Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-28T00:00:26.806Z Has data issue: false hasContentIssue false

STATIONARY REFLECTION AND THE FAILURE OF THE SCH

Part of: Set theory

Published online by Cambridge University Press:  27 October 2023

OMER BEN-NERIA
Affiliation:
EINSTEIN INSTITUTE OF MATHEMATICS THE HEBREW UNIVERSITY OF JERUSALEM JERUSALEM 91904, ISRAEL E-mail: omer.bn@mail.huji.ac.il
YAIR HAYUT*
Affiliation:
EINSTEIN INSTITUTE OF MATHEMATICS THE HEBREW UNIVERSITY OF JERUSALEM JERUSALEM 91904, ISRAEL
SPENCER UNGER
Affiliation:
DEPARTMENT OF MATHEMATICS UNIVERSITY OF TORONTO TORONTO, ON, CANADA E-mail: unger.the.aronszajn.trees@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

In this paper we prove that from large cardinals it is consistent that there is a singular strong limit cardinal $\nu $ such that the singular cardinal hypothesis fails at $\nu $ and every collection of fewer than $\operatorname {\mathrm {cf}}(\nu )$ stationary subsets of $\nu ^{+}$ reflects simultaneously. For $\operatorname {\mathrm {cf}}(\nu )> \omega $, this situation was not previously known to be consistent. Using different methods, we reduce the upper bound on the consistency strength of this situation for $\operatorname {\mathrm {cf}}(\nu ) = \omega $ to below a single partially supercompact cardinal. The previous upper bound of infinitely many supercompact cardinals was due to Sharon.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

We study stationary reflection at successors of singular cardinals and its connection with cardinal arithmetic. We start by recalling some basic definitions. For an ordinal $\delta $ and a set $S \subseteq \delta $ , we say that S is stationary if it meets every closed and unbounded subset of $\delta $ . If $\{S_i \mid i \in I\}$ is a collection of stationary subsets of a regular cardinal $\kappa $ , then we say that $\{S_i \mid i \in I\}$ reflects simultaneously if there is an ordinal $\delta $ such that $S_i \cap \delta $ is stationary for all $i \in I$ .

The consistency of stationary reflection at the successor of singular cardinal is already complex in the context of the generalized continuum hypothesis (GCH). A theorem of Magidor [Reference Magidor14] shows that it is consistent relative to the existence of infinitely many supercompact cardinals that every finite collection of stationary subsets of $\aleph _{\omega +1}$ reflects. Recently, the second and third authors [Reference Hayut and Unger13] were able to show the same result from an assumption below the existence of a cardinal $\kappa $ which is $\kappa ^{+}$ -supercompact. Both of these models satisfy GCH. Combining stationary reflection at the successor of a singular cardinal with the failure of $\operatorname {\mathrm {SCH}}$ presents additional difficulties.

For a singular cardinal $\nu $ , the singular cardinal hypothesis ( $\operatorname {\mathrm {SCH}}$ ) at $\nu $ is the assertion that if $\nu $ is strong limit, then $2^\nu = \nu ^{+}$ . The failure of the singular cardinal hypothesis at a singular cardinal $\nu $ is known to imply the existence of many nonreflecting objects. For instance, Foreman and Todorcevic [Reference Foreman and Todorcevic4] have shown that the failure of the singular cardinal hypothesis at $\nu $ implies that there are two stationary subsets of $[\nu ^{+}]^\omega $ which do not reflect simultaneously. This was improved by Shelah [Reference Shelah22] to obtain a single stationary subset of $[\nu ^{+}]^\omega $ which does not reflect. Reflection for stationary subsets of $[\nu ^{+}]^\omega $ has a different character than reflection for stationary subsets of ordinals.

In [Reference Poveda, Rinot and Sinapova19, Corollary 5.4], Poveda, Rinot, and Sinapova showed that the failure of $\operatorname {\mathrm {SCH}}$ at cardinal $\nu $ implies the failure of simultaneous reflection for collections of $\operatorname {\mathrm {cf}}\, (\nu )$ many stationary subsets of $\nu ^{+}$ . In his PhD thesis from 2005, Sharon [Reference Sharon21] proved that relative to the existence of infinitely many supercompact cardinals it is consistent that there is a singular cardinal $\nu $ of cofinality $\omega $ such that $\operatorname {\mathrm {SCH}}$ fails at $\nu $ and every stationary subset of $\nu ^{+}$ reflects. Sharon’s method is a tour de force construction, which builds on Gitik’s long extenders forcing [Reference Gitik6] for $\omega $ -sequence of hypermeasurable cardinals. As such, the construction does not extend to singular cardinals of uncountable cofinalities, and the question of whether the failure of $\operatorname {\mathrm {SCH}}$ at singular $\kappa $ of uncountable cofinality together with stationary reflection at $\kappa ^{+}$ is at all consistent.

This paper follows a study by the authors on stationary reflection at successors of singular cardinals at which $\operatorname {\mathrm {SCH}}$ fails. This study was prompted by two other recent studies. First, the work of second and third authors in [Reference Hayut and Unger13] on stationary reflection in Prikry forcing extensions from subcompactness assumptions. The arguments of [Reference Hayut and Unger13] show how to examine the stationary reflection in the extension by Prikry type forcing by studying suitable iterated ultrapowers of V. This approach and method has been highly effective in our situation and we follow it here. The second study is Gitik’s recent work [Reference Gitik8] for blowing up the power of a singular cardinal using a Mitchell order increasing sequence of overlapping extenders. The new forcing machinery of [Reference Gitik8] gives new models combining the failure of $\operatorname {\mathrm {SCH}}$ with reflection properties at successors of singulars. Moreover the arguments are uniform in the choice of cofinality. For example, Gitik has shown that in a related model the tree property holds at $\kappa ^{+}$ [Reference Gitik7] and that for all $\delta < \kappa $ there is a stationary subset of $\kappa ^{+}$ of ordinals of cofinality greater than $\delta $ which is not a member of $I[\kappa ^{+}]$ [Reference Gitik9].

Our first main theorem shows how to construct a model in which there is a singular cardinal $\nu $ such that stationary reflection holds at $\nu^+$ , SCH fails at $\nu $ and the cofinality of $\nu $ can be chosen in advance.

Theorem 1.1. Let $\eta < \lambda $ be regular cardinals. Suppose that there is an increasing sequence of cardinals $\langle \kappa _\alpha \mid \alpha < \eta \rangle $ with:

  1. (1) $\eta < \kappa _0$ ,

  2. (2) for each $\alpha < \eta $ , $\kappa _\alpha $ carries a $(\kappa _\alpha ,\lambda )$ -extender $E_\alpha $ and there is a supercompact cardinal between $\sup _{\beta <\alpha }\kappa _\beta $ and $\kappa _\alpha $ , and

  3. (3) the sequence $\langle E_{\alpha } \mid \alpha < \eta \rangle $ is Mitchell order increasing and coherent.

Then, there is a cardinal and cofinality preserving extension in which, setting $\kappa = \sup _{\alpha < \eta }\kappa _\alpha $ , $2^{\kappa } = \lambda $ and every collection of fewer than $\eta $ stationary subsets of $\kappa ^{+}$ reflects.Footnote 1

We have the following improvement of Gitik’s result about $I[\kappa ^{+}]$ .

Theorem 1.2. Let $\langle \kappa _\alpha \mid \alpha < \eta \rangle $ be as in Theorem 1.1 with the exception that we only assume there is a single supercompact cardinal $\theta <\kappa _0$ . There is a cardinal and cofinality preserving extension in which setting $\kappa = \sup _{\alpha <\eta }\kappa _\alpha $ , we have that $\kappa $ is strong limit, $2^\kappa =\lambda $ and there is a scale of length $\kappa ^{+}$ such that the set of nongood points of cofinality less than $\theta $ is stationary.

After a suitable preparation, the extension is obtained using the same forcing as in Theorem 1.1. Standard arguments show that in this model, the set of nongood points is stationary in cofinalities that are arbitrarily high below $\theta $ . Further, the same argument shows that we can take $\theta $ between $\sup _{\alpha <\beta }\kappa _\alpha $ and $\kappa _\beta $ and reach the same conclusion.

These two results continue a long line of research about the interaction between the failure of the singular cardinal hypothesis and weak square-like principles [Reference Gitik and Sharon11, Reference Neeman18, Reference Sinapova23Reference Sinapova and Unger26].

We do not know whether these results can also be obtained at small singular cardinals of uncountable cofinalities. For concreteness we suggest the following question.

Question 1.3. Is it consistent that $\operatorname {\mathrm {SCH}}$ fails $\aleph _{\omega _1}$ and every stationary subset of $\aleph _{\omega _1+1}$ reflects?

We also give another model for stationary reflection at $\nu ^{+}$ where $\nu $ is a singular cardinal of cofinality $\omega $ where the singular cardinal hypothesis fails. This construction replaces the supercompactness assumption in Sharon’s result and Theorem 1.1 with the weaker one of subcompactness together with hypermeasurability.

Theorem 1.4. Suppose that $\kappa $ is $\kappa ^{+}$ - $\Pi _1^1$ -subcompact and carries a $(\kappa ,\kappa ^{++})$ -extender. There is a forcing extension in which $\kappa $ is singular strong limit of cofinality $\omega $ , $2^\kappa =\kappa ^{++}$ and every finite collection of stationary subsets of $\kappa ^{+}$ reflects simultaneously.

In Theorem 6.22, we will also show how to collapse cardinals and make $\kappa $ in Theorem 1.4 to be $\aleph _{\omega }$ .Footnote 2

The construction and proof follows the lines of the second and third authors’ paper [Reference Hayut and Unger13], and the work of Merimovich [Reference Merimovich15] on generating generics for extender based forcing over iterated ultrapowers. The large cardinal assumption in the theorem is the natural combination of the assumption from [Reference Hayut and Unger13] and an assumption sufficient to get the failure of $\operatorname {\mathrm {SCH}}$ by extender based forcing. Unfortunately, we are unable to adapt the argument from the previous theorem to a singular cardinal of uncountable cofinality. We ask:

Question 1.5. It is possible to obtain the result of Theorem 1.1 without any supercompactness assumptions? Namely, is it possible to start with a single partially supercompact cardinal $\kappa $ and force $\kappa $ to be a strong limit singular of uncountable cofinality such that every stationary subset of $\kappa ^{+}$ reflects? Is it possible to obtain that together with the failure of $\operatorname {\mathrm {SCH}}$ at $\kappa $ ?

The paper is organized as follows. In Section 2, we define Gitik’s forcing from [Reference Gitik8], which will be used in our main theorem. In Section 3, we prove that in mild generic extensions of V we can find a generic for Gitik’s forcing over a suitable iterated ultrapower. In Section 4, we give the proof of the main theorem by arguing that stationary reflection holds in the generic extension of the iterated ultrapower constructed in the previous section. In Section 5, we prove Theorem 1.2. In Section 6, we give the proof of Theorem 1.4 and its variation for $\aleph _\omega $ .

2 Gitik’s forcing

In this section we give a presentation of Gitik’s forcing [Reference Gitik8] for blowing up the power of a singular cardinal with a Mitchell order increasing sequence of extenders.

Let us start with the following definitions:

Definition 2.1. Let $E_0, E_1$ be $(\kappa _0, \lambda _0)$ and $(\kappa _1, \lambda _1)$ -extenders respectively. We say that $E_0$ is less than $E_1$ in the Mitchell order, or $E_0 \trianglelefteq E_1$ , if $E_0 \in \operatorname {Ult}(V, E_1)$ . We say that $E_0$ coheres with $E_1$ if $j_{E_1}(E_0) \restriction \lambda _0 = E_0$ .

The existence of a long Mitchell increasing and pairwise coherent sequence of extenders $E_i$ , where $E_i$ is a $(\kappa _i, \lambda )$ -extender, follows from the existence of superstrong cardinal or even weaker large cardinal axioms.

Before we begin with the definition of the forcing, let us show a few basic facts about extenders and Mitchell order. We recall the notion of width from [Reference Cummings2].

Definition 2.2. Let $k\colon M \to N$ be an elementary embedding between transitive models of set theory and let $\mu $ be an ordinal. We say the embedding k has width ${\leq}\mu $ if every element of N is of the form $k(f)(a)$ for some $f \in M$ and $a \in N$ such that $M \models \vert \operatorname {\mathrm {dom}}(f)\vert \leq \mu $ .

Lemma 2.3. Let E be a $(\kappa ,\lambda )$ -extender and let $\leq _E$ be the Rudin–Keisler order of the extender E. Let $k\colon V \to M$ be an elementary embedding with width ${\leq}\kappa $ . Then the set $k " \operatorname {\mathrm {dom}} E$ is $\leq _{k(E)}$ -cofinal in $\operatorname {\mathrm {dom}} k(E)$ .

Proof By [Reference Gitik6], the Rudin–Keisler order $\leq _E$ is $\kappa ^{+}$ -directed.

Let $a \in \operatorname {\mathrm {dom}} k(E)$ . Then, by the definition of width, there is a function $f\colon \kappa \to \operatorname {\mathrm {dom}} E$ and some generator b such that $k(f)(b) = a$ . Let $a'$ be a $\leq _E$ upper bound of the image of f. Then, clearly, $k(a')$ is $\leq _{k(E)}$ above a.

Lemma 2.4. Let $E_0$ be a $(\kappa _0, \lambda )$ -extender and let $E_1$ be a $(\kappa _1,\lambda )$ -extender, where $\kappa _0 \leq \kappa _1$ . Let us assume that $E_0 \trianglelefteq E_1$ . Then the following diagram commutes:

where each arrow represents the ultrapower map using the indicated extender.

Proof First, since $E_0\in M_1$ , all maps are internally defined and in particular, all models are well founded.

Moreover,

$$\begin{align*}\left(j_{E_0}\right)^V \restriction M_1 = \left(j_{E_0}\right)^{M_1}.\end{align*}$$

In order to verify this, it is sufficient to show that those maps are the same on sets of ordinals. First, let us verify that they agree of the ordinals. Indeed, let us consider the class:

$$\begin{align*}\{(f,a) \mid f \colon \kappa_0^{<\omega} \to \mathrm{Ord}\},\end{align*}$$

ordered using the extender $E_0$ :

$$\begin{align*}(f_0,a_0) \leq (f_1, a_1) \iff j_{E_0}^V(f_0)(a_0) \leq j_{E_0}^V(f_1)(a_1).\end{align*}$$

Using the combinatorial definition of the extender ultrapower, we conclude that $M_1$ can compute this order correctly, and in particular, it computes correctly the ordertype of the elements below any constant function, or indeed below any element of the form $[f,a]$ that represents an ordinal. Indeed, as the ultrapower by $E_1$ is closed under $\kappa _0$ -sequences, it contains the above class. Since $E_0 \trianglelefteq E_1$ , $M_1$ contains the relevant measures and Rudin–Keisler projections.

For an arbitrary set of ordinals, y, one can easily compute whether $[f,a] \in [c_y, \{\kappa_0\}]$ using the combinatorial information of the extender.

Let us consider an element of $\operatorname {Ult}(M_1, E_0)$ . By the definition, it has the form: $x = j_{E_0}(g)(a_0)$ where $g\colon \kappa _0^{<\omega } \to M_1$ . Going backwards, we can find a function in V, f, and a generator $a_1$ such that for every z, $j_{E_1}(f)(z, a_1) = g(y)$ and therefore

$$\begin{align*}x = j_{E_0}\left( j_{E_1}(f)( - , a_1) \right)(a_0) = j_{j_{E_0}(E_1)} \big( j_{E_0} (f)\big)\left(a_0, j_{E_0}(a_1)\right).\end{align*}$$

This element in obviously in $\operatorname {Ult}(M_0, j_{E_0}(E_1))$ .

On the other hand, if $y \in \operatorname {Ult}(M_0, j_{E_0}(E_1))$ then there is some generator $a_1^{\prime } \in \operatorname {\mathrm {dom}} j_{E_0}(E_1)$ and a function g such that $y = j_{j_{E_0}(E_1)}(g)(a_1^{\prime })$ . By Lemma 2.3, we may assume that $a_1^{\prime } = j_{E_0}(a_1)$ . Let f be a function in V and let $a_0$ be some generator such that $g = j_{E_0}(f)(a_0)$ , then we have

$$\begin{align*}\begin{matrix} y & = & j_{j_{E_0}(E_1)}\left(j_{E_0}(f)(a_0, - )\right)(j_{E_0}(a_1)) \\ & = & j_{j_{E_0}(E_1)}\left(j_{E_0}\left(f(-, a_1)\right)(a_0)\right) \\ & = & j_{E_0}\big(j_{E_1}(f)( - , a_1)\big)(a_0), \end{matrix} \end{align*}$$

as wanted.

Let $\langle \kappa _\alpha \mid \alpha < \eta \rangle $ be a sequence of cardinals as in Theorem 1.1. Following Merimovich (see for example [Reference Merimovich15]), we can assume that the extenders $E_\alpha $ are of the form $\langle E_\alpha (d) \mid d \in [\lambda ]^{\kappa _\alpha } \rangle $ where for $X \subseteq \{f\mid \operatorname {\mathrm {dom}} f \subseteq d,\, \operatorname {\mathrm {range}} f \subseteq \kappa _\alpha ,\, |f| < \kappa _\alpha \}$ , $X \in E_\alpha (d)$ if and only if $\{ (j_{E_\alpha }(\xi ),\xi ) \mid \xi \in d \} \in j_{E_\alpha }(X)$ .

For $d\in [\lambda ]^{\kappa _\alpha }$ with $\kappa _\alpha \in d$ , it is easy to see that the measure $E_\alpha (d)$ concentrates on a set of order preserving functions $\nu $ with $\kappa \in \operatorname {\mathrm {dom}}(\nu )$ . So we may assume that the elements of every measure one set mentioned below are of this form. We also fix functions $\langle \ell _\alpha \mid \alpha < \eta \rangle $ so that for every $\alpha < \eta $ , $j_{E_\alpha }(\ell _\alpha )(\kappa _\alpha ) = \lambda $ . The existence of such functions can always be arranged by a simple preparatory forcing.

Using the set $d \in [\lambda ]^{\kappa _\alpha }$ to index the extenders $E_\alpha $ has the advantage that the projection maps from $E_\alpha (d')$ and $E_\alpha (d)$ for $d \subseteq d'$ can be made very explicit. The measure $E_\alpha (d^{\prime })$ concentrates on partial functions from $d'$ to $\kappa _\alpha $ with domain smaller than $\kappa _\alpha $ . Thus, the map $\nu \to \nu \restriction d$ is a projection from $E_\alpha (d^{\prime })$ to $E_\alpha (d)$ .

For every two cardinals $\mu < \lambda $ , let $\mathcal {A}(\lambda ,\mu )$ be the poset consisting of partial functions $f \colon \lambda \to \mu $ with $|f| \leq \mu $ and $\mu \in \operatorname {\mathrm {dom}}(f)$ . Therefore $\mathcal {A}(\lambda ,\mu )$ is isomorphic to Cohen forcing for adding $\lambda $ many subsets of $\mu ^{+}$ .

We let $\mathbb {P}$ be Gitik’s forcing from [Reference Gitik8] defined from the sequence of extenders $\langle E_\alpha \mid \alpha < \eta \rangle $ . We give a compact description of the forcing. A condition $p \in \mathbb {P}$ is a sequence $\langle p_\alpha \mid \alpha < \eta \rangle $ such that there is a finite set $s^p \subseteq \eta $ such that for each $\alpha < \eta $ , $p_\alpha = (f_\alpha ,\lambda _\alpha )$ if $\alpha \in s^p$ , and $p_\alpha = (f_\alpha ,A_\alpha )$ otherwise, and the following conditions hold.

  1. (1) $f_\alpha \in \mathcal {A}(\lambda _{\alpha ^*},\kappa _\alpha )$ where $\alpha ^*$ is the next element of $s^p$ above $\alpha $ if it exists and $f_\alpha \in \mathcal {A}(\lambda ,\kappa _\alpha )$ otherwise.

  2. (2) For all $\alpha \in s^p$ , $\lambda _\alpha $ is a cardinal and $\sup _{\beta < \alpha }\kappa _\beta < \lambda _\alpha < \kappa _\alpha $ .

  3. (3) For all $\alpha \in \eta \setminus s^p$ , if $\alpha> \max (s^p)$ , then $A_\alpha \in E_\alpha (\operatorname {\mathrm {dom}}(f_\alpha ))$ , otherwise if $\alpha ^*$ is the least element of $s^p$ above $\alpha $ , then $A_\alpha \in E_\alpha (\operatorname {\mathrm {dom}}(f_\alpha ))$ .

  4. (4) For $\alpha \notin s^p$ , $f_\alpha ^p(\kappa _\alpha ) =0$ .Footnote 3

  5. (5) The sequence $\langle \operatorname {\mathrm {dom}}(f_\alpha ) \mid \alpha < \eta \rangle $ is increasing.

We adopt the convention of adding a superscript $f_\alpha ^p$ , $A_\alpha ^p$ , etc. to indicate that each component belongs to p. When the value of $\lambda _{\alpha ^{*}}$ might behave non-trivially, we will add it as a third coordinate to the pairs $p_\alpha $ , where $\alpha \notin s^p$ . We call $\eta \setminus s^p$ the pure part of p and $s^p$ the non-pure part of p.

We briefly sketch the notion of extension. A condition p is a direct extension of q if $s^p = s^q$ and for all $\alpha $ , $f_\alpha ^p$ is stronger than $f_\alpha ^q$ and $A_\alpha ^p$ projects to a subset of $A_\alpha ^q$ using the natural Rudin–Keisler projection from $E_\alpha (\operatorname {\mathrm {dom}} f_\alpha ^p)$ to $E_\alpha (\operatorname {\mathrm {dom}} f_\alpha ^q)$ .

Let us describe now the one point extension. Suppose that $\nu \in A_\beta ^p$ . We let $q = p \frown \nu $ be the condition with $s^q = s^p \cup \{\beta \}$ and the following.

  1. (1) $f_\beta ^q = f_\beta ^p {}^\frown \nu $ is the overwriting of $f_\beta ^p$ by $\nu $ , that is,

    $$\begin{align*}(f_\beta^p{}^\frown \nu)(\tau) = \begin{cases} \nu(\tau), \mbox{ if } \tau \in \operatorname{\mathrm{dom}}(\nu),\\ f^p_\beta(\tau), \mbox{ otherwise. } \end{cases} \end{align*}$$
  2. (2) $\lambda _\beta ^q = \ell _\beta (\nu (\kappa _\beta ))$ .

  3. (3) For $\alpha \in [\max (s^p) \cap \beta ,\beta )$ , $f_\alpha ^q = f_\alpha ^p \circ \nu ^{-1}$ and $A_\alpha ^q = \{ \xi \circ \nu ^{-1} \mid \xi \in A_\alpha ^p \}$ if applicable.

The following analysis of dense open sets of $\mathbb {P}$ was established in [Reference Gitik8].

Lemma 2.5. For every condition $p \in \mathbb {P}$ and dense open set $D \subseteq \mathbb {P}$ , there are $p^* \geq ^* p$ and a finite subset $\{ \alpha _0, \dots , \alpha _{m-1}\}$ of the pure part of p, such that for every sequence $\vec {\nu } = \langle \nu _{\alpha _0},\dots , \nu _{\alpha _{m-1}}\rangle \in \prod _{i< m} A^{p^*}_{\alpha _i}$ , $p^* {}^\frown \vec {\nu } \in D$ .

For limit $\delta \leq \eta $ , we define $\bar {\kappa }_\delta = \sup _{\alpha <\delta }\kappa _\alpha $ . Gitik used the previous lemma to prove:

Theorem 2.6. In the generic extension by $\mathbb {P}$ , cardinals and cofinalities are preserved and for every limit $\delta \leq \eta $ , the singular cardinal hypothesis fails at $\bar {\kappa }_\delta $ .

For use later, we define $\vec {\mathcal {A}}$ to be the full-support product $\prod _{\alpha < \eta } \mathcal {A}(\lambda ,\kappa _\alpha )$ , and similarly, for each $\beta < \eta $ , $\vec {\mathcal {A}}_{\geq \beta } = \prod _{\beta \leq \alpha < \eta } \mathcal {A}(\lambda ,\kappa _\alpha )$ . We aim to show that if $\vec {H} = \langle H(\alpha ) \mid \alpha < \eta \rangle $ is $\vec {\mathcal {A}}$ generic over V, then there is a generic for the image of $\mathbb {P}$ in a suitable iterated ultrapower.

3 The iterated ultrapower $M_{\eta }$ and the generic filter $G^*$

We consider the following iterated ultrapower

$$\begin{align*}{\langle} M_\alpha,j_{\alpha,\beta} \mid \alpha \leq \beta \leq \eta {\rangle}\end{align*}$$

by the extenders in $\vec {E}$ . The iteration is defined by induction of $\alpha $ . Let $M_0 = V$ . For every $\alpha < \eta $ , given that $j_{0,\alpha } \colon M_0 \to M_\alpha $ has been defined, we take $E^\alpha _\alpha = j_{0,\alpha }(E_\alpha )$ , and set $j_{\alpha ,\alpha +1} \colon M_\alpha \to M_{\alpha +1} \cong \operatorname {Ult}(M_\alpha ,E_{\alpha }^\alpha )$ . At limit stage $\delta \leq \eta $ we take $M_\delta $ to be the direct limit of the system ${\langle } M_\alpha ,j_{\alpha ,\beta } \colon \alpha \leq \beta <\delta {\rangle }$ , and $j_{\alpha , \delta }$ to be the limit maps.

For every $\beta \leq \eta $ , we shall abbreviate and write $j_\beta $ for $j_{0,\beta }$ . For a given $\beta \leq \eta $ we shall denote $j_{\beta }(\kappa _\alpha ), j_{\beta }(E_\alpha )$ , $j_{\beta }(\lambda )$ by $\kappa _\alpha ^\beta $ , $E_\alpha ^\beta $ , and $\lambda ^\beta $ respectively. Similarly, we will denote $j_{\beta }(x) = x^\beta $ for objects $x \in V$ whose images along the iteration will be significant for our construction.

Lemma 3.1. Let $\beta \leq \eta $ . For every $x \in M_{\beta }$ there are $\beta _0< \beta _1 < \dots < \beta _{l-1}$ below $\beta $ , ordinals $\tau _0,\dots ,\tau _{l-1}$ below $\lambda $ , and $f \colon \prod _{i < l} \kappa _{\beta _i} \to V$ , such that

$$\begin{align*}x = j_{\beta}(f)\left(j_{\beta_0}(\tau_0), j_{\beta_1}(\tau_1), \dots, j_{\beta_{l-1}}(\tau_{l-1})\right).\end{align*}$$

Proof It is immediate from the definition of the iteration that for every $x \in M_\beta $ there are $\beta _0 < \dots < \beta _{l-1}$ below $\beta $ , finite subsets of ordinals $a_0,\dots , a_{l-1}$ , with each $a_i \in [\lambda ^{\beta _i}]^{<\omega }$ a generator of $E_{\beta _i}^{\beta _i}$ , and a function $g \colon \prod _{i < l} \kappa _{\beta _i}^{<\omega } \to V$ so that

$$\begin{align*}x = j_{0,\beta}(g)(j_{\beta_0+1,\beta}(a_0), \dots, j_{\beta_{l-1}+1,\beta}(a_{l-1})).\end{align*}$$

Note that for each $i < l$ , $j_{\beta _i+1,\beta }(a_i) = a_i$ , because

$$\begin{align*}\mathop{\mathrm{crit}}(j_{\beta_i+1, \beta}) = \kappa_{\beta_i+1}^{\beta_{i}+1}> \lambda^{\beta_i}.\end{align*}$$

It follows that

$$\begin{align*}x = j_{\beta}(g)(a_0, a_1, \dots, a_{l-1}).\end{align*}$$

We may further assume that each $a_i < \lambda ^{\beta _i}$ is an ordinal. Finally, we claim that we may replace each $a_i$ with an ordinal of the form $j_{\beta _i}(\tau _i)$ , for some $\tau _i < \lambda $ . By Lemma 2.3, in $M_{\beta _i}$ , the extender $E_{\beta _i}^{\beta _i} = j_{\beta _i}(E_{\beta _i})$ is generated by a subset of its measures, which are of the form $j_{\beta _i}(E_{\beta _i}(\tau _i)) = E_{\beta _i}^{\beta _i}(j_{\beta _i}(\tau _i))$ . Namely, for every $a_i \in \lambda ^{\beta _i}$ there exists some $\tau _i \in \lambda $ such that $a_i \leq ^{RK}_{E_{\beta _i}^{\beta _i}} j_{\beta _i}(\tau _i)$ .

In particular

$$\begin{align*}a_i = j_{\beta_i}(h)(b_1,\dots,b_k) \leq^{RK}_{E_{\beta_i}^{\beta_i}} j_{\beta_i}(\tau_i).\end{align*}$$

The claim follows.

Having defined the iterated ultrapower $M_{\eta }$ we proceed to introduce relevant conditions in $j_{\eta }(\mathbb {P})$ . Let $p = {\langle } p_\alpha \mid \alpha < \eta {\rangle }$ be a pure condition in $\mathbb {P}$ , that is, a condition with $s^p = \emptyset $ . For every $\alpha < \eta $ , we set $p_\alpha = \langle f_\alpha ,A_\alpha ,\lambda \rangle $ , where $f_\alpha \in A(\lambda ,\kappa _\alpha )$ and $A_\alpha \in E_\alpha (d_\alpha )$ where $d_\alpha = \operatorname {\mathrm {dom}}(f_\alpha ) \in [\lambda ]^{\leq \kappa _\alpha }$ .

Since $j_{\alpha ,\alpha +1} = j^{M_\alpha }_{E_\alpha ^\alpha }$ it follows that the function

$$\begin{align*}\nu^{p^\alpha}_{\alpha,\alpha+1} := (j_{\alpha,\alpha+1}\restriction d_\alpha^\alpha)^{-1} \end{align*}$$

belongs to $j_{\alpha ,\alpha +1}(A^\alpha _\alpha ) = A^{\alpha +1}_\alpha $ , which is the measure one set of the $\alpha $ -th component of $j_{\alpha +1}(p)$ . By applying $j_{\alpha +1,\eta }$ to $j_{\alpha +1}(p)$ , we conclude that the function

$$\begin{align*}\nu^{p^\alpha}_{\alpha,\eta} := j_{\alpha+1,\eta}(\nu^{p^\alpha}_{\alpha,\alpha+1}) = (j_{\alpha,\eta}\restriction d_\alpha^\alpha)^{-1}\end{align*}$$

belongs to $j_{\eta }(A_\alpha )$ , and thus $j_{\eta }(p)$ has an one point extension at the $\alpha $ -th coordinate of the form

$$\begin{align*}j_{\eta}(p) {}^\frown \langle \nu^{p^\alpha}_{\alpha,\eta}{\rangle}.\end{align*}$$

It is clear that for every finite sequence $\alpha _0 < \dots < \alpha _{m-1}$ of ordinals below $\eta $ , and pure condition $p \in \mathbb {P}$ , we have $j_{\eta }(p) {}^\frown {\langle } \nu ^{p^{\alpha _0}}_{\alpha _0,\eta }, \nu ^{p^{\alpha _1}}_{\alpha _1,\eta }, \dots , \nu ^{p^{\alpha _{m-1}}}_{\alpha _{m-1},\eta }{\rangle }$ is an extension of $j_{\eta }(p)$ in $j_{\eta }(\mathbb {P})$ . It is also clear that if $p,q \in \mathbb {P}$ are two pure conditions so that for all $\alpha < \eta $ , $f_\alpha ^p,f_\alpha ^q$ are compatible, then for every two sequences $\alpha _0 < \dots < \alpha _{m-1}$ and $\beta _0 < \dots < \beta _{l-1}$ of ordinals below $\eta $ , the conditions

$$\begin{align*}j_{\eta}(p) {}^\frown {\langle} \nu^{p^{\alpha_0}}_{\alpha_0,\eta}, \nu^{p^{\alpha_1}}_{\alpha_1,\eta}, \dots , \nu^{p^{\alpha_{m-1}}}_{\alpha_{m-1},\eta}{\rangle}\end{align*}$$

and

$$\begin{align*}j_{\eta}(q) {}^\frown {\langle} \nu^{q^{\beta_0}}_{\beta_0,\eta}, \nu^{q^{\beta_1}}_{\beta_1,\eta}, \dots , \nu^{q^{\beta_{l-1}}}_{\beta_{l-1},\eta}{\rangle}\end{align*}$$

are compatible in $j_{\eta }(\mathbb {P})$ . We remark that in the assumption that the sequence of extenders ${\langle } E_\alpha \mid \alpha < \eta {\rangle }$ is Mitchell order increasing is used in order to be able to permute the order in which the extensions are done, using a sequence of applications of Lemma 2.4.

Fix $\vec {H} = {\langle } H(\alpha ) \mid \alpha < \eta {\rangle }$ a V-generic for $\vec {\mathcal {A}} = \prod _{\alpha < \eta } \mathcal {A}(\lambda ,\kappa _\alpha )$ .

Definition 3.2. Define $G^* \subset j_{\eta }(\mathbb {P})$ to be the filter generated by all conditions

$$\begin{align*}j_{\eta}(p) {}^\frown {\langle} \nu^{p^{\alpha_0}}_{\alpha_0,\eta}, \nu^{p^{\alpha_1}}_{\alpha_1,\eta}, \dots , \nu^{p^{\alpha_{m-1}}}_{\alpha_{m-1},\eta}{\rangle},\end{align*}$$

where $p \in \mathbb {P}$ satisfy that $f^p_\alpha \in H(\alpha )$ for all $\alpha < \eta $ , $m < \omega $ , and $\alpha _0 < \dots < \alpha _{m-1}$ are ordinals below $\eta $ .

Our first goal is to prove that $G^*$ generates a $j_{\eta }(\mathbb {P})$ generic filter over $M_{\eta }$ .

Proposition 3.3. $G^*_{\eta }$ is $j_{\eta }(\mathbb {P})$ generic over $M_{\eta }$ .

Before moving to the proof of the proposition, we discuss finite subiterates of $M_{\eta }$ .

3.1 Finite sub-iterations $N^F$

The model $M_{\eta }$ can be seen as a directed limit of all its finite subiterates, $N^F$ , $F \in [\eta ]^{<\omega }$ . Given a finite set $F = \{ \beta _0,\dots ,\beta _{l-1} \} \in [\eta ]^{<\omega }$ , we define its associated iteration $\langle N^F_{i}, i^F_{m,n} \mid m \leq n \leq l\rangle $ by $N^F_0 = V$ , and $i^F_{n,n+1} \colon N^F_n \to N^F_{n+1} \cong \operatorname {Ult}(N^F_{n}, i^F_n(E_{\beta _{n}}))$ . Since for the moment we handle a single finite set at a time, we will suppress the mention of the finite set F and refer only to the iteration as $i_{m,n}\colon N_m \to N_n$ for $m \leq n \leq l$ and as usual we set $i_n = i_{0,n}$ for $n \leq l$ .

The proof of Lemma 3.1 shows that the elements of $N_l$ are of the form $i_l(f)\left (\tau _0, i_{1}(\tau _1), \dots , i_{l-1}(\tau _{l-1})\right )$ . Let $k\colon N_l \to M_{\eta }$ be the usual factor map defined by

$$\begin{align*}k(i_l(f)\left(\tau_0, i_{1}(\tau_1), \dots, i_{l-1}(\tau_{l-1})\right)) = j_{\eta}(f)\left(j_{\beta_0}(\tau_0), j_{\beta_1}(\tau_1), \dots, j_{\beta_{l-1}}(\tau_{l-1})\right).\end{align*}$$

It is routine to verify that k is well defined, elementary, and $j_{\eta } = k \circ i_l$ .

Moreover, following this explicit description of k, it is straightforward to verify that $k : N_l \to M_{\eta }$ is the resulting iterated ultrapower limit embedding, associated with the iteration of $N_l$ by the sequence

$$\begin{align*}{\langle} i_{0,n_\alpha}(E_\alpha) \mid \alpha \in \eta \setminus \{\beta_0, \dots \beta_{l-1}\}{\rangle},\end{align*}$$

where $n_\alpha $ is the minimal $n < l$ for which $\beta _n \geq \alpha $ , if exists, and $n_\alpha = l$ otherwise. This again uses the Mitchell order assumption of the sequence of extenders in order to get the desired commutativity.

Next, we observe that our assignment of generators $\nu ^{p^\alpha }_{\alpha ,\eta }$ , $\alpha < \eta $ , to $j_{\eta }(p)$ of conditions $p \in \mathbb {P}$ , in Definition 3.2, can be defined at the level of the finite subiterates.

Indeed, for a condition $p \in \mathbb {P}$ , $n < l$ , we temporarily define

$$\begin{align*}\nu^{i_n(p)}_{n,n+1} = (i_{n,n+1}\restriction d)^{-1} = \{ (i_{n,n+1}(\tau), \tau) \mid \tau \in d\},\end{align*}$$

where $d = \operatorname {\mathrm {dom}}(f^{i_n(p)})_{\beta _n}$ . We let $\nu ^{i_n(p)}_{n,l} = i_{n+1,l}(\nu ^{i_n(p)}_{n,n+1})$ be the natural push forward to $N_l$ . As with the conditions $j_{\eta }(p)$ , we have that

$$\begin{align*}i_l(p) {}^\frown \langle \nu^{i_0(p)}_{0,l}, \dots, \nu^{i_{l-1}(p)}_{l-1,l} \rangle\end{align*}$$

is a valid extension of $i_l(p)$ in $i_l(\mathbb {P})$ .

To record this definition (which depends on F), we make a few permanent definitions which explicitly mention F as well as objects from the discussion above:

Definition 3.4. Fix $i^F = i_l$ and $k^F=k$ and the model $N^F = N_l$ . Denote the condition defined above by $i^F(p) {}^\frown \vec {\nu }^{p,F}$ and refer to it as the natural non-pure extension of $i^F(p)$ .

Using the description of $k^F \colon N^F \to M_{\eta }$ , it is straightforward to apply Łoś’s Theorem and show that $k^F(\vec {\nu }^{p,F}) = {\langle } \nu ^{p^{\beta _n}}_{\beta _n,\eta }\mid n < l {\rangle }$ . We conclude that

$$\begin{align*}k^F\left( i^F(p) {}^\frown \vec{\nu}^{p,F} \right) = j_{\eta}(p) {}^\frown {\langle} \nu^{p^{\beta_0}}_{\beta_0,\eta}, \dots , \nu^{p^{\beta_{l-1}}}_{\beta_{l-1},\eta}{\rangle}. \end{align*}$$

Finally, we note that the forcing $\vec {\mathcal {A}} = \prod _{\alpha < \eta }\mathcal {A}(\lambda ,\kappa _\alpha )$ has a natural factorization, associated with F. Setting $\beta _{-1} = 0$ and $\beta _l = \eta $ , we have

$$\begin{align*}\vec{\mathcal{A}} = \prod_{n \leq l}\vec{\mathcal{A}}\restriction{[\beta_{n-1},\beta_{n})},\end{align*}$$

where for each $n \leq l$ ,

$$\begin{align*}\vec{\mathcal{A}}\restriction{[\beta_{n-1},\beta_{n})} = \prod_{\beta_{n-1} \leq \alpha < \beta_n} \mathcal{A}(\lambda,\kappa_\alpha).\end{align*}$$

For each $0\leq n \leq l$ , we define

$$\begin{align*}\vec{\mathcal{A}}^F_{n} = i_{n}(\vec{\mathcal{A}}\restriction{[\beta_{n-1},\beta_{n})}) \end{align*}$$

and denote the resulting product by $\vec {\mathcal {A}}^F = \prod _{n \leq l} \vec {\mathcal {A}}^F_n$ . Suppose that $\vec {H} \subseteq \vec {\mathcal {A}}$ is a V-generic filter. For each $n \leq l$ , the fact $\vec {\mathcal {A}}\restriction {[\beta _{n-1},\beta _{n})}$ is a $\kappa _{\beta _{n-1}}^{+}$ -closed forcing guarantees that $i_n"\vec {H}\restriction {[\beta _{n-1},\beta _{n})}) \subseteq \vec {\mathcal {A}}^F_n$ forms a generic filter over $N_n$ , hence also $N_l=N^F$ .Footnote 4 For each n, we denote the resulting $N_l$ generic for $\vec {\mathcal {A}}^F_n$ by $\vec {J}^F_n$ .

This is the crucial part: the main effect of taking a non-pure extension of a condition is the reflection of its Cohen part downwards. This reflection is exactly moving from $\vec {H}$ to $\vec {J}^F_n$ .

It is clear that the product $\prod _{0 \leq n \leq l} \vec {J}^F_n$ is $\vec {\mathcal {A}}^F$ generic over $N_l$ . We note that for each pure $p \in \mathbb {P}$ with $\vec {f}^p \in \vec {H}$ , if $\vec {f}$ is the Cohen part of the natural non-pure extension of $i^F(p)$ , then $\vec {f} \in \vec {\mathcal {A}}^F$ . Moreover the collection of such $\vec {f}$ forms an $\vec {\mathcal {A}}^F$ -generic filter over $N^F$ which is obtained by modifying $\prod _{0 \leq n \leq l} \vec {J}^F_n$ on the coordinates $\beta _n$ for $n<l$ by taking the natural non-pure extensions of $i^F(p)$ for $p\in \vec {J}^F$ (see Definition 3.4).

For each individual modified filter, its genericity follows from Woodin’s surgery argument [Reference Gitik5].Footnote 5 The fact that the product remains generic follows by induction on n from applications of Easton’s lemma, using the chain condition of the forcing notions with smaller indexes and the closure of the components with larger indexes. We denote by $\vec {H}^F$ the resulting $N^F$ -generic filter over $\vec {\mathcal {A}}^F$ .

Remark 3.5. $G^*$ can be constructed in $N^F[\vec {H}^F]$ in the same way that it was constructed in V using the fact that $M_\eta $ can be described as an iterated ultrapower of $N^F$ and starting with conditions q such that $s^q=F$ and $\vec {f}^q \in \vec {H}^F$ .

We turn to the proof of Proposition 3.3.

Proof It is clear that $G^* \subseteq j_{\eta }(\mathbb {P})$ is a filter. We verify that $G^*$ meets every dense open set $D \subseteq j_{\eta }(\mathbb {P})$ in $M_{\eta }$ . Since $M_\eta $ is the direct limit of its finite subiterates. There are $F = \{\beta _0, \dots \beta _{l-1}\} \in [\eta ]^{<\omega }$ and $\bar {D}$ such that $k^F(\bar {D}) = D$ .

Let $p'$ be the natural non-pure extension of $i^F(p)$ for some p with $\vec {f}^p \in \vec {H}$ . Now by its definition $\vec {f}^{p'} \in \vec {H}^F$ . Appealing to Lemma 2.5 and the genericity of $\vec {H}^F$ , we conclude that there exists a direct extension $p^* \geq ^* p'$ , with ${\langle } f^{p^*}_\alpha \mid \alpha <\eta {\rangle } \in \vec {H}^F$ and a finite set $\{ \alpha _0,\dots ,\alpha _{m-1}\} \subset (\eta \setminus F)$ , such that $p^* {}^\frown \vec {\nu } \in \bar {D}$ for every $\vec {\nu } = \langle \nu _{\alpha _0},\dots , \nu _{\alpha _{m-1}}\rangle \in \prod _{i<m} A^{p^*}_{\alpha _i}$ . By the elementarity of $k^F$ , $k^F(p^*) {}^\frown \vec {\nu } \in D$ for every $\vec {\nu } \in \prod _{i<m} k^F(A^{p^*})_{\alpha _i}$ .

In particular,

$$\begin{align*}k^F(p^*) {}^\frown {\langle} \nu^{(p^*)^{\alpha_0}}_{\alpha_0,\eta}, \dots , \nu^{(p^*)^{\alpha_{m-1}}}_{\alpha_{m-1},\eta}{\rangle} \in \mathcal{D}.\end{align*}$$

It remains to verify that the last condition belongs to $G^*$ . This is immediate from Remark 3.5 and the fact that $\vec {f}^{p^*} \in \vec {H}^F$ .

Proposition 3.6. $M_{\eta }[G^*]$ is closed under $\kappa _0 = \mathop {\mathrm {crit}}(j_{\eta })$ sequences of its elements in $V[\vec {H}]$ .

Proof Let ${\langle } x_\mu \mid \mu < \kappa _0{\rangle }$ be a sequence of elements in $M_{\eta }[G^*]$ . Since $M_\eta [G^*]$ is a model of $\mathrm {ZFC}$ , we may assume that all $x_\mu $ are ordinals. By Lemma 3.1, for each $x_\mu $ is of the form

$$\begin{align*}x_\mu = j_{\eta}(g_\mu)(j_{\beta^\mu_0}(\tau^\mu_0), \dots, j_{\beta^\mu_{l^\mu-1}}(\tau^\mu_{l^\mu-1})) \end{align*}$$

for some finite sequences $\beta ^\mu _0 < \dots , \beta ^\mu _{l^\mu -1} < \eta $ and $\tau ^\mu _0, \dots , \tau ^\mu _{l^\mu -1} < \lambda $ . Moreover, since the Rudin–Keisler order $\leq ^{RK}_{E_{\beta ^\mu _i}}$ is $\kappa _0^{+}$ -directed, we may assume that there exists some $\tau ^* < \lambda $ such that $\tau ^\mu _i = \tau ^*$ for all $\mu < \kappa _0$ and $i<l^\mu $ . Hence for ${\langle } x_\mu \mid \mu < \kappa _0{\rangle }$ to be a member of $M_{\eta }[G^*]$ , it suffices to verify that the sequences $\langle j_{\eta }(g_\mu ) \mid \mu < \kappa _0\rangle $ and ${\langle } j_{\alpha }(\tau ^*) \mid \alpha < \eta {\rangle }$ belong to $M_{\eta }[G^*]$ .

The first sequence already belongs to $M_{\eta }$ as $\mathop {\mathrm {crit}}(j_{\eta }) = \kappa _0$ implies that it is just $j_{\eta }(\vec {g})\restriction \kappa _0$ . The latter sequence $\langle j_{\alpha }(\tau ^*) \mid \alpha < \eta \rangle $ can be recovered from $G^*$ as follows. It follows from a simple density argument that for every $\alpha < \eta $ , there exists some $q \in G^*$ such that:

  1. (1) $\alpha $ is a non-pure coordinate of q, and

  2. (2) $j_{\eta }(\tau ^*) \in \operatorname {\mathrm {dom}}(f^q_\alpha )$ .

It is also clear that two conditions $q,q' \in G^*$ of this form must satisfy $f^{q}_\alpha (j_{\eta }(\tau ^*)) = f^{q'}_\alpha (j_{\eta }(\tau ^*))$ . We may therefore define in $M_{\eta }[G^*]$ a function $t_{j_{\eta }(\tau ^*)} \colon \eta \to \lambda ^{\eta }$ by $t_{j_{\eta }(\tau ^*)}(\alpha ) = f^{q}_\alpha (j_{\eta }(\tau ^*))$ for some condition $q \in G^*$ as above. We claim that $t_{j_{\eta }(\tau ^*)}(\alpha ) = j_{\alpha }(\tau ^*)$ for all $\alpha < \eta $ . Indeed, for every $\alpha < \eta $ , there exists a condition $p \in \mathbb {P}$ with ${\langle } f^p_\alpha \mid \alpha < \eta {\rangle } \in \vec {H}$ so that $\tau ^* \in \operatorname {\mathrm {dom}}(f^p_\alpha )$ , and clearly, the condition $q = j_{\eta }(p) {}^\frown \langle \nu ^{p^\alpha }_{\alpha ,\eta }\rangle \in G^*$ has $j_{\eta }(\tau ^*) \in \operatorname {\mathrm {dom}}(f^q_\alpha )$ . But $j_{\eta }(\tau ^*) \in \operatorname {\mathrm {dom}}(\nu ^{p^\alpha }_{\alpha ,\eta }) = j_{\alpha +1,\eta }"j_{\alpha }(\operatorname {\mathrm {dom}}(f^p_\alpha ))$ and $\nu ^{p^\alpha }_{\alpha ,\eta } = (j_{\alpha ,\eta }\restriction d_\alpha ^\alpha )^{-1}$ ; thus, it follows that

$$\begin{align*}f^{q}_\alpha(j_{\eta}(\tau^*)) = \nu^{p^\alpha}_{\alpha,\eta}(j_{\eta}(\tau^*)) = j_{\alpha}(\tau^*).\\[-32pt]\end{align*}$$

Fix some ordinal $\beta < \eta $ . The forcing $\vec {\mathcal {A}}$ naturally breaks into the product ${\vec {\mathcal {A}}\restriction \beta \times \vec {\mathcal {A}}_{\geq \beta }}$ , and we observe that:

  1. (1) The latter part $\vec {\mathcal {A}}_{\geq \beta }$ is $\kappa _\beta ^{+}$ -closed. Therefore, if $\vec {H}_{\geq \beta } \subset \vec {\mathcal {A}}_{\geq \beta }$ is V-generic, then by [Reference Cummings2, Proposition 15.1] its pointwise image $j_{\beta }"\vec {H}_{\geq \beta }$ generates an $M_{\beta }$ generic filter for the forcing

    $$\begin{align*}j_{\beta}(\vec{\mathcal{A}}_{\geq \beta}) = \prod_{\beta \leq \alpha < \eta} \mathcal{A}(\lambda^\beta,\kappa_\alpha^\beta)^{M_\beta}.\end{align*}$$
    We denote this generic by $\vec {H}^{\beta }_{\geq \beta }$ .
  2. (2) Using the same arguments as above, a V-generic filter $\vec {H}_\beta \subset \vec {\mathcal {A}}\restriction \beta $ generates an $M_\beta $ -generic filter $G^*_\beta $ for the $j_{\beta }(\mathbb {P}_{\vec {E}\restriction \beta })$ .

We conclude that in the model $V[\vec {H}]$ , we can form the generic extension $M_\beta [G^*_\beta \times \vec {H}^\beta _{\geq \beta }]$ of $M_\beta $ , with respect to the product $j_{\beta }(\mathbb {P}_{\vec {E}\restriction \beta }) \times j_{\beta }(\vec {\mathcal {A}}_{\geq \beta })$ .

We have the following.

Proposition 3.7. For each $\beta < \eta $ , the model $M_{\beta }[G^*_\beta \times \vec {H}^\beta _{\geq \beta }]$ can compute $M_{\eta }[G^*_{\eta }]$ and therefore

$$\begin{align*}\bigcap_{\beta<\eta}M_\beta[G^*_\beta\times \vec{H}^\beta_{\geq\beta}] \supseteq M_\eta[G^*].\end{align*}$$

It seems likely that equality holds, $\bigcap _{\beta <\eta }M_\beta [G^*_\beta \times \vec {H}^\beta _{\geq \beta }] = M_\eta [G^*]$ , but we will not need that in our argument.

4 Stationary reflection in $M_\eta [G^*_\eta ]$

We assume that for every $\alpha < \eta $ , there is a Laver indestructible supercompact cardinal $\theta _\alpha $ such that $\sup _{\beta <\alpha } \kappa _\beta < \theta _\alpha < \kappa _\alpha $ .

We prove the following which completes the proof of Theorem 1.1.

Theorem 4.1. In $M_{\eta }[G^*]$ , every collection of fewer than $\eta $ many stationary subsets of $j_{\eta }(\bar {\kappa }_{\eta }^{+})$ reflects.

We start by proving a stationary reflection fact that will be used as an intermediate step in the proof.

Claim 4.2. For every $\alpha <\eta $ , every collection of fewer than $\eta $ many stationary subsets of $j_{\alpha }(\bar {\kappa }_{\eta }^{+})$ with cofinalities bounded by $\sup _{\beta <\alpha }\kappa _{\beta }^\alpha $ reflects in $M_{\alpha }[G^*_\alpha \times \vec {H}^\alpha _{\geq \alpha }]$ .

Proof Let $T_i$ for $i<\mu $ be such a collection of stationary sets. By elementarity, $j_\alpha (\theta _\alpha )$ is an indestructible supercompact cardinal between $\sup _{\beta <\alpha }\kappa ^\alpha _\beta $ and $\kappa _\alpha ^\alpha $ . Recall that $\vec {H}^\alpha _{\geq \alpha }$ is generic for $(\kappa _\alpha ^\alpha )^{+}$ -directed closed forcing. By the indestructibility of $j_\alpha (\theta _\alpha )$ , there is a $j_\alpha (\bar {\kappa }_{\eta }^{+})$ -supercompact embedding $k\colon M_{\alpha }[\vec {H}^\alpha _{\geq \alpha }] \to N$ . Further, $G^*_\alpha $ is generic for $j_\alpha (\bar {\kappa }_\alpha ^{++})$ -cc forcing and $j_{\alpha }(\bar {\kappa }_\alpha ^{++}) < j_\alpha (\theta _\alpha )$ . By standard arguments, we can extend k to include $M_\alpha [G^*_\alpha \times \vec {H}^\alpha _{\geq \alpha }]$ in the extension by a $j_{\alpha }(\bar {\kappa }_\alpha ^{++})$ -cc forcing. Each $T_i$ remains stationary in this extension, so $\{k(T_i) \mid i < \mu \}$ reflects at $\sup k"j_\alpha (\bar {\kappa }_{\eta }^{+})$ . It follows that $\{T_i \mid i < \mu \}$ reflects in $M_\alpha [G^*_\alpha \times \vec {H}^\alpha _{\geq \alpha }]$ .

We begin the proof of Theorem 4.1. Suppose that for each $i<\mu $ , $S_i \in M_{\eta }[G^*]$ is a stationary subset of $j_{\eta }(\bar {\kappa }_{\eta }^{+})$ . We can assume that for all i, the cofinality of each ordinal in $S_i$ is some fixed $\gamma _i$ . It follows that there is some $\bar {\alpha }<\eta $ such that $\sup _{i<\mu } \gamma _i < \sup _{\beta <\bar {\alpha }}\kappa _{\beta }^{\bar {\alpha }}$ . For each $\alpha $ in the interval $[\bar {\alpha },\eta )$ , let $T_i^\alpha = \{ \delta < j_\alpha (\bar {\kappa }_{\eta }^{+}) \mid j_{\alpha ,\eta }(\delta ) \in S_i\}$ .

Claim 4.3. For $\alpha \geq \bar {\alpha }$ , if $\{T_i^{\alpha } \mid i<\mu \}$ reflects at an ordinal of cofinality less than $\kappa _\alpha ^\alpha $ in $M_\alpha [G_\alpha ^* \times \vec {H}^\alpha _{\geq \alpha }]$ , then $\{S_i \mid i<\mu \}$ reflects in $M_{\eta }[G^*]$ .

Proof Let $\delta < j_\alpha (\bar {\kappa }_{\eta }^{+})$ with $\operatorname {\mathrm {cf}}(\delta ) < \kappa _\alpha ^\alpha $ be a common reflection point of the collection $\{T_i^\alpha \mid i<\mu \}$ . We claim that each $S_i$ reflects at $j_{\alpha ,\eta }(\delta )$ . Let $D \subseteq j_{\alpha ,\eta }(\delta )$ be club in $M_{\eta }[G^*]$ , of order type $\operatorname {\mathrm {cf}}(\delta ) = \operatorname {\mathrm {cf}}(j_{\alpha ,\eta }(\delta ))$ .

Since $\mathop {\mathrm {crit}}{j_{\alpha ,\eta }}> \operatorname {\mathrm {cf}}(\delta )$ , $j_{\alpha ,\eta }$ is continuous at $\delta $ and $E = \{\gamma < \delta \mid j_{\alpha ,\eta }(\gamma ) \in D \}$ is a club in $\delta $ . Note that $E \in M_{\alpha }[G^*_\alpha \times \vec {H}^{\alpha }_{\geq \alpha }]$ . Since $T_i^\alpha $ reflects at $\delta $ , $T_i^\alpha \cap E \neq \emptyset $ and hence $S_i \cap D \neq \emptyset $ .

Combining the previous two claims if for some $\alpha \geq \bar {\alpha }$ , $\{T_i^\alpha \mid i < \mu \}$ consists of stationary sets, then $\{S_i \mid i<\mu \}$ reflects. So we assume for a contradiction that for each $\alpha {\kern-1pt}\geq{\kern-1pt} \bar {\alpha }$ , there are $i_\alpha {\kern-1pt}<{\kern-1pt}\mu $ and a club $C_\alpha {\kern-1pt}\in{\kern-1pt} M_\alpha [G_\alpha ^* {\kern-1pt}\times{\kern-1pt} \vec {H}^\alpha _{\geq \alpha }]$ such that $T_{i_\alpha }^\alpha \cap C_\alpha = \emptyset $ . We fix $J \subseteq \eta $ unbounded and $i^*<\mu $ such that for all $\alpha \in J$ , $i_\alpha = i^*$ .

Let $I_\eta $ be the ideal of bounded subsets of $\eta $ . For $\alpha \leq \eta $ , let $\vec {H}^\alpha /I_\eta $ be the generic for $j_\alpha (\vec {\mathbb {A}})/I_\eta $ derived from $j_\alpha "\vec {H}$ . Recall that $\vec {H}$ is generic for $\vec {\mathbb {A}}$ which is a product of Cohen posets, so this makes sense.

Claim 4.4. For $\alpha \geq \bar {\alpha }$ , there is a club subset of $C_\alpha $ in $M_\alpha [\vec {H}^\alpha /I_\eta ]$ .

Proof We start by showing that for all $\beta> \alpha $ below $\eta $ there is a club subset of $C_\alpha $ in $M_{\alpha }[\vec {H}^\alpha \restriction [\beta , \eta )]$ .

To see this, note that $M_\alpha [G^*_\alpha \times \vec {H}^\alpha _{\geq \alpha }]$ is $(\bar {\kappa }^\alpha _\beta )^{++}$ -cc extension of $M_\alpha [\vec {H}^\alpha \upharpoonright [\beta ,\eta )]$ and $\bar {\kappa }_\beta ^\alpha <j_\alpha (\bar {\kappa }_{\eta })$ . Indeed, this forcing decomposes into a part contributing $G^*_\alpha $ which have small cardinality and the product of forcing notions $j_{\alpha }(\mathcal {A}(\lambda , \kappa _\gamma ))$ for $\gamma < \beta $ , whose product has $(\bar \kappa _\beta ^\alpha )^{++}$ -cc. For the moment we let $\mathbb {Q}_\beta $ denote the $(\bar {\kappa }^\alpha _\beta )^{++}$ -cc poset used in order to move from $M_\alpha [\vec {H}^\alpha \restriction [\beta ,\eta )]$ to $M_\alpha [G_\alpha ^* \times \vec {H}^\alpha _{\geq \alpha }]$ .

We set take $C_{\alpha ,\beta }$ to be the set of closure points of the function assigning each $\gamma $ to the supremum of the set

$$\begin{align*}\{\gamma^* \mid \exists q\in \mathbb{Q}_\beta,\, q \Vdash \check{\gamma}^* = \min (\dot{C}_\alpha \setminus (\check\gamma + 1))\},\end{align*}$$

as computed in the model $M_\alpha [\vec {H}^\alpha \upharpoonright [\beta ,\eta )]$ . Clearly this is a club. Further if $\beta < \beta '$ , then $C_{\alpha ,\beta '} \subseteq C_{\alpha ,\beta }$ .

Since the clubs are decreasing, $\bigcap _{\beta> \alpha } C_{\alpha ,\beta }$ is definable in $M_\alpha [\vec {H}^\alpha /I_\eta ]$ as the set of ordinals $\gamma $ such that for some condition $\vec {a} \in j_\alpha (\vec {\mathbb {A}})$ , $\vec {a}/I_\eta \in \vec {H}^\alpha / I_\eta $ and for all sufficiently large $\beta $ , $\vec {a} \upharpoonright [\beta ,\eta )$ forces $\gamma \in C_{\alpha ,\beta }$ . So $\bigcap _{\beta>\alpha }C_{\alpha ,\beta }$ is as required for the claim.

Let $\dot {D}_\alpha $ be an $\vec {H}^\alpha /I_\eta $ -name for the club from the previous lemma.

Lemma 4.5. $\vec {H}^{\eta }/I_\eta \in M_{\eta }[G^*]$ .

Proof Let $p \in j_\eta(\vec{\mathbb{A}})$ . We will show how to decide in $M_\eta [G^*]$ whether $p \in \vec {H}^\eta /I_\eta $ or not.

Now, fix $q \in G^*$ with $\operatorname {\mathrm {dom}} q(\alpha ) \supseteq \operatorname {\mathrm {dom}} p(\alpha )$ . Let us claim that $p \in \vec {H}/I_\eta $ if and only if there is $\beta $ such that $f^q(\gamma )\supseteq p(\gamma )$ for all $\gamma \in [\beta , \eta )$ .

Indeed, without loss of generality, $q = j_\eta (\bar {q}) ^\smallfrown \langle \nu _0,\dots , \nu _{m-1}\rangle $ for $\bar {q} \in \vec {H}$ and $\nu _0,\dots , \nu _{m-1}$ generators.

Thus, for every $\beta $ which is larger than the index of the measures of $\nu _0,\dots ,\nu _{m-1}\ f^{\bar {q}(\beta )} \in H(\beta )$ and thus after applying $j_\eta $ we conclude that the condition which is defined by $\langle f^q(\gamma ) \mid \beta \leq \gamma < \eta \rangle $ belongs to $\vec {H}^\eta \restriction [\beta ,\eta )$ . Since it is compatible with p, $p \in \vec {H}^\eta /I_\eta $ .

In the other direction, one can take $\beta $ large enough so that $p \restriction [\beta ,\eta ) \in \vec {H}^\eta \restriction [\beta ,\eta )$ and then $p \restriction [\beta ,\eta )$ is the Cohen part of a condition in $G^*$ .

Claim 4.6. $\langle j_{\alpha ,\eta }(\dot {D}_\alpha )_{\vec {H}^{\eta }/I_\eta } \mid \bar {\alpha } \leq \alpha < \eta \rangle \in M_{\eta }[G^*]$ .

Proof By the previous lemma, $\vec {H}^{\eta }/I_\eta \in M_{\eta }[G^*]$ . Further by Proposition 3.6, $M_{\eta }[G^*]$ is closed under $\eta $ -sequences. So the sequence of names for clubs $\langle j_{\alpha ,\eta }(\dot {D}_\alpha ) \mid \alpha < \eta \rangle \in M_{\eta }[G^*]$ .

To get a contradiction it is enough to show that $\bigcap _{\bar {\alpha } \leq \alpha <\eta } j_{\alpha }(\dot {D}_\alpha )_{\vec {H}^{\eta }/I_\eta } \cap S_{i^*} = \emptyset $ . Suppose that there is some $\delta $ in the intersection. We can find some $\alpha \in J$ and $\bar {\delta }$ such that $j_{\alpha ,\eta }(\bar {\delta }) = \delta $ . However, by the definitions of $D_\alpha $ and $T_{i^*}^\alpha $ , we must have that $\bar {\delta } \in C_\alpha \cap T_{i^*}^\alpha $ , a contradiction.

This completes the proof of Theorem 1.1.

5 Bad scales

In this section we give the proof of Theorem 1.2. For this theorem we work with the forcing $\mathbb {P}$ as before and assume that there is an indestructibly supercompact cardinal $\theta < \kappa _0$ . Working in V, let $\vec {f}$ be a scale of length $\bar {\kappa }_\eta ^{+}$ in $\prod _{\alpha <\eta }\kappa _\alpha ^{+}$ . As before let $\vec {H}$ be generic for $\vec {\mathcal {A}}$ over V and let $G^*$ be the $j_\eta (\mathbb {P})$ generic over $M_\eta $ defined above.

Lemma 5.1. In $M_\eta [G^*]$ , $j_\eta (\vec {f})$ is a scale of length $j_\eta (\bar {\kappa }_\eta ^{+})$ in $\prod _{\alpha <\eta }j(\kappa _\alpha )^{+}$ .

Proof Let $g \in \left (\prod _{\alpha <\eta }j(\kappa _\alpha )^{+} \right )\cap M_\eta [G^*]$ . Clearly $g \in V[\vec {H}]$ . Since each $\kappa _\alpha ^{+}$ is a continuity point of $j_\eta $ , we can find an ordinal $\gamma _\alpha <\kappa _\alpha ^{+}$ such that $j_\eta (\gamma _\alpha )>g(\alpha )$ .

By the distributivity of $\vec {\mathbb {A}}$ , the sequence $\tilde {g} = \langle \gamma _\alpha \mid \alpha < \eta \rangle $ belongs to V. Pick (in V) an ordinal $\zeta $ such that $f_\zeta $ dominates $\tilde {g}$ . Then, $j_\eta (f_\zeta ) = j_\eta (f)_{j_\eta (\zeta )}$ dominates g.

Lemma 5.2. For $\delta <\kappa _\eta ^{+}$ with $\eta < \operatorname {\mathrm {cf}}(\delta )<\kappa _0$ , if $j_\eta (\delta )$ is a good point for $j(\vec {f})$ in $M_\eta [G^*]$ , then $\delta $ is a good point for $\vec {f}$ in $V[\vec {H}]$ .

Proof Suppose that $A \subseteq j_\eta (\delta )$ and $\alpha ^* < \eta $ witness that $j_\eta (\delta )$ is good for $\vec {f}$ in $M_\eta [G^*]$ . Note that $j_\eta "\delta $ is cofinal in $j(\delta )$ . By thinning A if necessary we let $B \subseteq j_\eta "\delta $ be an unbounded set such that each element $\gamma \in B$ has a greatest element of A less than or equal to it. For each $\gamma $ in B, let $\alpha _\gamma $ be such that for all $\alpha \geq \alpha _\gamma $ ,

$$\begin{align*}j(\vec{f})_{\max(A \cap (\gamma+1))}(\alpha) \leq j(\vec{f})_\gamma(\alpha) < j(\vec{f})_{\min(A \setminus (\gamma+1))}(\alpha).\end{align*}$$

Let $B'$ be an unbounded subset of B on which $\alpha _\gamma $ is fixed. Then $B'$ witnesses that $j_\eta (\delta )$ is good. Since $B' \subseteq j_\eta "\delta $ , we have that $\{ \gamma < \delta \mid j_\eta (\gamma ) \in B' \}$ witnesses that $\delta $ is good for $\vec {f}$ in $V[\vec {H}]$ .

In $V[\vec {H}]$ , let $S = \{ \delta < \bar {\kappa }_\eta ^{+} \mid \delta $ is nongood for $\vec {f}$ and $\eta < \operatorname {\mathrm {cf}}(\delta ) < \theta \}$ . We claim that S is stationary. Let $k\colon V[\vec {H}] \to N$ be an elementary embedding witnessing that $\theta $ is $\bar {\kappa }_\eta ^{+}$ -supercompact in $V[\vec {H}]$ . Standard arguments show that $\sup k"\bar {\kappa }_\eta ^{+}$ is a nongood point for $k(\vec {f})$ . It follows that S is stationary, since $\sup k"\bar {\kappa }_\eta ^{+} \in k(C)$ every club $C \subseteq \bar {\kappa }_\eta ^{+}$ in $V[\vec {H}]$ .

Now suppose that in $M_\eta [G^*]$ , there is a club D of good points for $j_\eta (\vec {f})$ . In $V[\vec {H}]$ , let $C = \{ \delta < \bar {\kappa }_\eta ^{+} \mid j_\eta (\delta ) \in D \}$ . By the previous lemma, C is a $<\ \kappa _0$ -club consisting of good points for $\vec {f}$ . However, $S \cap C$ is nonempty, a contradiction.

6 Countable cofinality

Our goal in this section is to show that given the assumptions of Theorem 1.4, there exists a forcing extension which adds a generic filter for the extender based Prikry forcing by a $(\kappa ,\kappa ^{++})$ -extender (in particular, forces $\operatorname {\mathrm {cf}}(\kappa ) = \omega $ and $2^\kappa = \kappa ^{++}$ ) and satisfies that every stationary subset of $\kappa ^{+}$ reflects.

A necessary step for obtaining the latter is to add a club $D \subseteq \kappa ^{+}$ disjoint from $S^{\kappa ^{+}}_{\kappa }$ , the set of ordinals $\alpha < \kappa ^{+}$ of cofinality $\kappa $ in the ground model. In [Reference Hayut and Unger13], the second and third authors address the situation for adding a single Prikry sequence to $\kappa $ . It is shown that under the subcompactness assumption of $\kappa $ , there is a Prikry-type forcing which both singularizes $\kappa $ and adds a club D as above, without generating new nonreflecting stationary sets.

An additional remarkable aspect of the argument of [Reference Hayut and Unger13] is that it presents a fertile framework in which the arguments address an extension $N_\omega [\mathcal {H}]$ of an iterated ultrapower $N_\omega $ of V, and assert directly that stationary reflection holds in $N_{\omega }[\mathcal {H}]$ without having to specify the poset by which $\mathcal {H}$ is added.

Let us briefly describe the key ingredients of the construction of [Reference Hayut and Unger13], to serve as a reference for our arguments in the context of the failure of SCH. In [Reference Hayut and Unger13], one starts from a normal measure U on $\kappa $ in V, and consider the $\omega $ -iterated ultrapower by U, given by

$$\begin{align*}N_0 = V\text{, }i_{n,n+1} \colon N_n \to N_{n+1} \cong \operatorname{Ult}(N_n,i_n(U)),\end{align*}$$

and the direct limit embedding $i_{\omega } \colon V \to N_\omega $ . It is well known that the sequence of critical points ${\langle } \kappa _n \mid n < \omega {\rangle }$ , $\kappa _{n+1} = i_{n,n+1}(\kappa _{n})$ is Prikry generic over $N_\omega $ and that $N_{\omega }[{\langle } \kappa _n \mid n<\omega {\rangle }] = \bigcap _n N_n$ [Reference Bukovský1, Reference Dehornoy3].

Let $\mathbb {Q}$ be the Prikry name for the forcing for adding a disjoint club from $(S^{\kappa ^{+}}_\kappa )^V = \kappa ^{+} \cap \operatorname {\mathrm {Cof}}^V(\kappa )$ . The forcing $\mathbb {Q}_{\omega } = i_\omega (\mathbb {Q})^{{\langle } \kappa _n \mid n < \omega {\rangle }} \in N_\omega [{{\langle } \kappa _n \mid n < \omega {\rangle }}]$ is isomorphic to the forcing for adding a $\kappa ^{+}$ -Cohen set over V, and likewise, to adding a $\kappa _n^{+}$ -Cohen set over $N_n$ , for each $n < \omega $ .

Moreover, taking a $\mathbb {Q}_{\omega }$ -generic filter $H $ over V, we have that both H and $i_{0,1}"H $ generate mutually generic filters over $N_1$ for $i_1(\mathbb {Q}_\omega ) = \mathbb {Q}_\omega $ . More generally, for each n, the sequence $H^n_0, \dots , H^n_n$ , where each $H^n_k$ is generated by $i_{k,n}"H \subset i_n(\mathbb {Q}_\omega ) = \mathbb {Q}_\omega $ for $1 \leq k \leq n$ , are mutually generic filters for $\mathbb {Q}_\omega $ over $N_n$ . With this choice of “shifts” of H, we obtain that for each $n < k$ , the standard iterated ultrapower map $i_{n,k} \colon N_n \to N_k$ extends to $i_{n,k}^* \colon N_n[{\langle } H^n_0,\dots , H^n_n {\rangle }] \to N_k[ {\langle } H^{k}_0, \dots , H^{k}_n{\rangle }] \subseteq N_k[ {\langle } H^{k}_0, \dots , H^{k}_k{\rangle }]$ . For each n, the sequence ${\langle } H^n_0, \dots , H^n_n {\rangle }$ is denoted by $\mathcal {H}_n$ . The final extension $N_\omega [\mathcal {H}]$ of $N_\omega $ is given by the sequence $\mathcal {H} = {\langle } H_n^\omega \mid n < \omega {\rangle }$ , where $H_n^\omega $ is the filter generated by $i_{n,\omega }"H$ , which achieves the critical equality

$$\begin{align*}N_\omega[\mathcal{H}] = \bigcap_n N_n[\mathcal{H}_n].\end{align*}$$

From this equality it follows at once that:

  1. (1) $N_\omega [\mathcal {H}]$ is closed under its $\kappa $ -sequences;

  2. (2) $\kappa _\omega = i_\omega (\kappa )$ is singular in $N_\omega [\mathcal {H}]$ , as ${\langle } \kappa _n \mid n<\omega {\rangle }$ belongs to each $N_n[\mathcal {H}_n]$ ; and

  3. (3) $H \in N_\omega [\mathcal {H}]$ , as $H = H^n_n$ for all $n < \omega $ .

Therefore every stationary subset S of $\kappa _\omega ^{+}$ in $N_\omega [\mathcal {H}]$ can be assumed to concentrate at some cofinality $\rho < \kappa _\omega $ . Say for simplicity that $\rho < \kappa _0$ , one shows that S reflects in $N_\omega [\mathcal {H}]$ by examining its pull backs $S_n = i_{n,\omega }^{-1}(S) \subseteq \kappa _n^{+}$ . If we have that $\kappa _n = i_n(\kappa )$ is $\Pi ^1_1$ -subcompact in the Cohen extension $N_n[\mathcal {H}_n]$ of $N_n$ , then we have that if $S_n$ is stationary then it must reflect at some $\delta < \kappa _n^{+}$ of cofinality $\delta <\kappa _n$ . Using $i_{n,\omega }$ , we conclude from this that S reflects at $i_{n,\omega }(\delta )$ in $N_\omega [\mathcal {H}]$ .

To rule out the other option, of having all $S_n \subseteq \kappa _n^{+}$ being nonstationary in $N_n[\mathcal {H}_n]$ , one takes witnessing disjoint clubs $C_n \subseteq \kappa _n^{+}$ and uses the fact that for each $n < \omega $ , $i_{n,\omega } \colon N_n \to N_\omega $ extends to $i_{n,\omega }^* \colon N_n[\mathcal {H}_n] \to N_\omega [ i_{n,\omega }"\mathcal {H}_n] \subset N_{\omega }[\mathcal {H}]$ . This allows us to show that the club $D_n = i_{n,\omega }^*(C_n) \subseteq \kappa _\omega ^{+}$ belongs to $N_\omega [\mathcal {H}]$ for each n. Since $N_\omega [\mathcal {H}]$ is closed under its $\kappa $ sequences, it computes ${\langle } D_n \mid n < \omega {\rangle }$ correctly and thus also $D = \bigcap _n D_n$ , that would have to be disjoint from S, a contradiction.

We turn now to the new construction and prove Theorem 1.4. Let $V'$ be a model which contains a $\kappa ^{+}$ - $\Pi ^1_1$ -subcompact cardinal $\kappa $ , which also carries a $(\kappa ,\kappa ^{++})$ -extender.

Recall that $\kappa $ is $\kappa ^{+}$ - $\Pi ^1_1$ -subcompact if for every set $A \subseteq H(\kappa ^{+})$ and every $\Pi ^1_1$ -statement $\Phi $ such that ${\langle } H(\kappa ^{+}), \in ,A\rangle \models \Phi $ , there are $\rho < \kappa $ , $B \subseteq H(\rho ^{+})$ , and an elementary embedding $j \colon {\langle } H(\rho ^{+}), \in ,B\rangle \to {\langle } H(\kappa ^{+}), \in ,A{\rangle }$ with $cp(j) = \rho $ , such that ${\langle } H(\rho ^{+}), \in ,B\rangle \models \Phi $ .

Let V be obtained from $V'$ by an Easton-support iteration of products $\mathop {\mathrm {Add}}(\alpha ^{+},\alpha ^{++})$ for inaccessible $\alpha \leq \kappa $ .

By [Reference Hayut and Unger13, Lemma 42], $\kappa $ remains $\kappa ^{+}$ - $\Pi _1^1$ -subcompact in V and even in a further extension by $\mathop {\mathrm {Add}}(\kappa ^{+},\kappa ^{++})$ . Moreover, by standard arguments it is routine to verify that $\kappa $ still carries a $(\kappa ,\kappa ^{++})$ -extender in V. We note that as a consequence of $\kappa ^{+}$ - $\Pi _1^1$ -subcompactness in V, simultaneous reflection holds for collections of fewer than $\kappa $ many stationary subsets of $S^{\kappa ^{+}}_{<\kappa }$ . Further this property is indestructible under $\mathop {\mathrm {Add}}(\kappa ^{+},\kappa ^{++})$ .

Working in V, let E be a $(\kappa ,\kappa ^{++})$ -extender. Let

$$\begin{align*}\langle j_{m,n}\colon M_m \to M_n \mid m \leq n \leq \omega \rangle\end{align*}$$

be the iteration by E and let

$$\begin{align*}\langle i_{m,n}\colon N_m \to N_m \mid m \leq n \leq \omega\rangle\end{align*}$$

be the iteration by the normal measure $E_\kappa $ where $V = M_0 = N_0$ . We write $j_n$ for $j_{0,n}$ and $i_n$ for $i_{0,n}$ .

We describe a generic extension of $M_\omega $ in which $j_\omega (\kappa ^{+})$ satisfies the conclusion of the theorem. It follows by elementarity that there is such an extension of V. We are able to isolate the forcing used, but this is not required in the proof.

We start by constructing a generic object for $j_\omega (\mathbb {P}_E)$ over $M_\omega $ , where $\mathbb {P}_E$ is the extender based forcing of Merimovich. Although for the most part, we will refer to Merimovich’s arguments in [Reference Merimovich15], our presentation follows a more up-do-date presentation of the forcing, given by Merimovich in [Reference Merimovich16, Section 2].

We recall that conditions $p \in \mathbb {P}_E$ are pairs of the form $p = {\langle } f, T{\rangle }$ , where $f \colon d \to [\kappa ]^{<\omega }$ is a partial function from $\kappa ^{++}$ to $[\kappa ]^{<\omega }$ with domain $d \in [\kappa ^{++}]^{\leq \kappa }$ with $\kappa \in d$ , and T is tree of height $\omega $ , whose splitting sets are all measure one with respect to a measure $E(d)$ on $V_\kappa $ , derived from the extender E. The generator of $E(d)$ is the function $\operatorname {\mathrm {mc}}(d) = \{ {\langle } j(\alpha ), \alpha {\rangle } \mid \alpha \in d\}$ . Therefore, a typical node in the tree is an increasing sequence of functions ${\langle } \nu _0,\dots ,\nu _{k-1}{\rangle }$ where each $\nu _i$ is a partial, order preserving function $\nu _i \colon \operatorname {\mathrm {dom}}(\nu _i) \to \kappa $ , with $\kappa \in \operatorname {\mathrm {dom}}(\nu _i)$ and $|\nu _i| = \nu _i(\kappa )$ . The sequence ${\langle } \nu _0,\dots , \nu _{k-1}{\rangle }$ is increasing in the sense that $\nu _i(\kappa ) < \nu _{i+1}(\kappa )$ for all i. When extending conditions $p = {\langle } f, T{\rangle } \in \mathbb {P}_E$ we are allowed to:

  1. (i) extend f in as Cohen conditions (namely, add points $\gamma < \kappa ^{++}$ to $\operatorname {\mathrm {dom}}(f)$ and arbitrarily choose $f(\gamma ) \in [\kappa ]^{<\omega }$ ), and modify the tree parts;

  2. (ii) shrink the tree T; and

  3. (iii) choose a point ${\langle } \nu {\rangle } \in \operatorname {\mathrm {succ}}_{\emptyset }(T)$ to extend $p = {\langle } f,T{\rangle }$ to $p_{{\langle } \nu {\rangle }} = {\langle } f_{{\langle } \nu {\rangle }}, T_{{\langle } \nu {\rangle }}{\rangle }$ , where $f_{{\langle } \nu {\rangle }}$ is defined by $\operatorname {\mathrm {dom}}(f_{{\langle } \nu {\rangle }}) = \operatorname {\mathrm {dom}}(f)$ and

    $$ \begin{align*} f_{{\langle} \nu {\rangle}}(\alpha) = \begin{cases} f(\alpha) \cup \{\nu(\alpha)\}, &\mbox{if } \alpha \in \operatorname{\mathrm{dom}}(\nu) \text{ and } \nu(\alpha)> \max(f(\alpha)),\\ f(\alpha), &\mbox{otherwise. } \end{cases} \end{align*} $$

Any extension of p is obtained by finite combination of (i)–(iii), and the direct extensions of p are those which are obtained by (i) and (ii). The poset $\mathbb {P}_E^*$ is the suborder of $\mathbb {P}_E$ whose extension consists only of the Cohen type extension (i). Clearly, $\mathbb {P}_E^*$ is isomorphic to $\mathop {\mathrm {Add}}(\kappa ^{+},\kappa ^{++})$ .

We turn to describe the construction of a $j_\omega (\mathbb {P}_E)$ generic following [Reference Merimovich15]. Let $G_0$ be $\mathbb {P}_E^*$ generic. We work by induction to define $G_n$ for $n < \omega $ . Suppose that we have defined $G_n \subseteq j_n(\mathbb {P}_E^*)$ for some $n<\omega $ . First, let $G_{n+1}'$ be the closure of the set $j_{n,n+1}"G_n$ . Then, we take $G_{n+1}$ to be obtained from $G_{n+1}'$ by adding the ordinal $\alpha $ to $G_{n+1}'$ at coordinate $j_{n,n+1}(\alpha )$ , for all $\alpha < j_n(\kappa ^{++})$ .

Claim 6.1. $G_{n+1}$ is $M_{n+1}$ -generic for $j_{n+1}(\mathbb {P}_E^*)$ .

Proof This is a straightforward application of Woodin’s surgery argument [Reference Gitik5], so we only sketch the proof. Let D be a dense open subset of $j_{n+1}(\mathbb {P}_E^*)$ . Let E be the set of all f in D such that all $j_{n}(\kappa )$ sized modifications of f are in D. E is dense using the closure of $j_{n+1}(\mathbb {P}_E^*)$ . Now the fact that $G_{n+1}'$ meets E implies that $G_{n+1}$ meets D, since each condition in $G_{n+1}$ is a $j_n(\kappa )$ sized modification of one in $G_{n+1}'$ .

We make a few remarks.

Remark 6.2. For each n, $G_{n}$ can be obtained directly from the upwards closure of $j_n"G_0$ by combining the alterations used in construction of $G_i$ for $i<n$ .

Since the proof of the previous claim can be repeated in any suitably closed forcing extension, we have the following.

Remark 6.3. If H is generic for $j_n(\kappa ^{+})$ -closed forcing and mutually generic to the upwards closure of $j_n"G_0$ , then H and $G_n$ are mutually generic.

Let $G_\omega $ be the $j_\omega (\mathbb {P}_E)$ -generic obtained from $G_0$ as in [Reference Merimovich15].

Claim 6.4. $M_\omega [G_\omega ]$ is closed under $\kappa $ -sequences.

Proof It is enough to show that $M_\omega [G_\omega ]$ is closed under $\kappa $ -sequences of ordinals. Let ${\langle } \gamma _\delta \mid \delta < \kappa {\rangle }$ be a sequence of ordinals. We can assume that each $\gamma _\delta $ is of the form $j_\omega (g_\delta )(\alpha _\delta , j(\alpha _\delta ),\dots j_{n_\delta -1}(\alpha _\delta ))$ for some $g_\delta \colon [\kappa ]^{n_\delta } \to \kappa ^{++}$ and $\alpha _\delta <\kappa ^{++}$ . We refer the reader to [Reference Merimovich15, Corollary 2.6] for a proof. Since ${\langle } j_\omega (g_\delta ) \mid \delta < \kappa {\rangle } = j_\omega ({\langle } g_\delta \mid \delta <\kappa {\rangle })\restriction \kappa \in M_\omega $ , it is enough to show that ${\langle } (\alpha _\delta , \dots j_{n_\delta -1}(\alpha _\delta )) \mid \delta < \kappa {\rangle } \in M_\omega [G_\omega ]$ . To see this note that $\{\alpha _\delta \mid \delta < \kappa \} = \operatorname {\mathrm {dom}}(f)$ for some $f \in G_0$ and that the sequence $f(\alpha _\delta ) \frown (\alpha _\delta , \dots j_{n_\delta -1}(\alpha _\delta ))$ is an initial segment of the $\omega $ -sequence with index $j_\omega (\alpha _\delta )$ in $G_\omega $ . This is enough to compute ${\langle } (\alpha _\delta , \dots j_{n_\delta -1}(\alpha _\delta )) \mid \delta < \kappa {\rangle }$ , since $j_\omega "f \in M_\omega $ .

The following lemma is proved in [Reference Hayut12, Section 4].

Lemma 6.5. $\bigcap _{n<\omega } M_n[G_n] = M_\omega [G_\omega ]$ .

Since we will need to prove a stronger version ahead, let us sketch the proof of this lemma.

Proof First, the inclusion $\bigcap _{n<\omega } M_n[G_n] \supseteq M_\omega [G_\omega ]$ is clear.

As in the case of the standard Bukovský–Dehornoy Theorem, in order to prove the other direction we would like to take a set of ordinals $X \in M_\omega [G_\omega ]$ and consider the sets

$$\begin{align*}Y_n = \{\alpha \mid j_{n,\omega}(\alpha) \in X\} \in M_n[G_n].\end{align*}$$

Note that the filter generated by $j_{n,\omega }" G_n$ is not a member of $M_\omega [G_\omega ]$ and thus we must be more careful about the way we push $Y_n$ into $M_\omega [G_\omega ]$ .

Let us pick for each n a name $\dot {y}_n \in M_n$ for $Y_n$ . As the sequence of names $\langle j_{n,\omega }(\dot {y}_n)\mid n < \omega \rangle $ is a member of $M_\omega [G_\omega ]$ (using its closure under countable sequence), we must verify that we can evaluate correctly, without access to the filters $j_{n,\omega }(G_n)$ , whether an ordinal $\gamma $ belongs to a tail of sets $j_{n,\omega }(\dot {y}_n)_{j_{n,\omega }(G_n)}$ .

This is done using Merimovich’s genericity criteria [Reference Merimovich17]. Indeed, for $a \subseteq j_{\omega }(\kappa ^{++})$ set of ordinals of cardinality $j_{\omega }(\kappa )$ , which is the set of ordinals of an elementary substructure of a sufficiently large initial segment of $M_\omega $ , one can consider a tree Prikry type forcing using $j_\omega (E)(a)$ , and verify that there is a generic sequence $\langle t_n \mid n < \omega \rangle $ for it in $M_\omega [G_\omega ]$ such that, $\exists q \in j_\omega (\mathbb {P}_E^*)$ such that for all n, $q^\smallfrown \langle t_0,\dots , t_{n-1}\rangle \in G_\omega $ . Moreover, this sequence is unique up to shifts and finite modifications (see [Reference Merimovich17, Lemma 3.4])—which is the crux of this argument.

As $j_{\omega }(\kappa ) \in a$ and we may assume that the Prikry sequence at this coordinate is the standard one, one can fix the shift and identify the sequence $\langle j_{\omega }(p)^\smallfrown \langle \nu _0, \dots , \nu _{n-1}\rangle \mid n < \omega \rangle $ up to a finite error.

In particular, by taking a to be large enough, we conclude that there is $p \in G_0$ such that for all large n,

$$\begin{align*}j_{\omega}(p)^\smallfrown \langle \nu_0, \dots, \nu_{n-1}\rangle \Vdash\check{\gamma} \in j_{n,\omega}(\dot{y}_n)\end{align*}$$

if and only if $\gamma \in X$ .

Therefore, $\gamma \in X$ if and only if there is a condition $q\in G_\omega $ with large domain a and a Prikry generic $\langle t_n \mid n < \omega \rangle $ such that the length of $q(j_\omega (\kappa ))$ is zero and for all large n

$$\begin{align*}q^\smallfrown \langle \nu_0, \dots, \nu_{n-1}\rangle \Vdash\check{\gamma} \in j_{n,\omega}(\dot{y}_n).\\[-32pt]\end{align*}$$

We are now ready to describe the generic extension of $M_\omega $ . We recall some of the basic ideas from [Reference Hayut and Unger13]. Let $\dot {\mathbb {Q}}$ be the canonical name in the Prikry forcing defined from $E_\kappa $ for the poset to shoot a club disjoint from the set of $\alpha < \kappa ^{+}$ such that $\operatorname {\mathrm {cf}}^V(\alpha ) = \kappa $ . Let $\mathbb {Q}_\omega $ be the forcing $i_\omega (\dot {\mathbb {Q}})$ as interpreted by the critical sequence $\langle i_n(\kappa ) \mid n<\omega \rangle $ . By the argument following Claim 39 of [Reference Hayut and Unger13], $\mathbb {Q}_\omega $ is equivalent to the forcing $\mathop {\mathrm {Add}}(\kappa ^{+},1)$ in V. In fact, for every $n<\omega $ , in $N_n$ it is equivalent to $\mathop {\mathrm {Add}}(i_n(\kappa ^{+}),1)$ .

Let $k_n\colon N_n \to M_n$ be the natural embedding given by

$$\begin{align*}k_n(i_n(f)(\kappa, \dots i_{n-1}(\kappa))) = j_n(f)(\kappa, \dots j_{n-1}(\kappa)). \end{align*}$$

We discussed the notion of a width of an elementary embedding in the previous section. We recall a fundamental result concerning lifting embeddings and their widths (see [Reference Cummings2]).

Lemma 6.6. Suppose that $k \colon N \to M$ has width $\mu $ , $\mathbb {Q} \in N$ is a $\mu ^{+}$ -distributive forcing, and $H \subseteq \mathbb {Q}$ is generic over N, then $k"H \subset k(\mathbb {Q})$ generates a generic filter ${\langle } k"H {\rangle }$ for M.

Claim 6.7. For all $1\leq n<\omega $ , $k_n$ has width $\leq (i_{n-1}(\kappa )^{++})^{N_n}$ .

Proof Let $x = j_n(g)(\alpha ,j(\alpha ), \dots j_{n-1}(\alpha ))$ be an element of $M_n$ , $g\colon [\kappa ]^n \to V$ and $\alpha < \kappa ^{++}$ . Consider $h = i_n(g) \restriction (i_{n-1}(\kappa )^{++})^{N_n}$ . Then

$$\begin{align*}k(h)(\alpha,j(\alpha), \dots j_{n-1}(\alpha)) = x.\end{align*}$$

Indeed, the domain of $k(h)$ is large enough for this evaluation to make sense.

We are now ready to define the analogs of $\mathcal {H}$ and $\mathcal {H}_n$ for our situation. Recall that for subset X of some poset, we write $< X>$ for the upwards closure in that poset. It will be clear from the context which poset we are working with, when taking this upwards closure.

Let H be generic for $\mathbb {Q}_\omega $ over $V[G_0]$ and recall that $\mathbb {Q}_\omega = i_n(\mathbb {Q}_\omega )$ is also a member of $N_n$ for every n, and is $i_n(\kappa ^{+})$ -closed. We note that clearly, $i_n(\kappa^+) \geq (i_{n-1}(\kappa)^{++})^{N_n} \geq \text{width}(k_n)$ .

Let

$$\begin{align*}\mathcal{H}_n = \langle {<}j_{m,n} \circ k_m"H{>} \mid m \leq n \rangle\end{align*}$$

and

$$\begin{align*}\mathcal{H} = \langle {<}j_{m,\omega} \circ k_m"H{>} \mid m < \omega \rangle.\end{align*}$$

Note that $\mathcal {H}_n$ is a subset of $\prod _{m \leq n} j_{m,n}(k_m(\mathbb {Q}_\omega ))$ . We do not yet know that it is generic.

We prove a sequence of claims about $\mathcal {H}_n$ and $\mathcal {H}$ . The first is straightforward.

Claim 6.8. For all $n<\omega $ ,

$$\begin{align*}\mathcal{H}_{n+1} = {<}j_{n,n+1} " \mathcal{H}_n{>} \frown {<}k_{n+1}"H{>}.\end{align*}$$

Let $\bar {\mathcal {H}}_n = \langle {<} i_{m,n}"H{>} \mid m \leq n \rangle $ .

Claim 6.9. $\mathcal {H}_n = {<} k_n " \bar {\mathcal {H}}_n{>}$ .

This is immediate from the following. For $m \leq n$ , we have

$$\begin{align*}{<}k_n " {<}i_{m,n}"H{>}{>} = {<}j_{m,n} " {<}k_m " H{>}{>}\end{align*}$$

since $k_n \circ i_{m,n} = j_{m,n} \circ k_m$ . In particular, since $\bar {\mathcal {H}}_n$ is generic for $i_n(\kappa ^{+})$ -closed forcing over $N_n$ and $k_n$ has width less than $i_n(\kappa )$ , $\mathcal {H}_n$ is generic over $M_n$ .

Claim 6.10. $\mathcal {H}_n$ is mutually generic with $G_n$ over $M_n$ .

Proof Since $G_0$ and H are mutually generic over V, we can repeat the argument from Claim 18 of [Reference Hayut and Unger13] to see that $\bar {\mathcal {H}}_n$ is generic over the model $N_n[{<}i_n"G_0{>}]$ . By the previous claim, ${<}j_n"G_0{>}$ and $\mathcal {H}_n$ are mutually generic over $M_n$ . By Remark 6.3, $\mathcal {H}_n$ is mutually generic with $G_n$ .

Lemma 6.11. $M_\omega [G_\omega ][\mathcal {H}]$ is closed under $\kappa $ -sequences.

Proof It is enough to show that every $\kappa $ -sequence of ordinals from the larger model $M_\omega [G_\omega ][\mathcal {H}]$ is in $M_\omega [G_\omega ]$ . Let $\vec {\alpha }$ be such a sequence. By construction, $\vec {\alpha } \in V[G_0][H]$ and hence in $V[G_0]$ . However, $M_\omega [G_\omega ]$ is closed under $\kappa $ -sequences in $V[G_0]$ , so $\vec {\alpha } \in M_\omega [G_\omega ]$ , as required.

Lemma 6.12. $\bigcap _{n<\omega }M_n[G_n][\mathcal {H}_n] = M_\omega [G_\omega ][\mathcal {H}]$ .

Proof The proof is similar to the proof of Lemma 6.5. First, note that the embedding $j_{n,\omega }\colon M_n \to M_\omega $ lifts to an elementary embedding

$$\begin{align*}j^*_{n,\omega}\colon M_n[\mathcal{H}_n] \to M_\omega[{<}j_{n,\omega}"\mathcal{H}_n{>}],\end{align*}$$

and $M_\omega [{<}j_{n,\omega }"\mathcal {H}_n{>}] \subseteq M_{\omega }[\mathcal {H}]$ . Further, the construction of $G_\omega $ is the same whether we start in $M_n$ or $M_n[\mathcal {H}_n]$ .

It is immediate that $\bigcap _{n<\omega }M_n[G_n][\mathcal {H}_n] \supseteq M_\omega [G_\omega ][\mathcal {H}]$ . For the other inclusion, we work as before and suppose that $x \in \bigcap _{n<\omega }M_n[G_n][\mathcal {H}_n]$ is a set of ordinals. For each n, we can define $y_n = \{ \gamma \mid j_{n,\omega }(\gamma ) \in x \}$ . Continuing as before, but working in $M_n[\mathcal {H}_n]$ , we have that $y_n = \dot {y}_n^{G_n}$ for some name $\dot {y}_n$ .

By the previous lemma, $M_\omega [G_\omega ][\mathcal {H}]$ has the sequence ${\langle } j^*_{n,\omega }(\dot {x}_n^*) \mid n<\omega {\rangle }$ . As in the proof of Lemma 6.5, we will now argue that for every suitable set of ordinals $a \subseteq j_\omega (\kappa ^{++})$ of cardinality $j_\omega (\kappa )$ , there is a unique (up to finite shift and finite modification) sequence $\langle t_n \mid n < \omega \rangle $ which is generic for the Prikry forcing for the measure $j_\omega (E)(a)$ and satisfies that $q ^\smallfrown \langle t_n \mid n < \omega \rangle \in G_\omega $ for all n.

Since there is $p\in G$ such that for $j_{\omega }(p)$ the sequence $\langle \nu _n \restriction a \mid n < \omega \rangle $ will be generic for the corresponding forcing, we conclude (by uniqueness) that for every such sequence, for all large n,

$$\begin{align*}q ^\smallfrown \langle t_n \mid n < \omega\rangle \in j_{n,\omega}(G_n)\end{align*}$$

assume that the length of q at $j_{\omega }(\kappa )$ is zero, so there is no shift relative to the indexing of the $G_n$ -s.

For each ordinal $\gamma $ one can take a large enough so that for all n, the truth value of $\gamma \in j_{n,\omega }(\dot {y}_n)$ is decided by $q ^\smallfrown \langle t_i \mid i < n\rangle $ . Therefore, $\gamma \in x$ if and only if there is $q \in G_\omega $ and $\langle t_n \mid n < \omega \rangle $ Prikry generic for $\operatorname {\mathrm {dom}} q$ as above such that for all large n,

$$\begin{align*}q ^\smallfrown \langle t_i \mid i < n\rangle \Vdash \check{\gamma} \in j_{n,\omega}(\dot{y}_n).\\[-36pt]\end{align*}$$

Let $k_\omega \colon N_\omega \to M_\omega $ be the natural elementary embedding. Note that $k_\omega $ naturally extends to an embedding between the two Prikry generic extensions,

$$\begin{align*}k^*_\omega\colon N_\omega[{\langle} i_m(\kappa) \mid m < \omega{\rangle}] \longrightarrow M_\omega[{\langle} j_m(\kappa) \mid m<\omega {\rangle}].\end{align*}$$

Similarly, for each $n<\omega $ we define $W_{n,\omega }$ to be the limit ultrapower obtained by starting in $M_n$ and iterating the measure $j_n(E_\kappa )$ , the normal measure of $j_n(E)$ . As with $N_\omega $ , denote the critical points of the iteration by $j_n(E_\kappa )$ and its images by ${\langle } \kappa ^n_m {\rangle }_{m<\omega }$ , and let $i^n_\omega \colon V \to W_{n,\omega }$ denote the composition of the finite ultrapower map $j \colon V \to M_n$ with the latter infinite iteration map by the normal measures. Therefore elements $y_n \in W_{n,\omega }$ are of the form $i^n_\omega (f)(\alpha _0,\dots ,\alpha _{n-1},\kappa ^n_0,\dots ,\kappa ^n_{m-1})$ for some $n,m < \omega $ , $f \colon \kappa ^{n+m} \to V$ in V, and $\alpha _{\ell } \in j_{\ell }(\kappa ^{++})$ for each $\ell < n$ .

The structures $W_{n,\omega }$ are naturally connected via maps $k_{n,r} \colon W_{n,\omega } \to W_{r,\omega }$ for $n \leq r < \omega $ , given by

$$\begin{align*}\begin{matrix} k_{n,r}(i^n_\omega(f)(\alpha_0,\dots,\alpha_{n-1},\kappa^n_0,\dots,\kappa^n_{m-1}) ) \\ = \,\, i^r_\omega(f)(\alpha_0,\dots,\alpha_{n-1},\kappa^r_0,\dots,\kappa^r_{m-1}). \end{matrix} \end{align*}$$

It is straightforward to verify that the limit of the directed system

$$\begin{align*}\{ W_{n,\omega}, k_{n,r} \mid n \leq r < \omega\}\end{align*}$$

is $M_\omega $ , and the direct limit maps $k_{n,\omega } \colon W_{n,\omega } \to M_\omega $ , which are defined by

$$\begin{align*}\begin{matrix} k_{n,r}\left( i^n_\omega(f)(\alpha_0,\dots,\alpha_{n-1},\kappa^n_0,\dots,\kappa^n_{m-1}) \right) & \\ \, \ = j_\omega(f)(\alpha_0,\dots,\alpha_{n-1}, j_{n}(\kappa), j_{n+1}(\kappa),\dots, j_{n+m-1}(\kappa))& \end{matrix} \end{align*}$$

naturally extend to the generic extensions by the suitable Prikry sequences. Indeed, denote $W_{n,\omega }[{\langle } j_m(\kappa ) \mid m<n {\rangle } {}^\frown {\langle } i_m(j_n(\kappa )) \mid m < \omega {\rangle }]$ by $W^*_{n,\omega }$ , for each $n <\omega $ . Then $k_{n,\omega }$ lifts to

$$\begin{align*}k^*_{n,\omega}\colon W_{n,\omega}^*\longrightarrow M_\omega[{\langle} j_m(\kappa) \mid m<\omega{\rangle} ].\end{align*}$$

Finally, we note that the following implies that $M_\omega [{\langle } j_m(\kappa ) \mid m < \omega {\rangle }]$ is the direct limit of the system of Prikry generic extensions

$$\begin{align*}{\langle} W_{n,\omega}^*, k^*_{n,k} \mid n \leq k < \omega{\rangle}.\end{align*}$$

Claim 6.13. ${<}k^*_\omega "H{>} \in M_\omega [G_\omega ][\mathcal {H}]$ is generic for $j_\omega (\dot {\mathbb {Q}})^{{\langle } j_n(\kappa ) \mid n<\omega {\rangle }}$ over $M_\omega [{\langle } j_n(\kappa )\mid n<\omega {\rangle }]$ .

Proof Let $D \in M_\omega [{\langle } j_n(\kappa ) \mid n < \omega {\rangle }]$ be a dense open subset of the forcing $j_\omega (\dot {\mathbb {Q}})^{{\langle } j_n(\kappa ) \mid n <\omega {\rangle }}$ . Then there are $n<\omega $ and $\bar {D} \in W_{n,\omega }^*$ such that $k^*_{n,\omega }(\bar {D}) = D$ . It follows from our arguments above that $<k_n"H>$ is generic for $k_n(\mathbb {Q}_\omega )$ over $M_n$ . Now $k_{n}(\mathbb {Q}_\omega ) \in W_{n,\omega }^*$ and $\bar {D}$ is a dense subset of it. So $<k_{n}"H>\cap \bar {D}$ is nonempty. Let $q \in H$ be such that $k_{n}(q) \in \bar {D}$ . It follows that $k^*_\omega (q) = k^*_{n,\omega }(k^*_n(q)) \in D$ . So we have shown that $<k^*_\omega "H>$ is generic over $M_\omega [{\langle } j_n(\kappa ) \mid n <\omega {\rangle }]$ . It remains to show that it is a member of $M_\omega [G_\omega ][\mathcal {H}]$ . Since $<k^*_\omega "H>= <k^*_{n,\omega }"<k^*_n"H>>$ , we have that $<k^*_\omega "H>\in M_n[\mathcal {H}_n]$ for all $n<\omega $ . By Claim 6.12, it follows that $<k^*_\omega "H>\in M_\omega [G_\omega ][\mathcal {H}]$ .

To complete the proof of Theorem 1.4 we need a finer control of the relationship between $\mathbb {P}$ and $\mathbb {P}^*$ names. To this end we make some definitions.

Definition 6.14. Let $f \in \mathbb {P}^*$ , $\alpha $ be an ordinal and let $\dot {C}$ be a $\mathbb {P}^*$ -name for a club subset of $\kappa ^{+}$ . We say that f stably forces $\check {\alpha } \in \dot {C}$ ( $f \Vdash ^s_{\mathbb {P}^*} \check {\alpha } \in \dot {C}$ ) if every alteration of f on fewer than $\kappa $ many coordinates forces $\check {\alpha } \in \dot {C}$ .

We have the following straightforward claims.

Claim 6.15. If g extends f in $\mathbb {P}^*$ and $f \Vdash ^s \check {\alpha } \in \dot {C}$ , then $g \Vdash ^s \check {\alpha } \in \dot {C}$ .

Claim 6.16. If $f \Vdash ^s \check {\alpha } \in \dot {C}$ , then for every finite sequence $\vec {\nu }$ from some tree associated with f, $f \frown \vec {\nu } \Vdash ^s \check {\alpha } \in \dot {C}$ . If in addition $f = f' \frown \vec {\nu }$ for some finite sequence $\vec {\nu }$ , then $f' \Vdash ^s \check {\alpha } \in \dot {C}$ .

We define $\dot {C}^s = \{ (\check {\alpha },f) \mid f \Vdash ^s \check {\alpha } \in \dot {C} \}$ . Clearly it is forced that $\dot {C}^s \subseteq \dot {C}$ . It is also straightforward to see that $\dot {C}^s$ is forced to be closed.

Claim 6.17. $\dot {C}^s$ is forced to be unbounded in $\kappa ^{+}$ .

Proof Fix some $f \in \mathbb {P}^*$ and $\alpha _0< \kappa ^{+}$ . Take some sufficiently large $\theta $ and some $N \prec H_\theta $ of size $\kappa $ such that $\mathbb {P}^*,f,\dot {C},\alpha _0 \in N$ and ${}^{<\kappa }N \subseteq N$ . Since $\mathbb {P}^*$ is $\kappa ^{+}$ -closed we can find a $(\mathbb {P}^*,N)$ -generic condition $f^* \geq f$ with $\operatorname {\mathrm {dom}}(f^*) = N \cap \kappa ^{++}$ . Let $\alpha = \sup (N \cap \kappa ^{++})$ .

Let $f'$ be any condition obtained by altering $f^*$ on a set of size less than $\kappa $ . Since ${}^{<\kappa }N \subseteq N$ , the alteration is a member of N. Hence a standard argument shows that $f'$ is also $(\mathbb {P}^*,N)$ -generic and so $f' \Vdash \check {\alpha } \in \dot {C}$ .

The name $\dot {C}^s$ behaves well when translated to a $\mathbb {P}$ -name. Let

$$\begin{align*}\dot{E} = \{ (\check{\alpha},p) \mid f^p \Vdash^s \check{\alpha} \in \dot{C} \}.\end{align*}$$

Claim 6.18. $\dot {E}$ is forced by $\mathbb {P}$ to be a club subset of $\kappa ^{+}$ .

Proof By the arguments of Claim 6.17, $\dot {E}$ is unbounded. Let us show first that $\dot {E}$ is forces to be closed.

Let $p = {\langle } f, T{\rangle }$ be a condition that forces that $\delta $ is an accumulation point of $\dot {E}$ . Therefore, for every $\alpha < \delta $ there are $\vec {\nu } = {\langle } \nu _0,\dots ,\nu _{n-1}{\rangle } \in T$ , $p' = {\langle } f' , T'{\rangle } \leq ^* p {}^\frown \vec {\nu }$ , and $\gamma < \delta $ such that $f' \Vdash ^s \gamma \in \dot {C}$ . Clearly, $f'$ is a Cohen extension of $f {}^\frown \vec {\nu }$ . Denote by $f^*$ the Cohen extension of f for which $f' = f^* {}^\frown \vec {\nu }$ . Since $f' \Vdash ^s \gamma \in \dot {C}$ , $f^* \Vdash ^s \gamma \in \dot {C}$ as well. It follows that the condition $p^* = {\langle } f^*,T^*{\rangle }$ , $T^* = \pi _{\operatorname {\mathrm {dom}}(f^*),\operatorname {\mathrm {dom}}(f)}^{-1}(T)$ is a direct extension of p and forces “ $\check {\gamma } \in \dot {E}$ ”.

Using the fact that the Cohen extension order on the function f is $\kappa ^{+}$ -closed and $\operatorname {\mathrm {cf}}(\delta ) \leq \kappa $ , we can repeat this process and construct a sequence of Cohen extensions ${\langle } f_i \mid i \leq \operatorname {\mathrm {cf}}(\delta ){\rangle }$ of f, and an increasing sequence ${\langle } \gamma _i \mid i < \operatorname {\mathrm {cf}}(\delta ){\rangle }$ cofinal in $\delta $ , such that $f_i \Vdash ^s \check {\gamma _i} \in \dot {C}$ . Let $\bar {f} = f_{\operatorname {\mathrm {cf}}(\delta )}$ . Since $\dot {C}$ is a Cohen name of a club, $\bar {f} \Vdash ^s \check {\delta } \in \dot {C}$ . Setting $\bar {T} = \pi _{\operatorname {\mathrm {dom}}(\bar {f}),\operatorname {\mathrm {dom}}(f)}^{-1}(T)$ and $\bar {p} = {\langle } \bar {f},\bar {T}{\rangle }$ , we have that $\bar {p}$ is a direct extension of p, and forces $\check {\delta } \in \dot {E}$ .

Remark 6.19. The proofs of the previous two claims work for any $\mathbb {P}^*$ -name for a club $\dot {C}$ in any $\kappa ^{+}$ -closed generic extension.

We are now ready to finish the proof of Theorem 1.4.

Claim 6.20. In $M_\omega [G_\omega ][\mathcal {H}]$ , every finite collection of stationary subsets of $j_\omega (\kappa ^{+})$ reflects at a common point.

Proof Let $S_i$ for $i < m$ be a sequence of stationary sets in $M_\omega [G_\omega ][\mathcal {H}]$ . By Claim 6.13, ${<}k_\omega "H{>}$ is a club in $j_\omega (\kappa ^{+})$ disjoint from $S_{j_\omega (\kappa )}^{j_\omega (\kappa ^{+})}$ as computed in $M_\omega $ and ${<}k_\omega "H{>} \in M_\omega [G_\omega ][\mathcal {H}]$ . So we can assume that each $S_i$ concentrates on a fixed cofinality below $j_\omega (\kappa )$ . For simplicity we assume that the cofinalities of the $S_i$ are bounded below $\kappa $ .

Let $T_n^i$ be the set of $\alpha $ such that $j_{n,\omega }(\alpha ) \in S_i$ . For a fixed $n<\omega $ , $T_n^i$ for $i < m$ is definable in $M_n[G_n][\mathcal {H}_n]$ . By the indestructibility of stationary reflection at $j_n(\kappa ^{+})$ in $M_n$ , if each $T_n^i$ for $i < m$ is stationary, then they reflect at a common point. Suppose that $\delta $ is this common reflection point. Then there are sets $A_i\in M_n$ for $i <k$ with $\mathop {\mathrm {ot}}(A)=\operatorname {\mathrm {cf}}(\delta )$ such that $A_i$ is stationary in $\delta $ and $A_i \subseteq T_n^i$ . It follows that for each $i<m$ , $j_{n,\omega }(A_i) = j_{n,\omega }"A_i \subseteq S_i$ and hence the collection of $S_i$ reflect at $j(\delta )$ .

We assume for the sake of a contradiction that for each $n<\omega $ at least one of the $T_n^i$ is nonstationary. Let $\dot {C}_n$ be a name whose interpretation by $G_n$ and $\mathcal {H}_n$ , $C_n$ , is a club disjoint from some $T_n^{i_n}$ . By the arguments above and Remark 6.19, we can work in $M_n[\mathcal {H}_n]$ and translate $\dot {C}_n$ to a $j_n(\mathbb {P})$ -name $\dot {E}_n$ for a club subset of $j_n(\kappa ^{+})$ . By the construction of $\dot {E}_n$ , we have that for $\alpha < j_n(\kappa ^{+})$ , $j_{n,\omega }(\alpha )$ is in $j_{n,\omega }(\dot {E}_n)^{G_\omega }$ if and only if it is in $j_{n,\omega }(\dot {C}_n)^{{<}j_{n,\omega }"G_n{>}}$ .

By the closure of $M_\omega [G_\omega ]$ under $\omega $ -sequences, $\langle j_{n,\omega }(\dot {E}_n) \mid n<\omega \rangle $ is in $M_\omega [G_\omega ]$ . Hence we can interpret it using $G_{\omega }$ and $j_{n,\omega }"\mathcal {H}_n$ . Let $D_n$ denote the resulting club. Let $i^*<m$ be such that $i^* = i_n$ for infinitely many n. We claim that $\bigcap _{n<\omega }D_n$ is disjoint from $S_{i^*}$ . Otherwise, we have $j_{n,\omega }(\alpha ) \in D_n \cap S_{i^*}$ for some n such that $i^* = i_n$ . It follows that $\alpha \in C_n \cap T_n^{i_n}$ , a contradiction.

6.1 Down to $\aleph _{\omega }$

In the paper [Reference Poveda, Rinot and Sinapova20] it is shown that it is consistent relative to infinitely many supercompact cardinals, that stationary reflection at $\aleph _{\omega +1}$ together with the failure of $\operatorname {\mathrm {SCH}}$ at $\aleph _{\omega }$ .

Let us show how to obtain the same result in our case, starting with the assumption of $\kappa ^{+}$ - $\Pi ^1_1$ -subcompact that carries a $(\kappa ,\kappa ^{++})$ -extender, as in the previous part.

Theorem 6.21. Let $\kappa $ be a $\kappa ^{+}$ - $\Pi ^1_1$ -subcompact cardinal and let us assume that it carries a $(\kappa ,\kappa ^{++})$ -extender. Then, there is a forcing extension in which $\kappa =\aleph _\omega $ , $2^{\aleph _{\omega }} = \aleph _{\omega +2}$ , $\operatorname {\mathrm {GCH}}$ holds below $\aleph _{\omega }$ and every finite collection of stationary subsets of $\aleph _{\omega +1}$ reflects simultaneously.

The proof of this theorem is essentially the same as the proof of Theorem 1.4. We follow the argument as in the previous subsection and explain the differences.

Let us prepare the universe so that $\operatorname {\mathrm {GCH}}$ holds and the subcompactness of $\kappa $ is indestructible under Cohen forcing $\mathop {\mathrm {Add}}(\kappa ^{+},\kappa ^{++})$ . Let $j_E \colon V \to M_1$ be the extender ultrapower embedding. Let us assume moreover that the approachability ideal is always maximal. Namely, for every regular uncountable cardinal $\rho $ , $\rho \in I[\rho ]$

We define the iterations $\langle j_{n,m} \colon M_n\to M_m \mid n < m \leq \omega \rangle $ , and $\langle \iota _{n,m} \colon N_n \to N_m \mid n < m \leq \omega \rangle $ as before.

By standard arguments (see for example [Reference Cummings2, Lemma 8.5 and Proposition 15.1], and use Claim 6.7 for the computation of the width), there is an $M_1$ -generic filter $K\subseteq \operatorname {Col}(\kappa ^{+3}, <j_E(\kappa ))$ . Moreover, by elementarity for every n, the filter $j_n(K)$ is $M_{n+1}$ -generic. Fix $K_0 \subseteq \operatorname {Col}(\omega _4, <\kappa )$ a V-generic and let

$$\begin{align*}\mathcal{K} = \langle K_0\rangle^\smallfrown \langle j_n(K) \mid n < \omega\rangle.\end{align*}$$

Let us define the filters $G_n, \mathcal {H}_n$ and the models $M_n$ for $n \leq \omega $ exactly as before and let us consider the model $M_{\omega }[G_\omega ][\mathcal {H}][\mathcal {K}]$ . Let $K_n = \mathcal {K} \restriction n + 1$ .

In [Reference Hayut12], it is shown that the parallel of Lemma 6.5, namely $M_\omega [G_\omega ][\mathcal {K}] = \bigcap M_n[G_n][K_n]$ , holds.

The next point in the proof in which we need to modify the argument slightly is Lemma 6.12, in which we need to show

$$\begin{align*}M_\omega[G_\omega][\mathcal{K}][\mathcal{H}] = \bigcap M_n[G_n][K_n][\mathcal{H}_n].\end{align*}$$

As there are elementary embeddings

$$\begin{align*}j_{n, \omega} \colon M_n[\mathcal{H}_n] \to M_\omega[<j_{n,\omega}"\mathcal{H}_n>] \subseteq M_\omega[\mathcal{H}],\end{align*}$$

the argument of [Reference Hayut12] for extender-based Prikry forcing with interleaved collapses works without further modifications.

The claims for the existence of a generic for the forcing that shoots a club disjoint from $j_\omega (S^{\kappa ^{+}}_{\kappa })$ , Claim 6.13, remains the same. We are now ready for the proof of the main theorem of this subsection.

Theorem 6.22. In $M_{\omega }[G_\omega ][\mathcal {H}][\mathcal {K}]\ \operatorname {\mathrm {SCH}}$ fails at $j_{\omega }(\kappa )$ which is $\aleph _{\omega }$ and every finite collection of stationary subsets of $j_{\omega }(\kappa ^{+})$ reflects simultaneously.

Proof Let us verify first that the only cardinals which are collapsed are the cardinals in the intervals $[\omega _4, \kappa ) \cup [\kappa ^{+3}, j_1(\kappa )) \cup \cdots [j_n(\kappa ^{+3}), j_{n+1}(\kappa )) \cup \cdots $ .

As mentioned above,

$$\begin{align*}M_{\omega}[G_\omega][\mathcal{H}][\mathcal{K}] = \bigcap M_{n}[G_n][\mathcal{H}_n][K_n]\end{align*}$$

and in particular, every $M_\omega $ -cardinal below $j_\omega (\kappa )$ which is collapsed in the extension must be collapsed in $M_{n}[G_n][\mathcal {H}_n][K_n]$ for all large n. For cardinals above $j_{\omega }(\kappa )$ one can use the continuity of the elementary embeddings in order to show that they do not change cofinality. Therefore, the arguments for the failure of $\operatorname {\mathrm {SCH}}$ remains the same.

Let us prove the stationary reflection. For simplicity, let us show that every stationary subset of $j_\omega (\kappa ^{+})$ reflects in this model. The general case is similar.

There are two components in the proof which are slightly different: First, given a stationary set $S \in M_{\omega }[G_\omega ][\mathcal {H}][\mathcal {K}]$ , one can pull it to a set $T_n \in M_n[G_n][\mathcal {H}_n][K_n]$ exactly as before and check whether it is stationary there. Note that the forcing adding $K_n$ is of cardinality $j_n(\kappa )$ . So, if $T_n$ is stationary in $M_n[G_n][\mathcal {H}_n][K_n]$ then there is a single condition $q \in K_n$ and stationary set $T^{\prime }_n \in M_n[G_n][\mathcal {H}_n]$ such that $q \Vdash T^{\prime }_n \subseteq T_n$ .

Then, by the $j_n(\kappa ^{+})$ -subcompactness of $j_n(\kappa )$ in $M_n[G_n][\mathcal {H}_n]$ one can conclude that $T_n^{\prime }$ has a reflection point $\delta $ of cofinality $<j_n(\kappa )$ and above $j_{n-1}(\kappa )^{+++}$ (see [Reference Hayut and Unger13] for details). This is where the assumption about the approachability plays a role: we conclude that this stationary subset of $\delta $ remains stationary in $M_n[G_n][\mathcal {H}_n][\mathcal K_n]$ . So, there is a stationary subset of $\delta $ of order type $\operatorname {\mathrm {cf}} \delta $ in $M_n$ , $T_n^{\prime \prime }\subseteq T_n \cap \delta $ . Since $j_{n,\omega }(T_n^{\prime \prime }) = j_{n,\omega } \text { " } T_n^{\prime \prime }$ , we conclude that $M_{\omega }[G_\omega ][\mathcal {H}][K_n]\models j_{n,\omega }(T_n^{\prime \prime }) \subseteq S$ .

By downwards absoluteness of stationarity and continuity of the embedding $j_{n,\omega }$ , $j_{n,\omega }(T_n^{\prime \prime })$ is stationary at $j_{n,\omega }(\delta )$ in $M_{\omega }[G_\omega ][\mathcal {H}][K_n]$ and thus S reflects at $j_{n,\omega }(\delta )$ in this model.

To summarize, if there is n such that $T_n$ is stationary, then S reflects in $M_\omega [G_\omega ][\mathcal {H}][\mathcal {K}]$ .

Otherwise, for all n, $T_n$ is non-stationary. In this case, for each n there is a club name for a club $\dot {C}_n$ disjoint from $T_n$ and we would like to push them (or subclubs of them) to $M_{\omega }[G_\omega ][\mathcal {H}][\mathcal {K}]$ .

For every n, the forcing introducing $K_n$ is $j_n(\kappa )$ -cc, and therefore, for every club $C_n \in M_n[G_n][\mathcal {H}_n][K_n]$ there is a subclub $C^{\prime }_n \in M_n[G_n][\mathcal {H}_n]$ . We are now in exactly the same situation of the proof of Theorem 1.4: we obtain for each n a stable name $\dot {E}_n$ which is forced to be a subclub of $C^{\prime }_n$ and thus there is in $M_{\omega }[G_\omega ][\mathcal {H}]$ a subclub of the set $\bigcap j_{n,\omega }(\dot {E}_n)$ which is forced to be disjoint from S.

Remark 6.23. The same method can be used to obtain arbitrary finite gap in the failure of $\operatorname {\mathrm {SCH}}$ .

Funding

The first author was partially supported by the Israel Science Foundation Grant 1832/19. The second author was partially supported by the FWF Lise Meitner Grant 2650-N35 and the Israel Science Foundation Grant 1967/21. The third author was partially supported by the NSF grant DMS-1700425 and an NSERC Discovery grant.

Footnotes

1 After proving this theorem, Gitik also gave an alternative proof for the same result (see [Reference Gitik10]).

2 This theorem did not appear in the original version of this paper.

3 This gives a clean way to distinguish between Cohen functions associated with members of $s^p$ and those which are not.

4 This is true by [Reference Cummings2, Proposition 15.1], since the width of the embedding is smaller than its distributivity.

5 Indeed, for every set $a \in N^F$ of cardinality $\leq \iota _k(\kappa _{\beta _k})$ , we modify the generic at most at $\kappa _{\beta _k}$ many coordinates.

References

Bukovský, L., Iterated ultrapower and Prikry’s forcing . Commentationes Mathematicae Universitatis Carolinae , vol. 18 (1977), no. 1, pp. 7785.Google Scholar
Cummings, J., Iterated forcing and elementary embeddings , Handbook of Set Theory , vols. 1–3 , Springer, Dordrecht, 2010, pp. 775883.CrossRefGoogle Scholar
Dehornoy, P., Iterated ultrapowers and Prikry forcing . Annals of Mathematical Logic , vol. 15 (1978), no. 2, pp. 109160.CrossRefGoogle Scholar
Foreman, M. and Todorcevic, S., A new Löwenheim–Skolem theorem . Transactions of the American Mathematical Society , vol. 357 (2005), no. 5, pp. 16931715.CrossRefGoogle Scholar
Gitik, M., The negation of the singular cardinal hypothesis from $o(\kappa )={\kappa}^{++}$ . Annals of Pure and Applied Logic , vol. 43 (1989), no. 3, pp. 209234.CrossRefGoogle Scholar
Gitik, M., Prikry-type forcings , Handbook of Set Theory , vols. 1–3 , Springer, Dordrecht, 2010, pp. 13511447.CrossRefGoogle Scholar
Gitik, M., Another model with the tree property and not SCH, preprint, 2018.Google Scholar
Gitik, M., Blowing up the power of a singular cardinal of uncountable cofinality, Journal of Symbolic Logic, vol. 84 (2019), no. (4), pp. 1722–1743.Google Scholar
Gitik, M., Another method for constructing models of not approachability and not SCH . Archive for Mathematical Logic , vol. 60 (2021), pp. 469475.CrossRefGoogle Scholar
Gitik, M., Reflection and not SCH with overlapping extenders . Archive for Mathematical Logic , vol. 61 (2022), nos. 5–6, pp. 591597.CrossRefGoogle Scholar
Gitik, M. and Sharon, A., On SCH and the approachability property . Proceedings of the American Mathematical Society , vol. 136 (2008), no. 1, pp. 311320.CrossRefGoogle Scholar
Hayut, Y., Prikry type forcings and the Bukovský–Dehornoy phenomena, preprint, 2023, arXiv:2308.05553.Google Scholar
Hayut, Y. and Unger, S., Stationary reflection . Journal of Symbolic Logic , vol. 85 (2020), no. 3, pp. 937959.Google Scholar
Magidor, M., Reflecting stationary sets, Journal of Symbolic Logic, vol. 47 (1982), no. 4, pp. 755–771 (1983).CrossRefGoogle Scholar
Merimovich, C., Prikry on extenders, revisited . Israel Journal of Mathematics , vol. 160 (2007), pp. 253280.CrossRefGoogle Scholar
Merimovich, C., Extender-based Magidor–Radin forcing . Israel Journal of Mathematics , vol. 182 (2011), pp. 439480.CrossRefGoogle Scholar
Merimovich, C., Mathias like criterion for the extender based Prikry forcing . Annals of Pure and Applied Logic , vol. 172 (2021), no. 9, Article no. 102994, 6 pp.CrossRefGoogle Scholar
Neeman, I., Aronszajn trees and failure of the singular cardinal hypothesis . Journal of Mathematical Logic , vol. 9 (2009), no. 1, pp. 139157.CrossRefGoogle Scholar
Poveda, A., Rinot, A., and Sinapova, D., Sigma–Prikry forcing I: The axioms . Canadian Journal of Mathematics , vol. 73 (2021), no. 5, pp. 12051238.CrossRefGoogle Scholar
Poveda, A., Rinot, A., and Sinapova, D., Sigma–Prikry forcing III: Down to ${\mathrm{\aleph}}_{\omega },\mathrm{preprint}$ , 2022, arXiv:2209.10501.CrossRefGoogle Scholar
Sharon, A., Weak squares, scales, stationary reflection and the failure of sch , Ph.D. thesis, Tel Aviv University, 2005.Google Scholar
Shelah, S., Reflection implies the SCH . Fundamenta Mathematicae , vol. 198 (2008), no. 2, pp. 95111.CrossRefGoogle Scholar
Sinapova, D., A model for a very good scale and a bad scale, Journal of Symbolic Logic, vol. 73 (2008), no. 4, pp. 1361–1372.Google Scholar
Sinapova, D., The tree property and the failure of SCH at uncountable cofinality . Archive for Mathematical Logic , vol. 51 (2012), nos. 5–6, pp. 553562.CrossRefGoogle Scholar
Sinapova, D., The tree property and the failure of the singular cardinal hypothesis at ${\mathrm{\aleph}}_{\omega^2},$ Journal of Symbolic Logic, vol. 77 (2012), no. 3, pp. 934–946.Google Scholar
Sinapova, D. and Unger, S., Scales at ${\mathrm{\aleph}}_{\omega }$ . Israel Journal of Mathematics , vol. 209 (2015), no. 1, pp. 463486.CrossRefGoogle Scholar