Processing math: 10%
World Scientific
Skip main navigation

Cookies Notification

We use cookies on this site to enhance your user experience. By continuing to browse the site, you consent to the use of our cookies. Learn More
×

Games with filters I

    https://doi.org/10.1142/S021906132450003XCited by:0 (Source: Crossref)

    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<γκ 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 κκ-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 RVκ there is a transitive set X and SX such that

    Vκ,,RX,,S.

    (2)

    If P(κ) is a κ-complete Boolean subalgebra with ||=κ and F is a κ-complete filter on , then F can be extended to a κ-complete ultrafilter on .

    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 :

    I𝒜0𝒜1𝒜α𝒜α+1
    IIU0U1UαUα+1

    The game proceeds for some length γ determined by the play. The sequence 𝒜δ:0δ<γ is an increasing sequence of κ-complete subalgebras of P(κ) of cardinality κ and Uδ:0δ< is sequence of uniform κ-complete filters, each Uα is a uniform ultrafilter on 𝒜α and α<α implies that UαUα. 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 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: γω 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=U𝒜α. 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 on a set X is precipitous if for all generic GP(X)/ 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)/. We begin by discussing a strong hypothesis

    κ-complete, uniform ideal  on κ such that the Boolean algebra P(κ)/ has the κ+-chain condition is called a saturated ideal.

    It follows from results of Solovay in [20] that if is a saturated ideal on κ then 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 γω+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)/. 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  on κ and a set D+ such that

    (1)

    (D,) is a downward growing tree of height γ,

    (2)

    D is closed under -decreasing sequences of length less than γ,

    (3)

    D is dense in P(κ)/.

    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 .

    Definition 1.3. Let be a κ-complete ideal on P(κ) and γ>ω be a regular cardinal. Then is γ-densely treed if there is a set D+ such that

    (1)

    (D,) is a downward growing tree,

    (2)

    D is closed under -decreasing sequences of length less than γ,

    (3)

    D is dense in P(κ)/.

    Note that this is weaker than the conclusions of Theorem 1.2.

    We will abuse notation slightly and say “D is dense in +” to mean that D is a dense subset of P(κ)/.

    We will say that an ideal is (κ,)-distributive if P(κ)/ 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 γκ 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 . Players I and II alternate playing

    IA0A1AnAn+1
    IIB0B1BnBn+1

    With AnBnAn+1 and each An,Bn+. Player II wins the game if nBn. We will call this game the Ideal Game for . They proved the following theorem.

    Theorem 2 ([11]). Let  be a countably complete ideal on a set X. Then  is precipitous if and only if Player I does not have a winning strategy in the ideal game for .

    In the proof of Theorem 1.1, we construct an ideal and show that Player II has a winning strategy in the ideal game for . 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<γκ 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 γ 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 γγ in the following sense: If γ,γκ are regular and γγ 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<γκ there is a uniform, normal γ-densely treed ideal 𝒥γ that is (κ+,)-distributive. The relationship between the ideals and strategies for different γs is as follows:

    (a*)

    There does not exist a uniform normal ideal 𝒥𝒥γ 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 γ be the ideal arising from the strategy 𝒮γ, there does not exist an ideal γ which is γ-densely treed as witnessed by a tree D+ of height γ, for any γ>γ.b

    In other words, the ideals in (c*) in Theorem 1.8 are like ideals 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)

    Sαα+cof(<α),

    (iii)

    Sαcξ= for all ξ

    whenever α is a cardinal.

    Let δ 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)

    δ is δ+-distributive.

    (b)

    If G is generic for ϱ over V and j:VM is an elementary embedding in some generic extension V of V which preserves κ+ then j(ϱ)/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 a κ-complete ideal on κ. We say that AB if A\B, and is the converse relation. The notations *, * are these notions when is the ideal of bounded subsets of κ. The notation A abbreviates the conjunction of AB and AB, where △ means symmetric difference.

    The ideal induces an equivalence relation on P(κ) by [A]=[B] if and only if AB. The notion induces a partial ordering on P(κ)/, we will sometimes call this and refer to the set of equivalence classes of P(κ) that don’t contain the empty set as +. We will force with (P(κ)/,*) 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. κ-complete sub-Boolean algebras of P(κ) that have cardinality κ are called  κ -algebras.

    If σ and τ are sequences we will use στ to mean the concatenation of σ and τ. We will abuse this slightly when τ has length one. For example, given σ=αi:i<β and δ we will write αi:i<βδ for the sequence of length β+1 whose first β elements coincide with σ and whose last element is δ.

    Usually our trees grow downwards, with longer branches extending shorter branches. A tree 𝒯 is γ-closed if when b is a branch through 𝒯 whose length has cofinality less than γ there is a node σ𝒯 such that σ is below each element of b. Occasionally, we will say <γ-closed to mean γ-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 κ is inaccessible and 𝒜 is a κ-algebra and B[κ]κ then 𝒜B generates a κ-complete subalgebra of P(κ) that has cardinality κ (i.e. another κ-algebra). The situation where κ is not inaccessible is quite different.

    Proposition 2.1. Suppose that κ is an infinite cardinal and either

    a singular strong limit cardinald or

    for some γ<κ, 2γ>κ but for all γ<γ, 2γ<κ.

    Then there is no Boolean subalgebra 𝒜P(κ) such that |𝒜|=κ, 𝒜 is κ-complete.

    Proof. In the first case, since κ is singular, if 𝒜 is κ-complete, it is κ+-complete. For δ<κ, let aδ={A𝒜:δA}. Then aδ is an atom of 𝒜 and each non-empty A𝒜 contains some aδ. Moreover distinct aδ’s are disjoint. Thus, the {aδ:δκ} generate 𝒜 as a κ-algebra. If there are κ many distinct aδ then |𝒜|=2κ. Otherwise, since κ is a strong limit, |𝒜|<κ.

    Assume now that γ<κ, 2γ>κ and for all γ<γ,2γ<κ. Since |𝒜|=κ, 𝒜 must have fewer than γ atoms. If aδ:δ<γ is the collection of these atoms, the κ-algebra generated by the atoms of 𝒜 has cardinality at most 2γ. Let B=δ<γaδ and C=κB. Since 2γ<κ and 𝒜 has cardinality κ, there is an a𝒜 that does not belong to . Since a is not in the set a=aδaδ. Hence, C is non-empty. Replacing 𝒜 with {AB:A𝒜} we get an atomless, κ-complete algebra on the set C.

    Since no element of 𝒜 is an atom we can write each A𝒜 as a disjoint union of non-empty elements A0,A1 of 𝒜. Build a binary splitting tree 𝒯 of elements of 𝒜 of height γ by induction on δ<γ as follows:e

    (𝒯)0 is the -maximal element C of 𝒜.

    Suppose (𝒯)δ is built. For each A𝒯δ, write A as the disjoint union A0A1 such that Ai𝒜, and Ai and let (𝒯)δ+1={A0,A1:A(𝒯)δ}.

    Suppose that δ is a limit ordinal. Let (𝒯)δ={b:b is a branch through (𝒯)<δ and b}.

    Note that for all δ<γ, each element cC determines a unique path through (𝒯)δ of length δ. Hence (𝒯)δ=C.

    Fix a cC and let b=Aδ:δ<γ be the branch through 𝒯 determined by c. For δ<γ, the tree splits Aδ=Aδ+1Aδ+1 with Aδ+1,Aδ+1(𝒯)δ+1. The sets Aδ+1:δ<γ each belong to 𝒜 and form a collection of disjoint subsets of C of size γ. By taking unions of these sets we see that |𝒜|2γ>κ. □

    In contrast to Proposition 2.1, we have

    Proposition 2.2. Let κ be infinite and 2γ=κ. Then P(γ)P(κ) and P(γ) is κ-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 κ is inaccessible.

    Theorem 2.3. Suppose that κ is an accessible infinite cardinal. Then either

    (1)

    There is no κ-algebra 𝒜P(κ) with |𝒜|=κ or

    (2)

    There is a κ-algebra 𝒜P(κ) with |𝒜|=κ but every κ-complete ultrafilter U on 𝒜 is principal

    Corollary 2.4. Consider the Welch game of length 1. Suppose that there is a κ-algebra 𝒜0 that is a legal move for Player I and that Player II has a winning strategy in 𝒢1W. Then κ is inaccessible.

    If κ 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 κ is inaccessible and Player II wins the Welch game of length 1. Then κ is weakly compact.

    Proof. To show κ is weakly compact, it suffices to show it has the tree property. Let 𝒯 be a κ-tree. For α<κ, let Aα be the set of β<κ such that α𝒯β. Let 𝒜 be the κ-algebra generated by {Aα:α<κ} and 𝒰 be a uniform κ-complete ultrafilter on 𝒜.

    For each γ<κ, κ=α(𝒯)γAαR where |R|<κ. It follows that for each γ there is an α(𝒯)γ such that Aα𝒰. But then {α:Aα𝒰} is a κ-branch through 𝒯. □

    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 κ. Assume F is a function with domain R such that for every rR the value F(r) is a sequence of length ξr of the form

    F(r)=𝒜ir,Uir|i<ξr(1)
    where for every i<ξr,

    (i)

    𝒜irP(κ) and

    (ii)

    Uir is a κ-complete ultrafilter on the κ-algebra of subsets of κ generated by ji𝒜jr.

    (iii)

    For all rR, the sequence Uir|i<ξr is monotonic with respect to the inclusion.

    (iv)

    (Density) For every rR, j<ξr and [P(κ)]κ there is sR such that F(r)j=F(s)j and 𝒜js.

    We will call functions F with the properties (i)–(iv) assignments.

    One can also formulate a variant with normal ultrafilters Uir. Denote the maximo-lexicographical ordering of κ×κ by <mlex. Let h:(κ×κ,<mlex)(κ,) be the natural isomorphism. For a set Aκ, let Ai=(h1[A])i be the ith section of h1(A). The sequence Ai|i<κ is associated to A. We will say that a κ-algebra 𝒜 of subsets of κ is normal if for all A𝒜, each Ai belongs 𝒜 and the diagonal intersection Δi<κAi also belongs to 𝒜. We will say that a sequence Ai|i<κ belongs to 𝒜 if it is associated to an element of 𝒜. Finally we say that an ultrafilter U on a normal κ-algebra 𝒜 is normal if and only if for every sequence Ai|i<κ𝒜,

    (i<κ)(AiU)Δi<κAiU.(2)
    A variant of this definition is an assignment with normal ultrafilters where we require, instead of (ii) above, that

    (ii)

    Uir is a κ-complete normal ultrafilter on the normal κ-algebra of subsets of κ generated by ji𝒜jr.

    If (ii) is satisfied we say that F is normal. Note that there is no need to modify clause (iv), as normal κ-algebras are able to decode κ-sequences of subsets of κ from other subsets of κ via the pairing function h introduced above. However, instead of families [P(κ)]κ, it is convenient in (iv) to consider sets BP(κ) that code .

    Definition 3.1. Given an assignment F, we define the ideal (F) as follows.

    (F)=the set of allAκsuch thatAUirfor anyi<ξrand anyrR.(3)
    The ideal (F) is called the hopeless ideal on P(κ) induced by F.

    Although in the above definition we say we are defining an ideal, an argument is needed to see that (F) is indeed an ideal. It follows immediately that (F) and (F) is downward closed under inclusion. The rest is given by the following proposition.

    Proposition 3.2. Given an assignment F, the ideal (F) is κ-complete. If all ultrafilters Uir are uniform then (F) is uniform. If additionally F is normal then (F) is normal. If F’ is an assignment on RR and FR=F, then (F)(F).

    Proof. We first verify κ-completeness of (F). We noted above that (F) and (F) is downward closed under inclusion; hence it suffices to check that (F) is closed under unions of cardinality <κ. If Aη|η<ξ is such that ξ<κ and A=η<ξAη(F), then there is some rR and some i<ξr such that AUir. By the density condition, there is some sR such that 𝒜js=𝒜jr and Ujs=Ujr for all ji, and {Aη|η<ξ}𝒜i+1s. In particular, AUisUi+1s and all sets Aη, ηξ are in the κ-algebra generated by ji+1𝒜j. By κ-completeness of Ui+1s then AηUi+1s for some η<ξ, hence Aη(F).

    The proof of normality of (F) for normal F is the same, with η<κAη in place of η<ξAη. The conclusion on uniformity of (F) follows by a straightforward argument from the definition of (F).

    Finally, if R,F are as in the statement of the proposition, then any A(F) trivially avoids all ultrafilters Uir where rR and i<ξr, so A(F). □

    Now assume 𝒢 is a two player game of perfect information, and 𝒮 is a strategy for Player II in 𝒢. Denote the set of all runs of 𝒢 according to 𝒮 by R𝒮 (by a run we mean a complete play). Assume every rR𝒮 is associated with a sequence of fragments 𝒜irP(κ) and ultrafilters Uir, in a way that makes the function

    F𝒮:r𝒜ir,Uir|i<lh(r)(4)
    an assignment/normal assignment with domain R𝒮. Here of course ξr=lh(r) when compared with (1). In all concrete situations we will consider, the rules of the game 𝒢 will guarantee that the function F𝒮 is really an assignment. As the strategy 𝒮 makes it clear which game is played, we suppress writing 𝒢 explicitly in our notation.

    Here are some examples. If 𝒢 is the Welch game 𝒢γW then F𝒮 is the identity function. In the following section, we introduce games 𝒢1,𝒢1 and 𝒢2. These games are defined relative to a sequence of models Nα|α<κ+ increasing with respect to the inclusion, and Player I plays ordinals α<κ+ which refer to these models. In the games 𝒢1 and 𝒢1 Player II plays uniform κ-complete ultrafilters on P(κ)Nα; in 𝒢1 these ultrafilters are required to be normal. Thus, if r is a run in one of these games according to 𝒮, say r=αir,Uir|i<lh(r) then

    F𝒮(r)=P(κ)Nαir,Uir|i<lh(r).
    In the game 𝒢2 Player II plays sets Yκ which determine uniform normal κ-complete ultrafilters U on P(κ)Nα defined by U={XP(κ)Nα|Y*X}. Thus, if r is a run in the game 𝒢2 according to 𝒮, say r=αir,Yir|i<lh(r) then
    F𝒮(r)=P(κ)Nαir,{XP(κ)Nαir|Yir*X}|i<lh(r).

    If P is a position in 𝒢 played according to 𝒮 we let

    R𝒮,P=the set of allrR𝒮extendingP(5)
    and
    F𝒮,P=F𝒮R𝒮,P.(6)
    We are now ready to define the central object of our interest.

    Definition 3.3. Assume 𝒢 is a game of perfect information played by two players, 𝒮 is a strategy for Player II in 𝒢, and F𝒮 is an assignment with domain R𝒮 as in (4). Consider a position P in 𝒢 according to 𝒮. We define

    (𝒮,P)=(F𝒮,P)

    to be the hopeless ideal with respect to𝒮 conditioned on P. Here we suppress mentioning the assignment F𝒮 in the notation, as in all situations we will consider it will be given by the strategy 𝒮 in a natural way. The ideal (𝒮,) is called the unconditional hopeless ideal with respect to 𝒮. We will write (𝒮) for (𝒮,).

    When the strategy 𝒮 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 𝒢 of limit length, a strategy 𝒮 for Player II in 𝒢 and a position P as in Definition 3.3, the ideal (𝒮,P) is κ-complete. If all ultrafilters Uir associated with F𝒮,P are uniform then (𝒮,P) is uniform. If moreover F𝒮,P is normal, then (𝒮,P) is normal as well.

    4. Games We Play

    In this section, we introduce a sequence of games 𝒢k closely related to Welch’s game 𝒢γW. The last one will be 𝒢2, and we will be able to show that if 𝒮 is a winning strategy for II in 𝒢2 of sufficient length then we can construct a winning strategy 𝒮* for Player II in 𝒢2 such that (𝒮*) is precipitous and more, depending on the length of the game and the payoff set.

    To unify the notation, we let 𝒢0 of length γ be the Welch game 𝒢γW. Thus, a run of the game continues until either Player II cannot play or else until γ rounds are played. The set of all runs of 𝒢0 of length γ is denoted by Rγ. As usual with these kinds of games, a set BRγ is called a payoff set. We say that Player II wins a run R of the game 𝒢0 of length γ with payoff set B if R has γ rounds and the resulting run is an element of B. We call this game 𝒢0(B). Thus, if B=Rγ then 𝒢0(B) is just the game 𝒢0. With this notation, the game 𝒢γ* is just the game 𝒢0(Qγ) of length γ where

    Qγ=The set of all runs𝒜i,Ui|i<γRγsuch that there isaκ-complete ultrafilter on theκ-algebra generated byi<γ𝒜iextending allUi,i<γ.(7)
    As already discussed in the introduction, the existence of a winning strategy for Player II in the game 𝒢0(Qγ) of length γ is a strengthening of the requirement that Player II has a winning strategy in 𝒢0 of length γ. This strengthening is among the weakest ones which increase the consistency strength in the case γ=ω. From the point of view of increasing the consistency strength, the case γ=ω is of primary interest, as follows from (TO1) combined with Corollary 1.6. Here are some trivial observations.

    (TO1)

    𝒢0(Qγ) is the same game as 𝒢0 whenever γ is a successor ordinal, so a winning strategy for Player II in 𝒢0(Qγ) gives us something new only when γ is a limit.

    (TO2)

    A winning strategy for Player II in 𝒢0(Qγ) is a winning strategy for Player II in 𝒢0, but the converse may not be true in general.

    (TO3)

    If 𝒮 is a winning strategy for Player II in 𝒢0 of length >γ then the restriction of 𝒮 to positions of length <γ is a winning strategy for Player II in 𝒢0(Qγ) of length γ.

    (TO4)

    Given ξ<κ and sequences 𝒜i|i<ξ and Ui|i<ξ where 𝒜iP(κ) and Ui is a κ-complete ultrafilter on the κ-algebra (respectively, normal κ-algebra) of subsets of κ generated by ji𝒜j such that UiUj whenever ij, there is at most one κ-complete (respectively, normal) ultrafilter U on the (normal) κ-algebra of subsets of κ generated by i<ξ𝒜i which extends all Ui. Thus, if we changed the rules of 𝒢0 to require that Player II goes first at limit stages then Player II has a winning strategy in this modified 𝒢0 if and only if Player II has a winning strategy in the original game 𝒢0.

    (TO5)

    Let 𝒮 be a winning strategy for Player II in 𝒢0 or 𝒢0(Qγ) and

    I𝒜0𝒜1𝒜α𝒜α+1
    IIU0U1UαUα+1

    be a play of the game 𝒢0 or 𝒢0(Qγ) according to 𝒮. Let i𝒜i be another sequence of κ-complete algebras. Then, the play

    I01αα+1
    IIU00U11UααUα+1α+1

    is a run of the game where Player II wins.

    In what follows, we will consider 𝜃 a regular cardinal much larger than κ, and fix a well-ordering of H𝜃 which we denote by <𝜃. 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 H𝜃. We will thus work with the structure (H𝜃,,<𝜃), but will frequently suppress the symbols denoting and <𝜃 in our notation.

    The common background setting for the games we are going to describe is an internally approachable sequence Nα|α<κ+ of elementary substructures of H𝜃. That is: a continuous sequence such that for all α<κ+ the following hold.

    (a)

    κ+1Nα and card(Nα)=κ,

    (b)

    <κNα+1Nα+1,

    (c)

    Nξ|ξαNα whenever α<α.

    The following are standard remarks:

    If we are playing any of the games 𝒢0,𝒢1,𝒢1 of 𝒢2 then the game has length γκ. Since κ+1N0, γNα for all α.

    If Nα:α<κ is an internally approachable sequence then there is a closed unbounded set Cκ+ such that for αC, Nακ=α.

    If 2κ=κ+, then there is a well ordering of P(κ) of order type κ+ in H𝜃. Hence, if Nα:α<κ+ is an internally approachable sequence then P(κ)=α<κ+(P(κ)Nα). Clearly, P(κ)P(κ)α<κ+Nα implies that 2κ=κ+, which we stated as an assumption in Theorems 1.1 and 1.2.

    Definition 4.1 (The Game 𝒢1−). The rules of the game 𝒢1 are as follows. Fix an ordinal γκ+.

    Player I plays an increasing sequence of ordinals αi<κ+.

    Player II plays an increasing sequence of uniform κ-complete ultrafilters Ui on P(κ)M where M=Nαi+1.

    Player I plays first at limit stages.

    A run of 𝒢1 continues until Player II cannot play or until it reaches length γ. Player II wins a run in 𝒢1 if and only if the length of the run is γ.

    Payoff sets Rγ and Qγ for 𝒢1 are defined analogously to the definition for the game 𝒢0. So Rγ consists of all runs of 𝒢1 of length γ, and Qγ consists of all runs αi,Ui|i<γRγ such that there is a κ-complete ultrafilter on the κ-algebra generated by P(κ)Nα, where α=supi<γαi, extending all Ui, i<γ.

    The symbols Rγ and Qγ have a double usage: They were also defined in connection with the game 𝒢0 and were different, but analogous to that in Definition 4.1. Thus, to determine the exact meaning of Rγ and Qγ one always needs to take into account which game is being considered.

    In the case where γ=κ+, if Player II has a winning strategy in any of the games then κ is measurable. So for the purposes of this paper we can assume that γκ, in particular γNα for every α.

    Remark. Let αi,Ui|i<ξ be a full or partial play of the game 𝒢1 and α=supi<ξαi.

    (1)

    If ξ has cofinality κ then P(κ)Nα is a κ-algebra.

    (2)

    If ξ=ζ+1, then α=αζ, and again P(κ)Nαζ+1 is a κ-algebra.

    (3)

    if ξ is a limit ordinal of cofinality less than κ, then the κ-algebra of sets generated by i<ξNαi is not a κ-algebra, the κ-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 𝒢0 and 𝒢0(Qγ) also hold for 𝒢1 and 𝒢1(Qγ).

    Proposition 4.2. Assuming 2κ=κ+ and γκ+ is an infinite regular cardinal, the following hold.

    (a)

    Player II has a winning strategy in 𝒢1 of length γ if and only if Player II has a winning strategy in 𝒢0 of length γ.

    (b)

    Player II has a winning strategy in 𝒢1(Qγ) of length γ if and only if Player II has a winning strategy in 𝒢0(Qγ) of length γ.

    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 𝒮0 is a winning strategy for Player II in 𝒢0 then 𝒮0 induces a winning strategy 𝒮1 for Player II in 𝒢1 the output of which at step i is the same as the output of 𝒮 at step i in the auxiliary game 𝒢0 where Player I plays P(κ)Nαi+1 at step i (where αi is the move of Player I in 𝒢1 at step i). For the converse, we proceed similarly. This time a winning strategy 𝒮1 for Player II in 𝒢1 induces a strategy 𝒮0 for Player II in 𝒢0 as follows. If Player I plays 𝒜i at step i in 𝒢0 then Player I plays

    αi=the leastα>αifor alli<isuch that𝒜iNα+1
    in the auxiliary game 𝒢1. Letting Ui be the output of 𝒮1 at step i, we let the output of 𝒮0 at step i to be Ui𝒜i. That 𝒮0 is a winning strategy for Player II in 𝒢0 is immediate.

    It is straightforward to verify that this choice of strategies also works in the case of games with payoff sets Qγ 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 𝒮0 is the <𝜃 least winning strategy for Player II in 𝒢0 and S1 be the strategy defined from 𝒮0 as in Proposition 4.2. Suppose that βγ and αi:i<β is a sequence of ordinals such that for all i,αi+1<α. Then Player II’s response to αi:i<β in 𝒢1 belongs to Nα+1.

    Proof. Because the sequence Nα:α<κ+ is internally approachable and αi<α, αi+1<α. Since we are taking γκ and Nα+1 is closed under <κ-sequences, the sequence of κ-algebras P(κ)Nαi+1:i<β belongs to Nα. Since 𝒮0 is <𝜃-least, the sequence of responses by Player II to P(κ)Nαi+1:i<β in 𝒢0 belongs to Nα+1, and hence the sequence of responses by Player II according to 𝒮1 as defined in Proposition 4.2 belongs to Nα+1. □

    Definition 4.4 (The Game 𝒢1). The rules of 𝒢1 are exactly the same as those of 𝒢1 with the only difference that the ultrafilters Ui played by Player II are required to be normal with respect to Nαi+1.

    As before, the payoff set Rγ is defined for 𝒢1 the same way as it was for 𝒢0 and 𝒢1, that is, Rγ consists of all runs of 𝒢1 of length γ. For 𝒢1 we define a payoff set Wγ as follows:

    Wγ=the set of allαi,Ui|i<γRγsuch that ifXi|i<γis a sequence satisfyingXiUifor alli<γtheni<γXi.

    Note that Wγ= whenever γκ, so the game 𝒢1(Wγ) is of interest only for γ<κ. The existence of a winning strategy for Player II in 𝒢1(Wω) of length ω seems to be exactly what is needed to run the proof of precipitousness of the hopeless ideal (𝒮*) 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 𝒢1(Qω) of length ω.

    In the case of 𝒢1 we will not make use of a payoff set for 𝒢1 that would be an analogue of what was Qγ for 𝒢0 and 𝒢1, so we will not introduce it formally. We note that Qγ is a subset of Wγ, so the winning condition for Player II is weaker using Wγ.

    Let us also note that the somewhat abstract notion of normality of an ultrafilter Ui on 𝒜i=P(κ)Nαi+1 introduced in Sec. 3 is identical with the usual notion of normality with respect to the model Nαi+1 where it is required that Ui is closed under diagonal intersections of sequences Aξ|ξ<κNαi+1 such that AξUi for all ξ<κ.

    Remark 4.5. If we have a strategy 𝒮 defined for either 𝒢1 or 𝒢1, then a play of the game according to this strategy is determined by Player I’s moves. Thus, if 𝒮 is clear from context we can save notation by referring to plays as sequences of ordinals αi:i<β. Similarly if 𝒮 is a partial strategy defined on plays of length at most β we can index these plays according to 𝒮 by αi:i<β*, where β*β. 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 𝒢1 and 𝒢1 hold.

    (a)

    Let γκ+ be an infinite regular cardinal. If Player II has a winning strategy in 𝒢1 of length γ then Player II has a winning strategy in 𝒢1 of length γ. (So in fact we haveif and only ifhere, as the converse holds trivially.)

    (b)

    If Player II has a winning strategy in 𝒢1(Qγ) of length γ then Player II has a winning strategy in 𝒢1(Wγ) of length γ.

    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 Mα be the transitive collapse of Nα. We will work with models Mα in place of Nα.

    Since κ+1Nα, we have

    P(κ)Nα=P(κ)Nα=P(κ)Mα=P(κ)Mα,

    so the games 𝒢1 and 𝒢1 can be equivalently defined using structures Mα instead of Nα.

    If U is an M-ultrafilter over κ we denote the internal ultrapower of M by U by Ult(M,U). Then Ult(M,U) is formed using all functions f:κM which are elements of M. If U is κ-complete then Ult(M,U) is well-founded, and we will always consider it transitive; moreover the critical point of the ultrapower map πU:MUlt(M,U) is precisely κ. Recall also that U is normal if and only if κ=[id]U, that is, κ is represented in the ultrapower by the identity map. As MZFC (by ZFC we mean ZFC without the power set axiom), the Łoś Theorem holds for all formulae, hence the ultrapower embedding πU is fully elementary. Finally recall that the M-ultrafilter derived from πU, which we denote by U*, is defined by

    XU*κπU(X)(8)
    and U* is normal with respect to M.

    Assume α<α and U is a κ-complete Mα-ultrafilter. Suppose that U=UMα. We have the following diagram:

    Here σ:MαMα is the natural map arising from collapsing the inclusion map from Nα to Nα, and σ is the natural embedding of the ultrapowers defined by
    [f]U[σ(f)]U.(10)
    Note that cr(σ)=(κ+)Mα. Using the Łoś theorem, it is easy to check that the diagram is commutative, σ is fully elementary, and σκ=idκ. It follows that
    σ(κ)κ.(11)
    Given a set XP(κ)Mα,
    XU*κπU(X)σ(κ)σ(πU(X))=πU(σ(X))=πU(X).(12)
    Thus, using (11) combined with (12),
    U*(U)*σ(κ)>κ.(13)

    The property that σ(κ)>κ can be restated as saying that if f represents κ in Ult(Mα,U) and g represents κ in Ult(Mα,U), then {δ:f(δ)>g(δ)}U. Equation (13) can also be rephrased in this way.

    Note that since U*,(U)* are ultrafilters on the respective models, the statement U*(U)* can be equivalently expressed as U*=(U)*Mα.

    Before we define the winning strategies for Player II in 𝒢1, 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 αn:n be an increasing sequence of ordinals between κ and κ+. Then there is no sequence of ultrafilters Un:n such that

    Un is a κ-complete ultrafilter on Mαn+1

    (Un)*(Un+1)*

    there is a countably complete ultrafilter V on n(P(κ)Mαn+1) with VUn for all n.

     □

    Proof. For each n let fnMαn+1 represent κ in Ult(Mαn+1,Un), and let σn be the map from Ult(Mαn+1,Un) to Ult(Mαn+1+1,Un+1) defined as in Eq. (10). Then there is a set Xn+1Un+1 such that for all δXn+1,fn(δ)>fn+1(δ). The Xn’s all belong to V and intersecting them we get a δκ such that for all n,fn(δ)>fn+1(δ), a contradiction. □

    We note that Lemma 4.7 implies that in a play αi,Ui:i<γ there is no infinite increasing sequence in:n such that (Uαin)*(Uαin+1)*.

    Let αi,Ui:i<β be a partial or complete play of the game 𝒢1 of limit length β. Suppose that N=i<βNαi+1. Then the transitive collapse of N is the direct limit of Mαi+1:i<β along the canonical functions σj,j:Mαj+1Mαj+1 in diagram (9). Denote the transitive collapse of N by M. Let U be a κ-complete ultrafilter defined on the κ-algebra generated by P(κ)M that extends i<βUi. The following lemma implies that if for some i<β,(Ui)*(U)* then for some j with i<j<β, (Ui)*(Uj)*.

    Lemma 4.8. Let MαMβMγ with α,β,γ members of the αi+1s. Let UαUβUγ be κ-complete ultrafilters on the respective P(κ)s of Mα,Mβ,Mγ. Suppose that (Uα)*(Uγ)*. Let XMαP(κ) be such that X(Uα)*,X(Uγ)*.

    Then we can choose fα,gα,fγ,gγ such that

    fα,gαMα,[fα]Uα=κ,[gα]Uα=X,fγ,gγMγ,[fγ]Uγ=κ,[gγ]Uγ=X.
    Suppose that fγ,gγMβ. Then (Uα)*(Uβ)*.

    Proof. If any of Uα,Uβ,Uγ are principal the hypothesis clearly fails. It follows that each of the ultrafilters is uniform.

    The point of the proof is showing that if fα,gα,fγ,gγ belong to Mβ, then X(Uβ)*. Since X(Uα)* but does belong to Mα, it follows that (κX)Uα*. So κX witnesses the conclusion of the lemma.

    Using the notation of diagram (9), since

    σ:Ult(Mβ,Uβ)Ult(Mγ,Uγ)
    is order preserving and [fγ]Uγ=κ, we must have [fγ]Uβ=κ.

    Since X(Uγ)*, [fγ]Uγ=κ and [gγ]γ=X, we must have {δ:fγ(δ)gγ(δ)}Uγ. Since fγ and gγ belong to Mβ and UβUγ we have {δ:fγ(δ)gγ(δ)}Uβ, and hence [gγ]Uβ(Uβ)*.

    To finish it suffices to show that [gγ]Uβ=X. Since {δ:sup(gγ(δ))=fγ(δ)}Uγ, we must have {δ:sup(gγ(δ))=fγ(δ)}Uβ. Thus, sup([g]Uβ)=κ.

    For α<κ, let cα:κκ be the constant function α. Then {δ:cα(δ)<fγ(δ)}Uβ, by κ-completeness. Using induction and the κ-completeness of Uβ, one proves that [cα]Uβ=α. But then

    α[gγ]Uβif and only if{δ:cα(δ)gγ(δ)}Uβif and only if{δ:cα(δ)gγ(δ)}Uγif and only ifα[gγ]Uγ.
    Since [gγ]Uγ=X we have [gγ]Uβ=X. □

    It follows from Lemmas 4.7 and 4.8 that if αi,Ui:i<β+k is a play of 𝒢1 where β is zero or a limit ordinal and kω, then there is a finite set i0=0<i1<i2<in=β such that for all 1m<n

    (A)

    for all i<j[im1,im), it holds that (Ui)*(Uj)*,

    (B)

    for all i[im1,im), (Ui)*(Uim)*.

    We will call the stages i1,in together with {0j<k1:(Uβ+j)*(Uβ+j+1)*} drops. Note that in clause (B), m<n so this does not imply that β is a drop.

    A position P of the game 𝒢1 has the form

    P=αiP,UiP|i<βP(14)
    where αiP are moves of Player I and UiP are moves of Player II, and we will not use the superscripts P if there is no danger of confusion. We will take β=0 as the length of the empty position. Given an infinite regular cardinal γ and a strategy 𝒮 for Player II in the game 𝒢1 of length γ, let 𝒵γ be the set of all positions in 𝒢1 of length <γ according to 𝒮 that have successor length, where the last move of Player I is a drop. As stated in Remark 4.5 we can index plays in 𝒵γ by increasing sequences of ordinals. On 𝒵γ we define a binary relation as follows. Given two positions P,Q𝒵γ, we let
    PQ(15)
    if and only if P properly extends Q.

    Claim 4.9. Assume one of the following holds

    (a)

    γ>ω is regular and 𝒮 is a winning strategy for Player II in 𝒢1 of length γ.

    (b)

    γ=ω and 𝒮 is a winning strategy for Player II in 𝒢1(Qγ) of length γ.

    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 𝒮 is a winning strategy in any of the variants of 𝒢1 of any length γ, then is well-founded. Note for well-foundedness the only relevant γ are limit ordinals. As stated, the Claim handles all of the cases relevant to the theorems we are proving.

    Now assume 𝒮 is as in (a) or (b) in Claim 4.9. For P𝒵γ, let iP be the largest drop in P if P does have a drop, and iP=0 otherwise. Fix a -minimal P𝒵γ. By the minimality of P, if P extends P then iP=iP; in other words, P has no drops above iP, hence (UiP)*(UiP)* whenever iPi<i.

    Let α*=αiP and V*=(Uα*)*.

    We define a winning strategy 𝒮P for Player II in 𝒢1 of length γ. Viewing 𝒮 as defined on sequences of ordinals αi:i<β, we define 𝒮P on such sequences αi:i<β by induction on their length β.

    For ordinals αi<α* played by the first player we assume inductively that the normal ultrafilter Vi played by the second player is (V)*Nαi+1.

    Suppose we have defined 𝒮P on sequences of length less than β, where β=0 corresponds to the empty position. Formally, to αi:i<β we inductively associate the play (αi,Vi):i<β where Vi is the response by Player II according to 𝒮P. We need to define 𝒮P on αi:i<βαβ.

    Case 1. αβα*. In this case

    𝒮P(αi:i<βαβ)=(V)*Nαβ+1.

    Case 2. αβ>α*.

    Let j be least such that αj>α*. Let

    𝒮P(αi:i<βαβ)=(𝒮(Pαi:jiβ))*.

    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 α* are sub-filters of V* and hence are legal plays and sub-filters of 𝒮(P)*. Going beyond P, the plays of 𝒮P are extensions of plays according to 𝒮 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 𝒮.

    From this we conclude that Player II wins the game of length γ in part (a) of Claim 4.9.

    We only prove (b) for γ=ω because that is the most relevant case for this paper. A straightforward generalization of this argument gives the result for general γ. The strategy 𝒮P is defined using a winning play by 𝒮 in the game 𝒢1(Qω). Since 𝒮 is a winning strategy in that game, if (αn,Un):n is that play according to 𝒮, there is a κ-complete ultrafilter UUn defined on the κ-algebra generated by nMαn. By Lemmas 4.7 and 4.8 and the remarks preceding them, (U)* extends Vn for all n. Part (b) follows.

    Remark 4.10. Arguing exactly as in Lemma 4.3, if αi:i<βNα is a sequence of ordinals and a -minimal position P in the game 𝒢1 belongs to Nα then the sequence of responses by Player II to αi:i<β using 𝒮P belongs to Nα. In particular, if β is a successor ordinal j+1 then 𝒮P’s responses belong to Nαj+2.

    Definition 4.11 (The Game 𝒢2). The rules of the game 𝒢2 are as follows.

    Player I plays an increasing sequence of ordinals αi<κ+ as before.

    Player II plays distinct sets Yiκ such that the following are satisfied.

    (i)

    Yj*Yi whenever i<j, and

    (ii)

    Letting Ui={XP(κ)Nαi+1|Yi*X}, the family Ui is a uniform normal ultrafilter on P(κ)Nαi+1.

    Player I goes first at limit stages.

    A run of 𝒢2 of length γκ+ continues until Player II cannot play or else until it reaches length γ.

    Payoff sets Rγ and Wγ for the game 𝒢2 are defined analogously to those for 𝒢1. So Rγ consists of all runs in 𝒢2 of length γ and Wγ consists of all those runs αi,Yi|i<γRγ such that if Xi|i<γ is a sequence satisfying XiNαi+1 and Yi*Xi for all i<γ then i<γXi.

    Note that YiNαi+1 in (ii). Note also that since the ultrafilters Ui are required to be uniform, the sets Yi are unbounded in κ. As with 𝒢1, we will not make any use of what would be an analogue of payoff set Qγ.

    Proposition 4.12. Assume γκ+ is an infinite regular cardinal.

    (a)

    Player II has a winning strategy in 𝒢1 of length γ if and only if Player II has a winning strategy in 𝒢2 of length γ.

    (b)

    Player II has a winning strategy in 𝒢1(Wγ) of length γ if and only if Player II has a winning strategy in 𝒢2(Wγ) of length γ.

    Proof. For (a), it is immediate that a winning strategy for Player II in 𝒢2 gives a winning strategy in 𝒢1: if Player II plays Yi at turn i, then Yi generates a normal ultrafilter on Nαi+1 which is Player II’s move in 𝒢1.

    For the non-trivial direction, assume Player II has a winning strategy 𝒮 in 𝒢1 of length γ. As noted before Definition 4.11, such a strategy exists in N0. We build a winning strategy 𝒮 for Player II in 𝒢2 of length γ by induction.

    Induction Hypothesis. Suppose that Player I plays αi:i<β in the game 𝒢2, and Ui:i<β is the play by Player II using 𝒮 in the game 𝒢1. Then Player II plays Yi:i<β where Yi is a definable diagonal intersection of the members of Ui.

    For each i, let Xξi|ξ<κ be the <𝜃-least enumeration of Ui of length κ (recall that <𝜃 is the well-ordering of H𝜃 fixed at the beginning of this section; see the paragraphs immediately above Definition 4.1). The induction hypothesis is that for all i<β, Player II’s responses according to the strategy 𝒮 to the sequence αi:iδ are Yi:iδ where

    Yi=Δξ<κXξi.

    This induction hypothesis is automatically preserved at limit stages. Suppose that it holds up to β and Player I plays αβ. Then Player II plays an ultrafilter Uβ on P(κ)Nαβ+1 in the game 𝒢1 using the strategy defined in Proposition 4.6. Then, as in Remark 4.10, Nαβ+2 contains the information that Uβ is Player II’s response as well as the <𝜃-least enumeration Xξβ:ξ<κ of Uβ. Let Yβ=Δξ<κXξβ and let Yβ be Player II’s response in 𝒢2 using 𝒮.

    Suppose now that αi:i<γ is a run of the game 𝒢2 according to 𝒮. Then, since Ui+1 is normal each Yi belongs to Ui+1. Since, Yj*X for all XUj, for i<j,Yj*Yi. Moreover, since Yi is a diagonal intersection of the ultrafilter Ui, clause (ii) in Definition 4.11 is immediate.

    Since the relevant ultrafilters are the same, whether II is playing by 𝒮 in 𝒢1 or 𝒮 in 𝒢2, clause (b) in Proposition 4.12 is immediate. □

    Remark 4.13. The definition of the winning strategy 𝒮 for Player II in the previous proof depends on the position P in 𝒢1, beyond which there are no drops. Suppose that Player I plays αi:i<β in the game 𝒢2 and Player II responds with Yi:i<β using the winning strategy 𝒮. Then for all j<β with PNj,

    YjNαj+1 because it induces an ultrafilter on Nαj+1,

    YjNαj+2 because Nαi:ijNαj+2 and Player II’s response to αi:ij according to 𝒮 is definable from Player II’s response to αi:ij according to the strategy 𝒮 for 𝒢1, which in turn is definable from P and Player II’s response according to her strategy in 𝒢1 and thus from the original strategy in 𝒢0.

    It follows that for all i<j, Yj*Yi and |YiYj|=κ. (Restating this Yj*Yi.)

    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 𝒮1 is a winning strategy for Player II in the game 𝒢1 of length γ and 𝒮2 is the winning strategy for Player II in the game 𝒢2 of length γ obtained as in Proposition 4.12. Then for every AP(κ),

    A(𝒮1)A(𝒮2).

    5. Strategies 𝒮* and 𝒮γ

    Consider a winning strategy 𝒮0 for Player II in 𝒢0 of length γ and a position P in 𝒢0 according to 𝒮0. Given a set X(𝒮0,P)+, there may exist different runs of 𝒢0 extending P which witness that X is (𝒮0,P)-positive. This causes difficulties in proving that (𝒮0,P) 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 𝒮* for Player II in 𝒢2 of length γ such that for each position Q in 𝒢2 according to 𝒮* and each X(𝒮*,Q) there is a unique run witnessing that X is (𝒮*,Q)-positive, and show that using 𝒮* one can prove the precipitousness of (𝒮*,) 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 κ, we always mean uniform κ-complete and κ+-saturated ideals over κ. The results in this section are formulated under the assumption of the non-existence of a normal saturated ideal over κ, 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 κ, is a consequence of the following standard proposition.

    Proposition 5.1. Given a regular cardinal κ>ω, the following are equivalent.

    (a)

    κ carries a saturated ideal.

    (b)

    κ carries a normal saturated ideal.

    Proof. A standard elementary argument shows that any uniform normal ideal over κ is κ-complete, hence (a) follows immediately from (b).

    To see that (b) follows from (a), assume is a saturated ideal over κ. Let be the partial ordering (+,) and U̇ be a -term for the normal V-ultrafilter over κ derived from the generic embedding jG:VMG associated with Ult(V,G) where G is (,V)-generic. Let *V be the ideal over κ defined by

    a*VǎU̇.
    Equivalently
    a*Vκ̌j(ǎ).
    A standard argument shows that * is a uniform normal ideal over κ. To see that * is saturated, we construct an incompatibility-preserving map e:(*)++. Let f:κκ be a function in V which represents κ in Ult(V,G) whenever G is (,V)-generic. Since is saturated, such a function can be constructed using standard techniques (see [20]). Let
    e(a)=def{ξ<κ|f(ξ)a}.

    Note that for every aP(κ)V and every (,V)-generic G

    aU̇GκjG(a)[f]G[ca]Ge(a)G.
    It follows from these equivalences that indeed e(a)+ whenever a(*)+. To see that e is incompatibility preserving, we prove the contrapositive. Assume e(a),e(b) are compatible, so e(a)e(b)+. Let G be (,V)-generic such that e(a)e(b)G. Then e(a),e(b)G, so a,bU̇G by the above equivalences. But then abU̇G, which tells us that ab(*)+. □

    We are now ready to formulate the main technical result of this section.

    Proposition 5.2. Assume 2κ=κ+ and there is no normal saturated ideal over κ. Let γκ+ be an infinite regular cardinal and 𝒮 be a winning strategy for Player II in 𝒢2 of length γ. Then there is a tree T(𝒮) which is a subtree of the partial ordering (P(κ),*) such that the following hold.

    (a)

    The height of T(𝒮) is γ and T(𝒮) is γ-closed.

    (b)

    If Y,YT(𝒮) are *-incomparable then Y,Y’ are almost disjoint.

    (c)

    There is an assignment YPY assigning to each YT(𝒮) a position PY in 𝒢2 of successor length according to 𝒮 in which the last move by Player II is Y; we denote the last move of Player I in PY by α(Y). The assignment YPY has the following property:

    Y*Yα(Y)<α(Y)andPYis an extension ofPY.

    (d)

    If b is a branch of T(𝒮) of length <γ, let Pb=YbPY. Then Pb is a position in 𝒢2 according to 𝒮, and the set of all immediate successors of b in T(𝒮) is of cardinality κ+. Moreover the assignment Yα(Y) is injective on this set.

    Finally, if A(𝒮)+ then it is possible to construct the tree T(𝒮) in such a way that

    AT(𝒮).(16)

    Clause (d) in the above definition treats both successor and limit cases for γ. The successor case in (d) simply says that if YT(𝒮) then the conclusions in (d) apply to the set of all immediate successors of Y in T(𝒮).

    Proof. The tree T(𝒮) is constructed by induction on levels. Limit stages of this construction are trivial: If γ̄<γ is a limit and we have already constructed initial segments Tγ* of T(𝒮) of height γ* for all γ*<γ̄ so that (b)–(d) hold with Tγ* in place of T(𝒮) and Tγ end-extends Tγ* whenever γ*<γ<γ̄ then it is easy to see that Tγ̄=γ*<γ̄Tγ* is a tree with tree ordering * end-extending all Tγ*, γ*<γ̄, and such that (b)–(d) hold with Tγ̄ in place of T(𝒮). We will thus focus on the successor stages of the construction.

    Assume γ̄<γ and T(𝒮) is constructed at all levels up to level γ̄; our task now is to construct the γ̄th level of T(𝒮). Let b be a cofinal branch through this initial segment of T(𝒮), so b is of length γ̄. We construct the set of immediate successors of b in T(𝒮), along with the assignment YPY on this set, as follows. As we are assuming there is no normal saturated ideal over κ, we can pick an antichain 𝒜 in (𝒮,Pb)+ of cardinality κ+. For each X𝒜 there is a position QX in 𝒢2 of successor length <γ according to 𝒮 extending Pb such that the last move by Player II in QX is almost contained in X. For the sake of definability we can let this position to be <𝜃-least, where recall that <𝜃 is the fixed well-ordering of H𝜃.

    Now construct the set Yξ|ξ<κ+ of all immediate successors of b in T(𝒮) recursively as follows. Assume ξ<κ+ and we have already constructed the set Yξ̄|ξ̄<ξ along with the assignment Yξ̄PYξ̄ with the desired properties. Since each model Nβ is of cardinality κ, we can pick the <𝜃-least set X𝒜 which is not an element of any Nα(Yξ̄)+1 where ξ̄<ξ. Now let Player I extend QX by playing the least ordinal α such that

    {X}{Yξ̄|ξ̄<ξ}Nα+1.(17)
    This is a legal move in 𝒢2 following QX. Let Y be the response of the strategy 𝒮 to QXα. We let Yξ be this Y and PY=QXα,Y. Note that Yξ*X, as Yξ, being played according to 𝒮, is almost contained in the last move by Player II in QX.

    We show

    Any two setsYYon theγ̄th level are almost disjoint.(18)
    If Y,Y are above two distinct cofinal branches then this follows immediately from the induction hypothesis: Letting Z, respectively, Z be the immediate successor of bb in b, respectively, b, we have Y*Z and Y*Z, and the induction hypothesis tells us that Z,Z are almost disjoint.

    Now assume Y,Y are above the same branch b; without loss of generality we may assume Y=Yξ and Y=Yξ in the above enumeration and ξ<ξ. Then we have X,X,PY,PY as in the construction, with Y*X and Y*X. Also α(Y)<α(Y).

    If Y*Y then Y*XX, thus witnessing XX(𝒮,Pb)+. This contradicts the fact that 𝒜 is an antichain in (𝒮,Pb)+. It follows that Y*Y. Now for every ZNα(Y)+1 the set Y is either almost contained in or almost disjoint from Z. As YNα(Y)+1 by our choice of α(Y) in (17), necessarily Y is almost disjoint from Y. 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 T(𝒮), 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 γ̄th level. But clause (b) follows from the combination of (18) with the induction hypothesis and the fact that every set on the γ̄th level is almost contained in some set on an earlier level. Finally, that adding the γ̄th level keeps * a tree ordering follows from clause (b). More generally, any collection 𝒳P(κ) which satisfies (b) with 𝒳 in place of T(𝒮) has the property that the set of all Y𝒳 which are *-predecessors of a set Y𝒳 is linearly ordered under *. What now remains is to see that clause (a) holds, but this is immediate once we have completed all γ steps of the construction.

    Finally, given a set A(S)+, to see that we can construct the tree T(𝒮) so that (16) holds, note that we can put A into the first level of T(𝒮) at the first step in the inductive construction. This involves a slight modification of the construction of the first level of T(𝒮), and is left to the reader. □

    The new strategy 𝒮* for Player II in 𝒢2 is now obtained by, roughly speaking, playing down the tree T(𝒮). More precisely:

    Definition 5.3. Assume γκ+ is an infinite regular cardinal, 𝒮 is a winning strategy for Player II in 𝒢2 of length γ, and T(𝒮) is a subtree of the partial ordering (P(κ),*) satisfying (a)–(d) in Proposition 5.2. We define a strategy 𝒮* for Player II in 𝒢2 of length γ associated with T(𝒮) recursively as follows.

    Assume

    P={(αi,Yi)|i<j}
    is a position in 𝒢2 of length j<γ according to 𝒮*. Denote the corresponding branch in T(𝒮) by bP, that is,
    bP={Yi|i<j}.
    If αj is a legal move of Player I in 𝒢2 at position P then
    𝒮*(Pαj)=the unique immediate successorYofbPinT(𝒮)with minimal possibleα(Y)αj.
    Here recall that α(Y) is the last move of Player I in PY.

    As an immediate consequence of the properties of T(𝒮) we obtain:

    Proposition 5.4. Let γκ+ be an infinite regular cardinal and assume T(𝒮) is as in Proposition 5.2. Then 𝒮* is a winning strategy for Player II in 𝒢2 of length γ.

    Moreover if,

    r*=αi,Yi|i<γ
    is a run of 𝒢2 of length γ according to 𝒮* then
    r=i<γPYi
    is a run of 𝒢2 of length γ according to 𝒮.

    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 𝒢i of length γ, and to which we will refer later.

    Corollary 5.5. Under the assumptions of Proposition 5.2, assume A(𝒮)+ and T(𝒮) is constructed in such a way that (16) holds, that is, AT(𝒮). Let 𝒮* be the winning strategy for Player II constructed as in Definition  5.3 using this T(𝒮). Then A(𝒮*)+.

    One of the main points of passing to 𝒮* is the following remark.

    Remark 5.6. For any position P of a partial run according to 𝒮* of successor length with Y being the last move by Player II, the conditional hopeless ideal (𝒮*,P) is equal to the unconditional hopeless ideal restricted to Y

    (𝒮*,P)=(S*)Y.

    We now turn to a proof of Theorem 1.1. If there is a normal saturated ideal over κ then there is nothing to prove. Otherwise Player II has a winning strategy in 𝒢2(Wω) of length ω, 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 κ. Let

    𝒮 be a winning strategy for Player II in 𝒢2(Wω) of length ω, and

    𝒮* be the winning strategy constructed from 𝒮 as in Definition 5.3.

    Then Player I does not have a winning strategy in the ideal game 𝒢((𝒮*)). Consequently, the ideal (𝒮*) is precipitous.

    Proof. Assume 𝒮 is a strategy for Player I in the ideal game 𝒢((𝒮*)). We construct a run in 𝒢((𝒮*)) according to 𝒮 which is winning for Player II. Odd stages in this run will come from positions in 𝒢2 played according to 𝒮*; more precisely, they will be tail-ends of sets on those positions. So suppose

    Q=X0,X1,X2,X3,X2n1
    is the finite run of 𝒢((𝒮*)) constructed so far, and
    β0,Z0,β1,Z1,βn1,Zn1
    is the associated auxiliary run of 𝒢2 according to 𝒮* such that Zi*X2i and
    X2i+1=the longest tail-end ofZithat is contained inX2i
    for all i<n. Let X2n be the response of 𝒮 to Q in 𝒢((𝒮*)). As X2n(𝒮*)+, there is a finite position in 𝒢2 according to 𝒮* where the last move of Player II is a set almost contained in X2n and, letting Zn be this set, we also have X2nNα(Zn)+1.

    As the sets Zn constitute an *-decreasing chain of nodes in T(𝒮), the positions PZn extend PZm whenever m<n. By Proposition 5.4

    r=nωPZn
    is a run in 𝒢2 of length ω according to 𝒮. Let
    r=αi,Yi|iω
    be this run. For each iω let
    Xi=X2nwherenis such thatlh(PZn)i<lh(PZn+1).
    Then
    nωXn=nωX2n=iωXi.
    Here the equality on the left comes from the fact that the sets Xn, nω constitute an -descending chain, and the inequality on the right follows from the fact that XiNαi+1 and Yi*Xi for all iω, and that 𝒮 is a winning strategy for Player II in 𝒢2(Wω) of length ω; see the last paragraph in Definition 4.11. □

    We remark that the proof of Proposition 5.7 shows Player II has a winning strategy in the ideal game (𝒮*).

    The following proposition gives a proof of Theorem 1.2. Recall that all background we have developed so far was under the assumption that κ is inaccessible and 2κ=κ+. 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 𝒢0 of length γ>ω then Player II has a winning strategy in 𝒢0(Qω) of length ω and in 𝒢2(Wω) of length ω, as well as in 𝒢2 of length γ whenever γ is regular. By a similar argument, if Player II has a winning strategy in 𝒢2 of length γ>ω then Player II has a winning strategy in 𝒢2(Wω) of length ω. 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 κ and 2κ=κ+. Let γκ+ be an uncountable regular cardinal. Assume further that 𝒮 and 𝒮* are strategies as in Proposition 5.7, with γ in place of ω.

    Then T(𝒮) is a γ-closed dense subset of (𝒮*)+. It follows that Player I does not have a winning strategy in the ideal game 𝒢((𝒮*)). Consequently, the ideal (𝒮*) is precipitous.

    Proof. That T(𝒮) is a γ-closed dense subset of (𝒮*)+ follows immediately from the properties of T(𝒮). If A(𝒮*)+, then there is a play of the game such that A is in the ultrafilter determined by some Yξ played by Player II using 𝒮*. But then Yξ*A. Since Yξ is on T(𝒮), we have shown that for every element of (𝒮*)+ there is an element of the tree below it. Hence, the tree is dense.

    To see that (𝒮*) 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 𝒢((𝒮*)). At stage n of the game suppose that Player I plays X2n. Player II chooses an X2n+1T(𝒮) (so X2n+1(𝒮*)+) and X2n+1(S*)*X2n.

    Let An(S*) be such that X2n+1AnX2n. Player II’s response to X2n in 𝒢((S*)) is X2n+1=defX2n+1An. Let A=nAn. Since (S*) is countably complete, A(S*). Let XT(S), with X*X2n+1 for all n. Then:

    nXnnXnA(𝒮*)*XA.

    It follows that there is a set B(S*) such that XnXB. Since X(S*), XB is not empty. Hence Xn. □

    Proof of Theorem 1.4

    Proof. Consider a uniform κ-complete ideal 𝒥 over κ such that P(κ)/𝒥 is (κ+,)-distributive and has a dense γ-closed set. Because of notational convenience we will work with the partial ordering 𝒥=(𝒥+,𝒥). (See also the partial ordering used in the proof of Proposition 5.1.) Since a[a]𝒥 is a dense embedding of 𝒥 onto P(κ)/𝒥, we can fix a dense γ-closed set D𝒥. We work inside H𝜃 for a sufficiently large 𝜃 and will use the fixed well-ordering <𝜃 introduced in Sec. 4 to define a winning strategy 𝒮γ for Player II in 𝒢γW. As usual, 𝒮γ is defined inductively on the length of runs.

    So assume

    𝒜0,U0,𝒜1,U1,,𝒜j,Uj,
    is a run of 𝒢γW according to 𝒮γ for j<i. Along the way, we define auxiliary moves Xj played by Player II; these moves are elements of D, constitute a descending chain in the ordering by 𝒥, and for each j<i,
    Xj𝒥Ġ𝒜̌j=Ǔj.(19)
    At step i<γ Player I plays a κ-algebra 𝒜i on κ of cardinality κ extending all 𝒜j, j<i. As D is γ-closed and i<γ, there is an element XD below all Xj in 𝒥, j<i. If G is a (𝒥,V)-generic filter such that XG then by (19), UjG whenever j<i. Since 𝒥 is (κ+,)-distributive and 𝒜iV is of cardinality κ, the intersection G𝒜i is an element of V, and is a uniform κ-complete ultrafilter on 𝒜i extending all Uj where j<i. This is then forced by some condition YG such that Y𝒥X, hence Y𝒥Xj for all j<i. As D is dense in 𝒥, Y can be chosen to be an element of D. The following is thus not vacuous. We define
    Xi=the<𝜃-least elementYofDsuch thatY𝒥Xjfor allj<iand there is aUVsatisfyingY𝒥Ġ𝒜̌i=Ǔ

    and

    Ui=the uniqueUVsuch thatXi𝒥Ġ𝒜̌i=Ǔ.
    Letting
    𝒮γ(𝒜j,Uj|j<i𝒜i)=Ui,
    it is straightforward to verify that 𝒮γ is a winning strategy for Player II in 𝒢γW. □

    6. The Model

    In this section, we give a construction of a model where the following holds.

    κis inaccessible and carries no saturated ideals(20)
    and
    For every regular uncountableγκthere is an ideal𝒥γonP(κ)as in Theorem1.5,that is,𝒥γis uniform, normal,γ-densely treed and(κ+,)-distributive.(21)
    The model is a forcing extension of a universe V in which the following are satisfied.

    (A)

    GCH.

    (B)

    U is a normal measure on κ.

    (C)

    Tα,ξ|ξ<α+ is a disjoint sequence of stationary subsets of α+cof(α) whenever ακ is inaccessible.

    (D)

    Assume V[K] is a generic extension via a set-size forcing which preserves κ+, and, in V[K]

    there is a definable class elementary embedding j:VM where M is transitive, and

    Letting

    Tα,ξ|ξ<α+|αj(κ) is inaccessible inM=j(Tα,ξ|ξ<α+|ακ is inaccessible)
    V,M agree on what Hκ+ is and Tκ,ξ=Tκ,ξ whenever ξ<κ+.

    We will informally explain the purpose of the sets Tα,ξ before we begin with the construction of the model. These sets are not needed for the construction of ideals 𝒥γ in Theorem 1.5, but only for the proof that κ 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 Tα,ξ 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 Kc-constructions and fully background certified constructions. If there is a proper class inner model with a measurable cardinal then any Kc-construction (see for instance [21] for Kc-constructions of models with Mitchell-Steel indexing of extenders, and [25] for Kc-constructions with Jensen’s λ-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 V.

    There is some similarity in the argument in Proposition 6.1 of the existence of a sequence of mutually disjoint stationary subsets Tκ,ξ of κ+ which behave nicely with respect to the ultrapower by a normal ultrafilter on κ to a similar claim in [10] where it is proved that one can have such sequence of stationary sets in L[U].

    A background certified construction as above gives rise to a model of the form L[E] where E=Eα|αOn is such that each Eα either codes an extender in a way made precise, or Eα=. 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 L[E] models with the properties needed for the construction in this paper that satisfy the statement

    There is a Woodin cardinal κ 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 L[E] 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 GCH, and is proved along the way the model is constructed.

    From now on assume W=L[E] is a fine structural extender model with indexing of extenders as in [21] or in [25, Chap. 9] (which covers all L[E] models discussed above). We often write EW in place of E to emphasize that E is the extender sequence of W.

    FS1

    W||α is the initial segment of W of height ωα with the top predicate, that is, W||α=(JαE,Eωα).

    FS2

    If α is a cardinal of W then Eα=. Thus, in this case W||α=(JαE,) and we identify this structure with JαE.

    FS3

    If μ is a cardinal of W then the structure W||μ calculates all cardinals and cofinalities μ the same way as W. This is a consequence of acceptability.

    FS4

    β(τ) is the unique β such that τ is a cardinal in W||β but not in W||(β+1).

    FS5

    ϱ1 stands for the first projectum; that ϱ1(W||β)α is equivalent to saying that there is a surjective partial map f:αJβE which is Σ1-definable over W||β with parameters.

    FS6

    (Coherence.) If i:WW is a Σ1-preserving map in possibly some outer universe of W such that κ is the critical point of i and τ=(κ+)W then EWτ=EWτ.

    FS7

    (Cores.) Assume α is a cardinal in W and N is a structure such that ϱ1(N)=α and there is a Σ1-preserving map π of N into a level of W such that πα=id. Let pN be the <*-least finite set of ordinals p such that there is a set of ordinals a which is Σ1(N)-definable in the parameter p and satisfies aαN. 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 Σ1-hull of α{pN} and σ:N̄N be the inverse of the collapsing isomorphism. Then ρ1(N̄)=α, the models N̄,N agree on what P(α) is, and π is Σ1-preserving and maps N̄ cofinally into N. In this situation, N̄ is called the core of N and σ is called the core map.

    FS8

    (Condensation lemma.) Assume α is a cardinal in W and N,N̄,π and σ are as in FS7. Then N̄ is a level of W, that is, N̄=W||β for some β.

    Proposition 6.1. There is a formula φ(u,v,w) in the language of extender models such that the following holds. If W=L[E] is a fine structural extender model, α is an inaccessible cardinal of W and ξ<α+, letting

    Tα,ξ={τα+cof(α)|W||(α+)Wφ(τ,α,ξ)},
    each Tα,ξ is a stationary subset of α+cof(α) in W, and Tα,ξTα,ξ= whenever ξξ. Moreover, the sequence (Tα,ξ|ξ<α+|ακ is inaccessible in W) satisfies clause (D) above with W in place of V.

    Proof. Since the definition of Tα,ξ|ξ<α+ takes place inside W||(α+)W, any two extender models W,W such that (α+)W=(α+)W and EWα+=EWα+ calculate this sequence the same way (here α+ stands for the common value of the cardinal successor of α in both models). Now if V=W and j is as in (D) above then

    Tα,ξ={τ(α+)Mcof(α)|M||(α+)Mφ(τ,α,ξ)},
    whenever αj(κ) is inaccessible in M, so to see that Tκ,ξ=Tκ,ξ for all ξ<κ+ it suffices to prove that (κ+)M=(κ+)V and EVκ+=EMκ+ (where again κ+ stands for the common value of the cardinal successor of κ in V and M). Regarding the former, the inequality (κ+)V(κ+)M is entirely general and follows from the fact that P(κV)P(κ)M. The reverse inequality follows from the assumption that the generic extension preserves κ+, so (κ+)V remains a cardinal in M. The latter is then a consequence of the coherence property FS6.

    It remains to come up with a formula φ such that the sets Tα,ξ are stationary in W for all α,ξ of interest, and pairwise disjoint. Here we make a more substantial use of the fine structure theory of W. Given an inaccessible α and a ξ<α+, letting

    Tα,ξ=defthe set of allτα+cof(α)such thatϱ1(W||β(τ))=αandW||β(τ)hasξ+1cardinals aboveα,(22)

    it is clear that Tα,ξTα,ξ= whenever ξξ. Then it suffices to show that

    Tα,ξis stationary inW,(23)
    as we can then take φ be the defining formula for the system (Tα,ξ)α,ξ.

    The first step toward the proof of (23) is the following observation.

    Assumeν>αis regular inW,pW||νandXis theΣ1-hull ofα{p}inW||ν.LetνX=sup(Xν).ThencofW(νX)=α.(24)

    Proof. Obviously, γ=cofW(νX)α. Assume for a contradiction that γ<α. Let νi|i<γ be an increasing sequence converging to νX such that νiX for every i<γ. For each such i pick a jiω and an ordinal ηi<α such that νi=hW||ν(ji,ηi,p) where hW||ν is the standard Σ1-Skolem function for W||ν. Here W||ν is of the form JνE, (see FS2), and we identify it with the structure JνE. The Skolem function hW||ν has a Σ1-definition of the form (w)ψ(w,u0,u1,v) where ψ is a Δ0-formula in the language of extender models. (The standard Σ1-Skolem function has a uniform Σ1-definition, which means that there is a Σ1-formula which defines a Σ1-Skolem function hN over every acceptable structure N. However, the argument below does not make use of uniformity of the definition.) Since ν>α is regular

    (ν̄)(Jν̄E(i<γ)(w)(v)ψ(w,ji,ηi,p,v)).(25)
    Since the statement in (25) is Σ1, there is some such ν̄ with Jν̄EX. To justify this note that the sequences ηi|i<γ and ji|j<γ are elements of X as JαEX, and we can view these sequences as parameters in the formula in (25). Fix such an ordinal ν̄. Now consider i<γ such that νi>ων̄. Using (25) pick z and ν* in Jν̄E such that Jν̄Eψ(z,ji,ηi,p,ν*). Since ψ is Δ0, we actually have JνEψ(z,ji,ηi,p,ν*), which tells us that ν*=hW||ν(ji,ηi,p)=νi. As νi>ων̄, this is a contradiction. This completes the proof of (24). □

    Now let C be a club subset of α+, X be the Σ1-hull of α{C,ξ,α+ξ+1} in W||α+ξ+2, N be the transitive collapse of X, and π:NW||α+ξ+2 be the inverse of the collapsing isomorphism. Let further τ=Xα+=cr(π). Then τ>ξ as α{ξ}X. It is a standard fact that cofW(τ)=cofW(sup(XOn)) (and can be proved similarly as (24) above). Now cofW(sup(XOn))=α by (24), hence cofW(τ)=α. Moreover τC as C is closed and τ is a limit point of C. Thus, the proof of (22) will be complete once we show that ϱ1(W||β(τ))=α and W||β(τ)) has ξ+1 cardinals above κ. We first look at the set of cardinals in N.

    By acceptability, the structures W||α+ξ+1 and W||α+ξ+2 agree on what is a cardinal below α+ξ+1. It follows that in W||α+ξ+2, the statement

    The order type of the set of cardinals in the interval(α,α+ξ+1)isξ

    can be expressed in a Σ1-way as

    The order type of the set of cardinals aboveαin the structureW||α+ξ+1isξ.(26)
    Since π is Σ1-preserving and cr(π)=τ, this Σ1-statement can be pulled back to N via π. Also by the Σ1-elementarity of π we have π1(α+ξ+1) is the largest cardinal in N. Then, using acceptability in N, we conclude:
    The order type of the set of cardinals aboveαinNisξ+1.(27)
    By construction, the Σ1-Skolem function of N induces a partial surjection of α onto N. Then ϱ1(N)α by FS5. Since α is a cardinal in W, we conclude ϱ1(N)=α. Let N̄ be the core of N and σ:N̄N be the core map. By FS7, ϱ1(N̄)=α and P(α)N̄=P(α)N, so in particular τ=(α+)N=(α+)N̄. By FS8, N̄=W||β for some β. Since ϱ1(N̄)=α, FS5 implies β=β(τ). To see that N̄=W||β(τ) has ξ+1 cardinals above α, first note that, since by FS7 the map σ is cofinal, the largest cardinal in N must be in the range of σ. This along with (27) provides a Σ1-definition of ξ in N from parameters in rng(σ). The point here is that we can reformulate the notion of cardinal in N below α+ξ+1 as the cardinal in the sense of the structure N||α+ξ+1, similarly as in (26). It follows that ξrng(σ), and since ξ<(α+)N we have ξ<cr(σ). Then, using the Σ1-reformulation of (27) one more time, we conclude that α+ηrng(σ) for every ηξ, which means that W||β(τ)=N̄ has ξ+1 cardinals above α. This completes the proof of (22) and thereby the proof of Proposition 6.1.

    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 λ we will often assume GCH even when it is known that λ<λ suffices. Since the models we work in satisfy the GCH this is not important for our results.

    Recall that if Sλ+ is a stationary set (where λ is a cardinal) then the club shooting partial ordering 𝕊(S) consists of closed bounded subsets of λ+ 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 λ is regular, λ<λ=λ and T is a subset of λ+ such that Tα is non-stationary in α whenever α<λ+, and (λ+cof(λ))\T is stationary. Then the following hold.

    (a)

    𝕊(λ+\T) is (λ+,)-distributive, that is, it does not add any new function f:λV. In particular, generic extensions of V via 𝕊(λ+\T) agree with V on all cardinals and cofinalities λ+, and on what Hλ+ is.

    (b)

    If γλ is regular and Tλ+cof(γ) then 𝕊(λ+\T) has a dense set which is γ-closed but if T is stationary then it does not have a dense set which is γ+-closed.

    (c)

    If G is (𝕊(λ+\T),V)-generic then CG=G is a closed unbounded subset of λ+ such that CGλ+\T.

    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 λ is an uncountable cardinal with λ<λ=λ, A1,A2 are disjoint stationary subsets of λ+ and that for all δ<λ+,A2δ is non-stationary. If G𝕊(λ+\A2) is generic, then A1 remains stationary in V[G].

    Proof. Let p𝕊(λ+\A2) force that is a closed unbounded subset of λ+ with A1=. Let 𝜃>(22λ) be a regular cardinal and let Nα:α<λ+ be an internally approachable sequence of elementary substructures of H𝜃,𝜖,<𝜃,{A1,A2,p,}. Then Nαλ+:α is a limit is a closed unbounded subset of λ+ and for each such α,Nα<cof(α)Nα.

    Choose a limit δ such that Nδλ+A1. Let γ=cof(δ) and Cδ(δ\A2) be a closed unbounded set of order type γ. Build a decreasing sequence of conditions pα:α<γ such that

    p0=p

    for each β<γ, pα:α<βNδ

    if i is the αth member of Cδ, then for some ordinal ξ<δ, with i<ξ

    pα+1ξ.

    sup(pα+1)>i.

    Such a sequence is possible to build, because Nδ<γNδ.

    But then sup(α<γpα)=δ and δA2, hence

    q=α<γpα{δ}𝕊(λ+\A2).
    Moreover qA1. This contradiction establishes Lemma 6.3. □

    Our application of the following definition and the following lemmas will be with μ=λ+ for a regular λ.

    Definition 6.4. Let μ be a regular cardinal.

    (a)

    The partial ordering (μ) for adding a non-reflecting stationary subset of μ consists of functions p:α{0,1} for some α<μ and letting

    Sp={ξ<α|p(ξ)=1},
    for every limit ᾱα there is a closed unbounded set Cᾱ such that SpC=.

    (b)

    Let γ<μ be regular. The partial ordering (μ,γ) for adding a non-reflecting stationary subset of μcof(γ) consists of those conditions p(μ) which concentrate on μcof(γ):

    p(ξ)=0whenevercof(ξ)γ.

    Let γ<μ be uncountable regular cardinals and define the map

    πγ:(μ)(μ,γ)
    by setting
    πγ(p)(ξ)=p(ξ)ifcof(ξ)=γ0otherwise.

    We will use the following lemma which relates (μ) with (μ,γ).

    Lemma 6.5. Let γ<μ be uncountable regular cardinals.

    (a)

    If G is generic for (μ), then

    SG=def{ξ<μ:for somepG,p(ξ)=1}
    is a non-reflecting stationary subset of μ.

    (b)

    If H is generic for (μ,γ) 

    SH=def{ξ<μ:for somepH,p(ξ)=1}
    is a non-reflecting stationary subset of μcof(γ).

    (c)

    If G(μ) is generic over V, and H=πγG, then H is generic over V for (μ,γ). (In other words the map πγ is a projection.)

    Proof. The first two items are immediate. For the third note that for all p(μ) and all q(μ,γ) with q(μ,γ)πγ(p), there is a p(μ)p with πγ(p)(μ,γ)q. This is the standard criterion for being a projection. □

    It is an easy remark that in (a) V[G]=V[SG] and in (b) V[H]=V[SH]. For this reason, we will frequently write V[S] 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 μ is of the form λ+. For this reason we formulate the next proposition for cardinals of the form λ+, although it is true for any regular μ>ω.

    Proposition 6.6 (See [4, 5, 6]). Assume γλ where γ is regular and λ<λ=λ. Then the following hold.

    (a)

    Both (λ+) and (λ+,γ) are strategically λ+-closed. In particular, both (λ+) and (λ+,γ) preserve stationarity of stationary subsets of λ+, are (λ+,)-distributive, so they do not add any new functions f:λV, and generic extensions of V via these partial orderings agree with V on all cardinals and cofinalities λ+ and on what Hλ+ is.

    (b)

    (λ+,γ) is γ-closed but not γ+-closed.

    (c)

    If G is ((λ+,γ),V)-generic then SG={Sp|pG} is a non-reflecting stationary subset of λ+cof(γ).

    (d)

    If G is ((λ+),V)-generic then SG={Sp|pG} is a non-reflecting stationary subset of λ+ such that SG has stationary intersection with each stationary subset of λ+ that lies in V. In particular, SGλ+cof(γ) is stationary for all regular γ<λ+.

    Proof. Only (d) is not explicitly proved in the earlier literature (though it was known). Let T be a stationary subset of λ+ in V. Let G(λ+) be generic and Sλ+ 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 γλ.

    If the claim fails let p(λ+) force over V that T= where is a term for a closed unbounded subset of λ+ in V[G].

    Let 𝜃>(22λ) be a regular cardinal and let Nα:α<λ+ be an internally approachable sequence of elementary substructures of

    H𝜃,𝜖,<𝜃,{,T,}.
    Then Nαλ+:α is a limit is a closed unbounded subset of λ+ and for each such α,Nα<cof(α)Nα. Choose a limit ordinal δ such that Nδλ+T and Nδλ+=δ. Then δ has cofinality γ. Let Cδδ be closed and unbounded in δ with order type γ such that every initial segment of Cδ belongs to Nδ.

    By recursion on β build a decreasing sequence of conditions pα:α<δ in (λ+) such that

    p0=p

    for each β<γ, pα:α<βNδ

    if i is the αth member of Cδ, then for some ordinal ζ<δ, with i<ζ

    pαζ.

    sup(dom(pα))>i.

    If β is a limit ordinal, then pβ=β̄<βpβ̄ and if δβ=sup(β̄<βdom(pβ̄)), then pβ+1 forces δβS.

    Let p*=β<δpβ. Then dom(p*) has supremum δ and forces that

    δ is non-stationary

    δ is cofinal in δ

    Extending p* by one point to get a condition q that forces δ gives a condition q(λ+) that forces δT. This contradiction shows that in the extension by (λ+), S intersects every stationary T. □

    Although both partial orderings (λ+,γ) and 𝕊(S) have a low degree of closure in general, the iteration (λ+,γ)*𝕊(λ+\) 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 λ is a cardinal, γλ is regular, and  is the canonical (λ+,γ)-term for the generic non-reflecting stationary subset of λ+cof(γ). Then the composition

    (λ+,γ)*𝕊(λ+\)
    has a dense λ+-closed subset DHλ+. In particular, this two-step iteration preserves stationarity of stationary subsets of λ+.

    Proof. Let D be the collection of all (p,ċ)(λ+,γ)*𝕊(λ+\)Hλ+ such that

    {ξ:p(ξ)=0} is closed (so has successor order type), and

    pċ=č for some closed unbounded set cdom(p) with p(ξ)=0 for all ξc.

    Then D is dense in (λ+,γ)*𝕊(λ+\). For details, see [4, 5] or [6]. □

    Fix a regular cardinal λ. At successor steps in the iteration used to prove Theorem 1.5, we will use an iteration of the form

    (λ+,γ)*𝕊(λ+\)*𝕊(λ+\),(28)
    where is a term for the generic non-reflecting stationary subset of λ+cof(γ) given by (λ+,γ) and will be a term for a certain subset of λ+cof(λ). We note in passing that the realization of is a non-reflecting stationary set. Since both and lie in V(λ+,γ), the following three forcing notions are equivalent:

    Version 1: (λ+,γ)*𝕊(λ+\)*𝕊(λ+\)

    Version 2: (λ+,γ)*(𝕊(λ+\)×𝕊(λ+\))

    Version 3: (λ+,γ)*𝕊(λ+\)*𝕊(λ+\).

    Lemma 6.8. Let =(λ+,γ)*𝕊(λ+\)*𝕊(λ+\). Then  has a dense set D such that

    (i)

    D has cardinality λ+,

    (ii)

    DHλ+,

    (iii)

    D is λ-closed, and

    (iv)

    D is (λ+,)-distributive.

    Proof. Proposition 6.7 shows that (λ+,γ)*𝕊(λ+\) has a dense λ+-closed subset. Since consists of ordinals of cofinality λ, 𝕊(λ+\) is λ-closed and (λ+,)-distributive. Since is isomorphic to (λ+,γ)*𝕊(λ+\)*𝕊(λ+\), items (iii) and (iv) follow. Now (i) is immediate, since 𝕊(λ+\) has a dense set of size λ+ 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 Hλ+. By Proposition 6.7 there is a dense subset of the first two steps that lies in Hλ+ and is λ+-closed. After forcing with (λ+,γ)*𝕊(λ+\) the conditions in 𝕊(λ+\) belong to Hλ+ and can be realized by elements of V using the closure of (λ+,γ)*𝕊(λ+\). Hence, there is a dense subset of Version 3 consisting of triples (p,c,d) where each coordinate belongs to Hλ+. Rearranging, we get (ii). □

    In the iteration, we will construct as a coding tool. Let {Tξ:ξ<λ+} be a sequence of disjoint stationary subsets of λ+cof(λ). Let Sλ+ and define

    T(S)=ξST2ξξST2ξ+1.(29)
    We will use T(S) for a set S that is V-generic for (λ+). When forcing with (λ+,γ) we will use the following variant:
    Tγ(S)=ξScof(γ)T2ξξcof(γ)(λ+\S)T2ξ+1.(30)

    Given an (λ+)-generic Sλ, and a sequence of sets Sγ for each regular uncountable γλ with Sγ=Scof(γ), the following holds:

    T(S)=γTγ(Sγ)(31)
    In particular, if δT(S) then δTγ(S).

    Proposition 6.9. Suppose λ is regular and the GCH holds. Let  be the partial ordering

    (λ+)*𝕊(λ+\T())*𝕊(λ+\)
    where  is the canonical (λ+)-term for the generically added non-reflecting stationary set S and T() is the canonical (λ+)-term for the set T(S). If G is generic then in V[G]:

    (a)

    If ξG, then T2ξ is non-stationary and T2ξ+1 is stationary.

    (b)

    If ξG, then T2ξ+1 is non-stationary, and T2ξ is stationary.

    Proof. Force with (λ+) to get a generic stationary set S and let Ċ be a term for the closed unbounded set added by 𝕊(λ+\T(S)). If H𝕊(λ+\T(S)) is V[S]-generic then ĊHT(S) is empty which shows the non-stationarity claims in both (a) and (b).

    What is left is to show that the appropriate Tη’s stationarity is preserved. The argument in each case is the same, so assume we argue for case (a). Since the partial ordering (λ+)*𝕊(λ+\) has a dense <λ+-closed subset, it preserves the stationarity of each Tξ.

    Suppose that ξG. Applying Lemma 6.3 in V[G] with A1=T2ξ+1 and A2=T(S) shows that T2ξ+1 is stationary in V[G][H]. □

    Essentially, the same proof shows:

    Proposition 6.10. Suppose λ is regular, γλ is regular and uncountable, and that the GCH holds. Let  be the partial letting

    (λ+,γ)*𝕊(λ+\Tγ())*𝕊(λ+\)
    where  and Tγ() are defined as in Proposition 6.9. If G is generic then in V[G]:

    (a)

    If ξGcof(γ), then T2ξ is non-stationary and T2ξ+1 is stationary.

    (b)

    If ξcof(γ)(λ+\G), then T2ξ+1 is non-stationary, and T2ξ is stationary.

    (c)

    If ξcof(γ), then Tξ is stationary.

    The point of this coding is that using the forcing in either Proposition 6.9 or 6.10, for ξ of the appropriate cofinality we have

    ξSif and only ifT2ξis non-stationary andT2ξ+1is stationary.

    Proposition 6.11. Under the hypotheses of Proposition 6.9 (or Proposition 6.10), the set S added by (λ+) (respectively, (λ+,γ)) remains stationary after forcing with 𝕊(λ+\T()) (respectively, 𝕊(λ+\Tγ())).

    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 G(λ+) be generic and Sλ+ be the generic stationary set constructed by G. By Proposition 6.6 item (d), in V[G], S has stationary intersection with each Tξ. Choose a ξ0 such that Tξ0T(S)=. Let A1=STξ0 and A2=T(S). The A1 and A2 satisfy the hypotheses of Lemma 6.3 for the forcing 𝕊(λ+\T(S)). Hence, STξ0 is stationary in the generic extension of V[S] by 𝕊(λ+\T(S)), and so S is stationary after the forcing (λ+)*𝕊(λ+\T(). □

    6.2. The construction

    Let U be the normal measure as in (B) above and

    j:VM(32)
    be the ultrapower embedding by U where M is transitive. Let κ be the critical point of j.

    The forcing will be an iteration of length κ+2 with Easton supports. If α<κ is inaccessible we will choose a regular uncountable γα 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 Tα,ξ. The last step is to shoot a club through the complement of the stationary set S created in the first step.

    At stage κ 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 α fix the stationary sets Tα,ξ:ξ<α+ from Proposition 6.1. Fix a regular uncountable γα. For this γ, let αγ be the partial ordering

    (α+,γ)*𝕊(α+\Tα,γ(α,γ))*𝕊(α+\α,γ),(33)
    defined as in Proposition 6.10, with α in place of λ, Tα,γ in place of Tγ, and α,γ in place of . (We will often suppress γ in the notation if γ is clear from the context, and write simply α.)

    The final partial ordering * will be an iteration with Easton supports of length κ+2. We define the initial segment of length κ, κ, as follows. κ will be the direct limit of the forcing iteration

    (α|ακ)
    satisfying the following.

    FI-1

    For inaccessible α, conditions in each α are partial functions p with dom(p) contained in inaccessibles below α such that dom(p)β is bounded in β whenever βα is inaccessible.

    FI-2

    If pα and ᾱdom(p) then

    p(ᾱ)=(γp(ᾱ),wp(ᾱ))
    is an ordered pair such that
    γp(ᾱ)Rᾱ={γᾱ|γis regular uncountable},

    and wp(ᾱ)Hα+ is a ᾱ-term for a condition in the three step forcing ᾱγp(ᾱ) defined in Eq. (33).f

    The ordering on α is defined in the standard way, that is,

    FI-3

    pq if and only if the following hold :

    (1)

    dom(p)dom(q) and

    (2)

    for every ᾱdom(q) :

    (a)

    γp(ᾱ)=γq(ᾱ) and

    (b)

    pᾱᾱwp(ᾱ) extends wq(ᾱ) in ̇ᾱγp(ᾱ)

    (where, by pᾱ we mean p(dom(p)ᾱ).)

    From Lemmas 6.6 to 6.8, we conclude that

    (i)

    For all inaccessible α,αVα

    (ii)

    For α Mahlo, α is α-c.c.

    (iii)

    If G is (α,V)-generic then in V[G] the partial ordering (̇αγp(ᾱ))G contains a dense α-closed set and is (α+,)-distributive.

    (iv)

    For α<κ, if κ=α*̇κα is the canonical factorization, and G is (α,V)-generic, then

    V[G](̇κα)Ghas anα-closed dense subset”.

    (v)

    For each inaccessible α<κ, if pκ then (p(ᾱ),p(ᾱ+1),p(ᾱ+3))Hᾱ+.

    (vi)

    For all cardinals α, α+3 preserves both α and α+.

    (vii)

    κ preserves all cardinals.

    Now define a partial ordering * as the κ+2 length iteration:

    *=κ*(κ+)*𝕊(κ+\T())(34)
    where and T() are as in Proposition 6.9, with κ in place of λ.

    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, κ remains inaccessible, and for each regular uncountable γκ there is a uniform normal (κ+,)-distributive ideal 𝒥γ such that P(κ)/𝒥γ has a dense γ-closed set, but no dense γ+-closed set.

    Proof. Fix a regular uncountable cardinal γκ.

    By GCH in V, any generic extension via κ satisfies 2κ=κ+, so in any such generic extension the partial ordering (κ+) has cardinality κ+. Using the strategic closure of (κ+) we conclude that 2κ=κ+ in the generic extension via κ*(κ+). Let S be the non-reflecting stationary set added by (κ+). Then 𝕊(κ+\T(S)) has cardinality κ+ in any such generic extension. All of this combined with the distributivity properties of (κ+) and 𝕊(κ+\T(S)), shows that 2κ=κ+. Similar arguments show that

    *preserves all cardinals and cofinalities and also theGCH.(35)

    Now return to the map j from (32). Let G be (κ,V)-generic. Because card(κ)=κ and κ is κ-c.c., M[G] is closed under κ-sequences in V[G] and the models M[G],V[G] agree on what Hκ+ is. It follows that the models M[G],V[G] agree on what (κ+) and (κ+,γ) are.

    Let G=G0*G1 be ((κ+)*𝕊(κ+\T()),V[G])-generic where is as in Eq. (34). It follows that S=G0=G0.

    Let Gκ,0=πγ(G0) where πγ is as in Lemma 6.5. Then Gκ,0 is generic for (κ,γ) over both M[G] and V[G]. Let be the term for the non-reflecting stationary set coming from G0. Then κ,γ=cof(γ). Denote Sκ,γ by Sκ.

    Since (κ+) and (κ+,γ) are (κ+,)-distributive in the models where they live

    M[G,Gκ,0]is closed underκ-sequences lying inV[G,G0].(36)
    In particular, M[G,Gκ,0] and V[G,G0] agree on what Hκ+ and 𝕊(κ+\Tκ,γ(Sκ)) are.

    Let CV[G,G] be the closed unbounded subset of κ+\T(S) associated with the generic ultrafilter G1 for 𝕊(κ+\T(S)) over V[G,G0].

    Note that Tκ,γ(Sκ)M[G,Gκ,0] and CTκ,γ(Sκ)= because Tκ,γ(Sκ)T(S). From the point of view of V[G,G0] there are only κ+ many dense subsets of 𝕊(κ+\Tκ,γ(Sκ)) which are in M[G,Gκ,0].

    We can construct a (𝕊(κ+\Tκ,γ(Sκ)),M[G,Gκ,0])-generic filter Gκ,1V[G,G] as follows. In V[G,G0] fix an enumeration Dβ|β<κ+ of dense subsets of 𝕊(κ+\Tκ,γ(Sκ)) which belong to M[G,Gκ,0]. Using recursion on β<κ+ construct a descending chain cβ,cβ|β<κ+ in 𝕊(κ+\Tκ,γ(Sκ)) as follows:

    Let c0=.

    Given cβ, pick cβDβ such that cβcβ in 𝕊(κ+\Tκ,γ(Sκ)).

    Given cβ, let cβ+1=cβ{δβ+1} where δβ+1 is the least element of C larger than max(cβ).

    If β is a limit let cβ=β̄<βcβ̄{δβ} where δβ=sup{max(cβ̄)|β̄<β}.

    To see that this works, note that for every β<κ+ both cβ and cβ are elements of M[G,Gκ,0], which is verified inductively on β. The only non-trivial step in the induction is to see that cβM[G,Gκ,0] for β limit. That the sequence cβ̄:β̄<β belongs to M[G,Gκ,0] follows from the κ-closure property of M[G,Gκ,0]. For the union to be a condition requires that the supremum δ of cβ does not belong to Tκ,γ(Sκ). However by Eq. (31), since δT(S) we know that δTκ,γ(Sκ).

    Now let Gκ,1 be the filter on 𝕊(κ+\Tκ,γ(Sκ)) generated by the sequence cβ|β<κ+; it is clear that Gκ,1V[G,G] and is (𝕊(κ+\Tκ,γ(Sκ)),M[G,Gκ,0])-generic. Finally, set Gκ=Gκ,0*Gκ,1.

    We note here that by Proposition 6.11, Sκ is stationary in V[G,G]. Thus in V[G,G], Sκ is a non-reflecting stationary set.

    Consider a (𝕊(κ+\Sκ),V[G,G])-generic filter H. Then the filter Gκ*H is ((κ+,γ)*𝕊(κ+\Tκ,γ(κ))*𝕊(κ+\κ),M[G])-generic. It follows that G*Gκ*H is (j(κ)(κ+3),M)-generic. By the Factor Lemma applied inside M[G,Gκ,H], the quotient j(κ)/G*Gκ*H is isomorphic to the iteration j(κ)κ+3 as calculated in M[G,Gκ,H]. Let μ be the least inaccessible of M above κ. Using (iv) in the list of the properties of the iteration stated below FI-3, we conclude that M[G,Gκ,H] satisfies the following:

    j(κ)/G*Gκ*Hhas a denseμ-closed subset.(37)
    Since (κ+)*𝕊(κ+\T(κ))*𝕊(κ+\κ) is (κ+,)-distributive in V[G],
    M[G,Gκ,H]is closed underκ-sequences inV[G,G,H].(38)
    Working in V[G,G,H]: since the cardinality of j(κ)κ+3 is κ+, we have an enumeration Dβ|β<κ+ of all dense subsets of j(κ)/G*Gκ*H which are in M[G,Gκ,H]. Using (iv) in the list of the properties of the iteration stated below FI-3, the sentences labeled (37), (38) above and the fact that μ>κ+, we can construct a descending sequence pβ|β<κ+ with each proper initial segment being an element of M[G,Gκ,H] and such that pβDβ for all β<κ+. Let K1 be the filter on j(κ)/G*Gκ*H generated by this sequence. Then K1 is (j(κ)/G*Gκ*H,M[G,Gκ,H])-generic and K1V[G,G,H]. Let K=G*Gκ*H*K1. Then K can be viewed as a (j(κ),M)-generic filter, so we can extend j to an elementary embedding
    jH,K:V[G]M[K]

    defined by setting jH,K(G)=j()K whenever V is a κ-term. Since K1 can be constructed inside V[G,G,H], there is a 𝕊(κ+\Sκ)-term K1̇V[G,G] such that K̇1H is (j(κ)/G*Gκ*H,M[G,Gκ,H])-generic whenever H is (𝕊(κ+\Sκ),V[G,G])-generic. In particular, there is a M-generic KHj() determined by forcing over V[G,Gκ] to get a generic H𝕊(κ+\Sκ).

    Changing notation slightly to emphasize the dependence on H, define jH be as follows:

    jH=jH,K̇H:V[G]M[K̇H].(39)

    We also have a 𝕊(κ+\Sκ)-term U̇V[G,G] such that U̇H is the normal V[G]-measure over κ derived from jH. That is,

    U̇H={xP(κ)V[G]|κjH(x)}(40)
    whenever H is a (𝕊(κ+\Sκ),V[G,G])-generic filter. It is a standard fact that
    M[K̇H]=Ult(V[G],U̇H)andjH:V[G]M[K̇H]is the associated ultrapower map.(41)
    Since the composition (κ+)*𝕊(κ+\T(κ))*𝕊(κ+\Sκ̇) is (κ+,)-distributive in V[G], the models V[G] and V[G,G] agree on what P(κ) is, so U̇H is also a normal V[G,G,H]-measure over κ. Since U̇HV[G,G,H] we record that
    κis measurable inV[G,G,H].(42)

    We now define the ideal 𝒥γ on P(κ) in V[G,G]. For every xP(κ)V[G,G],

    x𝒥γ:𝕊(κ+\Sκ)V[G,G]x̌U̇.(43)

    Note that this definition takes place in V[G,G] so 𝒥γV[G,G] and standard arguments show that 𝒥γ is a uniform normal ideal on P(κ) in V[G,G].

    Recall that Sκκ+cof(γ) where γ was fixed at the in V[G]. This is crucial for determining the closure properties of P(κ)/𝒥γ. The main tool for analyzing properties of 𝒥γ is the duality theory developed in [9]. Rather than simply citing theorems there, we show the following proposition:

    Proposition 6.13. In V[G,G] there is a dense

    e:𝕊(κ+\Sκ)P(κ)/𝒥γ.
     □

    Proof. In V, fix an assignment xfx where xM and fx:κV is such that

    x=[fx]U=j(fx)(κ).(44)
    The partial ordering 𝕊(κ+\Sκ) in the generic extension M[G,Gκ] can be viewed as the quotient (j(κ)(κ+1))/G*Gκ, so we can consider conditions in 𝕊(κ+\Sκ) as elements of M that are ordered the same way as conditions in j(κ). Hence each such condition p is represented in the ultrapower by U by the function fp.

    Next, recall that at each inaccessible α<κ, stages α,α+1 and α+2 of κ are a composition of three partial orderings where the last one is 𝕊(α+\Sα). The α+1,α+2,α+3 components of the generic filter G are then of the form Gα,0*Gα,1*h(α) where h(α) is (𝕊(α+\Sα),V[Gα*Gα,0*Gα,1])-generic. The function h is thus an element of V[G] and represents the filter H in the ultrapower by U̇H, that is, H=jH(h)(κ); see (41).

    Then for any p𝕊(κ+\Sκ) we have the following:

    pHjH(fp)(κ)jH(h)(κ)ap=def{α<κ|fp(α)h(α)}U̇H.(45)
    We show that in V[G,G], the map e:𝕊(κ+\Sκ)P(κ)/𝒥γ defined by
    e(p)=[ap]𝒥γ(46)
    is a dense embedding. The proof is a standard variant of the duality argument, which we include for the reader’s convenience. We write briefly [a] for [a]𝒥γ.

    To see that e is order-preserving, consider pq in 𝕊(κ+\Sκ). By the above remarks on the ordering of the quotient, we have pq in j(κ), hence j(fp)(κ)j(fq)(κ) in j(κ). It follows that

    bp,q=def{ξ<κ|fp(ξ)fq(ξ)}U,
    and so bp,qU̇H whenever H is a (𝕊(κ+\Sκ),V[G,G])-generic filter. It follows that κ\bp,q𝒥γ. Since ap\aqκ\bp,q, we have [ap]𝒥γ[aq].

    To see that the map e is incompatibility preserving, we prove the contrapositive. Assume p,q𝕊(κ+\Sκ) are such that apaq𝒥γ+. It follows that there is some (𝕊(κ+\Sκ),V[G,G])-generic filter H such that apaqU̇H. Then apU̇H and aqU̇H. Using (45) we conclude that p,qH. Hence p,q are compatible.

    To see that the range of e is dense, assume that a𝒥γ+. It follows that there is some (𝕊(κ+\Sκ),V[G,G])-generic filter H such that aU̇H. So there is some pH such that

    p𝕊(κ+\Sκ)V[G,G]ǎU̇.(47)
    Now for every (𝕊(κ+\Sκ),V[G,G])-generic filter H we have
    apU̇HpHaU̇H.
    Here the first implication follows from (45) and the second implication from (47). We thus conclude that ap\aU̇H whenever H is a (𝕊(κ+\Sκ),V[G,G])-generic filter, which means that ap\a𝒥γ, or equivalently, [ap]𝒥γ[a]. □

    We can now complete the proof of Proposition 6.12 by looking at the properties of the partial ordering 𝕊(κ+\Sκ) in V[G,G]. By Proposition 6.11, Sκ is stationary in V[G,G], so 𝕊(κ+\Sκ) is a standard forcing for killing a non-reflecting stationary subset of κ+. The (κ+,)-distributivity follows from Proposition 6.2(a). The existence of a dense γ-closed set as well as the non-existence of a dense γ+-closed set follows from Proposition 6.2(b) and the fact that Sκκ+cof(γ).

    The last major step toward the proof of Theorem 1.5 is the following proposition.

    Proposition 6.14. κ does not carry a saturated ideal in a generic extension via *.

    Proof. Assume for a contradiction that κ does carry a saturated ideal in V[G,G] where G,G are as above. Denote this ideal by , and let L be a (,V[G,G])-generic filter where is the partial ordering (+,) and

    j:V[G,G]N
    be the generic embedding associated with the ultrapower Ult(V[G,G],L). Letting M=j(V) and (K,K)=j(G,G), we have N=M[K,K]. The partial ordering ** preserves κ+, which allows us to refer to (D) at the beginning of this section. It follows that the models V,M and all transitive extensions of these models which are contained in V[G,G,L] have a common cardinal successor of κ, which we denote by κ+.

    Now look at the κth step of the iteration j(κ). Obviously j(κ)κ=κ and Kκ=G. Let γRκM=Rκ be the ordinal chosen by the generic filter K at step κ of the iteration j(κ) (see FI-2). Then steps κ,κ+1 and κ+2 are thus forcing with

    (κ+,γ)*𝕊(κ+\T(κ))*𝕊(κ+\κ)
    over M[G]. This composition of partial orderings is computed the same way in M[G] and V[G], as by (D) at the beginning of this section, the models V and M agree on what Hκ+ is, but we don’t use this directly. What is relevant is the agreement of the models on what κ+ is, along with the fact that Tκ,ξ=Tκ,ξ for all ξ<κ+ where the sets Tκ,ξ and Tκ,ξ are as in (D) quoted above.

    The κth component Kκ of K has the form Kκ,0*Kκ,1*Kκ,2. Let Sκ be the generic non-reflecting stationary subset of κ+cof(γ) added by Kκ,0 over M[G]. Since Kκ,2M[K]V[G,G,L] is a closed unbounded subset of κ+ disjoint from Sκ, the set Sκ is non-stationary in V[G,G,L].

    By elementarity, the generic filter Kκ,1 codes the set Sκ inside M[K] as follows. Given an ordinal ξκ+cof(γ),

    ξSκTκ,2ξ+1is stationary andTκ,2ξis non-stationary.
    By the agreement Tκ,ξ=Tκ,ξ coming from (D) and mentioned above
    ξSκTκ,2ξ+1is stationary andTκ,2ξis non-stationary
    for all such ξ. Recall that S is the subset of κ+ with characteristic function G0, and the generic filter G1 codes S in V[G,G] the same way as the generic filter Kκ,1 codes the set Sκ inside M[K], that is,
    ξSTκ,2ξ+1is stationary andTκ,2ξisnon-stationary
    whenever ξ<κ+. It follows that for every ξκ+cof(γ),
    ξSκTκ,2ξis non-stationary inM[K]Tκ,2ξis non-stationary inV[G,G,L]Tκ,2ξis non-stationary andTκ,2ξ+1is stationary inV[G,G,L]ξS
    Here the third implication follows from the fact that in V[G,G], if ξ<κ+ then exactly one of Tκ,2ξ, Tκ,2ξ+1 is stationary. As is κ+-c.c., for each ξ<κ+ exactly one of Tκ,2ξ, Tκ,2ξ+1 is stationary in V[G,G,L], namely, the one which is stationary in V[G,G]. Similarly we verify the implication ξSκξS whenever ξκ+cof(γ). Altogether we then conclude that Sκ=Scof(γ). But then, by Proposition 6.11, Sκ is stationary in V[G,G]. Then, again by the κ+-c.c. of , Sκ remains stationary in V[G,G,L], a contradiction. □

    Finally, we give a proof of incompatibility of strategies 𝒮γ from Corollary 1.6(a), as formulated at the end of Corollary 1.6.

    The point here is that in the construction of 𝒥γ, the ordinal γ at the κth stage in j(κ) is chosen before the generic filter H comes into play. Therefore, the set xγ is defined by

    xγ={α<κ|γp(α)=γfor some/allpGwithαdom(p)}ifγ<κ
    and
    xγ={α<κ|γp(α)=αfor some/allpGwithαdom(p)}ifγ=κ
    is an element of U̇H for all (𝕊(κ+\Sκ),V[G,G])-generic filters H, hence xγ is in the filter dual to 𝒥γ. Now if Player I plays 𝒜0 such that xγ,xγ𝒜0 and Player II responds with U0 according to 𝒮γ then xγU0, as U0=W𝒜0 for some (𝒥γ,V[G,G])-generic filter W. Similarly as above, xγU0 for the response U0 of 𝒮γ to 𝒜0. Since xγxγ=, we have U0U0.

    Remark 6.15. We could do the construction without the “lottery” aspect, aiming at a single γ. Indeed that works for that γ, but leaves open the problem of whether ideals exist with dense trees of height γ for γγ and for which γ 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 κ. 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 𝒥γ one can build a winning strategy 𝒮γ* for Player II in 𝒢γ. In turn, 𝒮γ* can used to build the ideal γ with the methods in Theorems 1.1 and 1.2

    𝒥γ𝒮γ*γ.
    Inspection of the proof shows that 𝒥γγ. Is there anything else one can say? For example, are the two ideals equal?

    An Ulam Game. Consider the following variant of the cut-and-choose game of length ω derived from games introduced by Ulam in [23] (see [15]).g

    I(A00,A10)(A01,A11)(A0n,A1n)(A0n+1,A1n+1)
    IIB0B1BnBn+1

    At stage 0, Player I plays a partition (A00,A10) of κ. At stage n0, Player II lets Bn be either A0n or A1n, and plays Bn. At stage n1, Player I plays a partition (A0n+1,A1n+1) of Bn. The winning condition for Player II is that |nωBn|2.

    These games generalize to lengths γ>ω as follows:

    (1)

    At successor stages α+1, Player I partitions Bα into two pieces and Player II chooses one of the pieces.

    (2)

    At limit stages α, let Bα=β<αBβ and then Player I partitions Bα 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 𝒢ω*, 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 𝒢ω* against the Boolean Algebras 𝒜n generated by {A0i,A1i:in}. In the game 𝒢ω*, she then plays as Bn whichever of A0n or A1n belongs to Un. By the winning condition on Gω*, nωBn belongs to a κ-complete, uniform filter. Hence |nBn|=κ>1.

    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 𝒢ω*” 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 ω2 by the Lévy collapse and is the ideal generated by the original normal measure on κ, then in the extension P(ω2)/ 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 κ such that for all γ<κ, 2γκ. At successor cardinals κ 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 κ, the Keisler–Tarski Theorem shows player I has a winning strategy in the game of length 1. But if κ is weakly compact, Player II has a winning strategy in the game of length ω.

    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 κ is weakly compact and that Player II has a winning strategy in the Ulam game of length γ (for γω), does Player II have a winning strategy in 𝒢γ*?

    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 ω. By work of Nielsen and Welch if II has a winning strategy in the game of length ω1, 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 γ, 𝒢γW 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 ZFC+GCH 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 Pκ(λ) 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 Pκ(λ) 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 κ-filter U on an Nα to its normal derivative U*. Normality presents an obstacle for Pκ(λ) 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 λ<ω one might consider games where Player I plays elements of λ<ω and sequences of κ-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 P2(κ) 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 ω-closed densely treed ideals on ω2 (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 P0 extends P then iP0=iP” should be replaced with “By the minimality of P, if P extends P then iP=iP”. This mistake has been corrected as of December 11, 2023.

    Notes

    a We could omit “uniform” and simply require 𝒜0 to include the co-<κ subsets of κ and U0 to extend the co-<κ-filter. As noted in Sec. 2, if κ 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 𝒥γ for these. The second is the new technique of hopeless ideals, built in Theorems 1.1 and 1.2 from the strategies 𝒮γ. These will be denoted by γ 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 (𝒯)α for level α of 𝒯.

    f We can view wp(ᾱ) as a triple (wp(ᾱ),wp(ᾱ+1),wp(ᾱ+2)) but the notation wp(ᾱ) is frequently more convenient.

    g Velickovic [24] calls these Mycielski games.