusted appropriately.
Lemma 9.2. The isomorphism relation on is Borel bireducible with for .
Proof. We begin with the case when . We have already shown that . Meanwhile a scattered order tree of rank 3 may be identified with a suborder of where each node is labeled with a natural number according to the order-type of its set of children, of which there are only countably many, and hence isomorphism of is bireducible with .
For , given a scattered order tree of rank , for any node of rank at least 4 we may first replace any of its children of rank at most 3 by a subset of using the case. We can then identify each remaining level with a subset of to produce a -tree of rank in an isomorphism-preserving way. Similarly, given a -tree of rank , considering each “missing” child of a non-terminal node as a trivial tree we may replace each node of rank 1 by a scattered order tree of rank 3 and order each level as a subset of to produce a scattered order tree of rank in an isomorphism-preserving way, and yielding the desired reductions. □
We remark that it follows from this together with Proposition 9.2 that the isomorphism relation on all scattered order trees is not Borel, but also not Borel complete.
Theorem 9.1. The isomorphism relation on scattered linear orders of rank is Borel bireducible with for .
Proof. We have is bireducible with by Lemma 9.1, the latter is bireducible with by Lemma 9.2, and the latter is bireducible with by Proposition 4.1. □
It follows using Corollary 5.1 that every Borel -action is reducible to some .
It also follows using Theorem 6.1 that the complexity of the classification of countable scattered linear orders increases strictly with the rank. For comparison, we note that Alvir and Rossegger have shown in [4] that the complexity of Scott sentences of scattered linear orders also increases strictly with the rank.
We now turn to the relationship between isomorphism of scattered linear orders and assignments of scattered linear orderings to equivalence classes of countable Borel equivalence relations. We recall the following notion from [22].
Definition 9.3. Let be a countable Borel equivalence relation on , and let be a class of -structures. We say that admits a Borel assignment of structures in to each equivalence class if there is an assignment assigning a structure in with universe to each -equivalence class , so that for each -ary the relation
A more general study of structurable equivalence relations may be found in [7]. Kechris has shown in [22] that every countable Borel equivalence relation which admits a Borel assignment of scattered linear orders to each equivalence class is amenable, and has asked whether the converse is true. It is also a long-standing open question whether every countable amenable Borel equivalence relation is hyperfinite. We give here a partial characterization of when a countable Borel equivalence relation admits a Borel assignment of scattered linear orders to each equivalence class.
Lemma 9.3. If a countable Borel equivalence relation admits a Borel assignment of scattered linear orders of rank to each equivalence class, then .
Proof. For , [11, Theorem 5.1] gives that admits an assignment of orders of rank 1 if and only if is hyperfinite, and hence reducible to . For , suppose is a countable Borel equivalence relation on which admits a Borel assignment of scattered linear orders of rank to each equivalence class. From Theorem 9.1, it will suffice to show that . Given , let be the order assigned to , and let be the finite-interval relation with respect to . Let iff and , so that is a subequivalence relation of and induces a scattered linear order of rank 1 on . Hence, we may fix a reduction from to . We now assign a scattered linear order to each as follows. We start with the -equivalence classes in , so that induces a scattered order of rank on these classes. For each we then replace by the ordering of rank 2 given by to produce an ordering of rank . Since will be encoded in this ordering via , this gives the desired reduction from to . □
Lemma 9.4. Let be a countable Borel equivalence relation. Then if and only if is -orderable, i.e. admits a Borel assignment of suborders of to each equivalence class.
Proof. Suppose first that is a countable Borel equivalence relation with . By Lemma 3.3 we have , so let be a reduction of to . Recall that denotes the product of -many copies of , so is a countable index subequivalence relation of and . Because of pairwise inequivalence, on the range of there is a Borel ordering of -classes with the property that the -classes within each -class are isomorphic to the ordering .
Let be the pullback of under , that is, iff . Then is a subequivalence relation of , and hence countable. Since , it follows from the Sixth Dichotomy Theorem (see Hjorth–Kechris, Theorem 7.1) that is hyperfinite. Hence there is AN assignment of suborders of to each -class. Moreover, letting be the pullback of , we have that the -classes within each -class are isomorphic to a suborder of . It follows that there is a Borel assignment of suborders of to the equivalence classes of .
The converse follows from the previous lemma. □
Note that this is analogous to [11, Theorem 5.1] which gives that a countable Borel equivalence relation is reducible to if and only if it admits a Borel assignment of suborders of to each equivalence class. We do not know whether analogous results for the forward direction hold for higher iterates of the -jump.
Question 9.1. For a countable Borel equivalence relation , does admit an assignment of scattered linear orders of rank iff
As observed in Corollary 5.2, if is a countable Borel equivalence relation with for some , then . In light of Lemma 9.3, Corollary 5.2, and boundedness, we have the following corollary.
Corollary 9.1. If is a countable Borel equivalence relation which admits a Borel assignment of scattered linear orders to each equivalence class, then .
We may then ask:
Question 9.2. If is a countable Borel equivalence relation with , is hyperfinite
Here it may be worth investigating a restricted form of the -jump using finitely-supported wreath products to preserve countability of the equivalence relations, instead of the full wreath product used in the -jump.
Using the above results, we can now establish the following corollary:
Corollary 9.2. .
Proof. Suppose towards a contradiction that . From Lemma 9.4, there would be a Borel assignment of suborders of to the equivalence classes of , and therefore by [22] we would have that was amenable. Since is not amenable, this contradiction completes the proof. □
Next, we consider a class of linear orders that is very closely related to the scattered linear orders, and show that this together with our results above leads to a model-theoretic corollary.
Definition 9.4. A linear order is said to be complete if for every , if has an upper bound then has a least upper bound.
Every countable complete linear order is scattered. Indeed, if is countable and complete then it has just countably many Dedekind cuts. It follows that there is no embedding of into , and that is scattered. On the other hand, one may verify that if is a countable scattered linear order then its Dedekind completion is a countable complete linear order. In the following, we let denote the subset of consisting of the countable complete linear orders.
Theorem 9.2. The statements of Lemma 9.1holds with replaced with . In particular for all we have that the relations , , and are Borel bireducible with one another.
Proof. In the proof of Lemma 9.1, given a sequence we defined a scattered order using the separator . Assuming the are complete, we can ensure that will be complete by using the separator . □
We close this section by mentioning an application to the classification of countable models of certain theories , which was pointed out to us by Ali Enayat. Let be a theory with built-in Skolem functions, and with a unary predicate and a linear ordering of . Further suppose that there is a model of with a proper elementary end extension , that is, is an elementary extension of and . We refer the reader to [27] for further context.
It follows directly from [27, Theorems 1.2 and 2.1(2)] that if is as above, then for any there is a Borel reduction from to the isomorphism relation on countable models of . In particular we have the following remark:
Remark 9.1. If is as above, then for any there is a Borel reduction from to the isomorphism relation on countable models of .
Acknowledgments
We would like to thank Shaun Allison, Ali Enayat, Aristoteles Panagiotopoulos, and Assaf Shani for helpful discussions about the content of this paper. We would also like to thank the referee for several helpful suggestions, in particular for pointing out that our original complexity bounds in Proposition 7.1 were not correct.
References
- 1. S. Allison, Non-Archimedean TSI Polish groups and their potential Borel complexity spectrum, preprint (2020), arXiv:2010.05085. Google Scholar
- 2. , Dynamical obstructions to classification by (co)homology and other TSI-group invariants, Trans. Amer. Math. Soc. 374(12) (2021) 8793–8811. Crossref, Web of Science, Google Scholar
- 3. S. Allison and A. Shani, Actions of tame abelian product groups, preprint (2021), arXiv:2105.05144. Google Scholar
- 4. , The complexity of Scott sentences of scattered linear orders, J. Symb. Log. 85(3) (2020) 1079–1101. Crossref, Web of Science, Google Scholar
- 5. , The Descriptive Set Theory of Polish Group Actions,
London Mathematical Society Lecture Note Series , Vol. 232 (Cambridge University Press, Cambridge, 1996). Crossref, Google Scholar - 6. , On isometry and isometric embeddability between ultrametric Polish spaces, Adv. Math. 329 (2018) 1231–1284. Crossref, Web of Science, Google Scholar
- 7. , Structurable equivalence relations, Fund. Math. 242(2) (2018) 109–185. Crossref, Web of Science, Google Scholar
- 8. , Relative primeness and Borel partition properties for equivalence relations, Trans. Amer. Math. Soc. 375(1) (2022) 111–149. Crossref, Web of Science, Google Scholar
- 9. , On the classification of vertex-transitive structures. Arch. Math. Log. 58(5–6) (2019) 565–574. Crossref, Web of Science, Google Scholar
- 10. , Essential countability of treeable equivalence relations, Adv. Math. 265 (2014) 1–31. Crossref, Web of Science, Google Scholar
- 11. , The structure of hyperfinite Borel equivalence relations, Trans. Amer. Math. Soc. 341(1) (1994) 193–225. Crossref, Web of Science, Google Scholar
- 12. , Borel and Baire reducibility, Fund. Math. 164(1) (2000) 61–69. Crossref, Web of Science, Google Scholar
- 13. , A Borel reducibility theory for classes of countable structures, J. Symb. Log. 54(3) (1989) 894–914. Crossref, Web of Science, Google Scholar
- 14. , Invariant Descriptive Set Theory,
Pure and Applied Mathematics , Vol. 293 (CRC Press, Boca Raton, FL, 2009). Google Scholar - 15. , A Glimm–Effros dichotomy for Borel equivalence relations, J. Amer. Math. Soc. 3(4) (1990) 903–928. Crossref, Google Scholar
- 16. , An absoluteness principle for Borel sets, J. Symb. Log. 63(2) (1998) 663–693. Crossref, Web of Science, Google Scholar
- 17. , Borel equivalence relations and classifications of countable models, Ann. Pure Appl. Log. 82(3) (1996) 221–272. Crossref, Web of Science, Google Scholar
- 18. , Recent developments in the theory of Borel reducibility, Fund. Math. 170(1–2) (2001) 21–52. Dedicated to the memory of Jerzy Łoś. Crossref, Web of Science, Google Scholar
- 19. , Borel equivalence relations induced by actions of the symmetric group, Ann. Pure Appl. Log. 92(1) (1998) 63–112. Crossref, Web of Science, Google Scholar
- 20. , Countable Borel equivalence relations, J. Math. Log. 2(1) (2002) 1–80. Link, Google Scholar
- 21. , Borel Equivalence Relations: Structure and Classification,
University Lecture Series , Vol. 44 (American Mathematical Society, Providence, RI, 2008). Crossref, Google Scholar - 22. , Amenable equivalence relations and Turing degrees, J. Symb. Log. 56(1) (1991) 182–194. Crossref, Web of Science, Google Scholar
- 23. , Classical Descriptive Set Theory,
Graduate Texts in Mathematics , Vol. 156 (Springer-Verlag, New York, 1995). Crossref, Google Scholar - 24. ,
On the reducibility order between Borel equivalence relations , in Logic, Methodology and Philosophy of Science IX,Studies in Logic and the Foundations of Mathematics , Vol. 134 (North-Holland, Amsterdam, 1994), pp. 151–155. Google Scholar - 25. , Cofinal families of Borel equivalence relations and quasiorders, J. Symb. Log. 70(4) (2005) 1325–1340. Crossref, Web of Science, Google Scholar
- 26. A. Shani, Strong ergodicity around countable products of countable equivalence relations, preprint (2019), arXiv:1910.08188. Google Scholar
- 27. , End extensions and numbers of countable models, J. Symb. Log. 43(3) (1978) 550–562. Crossref, Web of Science, Google Scholar
- 28. , Equivalence relations induced by actions of Polish groups, Trans. Amer. Math. Soc. 347(12) (1995) 4765–4777. Crossref, Web of Science, Google Scholar
- 29. , Borel complexity and potential canonical Scott sentences, Fund. Math. 239(2) (2017) 101–147. Crossref, Web of Science, Google Scholar
- 30. , Pinned equivalence relations, Math. Res. Lett. 18(3) (2011) 559–564. Crossref, Web of Science, Google Scholar