Games with filters I
Abstract
This paper has two parts. The first is concerned with a variant of a family of games introduced by Holy and Schlicht, that we call Welch games. Player II having a winning strategy in the Welch game of length ω on κ is equivalent to weak compactness. Winning the game of length 2κ is equivalent to κ being measurable. We show that for games of intermediate length γ, II winning implies the existence of precipitous ideals with γ-closed, γ-dense trees.
The second part shows the first is not vacuous. For each γ between ω and κ+, it gives a model where II wins the games of length γ, but not γ+. The technique also gives models where for all ω1<γ≤κ there are κ-complete, normal, κ+-distributive ideals having dense sets that are γ-closed, but not γ+-closed.
1. Introduction
Motivated by ideas of generalizing properties of the first inaccessible cardinal ω, Tarski [22] came up with the idea of considering uncountable cardinals κ such that ℒκκ-compactness holds for languages of size κ. This became the definition of a weakly compact cardinal. Hanf [12], showed that weakly compact cardinals are Mahlo. Work of Keisler [16] and Keisler and Tarski [17] showed the following theorem:
Theorem. Let κ be an uncountable inaccessible cardinal. Then the following are equivalent to weak compactness:
(1) | Whenever R⊆Vκ there is a transitive set X and S⊆X such that 〈Vκ,∈,R〉≺〈X,∈,S〉. | ||||
(2) | If ℬ⊆P(κ) is a κ-complete Boolean subalgebra with |ℬ|=κ and F is a κ-complete filter on ℬ, then F can be extended to a κ-complete ultrafilter on ℬ. |
Items (1) and (2) are clearly implied by their analogues for measurable cardinals:
(1′) | There is an elementary embedding of V into a transitive class M that has critical point κ. | ||||
(2′) | There is a non-atomic, κ-complete ultrafilter on P(κ). |
Holy–Schlicht Games. This paper concerns several of a genre of games originating in the paper [13] of Holy and Schlicht, which were modified and further explored by Nielsen and Welch [19]. The following small variant of the Holy–Schlicht–Nielsen–Welch games was suggested to us by Welch.
Players I and II alternate moves :
I | 𝒜0 | 𝒜1 | … | 𝒜α | 𝒜α+1 | … |
II | U0 | U1 | … | Uα | Uα+1 | … |
The game proceeds for some length ℓ≤γ determined by the play. The sequence 〈𝒜δ:0≤δ<ℓ≤γ〉 is an increasing sequence of κ-complete subalgebras of P(κ) of cardinality κ and 〈Uδ:0≤δ<ℓ〉 is sequence of uniform κ-complete filters, each Uα is a uniform ultrafilter on 𝒜α and α<α′ implies that Uα⊆Uα′. We assume without loss of generality that 𝒜0 contains all singletons. Player I goes first at limit stages. The game continues until either Player II can’t play or the play has length γ. If Player II can’t play, the game ends and ℓ is the length of the sequence already played.a We denote this game by 𝒢Wγ.
The winning condition. Player II wins if the game continues through all stages below γ.
There are two extreme cases: γ≤ω and γ=2κ. Using item (2) of the characterization of weakly compact cardinals, one sees easily that if κ is weakly compact then II wins the game of length ω.
The situation with the converse is slightly complicated. If κ is inaccessible and Player II can win the Welch game of length 2, then κ is weakly compact. If κ is not inaccessible, then either Player I does not have an opening move, or Player II loses. This follows from work in [1], though stated in a different way there. For completeness, it is proved in Sec. 2.
At the other extreme if κ is measurable one can fix in advance a κ-complete uniform ultrafilter 𝒰 on P(κ) and at stage α play Uα=U∩𝒜α. The converse is also immediate: if the second player has a winning strategy in the game of length 2κ, and the first player plays a sequence of algebras with ⋃α<2κ𝒜α=P(κ), then the union of the Uα’s in Player II’s responses gives a κ-complete ultrafilter on κ.
In [19], Nielsen–Welch proved that Player II having a winning strategy in the game of length ω+1 implies that there is an inner model with a measurable cardinal. This motivated the following:
Welch’s Question. Welch asked whether Player II having a winning strategy in the game of length ω1 implies the existence of a non-principal precipitous ideal.
For the readers’ convenience we recall the definition of precipitousness. An ideal ℐ on a set X is precipitous if for all generic G⊆P(X)/ℐ the generic ultrapower VX/G is well-founded. See [14] or [8] for details of the definition.
The main result of this paper is as follows:
Theorem. If κ is inaccessible, 2κ=κ+ and Player II can win the game of length ω+1 then there is a uniform normal precipitous ideal on κ.
In Sec. 2, we show that even the Welch game of length one is not meaningful if κ is not inaccessible.
We note here that for γ a limit, there is an intermediate property between “Player II wins the game of length γ” and “Player II wins the game of length γ+1”. It is the game 𝒢*γ of length γ that is played the same way as the original Welch game 𝒢Wγ, but with a different winning condition: For Player II to win, there must be an extension of ⋃α<γUα to a uniform κ-complete ultrafilter on the κ-complete subalgebra of P(κ) generated by ⋃α<γ𝒜α.
Precipitous ideals. We are fortunate that Welch’s question leads to a number of more refined results about the structure of the quotients of the Boolean algebras P(κ)/ℐ. We begin by discussing a strong hypothesis
A κ-complete, uniform ideal ℐ on κ such that the Boolean algebra P(κ)/ℐ has the κ+-chain condition is called a saturated ideal.
It follows from results of Solovay in [20] that if ℐ is a saturated ideal on κ then ℐ is precipitous. Thus to show that a property P implies that there is a non-principal precipitous ideal on κ it suffices to consider only the case where κ does not carry a saturated ideal.
The most direct answer to Welch’s question is given by the following theorem:
Theorem 1.1. Assume that 2κ=κ+ and that κ does not carry a saturated ideal. If Player II has a winning strategy in the game 𝒢*ω, then there is a uniform normal precipitous ideal on κ.
We recall that a normal uniform ideal on κ is κ-complete. As a corollary we obtain the following corollary:
Corollary. Under the assumptions of Theorem 1.1, if Player II has a winning strategy in either 𝒢*ω or 𝒢Wγ for any γ≥ω+1, then there is a uniform normal precipitous ideal on κ.
While this is the result with the simplest statement, its proof gives a lot of structural information about the quotient algebra P(κ)/ℐ. We prove the following theorem in Sec. 5:
Theorem 1.2. Assume that 2κ=κ+ and that κ does not carry a saturated ideal. Let γ>ω be a regular cardinal less than κ+. If Player II has a winning strategy in the Welch game of length γ, then there is a uniform normal ideal ℐ on κ and a set D⊆ℐ+ such that
(1) | (D,⊆ℐ) is a downward growing tree of height γ, | ||||
(2) | D is closed under ⊆ℐ-decreasing sequences of length less than γ, | ||||
(3) | D is dense in P(κ)/ℐ. |
In fact, it is possible to construct such a dense set D where (1) and (2) above hold with the almost containment ⊆* in place of ⊆ℐ.
Definition 1.3. Let ℐ be a κ-complete ideal on P(κ) and γ>ω be a regular cardinal. Then ℐ is γ-densely treed if there is a set D⊆ℐ+ such that
(1) | (D,⊆ℐ) is a downward growing tree, | ||||
(2) | D is closed under ⊆ℐ-decreasing sequences of length less than γ, | ||||
(3) | D is dense in P(κ)/ℐ. |
Note that this is weaker than the conclusions of Theorem 1.2.
We will abuse notation slightly and say “D is dense in ℐ+” to mean that D is a dense subset of P(κ)/ℐ.
We will say that an ideal ℐ is (κ,∞)-distributive if P(κ)/ℐ is a (κ,∞)-distributive Boolean Algebra.
In this language, Theorem 1.2 can be restated as saying that Player II having a winning strategy in the Welch game implies the existence of a normal γ-densely treed ideal and the tree has height γ.
We have a partial converse to Theorem 1.2:
Theorem 1.4. Let γ≤κ be uncountable regular cardinals and 𝒥 be a uniform κ-complete ideal over κ which is (κ+,∞)-distributive and has a dense γ-closed subset. Then Player II has a winning strategy in the game 𝒢Wγ which is constructed in a natural way from the ideal 𝒥, and which we denote by 𝒮γ(𝒥).
A proof of Theorem 1.4 is at the end of Sec. 5. We note that if κ carries a uniform, κ-complete ideal which is (κ+,∞)-distributive, then κ must be inaccessible.
How does precipitousness arise? In [11], Galvin et al. introduced the following game of length ω. Fix an ideal ℐ. Players I and II alternate playing
I | A0 | A1 | … | An | An+1 | … |
II | B0 | B1 | … | Bn | Bn+1 | … |
With An⊇Bn⊇An+1 and each An,Bn∈ℐ+. Player II wins the game if ⋂nBn≠∅. We will call this game the Ideal Game for ℐ. They proved the following theorem.
Theorem 2 ([11]). Let ℐ be a countably complete ideal on a set X. Then ℐ is precipitous if and only if Player I does not have a winning strategy in the ideal game for ℐ.
In the proof of Theorem 1.1, we construct an ideal ℐ and show that Player II has a winning strategy in the ideal game for ℐ. In Theorem 1.2, the existence of a dense set D closed under descending ω-sequences immediately gives that Player II has a winning strategy in the ideal game. (See [7] for some information about the relationship between games and dense closed subsets of Boolean Algebras.) The proofs of both Theorems 1.1 and 1.2 are in Sec. 5.
Is this vacuous? So far we haven’t addressed the question of the existence of strategies in the Welch games if κ is not measurable. We answer this with the following theorem. We use the terminology regarding closure and distributivity properties of forcing partial orderings from [4].
Theorem 1.5. Assume κ is measurable and V=L[E] is a fine structural extender model. Then there is a generic extension in which κ is inaccessible, carries no saturated ideals (in particular, κ is non-measurable) and for all regular γ with ω<γ≤κ there is a uniform, normal γ-densely treed ideal 𝒥γ on κ that is (κ+,∞)-distributive. The Boolean algebra P(κ)/𝒥γ does not contain a dense γ+-closed subset.
Corollary 1.6. It follows from Theorems 1.4and 1.5that in the forcing extension of Theorem 1.5
(a) | Player II has a winning strategy 𝒮γdef=𝒮(𝒥γ) in 𝒢Wγ. | ||||
(b) | There is an ideal ℐγ as in Theorem 1.2. |
It will follow from the proof of Theorem 1.5that the winning strategies 𝒮γ in (a) are incompatible with winning strategies 𝒮γ′ for Player II in 𝒢Wγ′ for γ′≠γ in the following sense: If γ,γ′≤κ are regular and γ≠γ′ then it is possible for Player I to play the first round 𝒜0 in such a way that the responses of 𝒮γ and 𝒮γ′ to 〈𝒜0〉 are distinct.
We give a proof of Theorem 1.5 in Sec. 6. The existence of winning strategies 𝒮γ as in (a) for Player II in 𝒢Wγ is a direct consequence of Theorem 1.4. A proof of the incompatibility of strategies 𝒮γ, as formulated at the end of Corollary 1.6, is at the end of Sec. 6.
Strengthenings of Theorem 1.5. We have two variants of Theorem 1.5 that are proved in Part II of this paper. The first deals with a single regular uncountable γ<κ, and shows that it is consistent that γ is the only cardinal such that there is a normal γ-densely treed ideal on κ. The second shows that it is consistent that for all such γ there is a normal γ-densely treed ideal 𝒥γ on κ but that they are all incompatible under inclusion.
Similar statements about the relevant strategies in the Welch games are also included. Explicitly
Theorem 1.7. Assume κ is a measurable cardinal, γ<κ is regular uncountable and V=L[E] is a fine structural extender model. Then there is a generic extension in which κ is inaccessible, carries no saturated ideals (in particular, κ is non-measurable) and there is a uniform, normal γ-densely treed ideal 𝒥γ on κ that is (κ+,∞)-distributive. Moreover, in the generic extension:
(a*) | There does not exist a uniform ideal 𝒥′ over κ such that P(κ)/𝒥′ has a dense γ′-closed subset for any γ′>γ. | ||||
(b*) | Player II does not have any winning strategy in 𝒢Wγ′ where γ′>γ. |
In particular, it is a consequence of (a*) that
(c*) | For all regular γ′>γ there is no uniform normal γ′-densely treed ideal on κ. |
Another modification of the proof of Theorem 1.5 which is based on Theorem 1.9 below yields the following variant of Theorem 1.5.
Theorem 1.8. Assume κ is a measurable cardinal, and V=L[E] is a fine structural extender model. Then there is a generic extension in which κ is inaccessible, carries no saturated ideals (in particular, κ is non-measurable) and for all regular γ with ω<γ≤κ there is a uniform, normal γ-densely treed ideal 𝒥γ that is (κ+,∞)-distributive. The relationship between the ideals and strategies for different γ’s is as follows:
(a*) | There does not exist a uniform normal ideal 𝒥′⊆𝒥γ over κ such that P(κ)/𝒥′ has a dense γ′-closed subset for any γ′>γ. | ||||
(b*) | The strategy 𝒮γdef=𝒮(𝒥γ) is not included in any winning strategy for Player II in 𝒢Wγ′ where γ′>γ. | ||||
(c*) | Letting ℐγ be the ideal arising from the strategy 𝒮γ, there does not exist an ideal ℐ⊆ℐγ which is γ′-densely treed as witnessed by a tree D⊆ℐ+ of height γ′, for any γ′>γ.b |
In other words, the ideals ℐ in (c*) in Theorem 1.8 are like ideals ℐ in Theorem 1.2, with γ′ in place of γ.
The models constructed in Theorems 1.7 and 1.8 require more sophisticated techniques than those used in the proof of Theorem 1.5. They involve the relationship between the fine structure in the base model and the forcing extension.
The most substantial difference is that the model in Theorem 1.5 is built by iteratively shooting clubs through the complements of non-reflecting stationary sets which have been added generically, however the proofs of Theorems 1.7 and 1.8 shoot club sets through non-reflecting stationary sets built from canonical square sequences constructed in the fine structural extender model. Unlike the partial orderings used in the construction of a model in the proof of Theorem 1.5, those partial orderings will have low closure properties, but high degree of distributivity. It is the proof of distributivity of the iterations of club shooting partial orderings which uses the significant fine structural properties of the extender model. Here is the result allowing the desired iteration.
Theorem 1.9. Assume V=L[E] is a fine structural extender model and κ is a measurable cardinal as witnessed by an extender on the extender sequence E. Assume further that
(i) | (cξ|ξ<α+) is a canonical square sequence,c | ||||
(ii) | Sα⊆α+∩cof(<α), | ||||
(iii) | Sα∩cξ=∅ for all ξ |
whenever α is a cardinal.
Let ℙδ be the Easton support iteration of length κ of club shooting partial orderings with initial segments where each active stage α is an inaccessible ≥δ and the club subset of α+ generically added at stage α is disjoint from Sα. Then there is an ordinal ϱ<κ such that for every inaccessible δ such that ϱ<δ<κ the following holds.
(a) | ℙδ is δ+-distributive. | ||||
(b) | If G is generic for ℙϱ over V and j:V→M is an elementary embedding in some generic extension V′ of V which preserves κ+ then j(ℙϱ)/G is κ+-distributive in V′. |
Although Theorem 1.9 is formulated for Easton support iterations with inaccessible active stages, variations which involve iterations with supports which are not necessarily Easton, but still sufficiently large, and with active stages that are not necessarily inaccessible can also be proved.
As the proof of Theorem 1.9 is of considerable length and (we believe) has broader applicability and is of interest on its own, we will postpone the proof to Part II of this paper.
Basic definitions and notation. We now present terminology and notation we use throughout the paper. We will use the phrases “ideal on κ” and “ideal on P(κ)” interchangeably. Perhaps ideals should be viewed as subsets of Boolean algebras, but the former phrase is the more common colloquialism.
Fix a regular cardinal κ and ℐ a κ-complete ideal on κ. We say that A⊆ℐB if A\B∈ℐ, and ⊇ℐ is the converse relation. The notations ⊆*, ⊇* are these notions when ℐ is the ideal of bounded subsets of κ. The notation A⊊ abbreviates the conjunction of and , where △ means symmetric difference.
The ideal induces an equivalence relation on by if and only if . The notion induces a partial ordering on , we will sometimes call this and refer to the set of equivalence classes of that don’t contain the empty set as . We will force with viewed either as a Boolean algebra, or removing the equivalence class of the empty set as a partial ordering. These are equivalent forcing notions. Occasionally, we will abuse language by saying “forcing with ” when we mean this forcing.
Definition 1.10. -complete sub-Boolean algebras of that have cardinality are called -algebras.
If and are sequences we will use to mean the concatenation of and . We will abuse this slightly when has length one. For example, given and we will write for the sequence of length whose first elements coincide with and whose last element is .
Usually our trees grow downwards, with longer branches extending shorter branches. A tree is -closed if when b is a branch through whose length has cofinality less than there is a node such that is below each element of b. Occasionally, we will say -closed to mean -closed.
2. Weak Compactness
In this section, we clarify the relationship between these games and weak compactness and discuss the role of inaccessibility in the work of Keisler and Tarski. It has been pointed out to us that these results appear in work of Abramson et al. [1] stated slightly differently and with different proofs. We include them here for completeness and because these techniques are relevant to the topics in this paper.
If is inaccessible and is a -algebra and then generates a -complete subalgebra of that has cardinality (i.e. another -algebra). The situation where is not inaccessible is quite different.
Proposition 2.1. Suppose that is an infinite cardinal and either
• | a singular strong limit cardinald or | ||||
• | for some , but for all , . |
Then there is no Boolean subalgebra such that , is -complete.
Proof. In the first case, since is singular, if is -complete, it is -complete. For , let . Then is an atom of and each non-empty contains some . Moreover distinct ’s are disjoint. Thus, the generate as a -algebra. If there are many distinct then . Otherwise, since is a strong limit, .
Assume now that , and for all . Since , must have fewer than atoms. If is the collection of these atoms, the -algebra generated by the atoms of has cardinality at most . Let and . Since and has cardinality , there is an that does not belong to . Since a is not in the set . Hence, C is non-empty. Replacing with we get an atomless, -complete algebra on the set C.
Since no element of is an atom we can write each as a disjoint union of non-empty elements of . Build a binary splitting tree of elements of of height by induction on as follows:e
• | is the -maximal element C of . | ||||
• | Suppose is built. For each , write A as the disjoint union such that , and and let . | ||||
• | Suppose that is a limit ordinal. Let is a branch through and . |
Note that for all , each element determines a unique path through of length . Hence .
Fix a and let be the branch through determined by c. For , the tree splits with . The sets each belong to and form a collection of disjoint subsets of C of size . By taking unions of these sets we see that . □
In contrast to Proposition 2.1, we have
Proposition 2.2. Let be infinite and . Then and is -complete.
Proof. Immediate. □
The upshot of Propositions 2.1 and 2.2 is the following theorem and corollary, which show that the Welch game is only interesting when is inaccessible.
Theorem 2.3. Suppose that is an accessible infinite cardinal. Then either
(1) | There is no -algebra with or | ||||
(2) | There is a -algebra with but every -complete ultrafilter U on is principal |
Corollary 2.4. Consider the Welch game of length 1. Suppose that there is a -algebra that is a legal move for Player I and that Player II has a winning strategy in . Then is inaccessible.
If is inaccessible we have the following result, which can also be deduced directly from the results of Abramson et al. [1]:
Theorem 2.5. Suppose that is inaccessible and Player II wins the Welch game of length 1. Then is weakly compact.
Proof. To show is weakly compact, it suffices to show it has the tree property. Let be a -tree. For , let be the set of such that . Let be the -algebra generated by and be a uniform -complete ultrafilter on .
For each , where . It follows that for each there is an such that . But then is a -branch through . □
3. Hopeless Ideals
In this section, we define the notion of a hopeless ideal in a general context, and toward the end of the section we will narrow our focus to the context of games. Fix an inaccessible cardinal . Assume F is a function with domain R such that for every the value is a sequence of length of the form
(i) | and | ||||
(ii) | is a -complete ultrafilter on the -algebra of subsets of generated by . | ||||
(iii) | For all , the sequence is monotonic with respect to the inclusion. | ||||
(iv) | (Density) For every , and there is such that and . |
We will call functions F with the properties (i)–(iv) assignments.
One can also formulate a variant with normal ultrafilters . Denote the maximo-lexicographical ordering of by . Let be the natural isomorphism. For a set , let be the section of . The sequence is associated to A. We will say that a -algebra of subsets of is normal if for all , each belongs and the diagonal intersection also belongs to . We will say that a sequence belongs to if it is associated to an element of . Finally we say that an ultrafilter U on a normal -algebra is normal if and only if for every sequence ,
(ii) | is a -complete normal ultrafilter on the normal -algebra of subsets of generated by . |
If (ii) is satisfied we say that F is normal. Note that there is no need to modify clause (iv), as normal -algebras are able to decode -sequences of subsets of from other subsets of via the pairing function h introduced above. However, instead of families , it is convenient in (iv) to consider sets that code .
Definition 3.1. Given an assignment F, we define the ideal as follows.
Although in the above definition we say we are defining an ideal, an argument is needed to see that is indeed an ideal. It follows immediately that and is downward closed under inclusion. The rest is given by the following proposition.
Proposition 3.2. Given an assignment F, the ideal is -complete. If all ultrafilters are uniform then is uniform. If additionally F is normal then is normal. If F’ is an assignment on and , then .
Proof. We first verify -completeness of . We noted above that and is downward closed under inclusion; hence it suffices to check that is closed under unions of cardinality . If is such that and , then there is some and some such that . By the density condition, there is some such that and for all , and . In particular, and all sets , are in the -algebra generated by . By -completeness of then for some , hence .
The proof of normality of for normal F is the same, with in place of . The conclusion on uniformity of follows by a straightforward argument from the definition of .
Finally, if are as in the statement of the proposition, then any trivially avoids all ultrafilters where and , so . □
Now assume is a two player game of perfect information, and is a strategy for Player II in . Denote the set of all runs of according to by (by a run we mean a complete play). Assume every is associated with a sequence of fragments and ultrafilters , in a way that makes the function
Here are some examples. If is the Welch game then is the identity function. In the following section, we introduce games and . These games are defined relative to a sequence of models increasing with respect to the inclusion, and Player I plays ordinals which refer to these models. In the games and Player II plays uniform -complete ultrafilters on ; in these ultrafilters are required to be normal. Thus, if r is a run in one of these games according to , say then
If P is a position in played according to we let
Definition 3.3. Assume is a game of perfect information played by two players, is a strategy for Player II in , and is an assignment with domain as in (4). Consider a position P in according to . We define
to be the hopeless ideal with respect to conditioned on P. Here we suppress mentioning the assignment in the notation, as in all situations we will consider it will be given by the strategy in a natural way. The ideal is called the unconditional hopeless ideal with respect to . We will write for .
When the strategy is clear from the context we suppress referring to it, and will talk briefly about the “hopeless ideal conditioned on P” and the “unconditional hopeless ideal”. By Proposition 3.2, we have the following as an immediate consequence.
Proposition 3.4. Given a game of limit length, a strategy for Player II in and a position P as in Definition 3.3, the ideal is -complete. If all ultrafilters associated with are uniform then is uniform. If moreover is normal, then is normal as well.
4. Games We Play
In this section, we introduce a sequence of games closely related to Welch’s game . The last one will be , and we will be able to show that if is a winning strategy for II in of sufficient length then we can construct a winning strategy for Player II in such that is precipitous and more, depending on the length of the game and the payoff set.
To unify the notation, we let of length be the Welch game . Thus, a run of the game continues until either Player II cannot play or else until rounds are played. The set of all runs of of length is denoted by . As usual with these kinds of games, a set is called a payoff set. We say that Player II wins a run R of the game of length with payoff set B if R has rounds and the resulting run is an element of B. We call this game . Thus, if then is just the game . With this notation, the game is just the game of length where
(TO1) | is the same game as whenever is a successor ordinal, so a winning strategy for Player II in gives us something new only when is a limit. | ||||||||||||||||||||||||||||||||
(TO2) | A winning strategy for Player II in is a winning strategy for Player II in , but the converse may not be true in general. | ||||||||||||||||||||||||||||||||
(TO3) | If is a winning strategy for Player II in of length then the restriction of to positions of length is a winning strategy for Player II in of length . | ||||||||||||||||||||||||||||||||
(TO4) | Given and sequences and where and is a -complete ultrafilter on the -algebra (respectively, normal -algebra) of subsets of generated by such that whenever , there is at most one -complete (respectively, normal) ultrafilter U on the (normal) -algebra of subsets of generated by which extends all . Thus, if we changed the rules of to require that Player II goes first at limit stages then Player II has a winning strategy in this modified if and only if Player II has a winning strategy in the original game . | ||||||||||||||||||||||||||||||||
(TO5) | Let be a winning strategy for Player II in or and
be a play of the game or according to . Let be another sequence of -complete algebras. Then, the play
is a run of the game where Player II wins. |
In what follows, we will consider a regular cardinal much larger than , and fix a well-ordering of which we denote by . We augment our language of set theory by a binary relation symbol denoting this well-ordering, and work in this language when taking elementary hulls of . We will thus work with the structure , but will frequently suppress the symbols denoting and in our notation.
The common background setting for the games we are going to describe is an internally approachable sequence of elementary substructures of . That is: a continuous sequence such that for all the following hold.
(a) | and , | ||||
(b) | , | ||||
(c) | whenever . |
The following are standard remarks:
• | If we are playing any of the games of then the game has length . Since , for all . | ||||
• | If is an internally approachable sequence then there is a closed unbounded set such that for , . | ||||
• | If , then there is a well ordering of of order type in . Hence, if is an internally approachable sequence then ). Clearly, implies that , which we stated as an assumption in Theorems 1.1 and 1.2. |
Definition 4.1 (The Game 𝒢1−). The rules of the game are as follows. Fix an ordinal .
• | Player I plays an increasing sequence of ordinals . | ||||
• | Player II plays an increasing sequence of uniform -complete ultrafilters on where . | ||||
• | Player I plays first at limit stages. |
A run of continues until Player II cannot play or until it reaches length . Player II wins a run in if and only if the length of the run is .
Payoff sets and for are defined analogously to the definition for the game . So consists of all runs of of length , and consists of all runs such that there is a -complete ultrafilter on the -algebra generated by , where , extending all , .
The symbols and have a double usage: They were also defined in connection with the game and were different, but analogous to that in Definition 4.1. Thus, to determine the exact meaning of and one always needs to take into account which game is being considered.
In the case where , if Player II has a winning strategy in any of the games then is measurable. So for the purposes of this paper we can assume that , in particular for every .
Remark. Let be a full or partial play of the game and .
(1) | If has cofinality then is a -algebra. | ||||
(2) | If , then , and again is a -algebra. | ||||
(3) | if is a limit ordinal of cofinality less than , then the -algebra of sets generated by is not a -algebra, the -algebra it generates is strictly larger. |
Finally, let us stress that remarks analogous to the remarks (TO1)–(TO5) that stated below formula (7) for games and also hold for and .
Proposition 4.2. Assuming and is an infinite regular cardinal, the following hold.
(a) | Player II has a winning strategy in of length if and only if Player II has a winning strategy in of length . | ||||
(b) | Player II has a winning strategy in of length if and only if Player II has a winning strategy in of length . |
Moreover, the analogues of the above equivalences (a) and (b) also hold for winning strategies for Player I in the respective games.
Although the last statement in the above proposition concerning winning strategies for Player I is not strictly relevant for this paper, we include it for the sake of completeness.
Proof. This is an easy application of auxiliary games. Regarding (a), if is a winning strategy for Player II in then induces a winning strategy for Player II in the output of which at step i is the same as the output of at step i in the auxiliary game where Player I plays at step i (where is the move of Player I in at step i). For the converse, we proceed similarly. This time a winning strategy for Player II in induces a strategy for Player II in as follows. If Player I plays at step i in then Player I plays
It is straightforward to verify that this choice of strategies also works in the case of games with payoff sets in (b).
Because we will not study winning strategies for Player I in the games we consider, we leave the proof of the last statement in the proposition concerning these strategies to the reader. The proof is based on the same ideas as the proof of (a), (b) above. □
We will use the following lemma:
Lemma 4.3. Suppose that is the least winning strategy for Player II in and be the strategy defined from as in Proposition 4.2. Suppose that and is a sequence of ordinals such that for all . Then Player II’s response to in belongs to .
Proof. Because the sequence is internally approachable and , . Since we are taking and is closed under -sequences, the sequence of -algebras belongs to . Since is -least, the sequence of responses by Player II to in belongs to , and hence the sequence of responses by Player II according to as defined in Proposition 4.2 belongs to . □
Definition 4.4 (The Game 𝒢1). The rules of are exactly the same as those of with the only difference that the ultrafilters played by Player II are required to be normal with respect to .
As before, the payoff set is defined for the same way as it was for and , that is, consists of all runs of of length . For we define a payoff set as follows:
Note that whenever , so the game is of interest only for . The existence of a winning strategy for Player II in of length seems to be exactly what is needed to run the proof of precipitousness of the hopeless ideal in Sec. 5; see Proposition 5.7. As we will see shortly, the existence of such a winning strategy follows from the existence of a winning strategy for Player II in of length .
In the case of we will not make use of a payoff set for that would be an analogue of what was for and , so we will not introduce it formally. We note that is a subset of , so the winning condition for Player II is weaker using .
Let us also note that the somewhat abstract notion of normality of an ultrafilter on introduced in Sec. 3 is identical with the usual notion of normality with respect to the model where it is required that is closed under diagonal intersections of sequences such that for all .
Remark 4.5. If we have a strategy defined for either or , then a play of the game according to this strategy is determined by Player I’s moves. Thus, if is clear from context we can save notation by referring to plays as sequences of ordinals . Similarly if is a partial strategy defined on plays of length at most we can index these plays according to by , where . This allows strategies to be defined by induction on the lengths of the plays.
Proposition 4.6 (Passing to normal measures). The following correspondences between the existence of winning strategies for and hold.
(a) | Let be an infinite regular cardinal. If Player II has a winning strategy in of length then Player II has a winning strategy in of length . (So in fact we have “if and only if” here, as the converse holds trivially.) | ||||
(b) | If Player II has a winning strategy in of length then Player II has a winning strategy in of length . |
We do not know whether there is an analogue of Proposition 4.6 with respect to strategies for Player I.
Proof. We begin with some conventions and settings. Let be the transitive collapse of . We will work with models in place of .
Since , we have
so the games and can be equivalently defined using structures instead of .
If U is an M-ultrafilter over we denote the internal ultrapower of M by U by . Then is formed using all functions which are elements of M. If U is -complete then is well-founded, and we will always consider it transitive; moreover the critical point of the ultrapower map is precisely . Recall also that U is normal if and only if , that is, is represented in the ultrapower by the identity map. As (by we mean without the power set axiom), the Łoś Theorem holds for all formulae, hence the ultrapower embedding is fully elementary. Finally recall that the M-ultrafilter derived from , which we denote by , is defined by
Assume and is a -complete -ultrafilter. Suppose that . We have the following diagram:

The property that can be restated as saying that if f represents in and g represents in , then . Equation (13) can also be rephrased in this way.
Note that since are ultrafilters on the respective models, the statement can be equivalently expressed as .
Before we define the winning strategies for Player II in , we prove two useful facts about the normalization process. The first says there can’t be an infinite sequence of ultrafilters that disagree on their normalizations.
Lemma 4.7. Let be an increasing sequence of ordinals between and . Then there is no sequence of ultrafilters such that
• | is a -complete ultrafilter on | ||||
• |
| ||||
• | there is a countably complete ultrafilter V on with for all n. |
Proof. For each n let represent in , and let be the map from to defined as in Eq. (10). Then there is a set such that for all . The ’s all belong to V and intersecting them we get a such that for all , a contradiction. □
We note that Lemma 4.7 implies that in a play there is no infinite increasing sequence such that .
Let be a partial or complete play of the game of limit length . Suppose that . Then the transitive collapse of is the direct limit of along the canonical functions in diagram (9). Denote the transitive collapse of by . Let be a -complete ultrafilter defined on the -algebra generated by that extends . The following lemma implies that if for some then for some j with , .
Lemma 4.8. Let with members of the ’s. Let be -complete ultrafilters on the respective ’s of . Suppose that . Let be such that .
Then we can choose such that
Proof. If any of are principal the hypothesis clearly fails. It follows that each of the ultrafilters is uniform.
The point of the proof is showing that if belong to , then . Since but does belong to , it follows that . So witnesses the conclusion of the lemma.
Using the notation of diagram (9), since
Since , and , we must have . Since and belong to and we have , and hence .
To finish it suffices to show that . Since , we must have . Thus, .
For , let be the constant function . Then , by -completeness. Using induction and the -completeness of , one proves that . But then
It follows from Lemmas 4.7 and 4.8 that if is a play of where is zero or a limit ordinal and , then there is a finite set such that for all
(A) | for all , it holds that , | ||||
(B) | for all , . |
We will call the stages together with drops. Note that in clause (B), so this does not imply that is a drop.
A position P of the game has the form
Claim 4.9. Assume one of the following holds
(a) | is regular and is a winning strategy for Player II in of length . | ||||
(b) | and is a winning strategy for Player II in of length . |
Then is a well-founded tree.
Proof. It is immediate that is a tree. The well-foundedness follows from the fact that there can be only finitely many drops along a play of the game. □
The proof of Claim 4.9 implies that if is a winning strategy in any of the variants of of any length , then is well-founded. Note for well-foundedness the only relevant are limit ordinals. As stated, the Claim handles all of the cases relevant to the theorems we are proving.
Now assume is as in (a) or (b) in Claim 4.9. For , let be the largest drop in P if P does have a drop, and otherwise. Fix a -minimal . By the minimality of P, if extends P then ; in other words, has no drops above , hence whenever .
Let and .
We define a winning strategy for Player II in of length . Viewing as defined on sequences of ordinals , we define on such sequences by induction on their length .
For ordinals played by the first player we assume inductively that the normal ultrafilter played by the second player is .
Suppose we have defined on sequences of length less than , where corresponds to the empty position. Formally, to we inductively associate the play where is the response by Player II according to . We need to define on .
Case 1. . In this case
Case 2. .
Let j be least such that . Let
Note that in Case 1, it is trivial that Player II’s move is a legal move. In Case 2, all of the filters played in response to ordinals less than are sub-filters of and hence are legal plays and sub-filters of . Going beyond P, the plays of are extensions of plays according to that have initial segment P. Since P is minimal there are no drops for those plays — in other words, there is inclusion of the normalized responses according to .
From this we conclude that Player II wins the game of length in part (a) of Claim 4.9.
We only prove (b) for because that is the most relevant case for this paper. A straightforward generalization of this argument gives the result for general . The strategy is defined using a winning play by in the game . Since is a winning strategy in that game, if is that play according to , there is a -complete ultrafilter defined on the -algebra generated by . By Lemmas 4.7 and 4.8 and the remarks preceding them, extends for all n. Part (b) follows.
Remark 4.10. Arguing exactly as in Lemma 4.3, if is a sequence of ordinals and a -minimal position P in the game belongs to then the sequence of responses by Player II to using belongs to . In particular, if is a successor ordinal then ’s responses belong to .
Definition 4.11 (The Game 𝒢2). The rules of the game are as follows.
• | Player I plays an increasing sequence of ordinals as before. | ||||||||||||||||
• | Player II plays distinct sets such that the following are satisfied.
| ||||||||||||||||
• | Player I goes first at limit stages. |
A run of of length continues until Player II cannot play or else until it reaches length .
Payoff sets and for the game are defined analogously to those for . So consists of all runs in of length and consists of all those runs such that if is a sequence satisfying and for all then .
Note that in (ii). Note also that since the ultrafilters are required to be uniform, the sets are unbounded in . As with , we will not make any use of what would be an analogue of payoff set .
Proposition 4.12. Assume is an infinite regular cardinal.
(a) | Player II has a winning strategy in of length if and only if Player II has a winning strategy in of length . | ||||
(b) | Player II has a winning strategy in of length if and only if Player II has a winning strategy in of length . |
Proof. For (a), it is immediate that a winning strategy for Player II in gives a winning strategy in : if Player II plays at turn i, then generates a normal ultrafilter on which is Player II’s move in .
For the non-trivial direction, assume Player II has a winning strategy in of length . As noted before Definition 4.11, such a strategy exists in . We build a winning strategy for Player II in of length by induction.
Induction Hypothesis. Suppose that Player I plays in the game , and is the play by Player II using in the game . Then Player II plays where is a definable diagonal intersection of the members of .
For each i, let be the -least enumeration of of length (recall that is the well-ordering of fixed at the beginning of this section; see the paragraphs immediately above Definition 4.1). The induction hypothesis is that for all , Player II’s responses according to the strategy to the sequence are where
This induction hypothesis is automatically preserved at limit stages. Suppose that it holds up to and Player I plays . Then Player II plays an ultrafilter on in the game using the strategy defined in Proposition 4.6. Then, as in Remark 4.10, contains the information that is Player II’s response as well as the -least enumeration of . Let and let be Player II’s response in using .
Suppose now that is a run of the game according to . Then, since is normal each belongs to . Since, for all , for . Moreover, since is a diagonal intersection of the ultrafilter , clause (ii) in Definition 4.11 is immediate.
Since the relevant ultrafilters are the same, whether II is playing by in or in , clause (b) in Proposition 4.12 is immediate. □
Remark 4.13. The definition of the winning strategy for Player II in the previous proof depends on the position P in , beyond which there are no drops. Suppose that Player I plays in the game and Player II responds with using the winning strategy . Then for all with ,
• | because it induces an ultrafilter on , | ||||
• | because and Player II’s response to according to is definable from Player II’s response to according to the strategy for , which in turn is definable from P and Player II’s response according to her strategy in and thus from the original strategy in . |
It follows that for all , and . (Restating this .)
We complete this section with a corollary which will be used in studying properties of the strategies constructed in Sec. 6.
Corollary 4.14. Assume is a winning strategy for Player II in the game of length and is the winning strategy for Player II in the game of length obtained as in Proposition 4.12. Then for every ,
5. Strategies and
Consider a winning strategy for Player II in of length and a position P in according to . Given a set , there may exist different runs of extending P which witness that X is -positive. This causes difficulties in proving that has strong properties like precipitousness or the existence of a dense subset with a high degree of closure. To address this issue, we construct a winning strategy for Player II in of length such that for each position Q in according to and each there is a unique run witnessing that X is -positive, and show that using one can prove the precipitousness of and the existence of a dense subset with a high degree of closure, thus proving Theorems 1.1 and 1.2.
Recall from the introduction that when we talk about saturated ideals over , we always mean uniform -complete and -saturated ideals over . The results in this section are formulated under the assumption of the non-existence of a normal saturated ideal over , as this allows to fit the results together smoothly. That the results actually constitute a proof of Theorem 1.2, which is stated under a seemingly stronger requirement on the non-existence of a saturated ideal over , is a consequence of the following standard proposition.
Proposition 5.1. Given a regular cardinal , the following are equivalent.
(a) | carries a saturated ideal. | ||||
(b) | carries a normal saturated ideal. |
Proof. A standard elementary argument shows that any uniform normal ideal over is -complete, hence (a) follows immediately from (b).
To see that (b) follows from (a), assume is a saturated ideal over . Let be the partial ordering and be a -term for the normal -ultrafilter over derived from the generic embedding associated with where G is -generic. Let be the ideal over defined by
Note that for every and every -generic G
We are now ready to formulate the main technical result of this section.
Proposition 5.2. Assume and there is no normal saturated ideal over . Let be an infinite regular cardinal and be a winning strategy for Player II in of length . Then there is a tree which is a subtree of the partial ordering such that the following hold.
(a) | The height of is and is -closed. | ||||
(b) | If are -incomparable then Y,Y’ are almost disjoint. | ||||
(c) | There is an assignment assigning to each a position in of successor length according to in which the last move by Player II is Y; we denote the last move of Player I in by . The assignment has the following property: | ||||
(d) | If b is a branch of of length , let . Then is a position in according to , and the set of all immediate successors of b in is of cardinality . Moreover the assignment is injective on this set. |
Finally, if then it is possible to construct the tree in such a way that
Clause (d) in the above definition treats both successor and limit cases for . The successor case in (d) simply says that if then the conclusions in (d) apply to the set of all immediate successors of Y in .
Proof. The tree is constructed by induction on levels. Limit stages of this construction are trivial: If is a limit and we have already constructed initial segments of of height for all so that (b)–(d) hold with in place of and end-extends whenever then it is easy to see that is a tree with tree ordering end-extending all , , and such that (b)–(d) hold with in place of . We will thus focus on the successor stages of the construction.
Assume and is constructed at all levels up to level ; our task now is to construct the th level of . Let b be a cofinal branch through this initial segment of , so b is of length . We construct the set of immediate successors of b in , along with the assignment on this set, as follows. As we are assuming there is no normal saturated ideal over , we can pick an antichain in of cardinality . For each there is a position in of successor length according to extending such that the last move by Player II in is almost contained in X. For the sake of definability we can let this position to be -least, where recall that is the fixed well-ordering of .
Now construct the set of all immediate successors of b in recursively as follows. Assume and we have already constructed the set along with the assignment with the desired properties. Since each model is of cardinality , we can pick the -least set which is not an element of any where . Now let Player I extend by playing the least ordinal such that
We show
Now assume are above the same branch b; without loss of generality we may assume and in the above enumeration and . Then we have as in the construction, with and . Also .
If then , thus witnessing . This contradicts the fact that is an antichain in . It follows that . Now for every the set Y is either almost contained in or almost disjoint from Z. As by our choice of in (17), necessarily Y is almost disjoint from . This proves (18).
To verify that (b)–(d) hold with the tree obtained by adding the immediate successors of a single branch b as described in the previous paragraph in place of , note that (c) and (d) immediately follow from the construction just described, so all we need to check is clause (b) and the fact that is still a tree ordering after adding the entire th level. But clause (b) follows from the combination of (18) with the induction hypothesis and the fact that every set on the th level is almost contained in some set on an earlier level. Finally, that adding the th level keeps a tree ordering follows from clause (b). More generally, any collection which satisfies (b) with in place of has the property that the set of all which are -predecessors of a set is linearly ordered under . What now remains is to see that clause (a) holds, but this is immediate once we have completed all steps of the construction.
Finally, given a set , to see that we can construct the tree so that (16) holds, note that we can put A into the first level of at the first step in the inductive construction. This involves a slight modification of the construction of the first level of , and is left to the reader. □
The new strategy for Player II in is now obtained by, roughly speaking, playing down the tree . More precisely:
Definition 5.3. Assume is an infinite regular cardinal, is a winning strategy for Player II in of length , and is a subtree of the partial ordering satisfying (a)–(d) in Proposition 5.2. We define a strategy for Player II in of length associated with recursively as follows.
Assume
As an immediate consequence of the properties of we obtain:
Proposition 5.4. Let be an infinite regular cardinal and assume is as in Proposition 5.2. Then is a winning strategy for Player II in of length .
Moreover if,
Before giving a proof of Theorem 1.1, we record the following obvious fact, which will be useful in Sec. 6 in studying properties of winning strategies for Player II in games of length , and to which we will refer later.
Corollary 5.5. Under the assumptions of Proposition 5.2, assume and is constructed in such a way that (16) holds, that is, . Let be the winning strategy for Player II constructed as in Definition 5.3 using this . Then .
One of the main points of passing to is the following remark.
Remark 5.6. For any position P of a partial run according to of successor length with Y being the last move by Player II, the conditional hopeless ideal is equal to the unconditional hopeless ideal restricted to Y
We now turn to a proof of Theorem 1.1. If there is a normal saturated ideal over then there is nothing to prove. Otherwise Player II has a winning strategy in of length , as follows from Propositions 4.2(b), 4.6(b) and 4.12(b). The conclusion in Theorem 1.1 then follows from a more specific fact we prove, namely, Proposition 5.7 in what follows. In the proof of this proposition, we will make use of the criterion for precipitousness in terms of the ideal game, see Sec. 1.
Proposition 5.7. Assume there is no normal saturated ideal over . Let
• | be a winning strategy for Player II in of length , and | ||||
• | be the winning strategy constructed from as in Definition 5.3. |
Then Player I does not have a winning strategy in the ideal game . Consequently, the ideal is precipitous.
Proof. Assume is a strategy for Player I in the ideal game . We construct a run in according to which is winning for Player II. Odd stages in this run will come from positions in played according to ; more precisely, they will be tail-ends of sets on those positions. So suppose
As the sets constitute an -decreasing chain of nodes in , the positions extend whenever . By Proposition 5.4
We remark that the proof of Proposition 5.7 shows Player II has a winning strategy in the ideal game .
The following proposition gives a proof of Theorem 1.2. Recall that all background we have developed so far was under the assumption that is inaccessible and . Also recall that by trivial observation (TO3) at the beginning of Sec. 4 and results in Sec. 4, if Player II has a winning strategy in of length then Player II has a winning strategy in of length and in of length , as well as in of length whenever is regular. By a similar argument, if Player II has a winning strategy in of length then Player II has a winning strategy in of length . Thus, under the assumptions of Theorem 1.2, the assumptions of Proposition 5.8 below are not vacuous.
Proposition 5.8. Assume there is no normal saturated ideal over and . Let be an uncountable regular cardinal. Assume further that and are strategies as in Proposition 5.7, with in place of .
Then is a -closed dense subset of . It follows that Player I does not have a winning strategy in the ideal game . Consequently, the ideal is precipitous.
Proof. That is a -closed dense subset of follows immediately from the properties of . If , then there is a play of the game such that A is in the ultrafilter determined by some played by Player II using . But then . Since is on , we have shown that for every element of there is an element of the tree below it. Hence, the tree is dense.
To see that is precipitous, we use an argument originally due to Laver. It follows the idea of Proposition 5.7 and shows that Player II has a winning strategy in the game . At stage n of the game suppose that Player I plays . Player II chooses an (so ) and .
Let be such that . Player II’s response to in is . Let . Since is countably complete, . Let , with for all n. Then:
It follows that there is a set such that . Since , is not empty. Hence . □
Proof of Theorem 1.4
Proof. Consider a uniform -complete ideal over such that is -distributive and has a dense -closed set. Because of notational convenience we will work with the partial ordering . (See also the partial ordering used in the proof of Proposition 5.1.) Since is a dense embedding of onto , we can fix a dense -closed set . We work inside for a sufficiently large and will use the fixed well-ordering introduced in Sec. 4 to define a winning strategy for Player II in . As usual, is defined inductively on the length of runs.
So assume
and
6. The Model
In this section, we give a construction of a model where the following holds.
(A) | . | ||||||||||||||||
(B) | U is a normal measure on . | ||||||||||||||||
(C) | is a disjoint sequence of stationary subsets of whenever is inaccessible. | ||||||||||||||||
(D) | Assume is a generic extension via a set-size forcing which preserves , and, in
|
We will informally explain the purpose of the sets before we begin with the construction of the model. These sets are not needed for the construction of ideals in Theorem 1.5, but only for the proof that does not carry a saturated ideal in our model. To understand this proof, it suffices to accept (D) as a black box, that is, it is not necessary to understand how the system of sets is constructed.
Proper class models satisfying (A)–(D) are known to exist, and can be produced via the so-called background certified constructors. The two most used background certified constructions are -constructions and fully background certified constructions. If there is a proper class inner model with a measurable cardinal then any -construction (see for instance [21] for -constructions of models with Mitchell-Steel indexing of extenders, and [25] for -constructions with Jensen’s -indexing) performed inside such a model gives rise to a fine structural proper class model satisfying (A)–(D). We will sketch a proof of this fact below in Proposition 6.1. Similar conclusions are true of fully background certified constructions, but one needs to assume that a measurable cardinal exists in .
There is some similarity in the argument in Proposition 6.1 of the existence of a sequence of mutually disjoint stationary subsets of which behave nicely with respect to the ultrapower by a normal ultrafilter on to a similar claim in [10] where it is proved that one can have such sequence of stationary sets in .
A background certified construction as above gives rise to a model of the form where is such that each either codes an extender in a way made precise, or . Additionally, a model of this kind admits a detailed fine structure theory. There is an entire family of such models, so-called fine structural models; the internal first-order theory of these models is essentially the same, up to the large cardinal axioms. There are models with the properties needed for the construction in this paper that satisfy the statement
There is a Woodin cardinal that is a limit of Woodin cardinals,
as is shown in [18].
We now list some notation, terminology and general facts which will be used for the proof of (C) and (D). Clauses (A) and (B) follow from the construction of the model, and their proofs can be found in [21] or [25]. In fact, each proper initial segment of the model is acceptable in the sense of fine structure theory. We omit the technical definition here and merely say that acceptability is a local form of , and is proved along the way the model is constructed.
From now on assume is a fine structural extender model with indexing of extenders as in [21] or in [25, Chap. 9] (which covers all models discussed above). We often write in place of E to emphasize that E is the extender sequence of W.
FS1 | is the initial segment of W of height with the top predicate, that is, . | ||||
FS2 | If is a cardinal of W then . Thus, in this case and we identify this structure with . | ||||
FS3 | If is a cardinal of W then the structure calculates all cardinals and cofinalities the same way as W. This is a consequence of acceptability. | ||||
FS4 | is the unique such that is a cardinal in but not in . | ||||
FS5 | stands for the first projectum; that is equivalent to saying that there is a surjective partial map which is -definable over with parameters. | ||||
FS6 | (Coherence.) If is a -preserving map in possibly some outer universe of W such that is the critical point of i and then . | ||||
FS7 | (Cores.) Assume is a cardinal in W and N is a structure such that and there is a -preserving map of N into a level of W such that . Let be the -least finite set of ordinals p such that there is a set of ordinals a which is -definable in the parameter p and satisfies . Here is the usual well-ordering of finite sets of ordinals, that is, finite sets of ordinals are viewed as descending sequences and is the lexicographical ordering of these sequences. Let X be the -hull of and be the inverse of the collapsing isomorphism. Then , the models agree on what is, and is -preserving and maps cofinally into N. In this situation, is called the core of N and is called the core map. | ||||
FS8 | (Condensation lemma.) Assume is a cardinal in W and and are as in FS7. Then is a level of W, that is, for some . |
Proposition 6.1. There is a formula in the language of extender models such that the following holds. If is a fine structural extender model, is an inaccessible cardinal of W and , letting
Proof. Since the definition of takes place inside , any two extender models such that and calculate this sequence the same way (here stands for the common value of the cardinal successor of in both models). Now if and j is as in (D) above then
It remains to come up with a formula such that the sets are stationary in W for all of interest, and pairwise disjoint. Here we make a more substantial use of the fine structure theory of W. Given an inaccessible and a , letting
it is clear that whenever . Then it suffices to show that
The first step toward the proof of (23) is the following observation.
Proof. Obviously, . Assume for a contradiction that . Let be an increasing sequence converging to such that for every . For each such i pick a and an ordinal such that where is the standard -Skolem function for . Here is of the form (see FS2), and we identify it with the structure . The Skolem function has a -definition of the form where is a -formula in the language of extender models. (The standard -Skolem function has a uniform -definition, which means that there is a -formula which defines a -Skolem function over every acceptable structure N. However, the argument below does not make use of uniformity of the definition.) Since is regular
Now let C be a club subset of , X be the -hull of in , N be the transitive collapse of X, and be the inverse of the collapsing isomorphism. Let further . Then as . It is a standard fact that (and can be proved similarly as (24) above). Now by (24), hence . Moreover as C is closed and is a limit point of C. Thus, the proof of (22) will be complete once we show that and has cardinals above . We first look at the set of cardinals in N.
By acceptability, the structures and agree on what is a cardinal below . It follows that in , the statement
can be expressed in a -way as
6.1. The tools
Two main tools we will use to construct the forcing used to build our model are club shooting with initial segments, and adding non-reflecting stationary sets with initial segments. We then use variations of standard techniques for building ideals using elementary embeddings. The background information on the first two can be found in [4, 5, 6] and on ideal constructions in [9], but we review the relevant facts for the reader’s convenience. When discussing the successor of a regular cardinal we will often assume even when it is known that suffices. Since the models we work in satisfy the this is not important for our results.
Recall that if is a stationary set (where is a cardinal) then the club shooting partial ordering consists of closed bounded subsets of which are contained in S, and is ordered by end-extension. In general, this partial ordering may not have good preservation properties, but if S is sufficiently large then it is known to be highly distributive. The following is standard.
Proposition 6.2 (See [3, 5, 6]). Assume is regular, and T is a subset of such that is non-stationary in whenever , and is stationary. Then the following hold.
(a) | is -distributive, that is, it does not add any new function . In particular, generic extensions of via agree with on all cardinals and cofinalities , and on what is. | ||||
(b) | If is regular and then has a dense set which is -closed but if T is stationary then it does not have a dense set which is -closed. | ||||
(c) | If G is -generic then is a closed unbounded subset of such that . |
To show that there is no saturated ideal in the model of Theorem 1.5 and Corollary 1.6 we will need to see that the forcing for shooting a closed unbounded set through the complement of a non-reflecting stationary set A preserves stationary sets disjoint from A. This is the content of the following proposition that appears in [3, 5, 6]. We give the proof here for the reader’s convenience.
Lemma 6.3. Assume is an uncountable cardinal with , are disjoint stationary subsets of and that for all is non-stationary. If is generic, then remains stationary in .
Proof. Let force that is a closed unbounded subset of with . Let be a regular cardinal and let be an internally approachable sequence of elementary substructures of . Then is a limit is a closed unbounded subset of and for each such .
Choose a limit such that . Let and be a closed unbounded set of order type . Build a decreasing sequence of conditions such that
• |
| ||||
• | for each , | ||||
• | if i is the member of , then for some ordinal , with | ||||
• | . |
Such a sequence is possible to build, because .
But then and , hence
Our application of the following definition and the following lemmas will be with for a regular .
Definition 6.4. Let be a regular cardinal.
(a) | The partial ordering for adding a non-reflecting stationary subset of consists of functions for some and letting | ||||
(b) | Let be regular. The partial ordering for adding a non-reflecting stationary subset of consists of those conditions which concentrate on : |
Let be uncountable regular cardinals and define the map
We will use the following lemma which relates with .
Lemma 6.5. Let be uncountable regular cardinals.
(a) | If G is generic for , then | ||||
(b) | If H is generic for | ||||
(c) | If is generic over V, and , then H is generic over V for . (In other words the map is a projection.) |
Proof. The first two items are immediate. For the third note that for all and all with , there is a with . This is the standard criterion for being a projection. □
It is an easy remark that in (a) and in (b) . For this reason, we will frequently write for the extension, when it is clear from context whether we are in case (a) or (b).
We will make use of these partial orderings in the special case where is of the form . For this reason we formulate the next proposition for cardinals of the form , although it is true for any regular .
Proposition 6.6 (See [4, 5, 6]). Assume where is regular and . Then the following hold.
(a) | Both and are strategically -closed. In particular, both and preserve stationarity of stationary subsets of , are -distributive, so they do not add any new functions , and generic extensions of via these partial orderings agree with on all cardinals and cofinalities and on what is. | ||||
(b) | is -closed but not -closed. | ||||
(c) | If G is -generic then is a non-reflecting stationary subset of . | ||||
(d) | If G is -generic then is a non-reflecting stationary subset of such that has stationary intersection with each stationary subset of that lies in V. In particular, is stationary for all regular . |
Proof. Only (d) is not explicitly proved in the earlier literature (though it was known). Let T be a stationary subset of in V. Let be generic and be the generic stationary set added by G. We claim that S has stationary intersection with T. We assume without loss of generality that every ordinal in T has the same cofinality .
If the claim fails let force over V that where is a term for a closed unbounded subset of in .
Let be a regular cardinal and let be an internally approachable sequence of elementary substructures of
By recursion on build a decreasing sequence of conditions in such that
• |
| ||||
• | for each , | ||||
• | if i is the member of , then for some ordinal , with | ||||
• | . | ||||
• | If is a limit ordinal, then and if , then forces . |
Let . Then has supremum and forces that
• | is non-stationary | ||||
• | is cofinal in |
Extending by one point to get a condition q that forces gives a condition that forces . This contradiction shows that in the extension by , S intersects every stationary T. □
Although both partial orderings and have a low degree of closure in general, the iteration that generically adds a non-reflecting stationary set S followed by adding a closed unbounded subset of the complement of S does have a high degree of closure.
Proposition 6.7. Assume is a cardinal, is regular, and is the canonical -term for the generic non-reflecting stationary subset of . Then the composition
Proof. Let D be the collection of all such that
• | is closed (so has successor order type), and | ||||
• | for some closed unbounded set with for all . |
Then D is dense in . For details, see [4, 5] or [6]. □
Fix a regular cardinal . At successor steps in the iteration used to prove Theorem 1.5, we will use an iteration of the form
Version 1:
Version 2:
Version 3: .
Lemma 6.8. Let . Then has a dense set D such that
(i) | D has cardinality , | ||||
(ii) | , | ||||
(iii) | D is -closed, and | ||||
(iv) | D is -distributive. |
Proof. Proposition 6.7 shows that has a dense -closed subset. Since consists of ordinals of cofinality , is -closed and -distributive. Since is isomorphic to , items (iii) and (iv) follow. Now (i) is immediate, since has a dense set of size after forcing with the first two partial orderings.
To see (ii), use Version 3 of the partial ordering . The first step is clearly a subset of . By Proposition 6.7 there is a dense subset of the first two steps that lies in and is -closed. After forcing with the conditions in belong to and can be realized by elements of V using the closure of . Hence, there is a dense subset of Version 3 consisting of triples where each coordinate belongs to . Rearranging, we get (ii). □
In the iteration, we will construct as a coding tool. Let be a sequence of disjoint stationary subsets of . Let and define
Given an -generic , and a sequence of sets for each regular uncountable with , the following holds:
Proposition 6.9. Suppose is regular and the holds. Let be the partial ordering
(a) | If , then is non-stationary and is stationary. | ||||
(b) | If , then is non-stationary, and is stationary. |
Proof. Force with to get a generic stationary set S and let be a term for the closed unbounded set added by . If is -generic then is empty which shows the non-stationarity claims in both (a) and (b).
What is left is to show that the appropriate ’s stationarity is preserved. The argument in each case is the same, so assume we argue for case (a). Since the partial ordering has a dense -closed subset, it preserves the stationarity of each .
Suppose that . Applying Lemma 6.3 in with and shows that is stationary in . □
Essentially, the same proof shows:
Proposition 6.10. Suppose is regular, is regular and uncountable, and that the holds. Let be the partial letting
(a) | If , then is non-stationary and is stationary. | ||||
(b) | If , then is non-stationary, and is stationary. | ||||
(c) | If , then is stationary. |
The point of this coding is that using the forcing in either Proposition 6.9 or 6.10, for of the appropriate cofinality we have
Proposition 6.11. Under the hypotheses of Proposition 6.9 (or Proposition 6.10), the set S added by (respectively, remains stationary after forcing with (respectively, ).
Proof. We prove it with the hypotheses of Proposition 6.9, the proof using the hypotheses of Proposition 6.10 is essentially the same.
Let be generic and be the generic stationary set constructed by G. By Proposition 6.6 item (d), in , S has stationary intersection with each . Choose a such that . Let and . The and satisfy the hypotheses of Lemma 6.3 for the forcing . Hence, is stationary in the generic extension of by , and so S is stationary after the forcing . □
6.2. The construction
Let U be the normal measure as in (B) above and
The forcing will be an iteration of length with Easton supports. If is inaccessible we will choose a regular uncountable and do a three step forcing. First we add a non-reflecting stationary set S. We then force to code the non-reflecting stationary set using the stationary sets . The last step is to shoot a club through the complement of the stationary set S created in the first step.
At stage we do the analogous forcing except that we only use the first two steps.
Description of the Forcing. We now formally define the partial orderings used in the construction. For an inaccessible cardinal fix the stationary sets from Proposition 6.1. Fix a regular uncountable . For this , let be the partial ordering
The final partial ordering will be an iteration with Easton supports of length . We define the initial segment of length , , as follows. will be the direct limit of the forcing iteration
FI-1 | For inaccessible , conditions in each are partial functions p with contained in inaccessibles below such that is bounded in whenever is inaccessible. | ||||
FI-2 | If and then and is a -term for a condition in the three step forcing defined in Eq. (33).f |
The ordering on is defined in the standard way, that is,
FI-3 | if and only if the following hold :
|
From Lemmas 6.6 to 6.8, we conclude that
(i) | For all inaccessible | ||||
(ii) | For Mahlo, is -c.c. | ||||
(iii) | If G is -generic then in the partial ordering contains a dense -closed set and is -distributive. | ||||
(iv) | For , if is the canonical factorization, and G is -generic, then | ||||
(v) | For each inaccessible , if then . | ||||
(vi) | For all cardinals , preserves both and . | ||||
(vii) | preserves all cardinals. |
Now define a partial ordering as the length iteration:
We claim that any generic extension via produces a model as in Theorem 1.5. We will first focus on the proof of the following proposition.
Proposition 6.12. In any generic extension via all cardinals and cofinalities are preserved, remains inaccessible, and for each regular uncountable there is a uniform normal -distributive ideal such that has a dense -closed set, but no dense -closed set.
Proof. Fix a regular uncountable cardinal .
By in , any generic extension via satisfies , so in any such generic extension the partial ordering has cardinality . Using the strategic closure of we conclude that in the generic extension via . Let S be the non-reflecting stationary set added by . Then has cardinality in any such generic extension. All of this combined with the distributivity properties of and , shows that . Similar arguments show that
Now return to the map j from (32). Let G be -generic. Because and is -c.c., is closed under -sequences in and the models agree on what is. It follows that the models agree on what and are.
Let be -generic where is as in Eq. (34). It follows that .
Let where is as in Lemma 6.5. Then is generic for over both and . Let be the term for the non-reflecting stationary set coming from . Then . Denote by .
Since and are -distributive in the models where they live
Let be the closed unbounded subset of associated with the generic ultrafilter for over .
Note that and because . From the point of view of there are only many dense subsets of which are in .
We can construct a -generic filter as follows. In fix an enumeration of dense subsets of which belong to . Using recursion on construct a descending chain in as follows:
• | Let . | ||||
• | Given , pick such that in . | ||||
• | Given , let where is the least element of C larger than . | ||||
• | If is a limit let where . |
To see that this works, note that for every both and are elements of , which is verified inductively on . The only non-trivial step in the induction is to see that for limit. That the sequence belongs to follows from the -closure property of . For the union to be a condition requires that the supremum of does not belong to . However by Eq. (31), since we know that .
Now let be the filter on generated by the sequence ; it is clear that and is -generic. Finally, set .
We note here that by Proposition 6.11, is stationary in . Thus in , is a non-reflecting stationary set.
Consider a -generic filter H. Then the filter is -generic. It follows that is -generic. By the Factor Lemma applied inside , the quotient is isomorphic to the iteration as calculated in . Let be the least inaccessible of M above . Using (iv) in the list of the properties of the iteration stated below FI-3, we conclude that satisfies the following:
defined by setting whenever is a -term. Since can be constructed inside , there is a -term such that is -generic whenever H is -generic. In particular, there is a M-generic determined by forcing over to get a generic .
Changing notation slightly to emphasize the dependence on H, define be as follows:
We also have a -term such that is the normal -measure over derived from . That is,
We now define the ideal on in . For every ,
Note that this definition takes place in so and standard arguments show that is a uniform normal ideal on in .
Recall that where was fixed at the in . This is crucial for determining the closure properties of . The main tool for analyzing properties of is the duality theory developed in [9]. Rather than simply citing theorems there, we show the following proposition:
Proposition 6.13. In there is a dense
Proof. In , fix an assignment where and is such that
Next, recall that at each inaccessible , stages and of are a composition of three partial orderings where the last one is . The components of the generic filter G are then of the form where is -generic. The function h is thus an element of and represents the filter H in the ultrapower by , that is, ; see (41).
Then for any we have the following:
To see that e is order-preserving, consider in . By the above remarks on the ordering of the quotient, we have in , hence in . It follows that
To see that the map e is incompatibility preserving, we prove the contrapositive. Assume are such that . It follows that there is some -generic filter H such that . Then and . Using (45) we conclude that . Hence are compatible.
To see that the range of e is dense, assume that . It follows that there is some -generic filter H such that . So there is some such that
We can now complete the proof of Proposition 6.12 by looking at the properties of the partial ordering in . By Proposition 6.11, is stationary in , so is a standard forcing for killing a non-reflecting stationary subset of . The -distributivity follows from Proposition 6.2(a). The existence of a dense -closed set as well as the non-existence of a dense -closed set follows from Proposition 6.2(b) and the fact that .
The last major step toward the proof of Theorem 1.5 is the following proposition.
Proposition 6.14. does not carry a saturated ideal in a generic extension via .
Proof. Assume for a contradiction that does carry a saturated ideal in where are as above. Denote this ideal by , and let L be a -generic filter where is the partial ordering and
Now look at the th step of the iteration . Obviously and . Let be the ordinal chosen by the generic filter K at step of the iteration (see FI-2). Then steps and are thus forcing with
The th component of K has the form . Let be the generic non-reflecting stationary subset of added by over . Since is a closed unbounded subset of disjoint from , the set is non-stationary in .
By elementarity, the generic filter codes the set inside as follows. Given an ordinal ,
Finally, we give a proof of incompatibility of strategies from Corollary 1.6(a), as formulated at the end of Corollary 1.6.
The point here is that in the construction of , the ordinal at the th stage in is chosen before the generic filter H comes into play. Therefore, the set is defined by
Remark 6.15. We could do the construction without the “lottery” aspect, aiming at a single . Indeed that works for that , but leaves open the problem of whether ideals exist with dense trees of height for and for which strategies exist in the Welch game. These questions are thorny and are left to the second part of this paper. The solutions there use extensive fine structural arguments.
7. Open Problems
In this section, we raise questions we don’t know the answer to. We do not guarantee any of these questions are deep, difficult or even make sense.
Open Problem 1. Removing Hypotheses: Theorem 1.2 requires the GCH and the non-existence of saturated ideals on . Are either of these hypotheses necessary? Can some variant of the proof work without those hypotheses?
Open Problem 2. What can be said about correspondence between ideals and strategies? Theorem 1.4 says that starting with a nice ideal one can build a winning strategy for Player II in . In turn, can used to build the ideal with the methods in Theorems 1.1 and 1.2
An Ulam Game. Consider the following variant of the cut-and-choose game of length derived from games introduced by Ulam in [23] (see [15]).g
I | ( | … | … | |||
II | … | … |
At stage 0, Player I plays a partition of . At stage , Player II lets be either or , and plays . At stage , Player I plays a partition of . The winning condition for Player II is that .
These games generalize to lengths as follows:
(1) | At successor stages , Player I partitions into two pieces and Player II chooses one of the pieces. | ||||
(2) | At limit stages , let and then Player I partitions into two pieces, and Player II chooses one of the pieces. | ||||
(3) | The winning condition is the same: the intersection of the pieces that player II chooses has to have at least two elements. |
Observation. If Player II has a winning strategy in the game , then Player II has a winning strategy in the Ulam game.
This is immediate: Player II follows her strategy in an auxiliary play of the game against the Boolean Algebras generated by . In the game , she then plays as whichever of or belongs to . By the winning condition on , belongs to a -complete, uniform filter. Hence .
Silver and Solovay (see [15, p. 249]) showed that if Player II wins the Ulam game, then there is an inner model with a measurable cardinal. This provides an alternate proof that the consistency strength of the statement “Player II has a winning strategy in ” is that of a measurable cardinal.
What is unclear is the exact relationship between the Ulam Game and the Welch Game. Laver showed that if a measurable cardinal is collapsed to by the Lévy collapse and is the ideal generated by the original normal measure on , then in the extension has a dense countably closed subset ([9]). He showed that it follows from this that Player II has a winning strategy in the Ulam game.
In Sec. 2, it is shown that the Welch games only make sense at regular cardinals such that for all , . At successor cardinals there is a single play by Player I (the algebra in part (2) of Theorem 2.3) that defeats Player II in the game of length 1. Moreover at non-weakly compact inaccessible cardinals , the Keisler–Tarski Theorem shows player I has a winning strategy in the game of length 1. But if is weakly compact, Player II has a winning strategy in the game of length .
The upshot of this discussion is that a comparison between the Ulam games and the Welch games should occur at weakly compact cardinals.
Open Problem 3. Suppose that is weakly compact and that Player II has a winning strategy in the Ulam game of length (for ), does Player II have a winning strategy in ?
Determinacy of the Welch Games. The discussion in the paragraphs before Problem 3 (based on Sec. 2 of this paper) shows that questions about the determinacy of Welch Games really only make sense at inaccessible cardinals. Moreover at non-weakly-compact inaccessible cardinals Player I wins the game of length 1 and at weakly compact cardinals Player II wins the game of length . By work of Nielsen and Welch if II has a winning strategy in the game of length , then there is an inner model with a measurable cardinal — so Player II can’t have such a winning strategy in L. (Theorem 1.1 in this paper also gives this result.) Welch showed that for all regular , is determined in L (this also follows immediately from [13, Theorem 5.6]).
However the following seems to be an open problem:
Open Problem 4. Is there a model of with a measurable cardinal where the Welch games are determined? With a supercompact cardinal?
Welch Games on Larger cardinals. In this paper, the Welch games are shown to provide intermediary properties between weakly compact cardinals and measurable cardinals. What is the analogue for cardinals that are at least measurable? Perhaps the most interesting question is the following:
Open Problem 5. Are there versions of the game?
It is not trivial to even formulate a reasonable analogue of Welch games on supercompact cardinals. The classical ultrafilter extension properties on that follow from large cardinals suggest one, but it is not clear how to proceed.
Another technical obstacle that would have to be overcome is the following: in the proofs in this paper one passes from a -filter U on an to its normal derivative . Normality presents an obstacle for because this is the crucial difference between supercompact and strongly compact cardinals.
In [2], Buhagiar and Dzamonja found analogies of strongly compact cardinals that Dzamonja suggested might be candidates for this game.
Extender Algebras. Large cardinals whose embeddings are determined by Extender Algebras also form candidates for places games like this can be played. If E is an extender with generators one might consider games where Player I plays elements of and sequences of -algebras in a coherent way, and player II plays ultrafilters on the associated algebras.
In this manner, one might hope to extend these results to or further.
Games on accessible cardinals:
Open Problem 6. Are there small cardinal versions of these games?
The results in Sec. 2 limit the Welch games to inaccessible cardinals. However one might hope that there is some version of these games that end up creating ideals on cardinals that are not weakly compact. A random suggestion is to require Player II to play ideals with some combinatorial property at each stage (rather than ultrafilters). One target would be to define a game similar to the Welch games that gives -closed densely treed ideals on (the original Laver ideals).
Acknowledgments
Matthew Foreman was supported in part by NSF grant DMS-2100367. Menachem Magidor was supported by ISF grant 684-17.
Publisher’s Note
Due to an error during production, there was a mistake in the online published version of this article: on page 21, line 13, “By the minimality of P, if extends P then ” should be replaced with “By the minimality of P, if extends P then ”. This mistake has been corrected as of December 11, 2023.
Notes
a We could omit “uniform” and simply require to include the co- subsets of and to extend the co--filter. As noted in Sec. 2, if is inaccessible, then Player I always has a legal play in the Welch game.
b There are two general techniques used in this paper for building ideals. One is the conventional method of starting with a large cardinal embedding and extending it generically. We use the notation for these. The second is the new technique of hopeless ideals, built in Theorems 1.1 and 1.2 from the strategies . These will be denoted by or very similar notation.
c By a canonical square sequence we mean a square sequence obtained by a slight variation of Jensen’s fine structural construction, generalized to extender models. This is made precise in Part II of this paper.
d We would like to thank James Cummings for giving significant help in understanding this case.
e We use the notation for level of .
f We can view as a triple but the notation is frequently more convenient.
g Velickovic [24] calls these Mycielski games.