Hostname: page-component-78c5997874-fbnjt Total loading time: 0 Render date: 2024-10-28T15:38:38.732Z Has data issue: false hasContentIssue false

$\mathscr {I}$-ULTRAFILTERS IN THE RATIONAL PERFECT SET MODEL

Part of: Set theory

Published online by Cambridge University Press:  12 December 2022

JONATHAN CANCINO-MANRÍQUEZ*
Affiliation:
INSTITUTE OF MATHEMATICS, CZECH ACADEMY OF SCIENCES, ŽITNÁ 25, 115 67 PRAHA 1, CZECH REPUBLIC; CENTRO DE CIENCIAS MATEMÁTICAS, UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO, A.P. 61-3, XANGARI, MORELIA, MICHOACÁN MÉXICO E-mail: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We give a new characterization of the cardinal invariant $\mathfrak {d}$ as the minimal cardinality of a family $\mathcal {D}$ of tall summable ideals such that an ultrafilter is rapid if and only if it has non-empty intersection with all the ideals in the family $\mathcal {D}$. On the other hand, we prove that in the Miller model, given any family $\mathcal {D}$ of analytic tall p-ideals such that $\vert \mathcal {D}\vert <\mathfrak {d}$, there is an ultrafilter $\mathcal {U}$ which is an $\mathscr {I}$-ultrafilter for all ideals $\mathscr {I}\in \mathcal {D}$ at the same time, yet $\mathcal {U}$ is not a rapid ultrafilter. As a corollary, we obtain that in the Miller model, given any analytic tall p-ideal $\mathscr {I}$, $\mathscr {I}$-ultrafilters are dense in the Rudin–Blass ordering, generalizing a theorem of Bartoszyński and S. Shelah, who proved that in such model, Hausdorff ultrafilters are dense in the Rudin–Blass ordering. This theorem also shows some limitations about possible generalizations of a theorem of C. Laflamme and J. Zhu.

Type
Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

Let us recall some basic definitions. Given $f,g\in \omega ^{\omega }$ , we say that $f\leq ^*g$ if the set $\{n\in \omega : g(n)<f(n)\}$ is finite. Inclusion modulo finite is denoted by $\subseteq ^*$ : $A\subseteq ^*B$ if and only if $A\setminus B$ is finite. A filter on $\omega $ is a family $\mathcal {F}\subseteq \mathcal {P}(\omega )$ which is closed under finite intersections and supersets. An ultrafilter is a maximal filter with respect to the inclusion relation $\subseteq $ . An ideal is a family $\mathscr {I}\subseteq \mathcal {P}(\omega )$ which is closed under finite unions and under subsets. We assume that all our filters are free, that is, for any filter $\mathcal {F}$ , $\bigcap \mathcal {F}=\emptyset $ . On the other hand, we assume all our ideals are tall, which means that for any $A\in [\omega ]^{\omega }$ , there is an infinite $B\subseteq A$ such that $B\in \mathscr {I}$ . Also, we require that any ideal contains all the finite subsets of $\omega $ . If for any $\langle A_n:n\in \omega \rangle \subseteq \mathscr {I}$ there is $B\in \mathscr {I}$ such that $A_n\subseteq ^*B$ for all $n\in \omega $ , we say that $\mathscr {I}$ is a p-ideal.

The notion of filter and ideal are dual to each other in the following sense: let $\mathcal {F}$ be a filter on $\omega $ , then $\mathcal {F}^*=\{\omega \setminus A:A\in \mathcal {F}\}$ is an ideal. On the other hand, if $\mathscr {I}$ is an ideal on $\omega $ , then $\mathscr {I}^*=\{\omega \setminus A:A\in \mathscr {I}\}$ is a filter. For any ideal $\mathscr {I}$ , the family of $\mathscr {I}$ -positive sets, denoted by $\mathscr {I}^+$ , is the collection of all sets which are not elements of $\mathscr {I}$ , that is, $\mathscr {I}^+=\mathcal {P}(\omega )\setminus \mathscr {I}$ .

An ultrafilter $\mathcal {U}$ on $\omega $ is rapid if for any function $f\in \omega ^{\omega }$ , there is $A\in \mathcal {U}$ such that $f\leq ^* e_A$ , where $e_A$ is the increasing enumeration of A. An ideal $\mathscr {I}\subseteq \mathcal {P}(\omega )$ is summable if and only if there is a function $g:\omega \to \mathbb {R}^+$ such that for all $A\in \mathcal {P}(\omega )$ , $A\in \mathscr {I}$ if and only of $\sum _{n\in A}g(n)<\infty $ .

Let us recall the following nice characterization of rapid ultrafilters due to P. Vojtáš:

Theorem 1.1 (Vojtáš, see [Reference Vojtáš15]).

An ultrafilter $\mathcal {U}$ on $\omega $ is rapid if and only if $\mathcal {U}$ has non-empty intersection with every summable ideal.

In [Reference Flašková8], among several results, J. Flašková proved the following refinement of Vojtáš theorem:

Theorem 1.2 (Flašková, see [Reference Flašková8]).

There is a family $\mathcal {D}$ of summable ideals with cardinality $\mathfrak {d}$ Footnote 1 such that an ultrafilter is rapid if and only if it has non-empty intersection with every ideal from $\mathcal {D}$ .

Then she asked the following question:

Question 1.3 (Flašková, see [Reference Flašková8]).

What is the minimal size of a family $\mathcal {D}$ of summable ideals such that an ultrafilter $\mathcal {U}$ is rapid if and only if it has non-empty intersection with each ideal in $\mathcal {D}$ ?

Let $\mathscr {I}$ be an ideal on $\omega $ . Given an ultrafilter $\mathcal {U}$ on $\omega $ , we say that $\mathcal {U}$ is an $\mathscr {I}$ -ultrafilter if for any function $f:\omega \to \omega $ there is $A\in \mathcal {U}$ such that $f[A]\in \mathscr {I}$ . If we only require the previous property to hold only for finite to one functions, we say that $\mathcal {U}$ is a weak $\mathscr {I}$ -ultrafilter. The definition of $\mathscr {I}$ -ultrafilter was introduced by J. Baumgartner in [Reference Baumgartner3], and it has proved to be very useful in the classification of combinatorial properties of ultrafilters.

In relation to $\mathscr {I}$ -ultrafilters, J. Flašková makes the following question which is a strengthening of the previous one. $\mathcal {D}$ denotes a family of summable ideals.

Question 1.4 (Flašková, see [Reference Flašková8]).

Is it true that whenever the cardinality of $\mathcal {D}$ is less than $\mathfrak {d}$ , then there exist an ultrafilter on the natural numbers which is an $\mathscr {I}_g$ -ultrafilter for every $\mathscr {I}_g\in \mathcal {D}$ , but not a rapid ultrafilter?

By Theorem 1.2, we know that the answer to Question 1.3 is at most $\mathfrak {d}$ . In Section 3 of this paper we prove that in fact the equality holds. In Sections 46, we establish the consistency in the positive way of Question 1.4, showing that in the Rational Perfect set model, for any family $\mathcal {D}$ of summable ideals with cardinality less than $\mathfrak {d}$ , there is an ultrafilter which is an $\mathscr {I}$ -ultrafilter for every $\mathscr {I}\in \mathcal {D}$ , although there is no rapid ultrafilter in such model. This is done in two steps: the first one is proving that after forcing with the Rational Perfect set forcing, p-points which are $\mathscr {I}$ -ultrafilters continue to still generate p-point which are $\mathscr {I}$ -ultrafilters, whenever the ideal $\mathscr {I}$ is an analytic p-ideal. The second step corresponds to prove a preservation theorem along countable support iterations of proper forcings which preserves p-points which are $\mathscr {I}$ -ultrafilters, for any $F_{\sigma } \ p$ -ideal $\mathscr {I}$ . It turns out that the answer to Question 1.4 is independent of $\mathsf {ZFC}$ : in [Reference Cancino-Manríquez7], it has been proved that it is relatively consistent with $\mathsf {ZFC}$ that there is no $\mathscr {I}$ -ultrafilter for any $F_{\sigma }$ ideal $\mathscr {I}$ , which implies the consistency of a negative answer to Question 1.4.

As a consequence of our results, we generalize a theorem of T. Bartoszyński and S. Shelah. They proved that in the Rational Perfect set model, Hausdorff ultrafilters are dense in the Rudin–Blass ordering (see [Reference Bartoszyński and Shelah2]). We prove that in the same model, $\mathscr {I}$ -ultrafilters are dense, for any analytic p-ideal $\mathscr {I}$ . This result has the interesting consequence that it can not be proved, in $\mathsf {ZFC}$ alone, the existence of an ultrafilter $\mathcal {U}$ none of whose Rudin–Blass predecessors are $\mathscr {I}$ -ultrafilters for any analytic p-ideal. This result contrasts with a theorem of C. Laflamme and J. Zhu, which states, in $\mathsf {ZFC}$ , the existence of an ultrafilter $\mathcal {U}$ none of whose Rudin–Blass predecessors are rapid ultrafilters (see [Reference Laflamme and Zhu11]). In particular, none of them are weak $\mathcal {ED}_{fin}$ -ultrafilters, i.e., q-points. In Section 7, we discuss these facts in more detail.

2 Preliminaries

Let us recall that a family of functions $\mathcal {D}\subseteq \omega ^{\omega }$ is a dominating family, if for any $g\in \omega ^{\omega }$ , there is $f\in \mathcal {D}$ such that $g\leq ^*f$ . The dominating number is defined as the minimal possible cardinality of a dominating family:

$$ \begin{align*} \mathfrak{d}=\min\{\vert\mathcal{D}\vert:(\mathcal{D}\subseteq\omega^{\omega})(\mathcal{D} \text{ is a dominating family})\}. \end{align*} $$

An ideal $\mathscr {I}$ on $\omega $ is a p-ideal if for any $\langle A_n:n\in \omega \rangle \subseteq \mathscr {I}$ there is $B\in \mathscr {I}$ such that for all $n\in \omega $ , $A_n\subseteq ^* B$ .

An ultrafilter on $\omega $ is called a p-point if for any sequence $\langle A_n:n\in \omega \rangle $ of elements from $\mathcal {U}$ , there is $A\in \mathcal {U}$ such that for all $n\in \omega $ , $A\subseteq ^*A_n$ . There is a nice characterization of p-points by mean of infinite games which we will make use of:

Definition 2.1 (p-point game).

Given an ultrafilter $\mathcal {U}$ , the two players game $G(\mathcal {U})$ is defined as a sequence of choices were, at round n, Player I chooses an element from the ultrafilter, $A_n\in \mathcal {U}$ , and Player II chooses a finite subset $F_n\in [A_n]^{<\omega }$ , and Player II wins if and only if $\bigcup _{n\in \omega } F_n\in \mathcal {U}$ :

The following well known characterization of p-points will be used in the proof of Theorem 4.8.

Lemma 2.2 (Laflamme, see [Reference Laflamme10]).

Let $\mathcal {U}$ be an ultrafilter on $\omega $ . Then $\mathcal {U}$ is a p-point if and only if Player I has no winning strategy in the p-point game.

Proof The reader can find a proof in Laflamme’s article [Reference Laflamme10]. Another good reference is found in Wohofsky master’s thesis [Reference Wohofsky16, pp. 53–59].

Let us recall the Rational Perfect set forcing. The family of all finite sequences of natural numbers will be denoted by $\omega ^{<\omega }$ . For any $s\in \omega ^{<\omega }$ , the length of the sequence s is defined as its cardinaliy, and will be denoted by $\vert s\vert $ . For $s\in \omega ^{<\omega }$ and $i\in \omega $ , denote by $s^{\frown } i$ the sequence obtained by adding i to the end of the sequence s. A subset $T\subseteq \omega ^{<\omega }$ is a tree if for any $s\in T$ and $i\leq \vert s\vert $ , we have $s\upharpoonright i\in T$ . For a tree $T\subseteq \omega ^{<\omega }$ and $s\in T$ , we say that $i\in \omega $ is an immediate successor of s in T if $s^{\frown } i\in T$ . The set of all immediate successors of s in T will be denoted by $succ_{T}(s)=\{i\in \omega :i\text { is an immediate successor of}\ s \text { in } T\}$ . A node $s\in T$ is an splitting node if it has more than one successor in T, that is $\vert succ_T(s)\vert>1$ , and the set of all splitting nodes of T is denote by $split(T)$ . For a tree T and a node $s\in T$ , $T\upharpoonright s$ denotes the tree of all nodes in T which are $\subseteq $ -comparable with s.

A superperfect tree tree $T\subseteq \omega ^{<\omega }$ is a non-empty tree satisfying the following properties:

  1. (1) For all $s\in T$ , there is $t\in split(T)$ such that $s\subseteq t$ .

  2. (2) For all $s\in split(T)$ , s has infinitely many immediate successors.

The Rational Perfect set forcing, denoted by $\mathbf {PT}$ , is the partial order whose members are all the superperfect trees, with the order given by set inclusion, that is $S\leq T$ if and only if $S\subseteq T$ .

For $T\in \mathbf {PT}$ , the stem of T, denoted by $st(T)$ , is defined as the unique node in $split(T)$ of minimal length.

Given a superperfect tree T, define an order isomorphism between $\omega ^{<\omega }$ and $split(T)$ , which we denote by $\varphi _T$ , as follows: $\varphi _T(\emptyset )=st(T)$ ; assume $\varphi _T(s)$ is defined and let $\{i_k:k\in \omega \}$ be the increasing enumeration of $succ_{T}(\varphi _T(s))$ , for each $k\in \omega $ let $r_{s,k}\in split(T)$ be the node of minimal length extending $\varphi _T(s)^{\frown } i_k$ , and define $\varphi _T(s^{\frown } k)=r_{s,k}$ .

An ideal $\mathscr {I}$ on $\omega $ is a summable ideal if there exists $g:\omega \to \mathbb {R}^+$ such that $A\in \mathscr {I}$ if and only if $\sum _{n\in A}g(n)<\infty $ . For a given $g:\omega \to \mathbb {R}^+$ , we denote by $\mathscr {I}_g$ the summable ideal defined by g. Note that since we are assuming all our ideals are tall, it should hold that $\lim _{n\to \infty }g(n)=0$ whenever g defines a summable ideal.

Recall that a function $\varphi :\mathcal {P}(\omega )\to [0,\infty ]$ is a lower semicontinuous submeasure, lscsm for short, if it satisfies the following properties:

  1. (1) $\varphi (\emptyset )=0$ .

  2. (2) For all $A,B\in \mathcal {P}(\omega )$ , if $A\subseteq B$ , then $\varphi (A)\leq \varphi (B)$ .

  3. (3) For all $A,B\in \mathcal {P}(\omega )$ , $\varphi (A\cup B)\leq \varphi (A)+\varphi (B)$ .

  4. (4) For all $A\in \mathcal {P}(\omega )$ , $\lim _{n\to \infty }\varphi (A\cap n)=\varphi (A)$ .

Given $\varphi $ a lscsm, there are two ideals associated with $\varphi $ :

  1. (1) $Fin(\varphi )=\{A\in \mathcal {P}(\omega ):\varphi (A)<\infty \}$ .

  2. (2) $Exh(\varphi )=\{A\in \mathcal {P}(\omega ):\lim _{n\to \infty }\varphi (A\setminus n)=0\}$ .

The following well known theorem will be necessary in Sections 46:

Theorem 2.3.

  1. (1) (Solecki, see [Reference Solecki14]) An ideal $\mathscr {I}$ is an analytic p-ideal if and only if there is a lscsm $\varphi $ such that $\mathscr {I}=Exh(\varphi )$ .

  2. (2) (Mazur, see [Reference Mazur12]) An ideal $\mathscr {I}$ is an $F_{\sigma }$ ideal if and only if there is a lscsm $\varphi $ such that $\mathscr {I}=Fin(\varphi )$ .

  3. (3) (Solecki, see [Reference Solecki14]) An ideal $\mathscr {I}$ is an $F_{\sigma } \ p$ -ideal if and only if there is a lscsm $\varphi $ such that $\mathscr {I}=Fin(\varphi )=Exh(\varphi )$ .

Note that given a lscsm $\varphi :\mathcal {P}(\omega )\to [0,\infty ]$ , the lower semicontinuity implies that $\varphi $ is determined by its values on the finite sets, in the sense that for any $A\in [\omega ]^{\omega }$ , $\varphi (A)=\lim _{n\to \infty }\varphi (A\cap n)$ . In this way, if $\varphi $ is defined in the ground model V, and $V[\dot {G}]$ is a generic extension of V, we can extend $\varphi $ to the whole $\mathcal {P}(\omega )^{V[\dot {G}]}$ by setting $\varphi (A)=\lim _{n\to \infty }\varphi (A\cap n)$ , for any $A\in \mathcal {P}(\omega )^{V[\dot {G}]}$ . Then, in $V[\dot {G}]$ , $\varphi $ defines an $F_{\sigma }$ ideal which extends $Fin(\varphi )^{V}$ , and moreover, the $\varphi $ -measure computed in V and $V[\dot {G}]$ agrees for any $A\in \mathcal {P}(\omega )^{V}$ . A similar remark holds for $F_{\sigma \delta }$ ideals of the form $Exh(\varphi )$ . Following this remark, whenever we have an $F_{\sigma }$ (resp. $F_{\sigma \delta }$ ) ideal (resp. p-ideal) $\mathscr {I}$ in the ground model, and we talk about such ideal in a generic extension, we mean the ideal computed by any lscsm which defines $\mathscr {I}$ . A similar remark holds for ideals arising along a forcing iteration.

Finally, let us recall some concepts and results about the theory of proper forcing.

Definition 2.4. Let $\mathbb {P}$ be a forcing and $\theta>2^{\vert \mathbb {P}\vert }$ a cardinal. Let $\mathcal {M}\prec H(\theta )$ be countable elementary submodel such that $\mathbb {P}\in \mathcal {M}$ . A condition $q\in \mathbb {P}$ is called an $(\mathcal {M},\mathbb {P})$ -generic condition if $\mathbb {P}\cap \mathcal {M}$ is predense below q.

Definition 2.5. Let $\mathbb {P}$ be a forcing notion. $\mathbb {P}$ is proper if for any $\theta> 2^{\vert \mathbb {P}\vert }$ and any countable elementary submodel $\mathcal {M}\prec H(\theta )$ such that $\mathbb {P}\in \mathcal {M}$ , any condition $p\in \mathcal {M}$ has an $(\mathcal {M},\mathbb {P})$ -generic extension $q\leq p$ .

The following well known results will be used in the proofs of Lemma 5.5 and Theorem 5.1 without explicit mention.

Lemma 2.6. Let $\mathbb {P}*\dot {\mathbb {Q}}$ be a two step iteration forcing, $\theta $ a sufficiently large cardinal, and let $\mathcal {M}\prec H(\theta )$ be countable elementary submodel such that $\mathbb {P}*\dot {\mathbb {Q}}\in \mathcal {M}$ . A condition $(p,\dot {q})$ is an $(\mathcal {M},\mathbb {P}*\dot {\mathbb {Q}})$ -generic condition if and only if p is $(\mathcal {M},\mathbb {P})$ -generic and

$$ \begin{align*} p\Vdash` `\dot{q}\text{ is }(\mathcal{M}[\dot{G}],\dot{\mathbb{Q}})\text{-generic}", \end{align*} $$

where $\dot {G}$ is the canonical name for the generic filter for $\mathbb {P}$ over V.

Theorem 2.7. Let $\mathbb {P}=\langle \mathbb {P}_{\alpha },\dot {\mathbb {Q}}_{\alpha }:\alpha <\gamma \rangle $ be a countable support iteration such that $\mathbb {P}_0$ is proper and for all $\alpha <\gamma $ , $\mathbb {P}_{\alpha }\Vdash ` `\dot {\mathbb {Q}}_{\alpha }\text { is proper}"$ . Then $\mathbb {P}$ is proper.

Theorem 2.8 (Blass and Shelah, see [Reference Blass and Shelah5]).

Let $\kern1.5pt\mathcal {U}$ be a p-point, and let $\mathbb {P}=\langle \mathbb {P}_{\alpha },\dot {\mathbb {Q}}_{\alpha }:\alpha <\gamma \rangle $ be a countable support iteration of proper forcings such that for all $\alpha <\gamma $ , $\mathbb {P}_{\alpha }\Vdash ` `\dot {\mathbb {Q}}_{\alpha }\text { forces that } \mathcal {U} \text { generates a } p\text {-point}"$ . Then after forcing with $\mathbb {P}$ , $\mathcal {U}$ generates a p-point.

3 $\mathfrak {d}$ is the best

For the sake of simplicity, let us say that a family $\mathcal {D}$ of summable ideals is a Vojtáš family if an ultrafilter is rapid if and only if it has non-empty intersection with each member from $\mathcal {D}$ . The main theorem of this section can be stated as follows:

Theorem 3.1. $\mathfrak {d}$ is equal to the minimum cardinality of a Vojtáš family.

Since Theorem 1.2 means that $\mathfrak {d}$ is an upper bound for the minimum cardinality of a Vojtáš family, we only have to prove the following:

Proposition 3.2. There is no Vojtáš family $\mathcal {D}$ of cardinality strictly smaller than $\mathfrak {d}$ .

Proof It is enough to prove that given any non-empty family of summable ideals of cardinality less than $\mathfrak {d}$ , there is an ultrafilter meeting each ideal in the family, but still there is a tall summable ideal having empty intersection with the ultrafilter. Fix $\mathcal {D}=\{\mathscr {I}_{\alpha }:\alpha <\lambda \}$ an arbitrary family of tall summable ideals such that $\lambda <\mathfrak {d}$ , and for $\alpha \in \lambda $ , let $g_{\alpha }:\omega \to \mathbb R$ be a function such that $\mathscr {I}_{\alpha }=\{A\in \mathcal {P}(\omega ):\sum _{n\in A}g_{\alpha }(n)<\infty \}$ . Let E be the set of even natural numbers. For each $F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}$ , define a function $\psi _F$ on E as follows:

$$ \begin{align*} &\qquad\qquad\qquad\qquad\qquad\qquad \psi_F(0)=0,\\ &\psi_{F}(n+2)=\min\{k\in\omega:k>\psi_{F}(n)\ \text{and}\ (\forall\alpha\in F)(\forall i\ge k)\\ &\qquad\qquad\qquad\qquad\quad(g_{\alpha}(i)\leq1/2^{n+2}(n+2)^{2})\}. \end{align*} $$

Since $\{\psi _{F}:F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}\}$ has cardinality less than $\mathfrak {d}$ , there is a function $f:E\to \omega $ which is not dominated by $\{\psi _{F}:F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}\}$ . We can assume that for all $n\in E$ , $f(n+2)-f(n)>n^{2}$ , and we can define the following function:

$$ \begin{align*} \tilde{f}(n)=\begin{cases} f(n),\text{ if }n\in E,\\ f(n-1)+(n-1)^{2}, \quad \text{if }n\notin E. \end{cases} \end{align*} $$

For each $F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}$ , let $A_{F}$ be the set where $\psi _{F}$ is dominated by f, that is $A_{F}=\{n\in E:\psi _{F}(n)\leq f(n)\}$ . Note that the family $\{A_{F}:F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}\}$ is a centered family. Indeed, pick $F_{0},\ldots ,F_{n}\in [\lambda ]^{<\omega }\setminus \{\emptyset \}$ . By the definition of the functions $\psi _{F}$ we have that if $F\subseteq G$ , then $\psi _{F}\leq \psi _{G}$ , so in particular, if $\psi _{F_{0}\cup \cdots \cup F_{n}}(k)\leq f(k)$ , then we have $\psi _{F_{i}}(k)\leq f(k)$ for $i=0,\ldots ,n$ , and so $k\in A_{0}\cap \cdots \cap A_{n}$ .

Now, for $F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}$ let $B_{F}$ be the set $\bigcup _{k\in A_{F}}[\tilde {f}(k),\tilde{f}(k+1))$ . Finally, take the family $\mathcal {G}=\{B_{F}:F\in [\lambda ]^{<\omega }\setminus \{\emptyset \}\}$ . It follows that $\mathcal {G}$ is a centered family, by the previous paragraph.

Let us see that $\mathcal {G}$ has non-empty intersection with each ideal in $\mathcal {D}$ . For $\alpha \in \lambda $ , we claim that $B_{\{\alpha \}}\in \mathscr {I}_{\alpha }$ . This follows easily from the definition of ${\psi _{\{\alpha \}}}$ and ${A_{\{\alpha \}}}$ . We have that for all ${n\in A_{\{\alpha \}}}$ , ${\psi _{\{\alpha \}}(n)\leq f(n)=\tilde {f}(n)}$ , and by the definition of ${\psi _{\{\alpha \}}}$ , for every ${k\ge \psi _{\{\alpha \}}(n)}$ , ${g_{\alpha }(k)\leq \frac {1}{2^{n}n^{2}}}$ , so in particular, for every ${n\in A_{\{\alpha \}}}$ and every ${k\ge \tilde {f}(n)},{\ g_{\alpha }(k)\leq \frac {1}{2^{n}n^{2}}}$ . We have the following:

$$ \begin{align*} \sum_{n\in B_{\{\alpha\}}} g_{\alpha}(n)&=\sum_{n\in A_{\{\alpha\}}} \left[\sum_{m\in[\tilde{f}(n),\tilde{f}(n+1))}g_{\alpha}(m)\right]\\ &\leq \sum_{n\in A_{\{\alpha\}}}\left[\sum_{m\in[\tilde{f}(n),\tilde{f}(n+1))} \frac{1}{2^n n^2}\right]\\ & = \sum_{n\in A_{\{\alpha\}}}\frac{n^2}{2^n n^2}\leq \sum_{n\in A_{\{\alpha\}}}\frac{1}{2^n}\leq 1. \end{align*} $$

What remains is to find a tall summable ideal $\mathscr {I}$ which has empty intersection with some ultrafilter extending $\mathcal G$ . Consider the following function from $\omega $ to $\mathbb {R}$ :

$$ \begin{align*} h(n)=\begin{cases} 1, \quad \text{if } n\in [0,\tilde{f}(0)),\\ \frac{1}{k+1}, \quad \text{if } n\in [\tilde{f}(k),\tilde{f}(k+1)). \end{cases} \end{align*} $$

Let $\mathscr {I}_{h}$ be the corresponding summable ideal. We claim that $\mathcal {G}\subseteq \mathscr {I}_{h}^{+}$ . Pick any $B_{F}\in \mathcal {G}$ . Then,

$$ \begin{align*} \sum_{n\in B_F} h(n) &= \sum_{n\in A_{F}}\left[\sum_{m\in[\tilde{f}(n),\tilde{f}(n+1))}h(m)\right]\\ & = \sum_{n\in A_{F}}\left[\sum_{m\in[\tilde{f}(n),\tilde{f}(n+1))}\frac{1}{n+1}\right]\\ & = \sum_{n\in A_{F}}\frac{n^2}{n+1}\longrightarrow\infty. \end{align*} $$

Then we have $\mathcal G\subseteq \mathscr {I}_h^+$ , so $\mathcal {G}\cup \mathscr {I}_h^*$ can be extended to an ultrafilter $\mathcal U$ . Obviously $\mathscr {I}_h$ and $\mathcal U$ satisfy our requirements.

Now, Theorem 3.1 follows immediately by Theorem 1.2 and Proposition 3.2.

4 Rational Perfect set forcing and p-ideals

Definition 4.1. Let $\mathscr {I}$ be an ideal on $\omega $ and let $\mathcal {U}$ be an ultrafilter on $\omega $ . We say that $\mathcal {U}$ is an $(\mathscr {I},p)$ -point provided $\mathcal {U}$ is an $\mathscr {I}$ -ultrafilter and a p-point at the same time.

Remark. One could also define weak $(\mathscr {I},p)$ -points as those ultrafilters $\mathcal {U}$ which are weak $\mathscr {I}$ -ultrafilters and p-points. However, it is easy to see that $(\mathscr {I},p)$ -points and weak $(\mathscr {I},p)$ -points are exactly the same, so for checking that $\mathcal {U}$ is an $(\mathscr {I},p)$ -point we need to check that $\mathcal {U}$ is a p-point and a weak $\mathscr {I}$ -ultrafilter. This remark will be used implicitly in the proofs of Theorems 4.8 and 5.1.

Recall from Section 2, the definition of the Rational Perfect set forcing, denoted by $\mathbf {PT}$ . Also, recall the definition of the isomorphism $\varphi _T:\omega ^{<\omega }\to split(T)$ . Recall that a forcing $\mathbb {P}$ satisfies the Axiom A if there exists a family of orderings $\{\leq _n:n\in \omega \}$ such that:

  1. (1) For all $n\in \omega $ , $\leq _{n+1}\subseteq \leq _n$ and $\leq _0\subseteq \leq $ .

  2. (2) For any maximal antichain $\mathcal {A}$ , any condition $p\in \mathbb {P}$ and any natural number $n\in \omega $ , there is $q\leq _n p$ such that the set $\{a\in \mathcal {A}:a\text { is compatible with }q\}$ is countable.

  3. (3) For any sequence $\langle p_n:n\in \omega \rangle $ of condition in $\mathbb {P}$ such that for all $n\in \omega $ it holds that $p_{n+1}\leq _{n}p_n$ , there is a condition $p_{\omega }$ such that for all $n\in \omega $ , $p_{\omega }\leq _n p_n$ .

Definition 4.2. Let $n\in \omega $ be a natural number, and let $T,S\in \mathbf {PT}$ be two conditions. We say that $S\leq _n T$ if and only if:

  1. (1) $S\leq T$ .

  2. (2) $\varphi _S(n^{\leq n})=\varphi _T(n^{\leq n})$ .

The following lemmas are well known properties of the Rational Perfect set forcing. We refer the reader to [Reference Miller13] for details about the proofs.

Lemma 4.3. The orders from Definition 4.2 give an Axiom A structure to the Rational Perfect set forcing. In particular, $\mathbf {PT}$ is proper.

Lemma 4.4. Let $\dot {f}$ be an $\mathbf {PT}$ -name for a function from $\omega $ to $\omega $ , $T\in \mathbf {PT}$ a condition and $g\in \omega ^{\omega }$ such that $T\Vdash ` `\dot {f}\leq g"$ . Then there is $T'\leq T$ , and a family $\{h_s:s\in split(T')\}\subseteq \omega ^{\omega }$ such that for all $s\in split(T')$ , for all $n\in \omega $ there is $k_n\in \omega $ such that for all $k\in succ_T(s)$ such that $k\ge k_n$ , it holds that $T'\upharpoonright s^{\frown } k\Vdash ` `\dot {f}\upharpoonright (\vert s\vert +n)=h_s\upharpoonright (\vert s\vert +n)"$ .

Lemma 4.5. Let $T\in \mathbf {PT}$ be a condition, and $A, B$ a partition of the splitting nodes of T. There is a stronger condition $T'\leq T$ such that $split(T')\subseteq A$ or $split(T')\subseteq B$ .

Theorem 4.6. $\mathbf {PT}$ preserves p-points.

The following lemma will be useful in the proof of Theorem 4.8, since it will allow us to use Lemma 4.4. For $f,g\in \omega ^{\omega }$ , the notation $f\leq g$ means that for all $n\in \omega $ , $f(n)\leq g(n)$ .

Lemma 4.7. Let $\mathscr {I}$ be an analytic p-ideal. Then there exists a function $g:\omega \to \omega $ such that for any ultrafilter $\mathcal {U}$ , $\mathcal {U}$ is a weak $\mathscr {I}$ -ultrafilter if and only if for all finite to one $f\leq g$ there is $A\in \mathcal {U}$ such that $f[A]\in \mathscr {I}$ .

Proof Let $\varphi $ be a lscsm such that $\mathscr {I}=Exh(\varphi )$ . Note that for any $n\in \omega $ there is $k\in \omega $ such that for all $i\ge k$ , $\varphi (\{i\})\leq \frac {1}{2^n}$ (recall that $\mathscr {I}$ is a tall ideal). Recursively construct a sequence $\langle k_n:n\in \omega \rangle $ such that for all $i\ge k_n$ , $\varphi (\{i\})\leq \frac {1}{2^n},$ and $k_n<k_{n+1}$ . Then define $g_{\varphi }(n)=k_n$ . Let us see that $g_{\varphi }$ is a witness for the lemma. It is clear that if $\mathcal {U}$ is a weak $\mathscr {I}$ -ultrafilter, then for all finite to one $f:\omega \to \omega $ such that $f\leq g_{\varphi }$ there is $A\in \mathcal {U}$ such that $f[A]\in \mathscr {I}$ . So let us assume that for all finite to one $f\leq g_{\varphi }$ there is $A\in \mathcal {U}$ such that $f[A]\in \mathscr {I}$ . Let $f\in \omega ^{\omega }$ be an arbitrary finite to one function. If $D=\{i\in \omega :{g}_{\varphi }(i)\leq f(i)\}\in \mathcal {U}$ , note that for any $i\in D$ , $\varphi (\{f(i)\})\leq 1/2^i$ , so for any $n\in \omega $ ,

$$ \begin{align*} \varphi(f[D]\cap n)\leq \sum_{j\in f[D]\cap n}\varphi(\{j\})&\leq \\ &\quad \sum_{i\in D,f(i)<n}\varphi(\{f(i)\})\leq\sum_{i\in D,f(i)<n}\frac{1}{2^i}\leq \sum_{i\in D}\frac{1}{2^i}. \end{align*} $$

Then, by the lower semicontinuity of $\varphi $ , it follows that $\varphi (f[D])\leq \sum _{i\in D}\frac {1}{2^i}$ . This same argument can be used to proved that $\varphi (f[D]\setminus n)\leq \sum _{i\in D,f(i)\ge n}\frac {1}{2^i}$ , which implies $\lim _{n\to \infty }\varphi (f[D]\setminus n)=0$ , so $f[D]\in \mathscr {I}$ . On the other hand, if $\omega \setminus D=\{n\in \omega :f(n)<g_{\varphi }(n)\}\in \mathcal {U}$ , then there is a finite to one function $h\in \omega ^{\omega }$ such that $h\leq g$ and $\omega \setminus D=\{n\in \omega :f(n)=h(n)\}\in \mathcal {U}$ (just define $h\upharpoonright (\omega \setminus D)=f\upharpoonright (\omega \setminus D)$ , and extend it to an arbitrary finite to one function h such that $h\leq g_{\varphi }$ ). By hypothesis, there is $A\in \mathcal {U}$ such that $h[A]\in \mathscr {I}$ . Define $B=A\cap (\omega \setminus D)$ , then we have $f[B]=h[B]\in \mathscr {I}$ .

Now we are ready to prove the following theorem.

Theorem 4.8. Let $\mathscr {I}$ be an analytic p-ideal and $\mathcal {U}$ an $(\mathscr {I},p)$ -point. Then in the generic extension by the Rational Perfect set forcing $\mathcal U$ generates an $(\mathscr {I},p)$ -point.

Proof Fix $\mathcal {U}$ an $(\mathscr {I},p)$ -point. Let $\dot {f}$ be a $\mathbf {PT}$ -name and $p\in \mathbf {PT}$ a condition such that $p\Vdash ` `\dot {f}\in \omega ^{\omega }"$ . By Lemma 4.7, we can assume that $p\Vdash ` `\dot {f}\leq g_{\varphi }"$ (note that by absoluteness the same $g_{\varphi }$ from the ground model works in any generic extension), so we are going to prove that in the generic extension, for any $\dot {f}\leq g_{\varphi }$ there is $A\in \mathcal {U}$ such that $\dot {f}[A]\in \mathscr {I}$ . By Theorem 4.6, $\mathcal U$ remains as a p-point in the generic extension, so we can assume that $p\Vdash ` `\dot {f} \textit {is finite to one}"$ (the case when $\dot {f}$ has a restriction on which it is constant follows immediately). By an application of Lemma 4.4, we can assume that for any $s\in split(p)$ there is a function $h_s\in \omega ^{\omega }$ such that for all $n\in \omega $ and for all but finitely many $k\in succ_{p}(s)$ , $p\upharpoonright s^{\frown } k\Vdash ` `\dot {f}\upharpoonright (\lvert s\rvert +n)=h_s\upharpoonright (\lvert s\rvert +n)"$ . Since $\mathcal U$ is an $\mathscr {I}$ -ultrafilter, for every $s\in split(p)$ we find $A_s\in \mathcal U$ such that $h_s[A_s]\in \mathscr {I}$ . Note that $h_s$ is not necessarily a finite to one function, but since $\mathcal U$ is a p-point, we can assume that $h_s\upharpoonright A_s$ is either constant or finite to one. Now, by Lemma 4.5, we can as well assume that the condition p has one of the following properties:

  1. a) For all $s\in split(p)$ the restriction $h_s\upharpoonright A_s$ is finite to one.

  2. b) For all $s\in split(p)$ the restriction $h_s\upharpoonright A_s$ is a constant $a_s$ .

Let us consider first the case a). Since $\mathscr {I}$ is a p-ideal and for all $s\in split(p)$ we have $h_s[A_s]\in \mathscr {I}$ , then there is $Z\in \mathscr {I}$ such that for all $s\in split(p)$ , $h_s[A_s]\subseteq ^*Z$ . Removing finitely many elements of each $A_s$ , we can assume that for all $s\in split(p)$ we have $h_s[A_s]\subseteq Z$ (recall that we are dealing with case a)). Making use of the p-point game we will construct two sequences $\{F_n:n\in \omega \}$ and $\{T_n:n\in \omega \}$ such that:

  1. (1) $T_0\leq _0 p$ .

  2. (2) For all $n\in \omega $ , $T_{n+1}\Vdash ` `\dot {f}[F_n]\subseteq Z"$ , and $T_{n+1}\leq _{n+1}T_n$ .

  3. (3) $\bigcup _{n\in \omega } F_n\in \mathcal {U}$ .

  4. (4) $T_{\omega }=\bigcap _{n\in \omega } T_n$ is a condition.

The construction is as follows:

  1. (i) Player I starts playing $A_0=A_{st(p)}$ , and define $T_0=p$ .

  2. (ii) Suppose Player II has answered with a set $F_0\subseteq A_0$ . Let $k_0\in \omega $ be big enough so that for all $i\in succ_{T_0}(st(T_0))\setminus k_0$ it holds that $T\upharpoonright st(T_0)^{\frown } i\Vdash ` `\dot {f}\upharpoonright (\max (F_0)+1)=h_{st(T_0)}\upharpoonright (\max (F_0)+1)$ ”. Note that this implies that for all $i\ge k_0$ , $T_0\upharpoonright st(T_0)^{\frown } i\Vdash ` `\dot {f}[F_0]\subseteq Z"$ . Then define

    $$ \begin{align*} T_1=\bigcup_{\substack{i\ge k_0,\\ i\in succ_{T_0}(st(T_0))}}T_0\upharpoonright st(T_0)^{\frown} i. \end{align*} $$

    Then Player I plays the set $A_1=\bigcap _{s\in {1^{\leq 1}}}A_{\varphi _{T_1}(s)}\setminus (\max (F_0)+1)$ .

  3. (iii) Suppose at round number n Player I has played the set

    $$ \begin{align*} A_n=\bigcap_{s\in n^{\leq n}} A_{\varphi_{T_n}(s)}\setminus \left(\max\left(\bigcup_{k<n}F_k\right)+1\right). \end{align*} $$

    Suppose Player II responds with a finite set $F_n\subseteq A_n$ . Then, let $k_n\in \omega $ be big enough such that for all $r\in \varphi _{T_n}(n^{\leq n})$ and all $i\in succ_{T_n}(r)\setminus k_n$ , it holds that $T_n\upharpoonright r^{\frown } i\Vdash ` `\dot {f}\upharpoonright (\max \left (\bigcup _{k\leq n}F_n\right )+1)=h_{r}\upharpoonright (\max \left (\bigcup _{k\leq n}F_n\right )+1)"$ . Note that this implies that for each $r\in \varphi _{T_n}(n^{\leq n})$ and $i\in succ_{T_n}(r)\setminus k_n$ , $T_n\upharpoonright r^{\frown } i\Vdash ` `\dot {f}[F_n]\subseteq Z"$ . Then define $T_{n+1}$ as follows:

    $$ \begin{align*} T_{n+1}=\bigcup_{s\in n^{\leq n}}\bigcup_{\substack{i\ge k_n,\\ i\in succ_{T_n}(\varphi_{T_n}(s))}}T\upharpoonright \varphi_{T_n}(s)^{\frown} i. \end{align*} $$

    By construction it follows that $T_{n+1}\Vdash ` `\dot {f}[F_n]\subseteq Z"$ . Let Player I play the set

    $$ \begin{align*} A_{n+1}=\bigcap_{s\in(n+1)^{\leq n+1}} A_{\varphi_{T_{n+1}}(s)}\setminus\left(\max\left(\bigcup_{k\leq n}F_k\right)+1 \right). \end{align*} $$

By Lemma 2.2, this is not a winning strategy for Player I, so there is a play where Player II wins the game. Let $\{F_n:n\in \omega \}$ and $\{T_n:n\in \omega \}$ be the sequences constructed by Player I in this play, and define $T_{\omega }=\bigcap _{n\in \omega }T_n$ .

Claim. $T_{\omega }\Vdash ` `\dot {f}[\bigcup _{n\in \omega }F_n]\subseteq Z"$ . As pointed out in the construction of the two sequences, it holds that $T_{n+1}\Vdash ` `\dot {f}[F_n]\subseteq Z"$ . But for all n, $T_{\omega }\leq T_n$ , so the claim is true for $T_{\omega }$ .

Now let us deal with case b). Fist note that there is no condition $q\leq p$ such that the set $\{a_s:s\in splitt(q)\}$ is finite. To prove this assume it is false and apply Lemma 4.5 finitely many times, then recall that the function $\dot {f}$ is a finite to one function to arrive at a contradiction. By using this remark it is possible to find a stronger condition $q\leq p$ such that $Z=\{a_s:s\in split(q)\}\in \mathscr {I}$ . The rest follows the same lines of the previous case. Using the p-point game again we construct two sequences $\{F_n:n\in \omega \}$ and $\{T_n:n\in \omega \}$ such that conditions (2)–(5) of the construction for case a) hold, and condition (1) is changed to (1’) $T_0\leq q$ . The strategy for player I is defined in the same way as it was in the previous case. Then the condition $T_{\omega }$ and the set $\bigcup _{n\in \omega }F_n$ constructed in this way satisfy $T\Vdash ` `\dot {f}[\bigcup _{n\in \omega }F_n]\subseteq Z"$ .

5 A preservation theorem for $(\mathscr {I},p)$ -points

In this section we prove the following preservation theorem for $(\mathscr {I},p)$ -points. We say that a forcing $\mathbb {P}$ preserves $(\mathscr {I},p)$ -points if any $(\mathscr {I},p)$ -point from the ground model generates an $(\mathscr {I},p)$ -point after forcing with $\mathbb {P}$ .

Theorem 5.1. Let $\mathscr {I}$ be an analytic p-ideal such that $\mathscr {I}=Fin(\varphi )=Exh(\varphi )$ , for some lscsm $\varphi $ . Let $\langle \mathbb {P}_{\beta },\dot {\mathbb {Q}}_{\beta }:\beta <\alpha \rangle $ be a countable support iteration of proper forcing notions such that for all $\beta <\alpha $ , $\mathbb {P}_{\beta }$ is proper, preserves $(\mathscr {I},p)$ -points and $\mathbb {P}_{\beta }\Vdash ` `\dot {\mathbb {Q}}_{\beta } \textit { is proper and preserves }(\mathscr {I},p)\textit {-points}"$ . Then $\mathbb {P}_{\alpha }$ preserves $(\mathscr {I},p)$ -points.

Recall that given a $\mathbb {P}$ -name $\dot {f}$ for a real, a pair $(\langle p_n:n\in \omega \rangle ,h)$ is an interpretation for $\dot {f}$ if $\langle p_n:n\in \omega \rangle $ is a decreasing sequence and for all $n\in \omega $ , $p_n\Vdash ` `\dot {f}\upharpoonright n=h\upharpoonright n"$ . In such case, we say that $\langle p_n:n\in \omega \rangle $ interprets $\dot {f}$ as h. We need a strengthening of this property.

Definition 5.2. Let $\dot {f}$ be a $\mathbb P$ -name for a finite to one function from $\omega $ to $\omega $ and $p\in \mathbb {P}$ a condition. A fine interpretation for $\dot f$ below p is a pair $(\langle p_n:n\in \omega \rangle ,h)$ such that $\langle p_n:n\in \omega \rangle $ is a decreasing sequence, $p_0\leq p$ , h is a finite to one function and for all $n\in \omega $ , $p_n\Vdash ` `\dot {f}\upharpoonright n=h\upharpoonright n"$ and $p_n$ decides the value of $\dot {f}^{-1}(n)$ .

It is easy to see that if $\dot f$ is forced to be a finite to one function, then given a condition p, we can find a fine interpretation for $\dot {f}$ below p. In order to prove Theorem 5.1, we need a slight modification of the notion of derived sequence defined in [Reference Abraham, Foreman and Kanamori1]. There are many different ways to proceed, but we choose to recall the usual notion of derived sequence, and then make use of Lemma 5.4 to state the existence of the objects we need. The notion we work with is that of fine derived sequence, and it is stated just before Lemma 5.4. Alternatively, it is possible to state everything we need in only one definition.

Definition 5.3. Let $\mathbb {P}*\dot {\mathbb {Q}}$ be a two step forcing iteration, and let $\dot {f}$ be a $\mathbb {P}*\dot {\mathbb {Q}}$ -name and $(p,\dot {q})\in \mathbb {P}*\dot {\mathbb {Q}}$ a condition such that $(p,\dot {q})\Vdash ` `\dot {f}\in \omega ^{\omega }"$ . Let $\sqsubseteq $ be a well ordering of $\mathbb {P}*\dot {\mathbb {Q}}$ . Let $(\langle r_n:n\in \omega \rangle ,h)$ be an interpretation of $\dot {f}$ below $(p,\dot {q})$ , where $r_n=(p_n,\dot {q}_n)$ . Suppose $G_{\mathbb {P}}$ is a generic filter for $\mathbb {P}$ over V. Then, in $V[G_{\mathbb {P}}]$ , define the following decreasing sequence of conditions $\langle t_n:n\in \omega \rangle $ in $\dot {\mathbb {Q}}[G_{\mathbb {P}}]$ which interprets $\dot {f}$ :

  1. (1) Define $n^*=\sup (\{n\in \{-1\}\cup \omega :p_n\in G_{\mathbb {P}}\})$ , where $r_{-1}=(p_{-1},\dot {q}_{-1})=(1_{\mathbb {P}},1_{\dot {\mathbb {Q}}})$ .

  2. (2) For $n\in \{-1\}\cup \omega $ , if $n\leq n^*$ , define $t_n=\dot {q}_n[{G}_{\mathbb {P}}]$ .

  3. (3) For $n\in \omega $ , if $n>n^*$ , then define $t_n$ to be $\dot {s}_n[G_{\mathbb {P}}]$ , where $\dot {s}_n$ is such that there is $u_n\in \mathbb {P}$ such that $(u_n,\dot {s}_n)\in \mathbb {P}*\dot {\mathbb {Q}}$ is an extension of $(u_{n-1},\dot {s}_{n-1})$ such that $u_{n}\in G_{\mathbb {P}}$ and $\dot {s}_n[G_{\mathbb {P}}]$ decides the value of $\dot {f}\upharpoonright n$ , and $(u_n,\dot {s}_n)$ is $\sqsubseteq $ -minimal with this property.

The sequence $\langle t_i:i\in \omega \rangle $ defined in $V[G_{\mathbb {P}}]$ is said to be the derived sequence from $\langle r_n:n\in \omega \rangle $ , $G_{\mathbb {P}}$ , $\sqsubseteq $ and $\dot {f}$ , and we write $\delta _{G_{\mathbb {P}}}(\langle r_n:n\in \omega \rangle ,\dot {f})$ to denote this sequence, and a $\mathbb {P}$ -name of this sequence will be denoted by $\tilde {\delta }(\langle r_n:n\in \omega \rangle ,\dot {f})$ . Note that by clause (3), in $V[G_{\mathbb {P}}]$ there is a function g such that $\delta _{G_{\mathbb {P}}}(\langle r_n:n\in \omega \rangle ,\dot {f})$ interprets $\dot {f}$ as g. If $\tilde {\delta }(\langle r_n:n\in \omega \rangle ,\dot {f})$ is the $\mathbb {P}$ -name for the derived sequence in $V[G_{\mathbb {P}}]$ , then define $\textrm{int}(\tilde {\delta }(\langle r_n:n\in \omega \rangle ,\dot {f}),\dot {f})$ to be a $\mathbb {P}$ -name for the function g such that in $V[G_{\mathbb {P}}]$ , the pair $(\delta _{G_{\mathbb {P}}}(\langle r_n:n\in \omega \rangle ,\dot {f}),\dot {g})$ is an interpretation of $\dot {f}$ .

It is easy to see that if $\dot {g}=\textrm{int}(\tilde {\delta }(\bar {r},\dot {f}),\dot {f})$ , and $\bar {r}$ interprets $\dot {f}$ as h, then $\bar {p}$ interprets $\dot {g}$ as h. Also note that clause (3) of the previous definition allows some flexibility in the choice of $t_n$ , so we can add more requirements to be satisfied by $t_n$ in relation to $\dot {f}$ . This is where our modification takes place: if $\dot {f}$ is forced to be a finite to one function, we could additionally required in clause (3) that $t_n$ determines not only the initial segment $\dot {f}\upharpoonright n$ , but also $\dot {f}^{-1}(n)$ . In this way, $\dot {g}=\textrm{int}(\tilde {\delta }(\bar {r},\dot {f}),\dot {f})$ is forced to be finite to one. Derived sequences having this property will be called fine derived sequences. This is the notion we work with in Lemma 5.4 and the proof of Theorem 5.1. On the other hand, strictly speaking, the derived sequence $\tilde {\delta }(\langle r_n:n\in \omega \rangle ,\dot {f})$ and the $\mathbb {P}$ -name $\textrm{int}(\tilde {\delta }(\bar {r},\dot {f}),\dot {f})$ depend not only on $\mathbb {P}*\dot {\mathbb {Q}}$ , $\dot {f}$ , $\langle r_n:n\in \omega \rangle $ and G, but also on $\sqsubseteq $ , and is uniquely determined by this five parameters. With exception of Lemma 5.4, we will usually not explicitly mention the well ordering $\sqsubseteq $ , and when we mention the expressions a derived sequence or a fine derived sequence, we assume that a well ordering $\sqsubseteq $ of the forcing has been fixed for the whole proof, and implicitly assume that the (fine)derived sequence at hand depends on such well ordering.

Lemma 5.4. Let $\mathbb {P}*\dot {\mathbb {Q}}$ be a forcing iteration, $\dot f$ a $\mathbb {P}*\dot {\mathbb {Q}}$ -name for a finite to one function, and $(p,\dot {q})\in \mathbb {P}*\dot {\mathbb {Q}}$ a condition. Let $(\langle r_n:n\in \omega \rangle ,h)$ be a fine interpretation for $\dot f$ below $(p,\dot {q})$ , where $r_n=(p_n,\dot {q}_n)$ . Let $\sqsubseteq $ be a well ordering of $\mathbb {P}*\dot {\mathbb {Q}}$ . Then there is a $\mathbb {P}$ -name for a fine derived sequence $\tilde {\delta }=\tilde {\delta }(\langle r_n:n\in \omega ,\dot f\rangle )$ , and the pair $(\tilde {\delta },\textrm{int}(\tilde {\delta },\dot {f}))$ is forced to be a fine interpretation of $\dot f$ in $V[\dot {G}_{\mathbb P}]$ . Moreover, $(\langle p_n:n\in \omega \rangle ,h)$ is a fine interpretation of $\textrm{int}(\tilde {\delta },\dot {f})$ .

Proof We follow the same notation as in Definition 5.3. Let $G_{\mathbb {P}}\subseteq \mathbb {P}$ be a generic filter over V, define $r_{-1}=(p_{-1},\dot {q}_{-1})=(1_{\mathbb {P}},1_{\dot {\mathbb {Q}}})$ and $n^*=\sup (\{n\in \{-1\}\cup \omega :p_n\in G_{\mathbb {P}}\})$ . Define $\tilde {\delta }$ as a $\mathbb {P}$ -name for the sequence $\langle t_n:n\in \omega \rangle $ of conditions in $\dot {\mathbb {Q}}[G_{\mathbb {P}}]$ defined as follows:

  • For all $n\in \omega $ , if $n\leq n^*$ , then $t_n=\dot {q}_n[G_{\mathbb {P}}]$ , and for all natural number $n> n^*$ , define $t_n$ to be $\dot {s}_n[G_{\mathbb {P}}]$ , where $\dot {s}_n$ is such that there is $u_n\in \mathbb {P}$ such that $(u_n,\dot {s}_n)\in \mathbb {P}*\dot {\mathbb {Q}}$ is the $\sqsubseteq $ -first extension of $(u_{n-1},\dot {s}_{n-1})$ such that $u_n\in G_{\mathbb {P}}$ , and $\dot {s}_n[G_{\mathbb {P}}]$ decides the value of $\dot {f}^{-1}(n)$ and $\dot {f}\upharpoonright n$ . Let $g\in \omega ^{\omega }$ be the function such that $t_n\Vdash ` `\dot {f}\upharpoonright n=g\upharpoonright n"$ . Define $\textrm{int}(\tilde {\delta },\dot {f})$ as a $\mathbb {P}$ -name for g.

By definition of $\tilde {\delta }$ and the definition of $\textrm{int}(\tilde {\delta },\dot {f})$ , it follows that $\tilde {\delta }$ is a fine derived sequence and the pair $(\tilde {\delta },\textrm{int}(\tilde {\delta },\dot {f}))$ is forced to a fine interpretation of $\dot {f}$ in $V[\dot {G}_{\mathbb {P}}]$ , since $\textrm{int}(\tilde \delta ,\dot {f})$ is actually a $\mathbb {P}$ -name of the function g. It is easy to see that $(\langle p_n:n\in \omega \rangle ,h)$ is a fine interpretation of $\textrm{int}(\tilde {\delta },\dot {f})$ .

The next lemma can be seen as an intermediate step in the proof of Theorem 5.1, but we consider that it is more readable to write it down as a separated lemma.

Lemma 5.5. Let $\mathscr {I}$ be an ideal such that $\mathscr {I}=Fin(\varphi )=Exh(\varphi )$ , for some lscsm $\varphi $ . Let $\mathcal {U}$ be an $(\mathscr {I},p)$ -point. Let $\mathbb {P}*\dot {\mathbb {Q}}$ be a two step iteration of proper forcings such that $\mathbb {P}$ preserves $(\mathscr {I},p)$ -points. Let $\dot {f}$ be a $\mathbb {P}*\dot {\mathbb {Q}}$ -name for a finite to one function from $\omega $ to $\omega $ . Let $(p,\dot {q})\in \mathbb {P}*\dot {\mathbb {Q}}$ be a condition and $(\langle r_n:n\in \omega \rangle , h_0)$ be a fine interpretation of $\dot {f}$ below the condition $(p,\dot {q})$ . We can assume that $r_n=(p_n,\dot {q}_n)$ . Let $\mathcal {M}\prec H(\theta )$ be a countable elementary submodel such that $p,\dot {f},\mathbb {P}*\dot {\mathbb {Q}},\mathcal {U},(\langle r_n:n\in \omega \rangle , h_0)\in \mathcal {M}$ , and let $Z\in \mathcal {U}$ be a pseudointersection of $\mathcal {U}\cap \mathcal {M}$ . Let $m\in \omega $ be such that $\varphi (h_0[Z])<m$ . Then for all $n\in \omega $ , there are the following:

  1. (1) An $(\mathcal {M},\mathbb {P})$ -generic condition $p_{n,\mathcal {M}}$ .

  2. (2) A $\mathbb {P}$ -name $\tilde {s}_n$ .

  3. (3) A $\mathbb {P}$ -name $\tilde {h}_n$ for a finite to one function from $\omega $ to $\omega $ .

  4. (4) A $\mathbb {P}$ -name for a sequence of conditions in $\dot {\mathbb {Q}}[\dot {G}_{\mathbb {P}}]$ .

Such that:

  1. (i) $p_{n,\mathcal {M}}\leq p_0$ .

  2. (ii) $p_{n,\mathcal {M}}\Vdash ` `\tilde {s}_n\leq \dot {q}"$ and $p_{n,\mathcal {M}}\Vdash ` `\tilde {s}_n\in \dot {\mathbb {Q}}[\dot {G}_{\mathbb {P}}]\cap \mathcal {M}[\dot {G}_{\mathbb {P}}]"$ .

  3. (iii) $p_{n,\mathcal {M}}$ forces that is a fine interpretation of $\dot {f}$ below $\tilde {s}_n$ and such that $\varphi (\tilde {h}_n[Z])<m$ .

  4. (iv) $p_{n,\mathcal {M}}\Vdash ` `\tilde {s}_n\Vdash ` `\dot {f}\upharpoonright n=\tilde {h}_n\upharpoonright n""$ .

Proof Fix a natural number $n\in \omega $ . We need to find the objects in clauses 1)–4). First note that by Lemma 5.4, we can find inside $\mathcal {M}$ a $\mathbb {P}$ -name for a fine derived sequence $\dot {\delta }=\dot {\delta }(\langle r_n:n\in \omega \rangle ,\dot {f})$ such that $\textrm{int}(\dot {\delta },\dot {f})$ is interpreted by $\langle p_n:n\in \omega \rangle $ as $h_0$ . For short, denote by $\dot {h}$ the name $\textrm{int}(\dot {\delta },\dot {f})$ .

Let $G\subseteq \mathbb {P}$ be a generic filter over V. By hypothesis, $\mathcal {U}$ remains as an $(\mathscr {I},p)$ -point in $V[G]$ , and by elementarity, $\mathcal {M}[G]\vDash \mathcal {U}\text { is an } (\mathscr {I},p)\text {-point}$ . So for each $k\in \omega $ , there are $u_k\leq p_k$ and $X_k\in \mathcal {U}$ such that $u_k\Vdash ` `\varphi (\dot {h}[X_k])<1"$ , since $\mathscr {I}=Exh(\varphi )$ . Clearly this can by done inside $\mathcal {M}$ , so we assume these two sequences are in $\mathcal {M}$ . Let $X\in \mathcal {M}\cap \mathcal {U}$ be a pseudointersection of $\langle X_k:k\in \omega \rangle $ . Now, working in $\mathcal {M}$ , define by recursion a sequence of natural numbers $\langle k_l:l\in \omega \rangle $ as follows:

  1. (1) $k_0=0.$

  2. (2) For all $i\leq k_l$ , $X\setminus k_{l+1}\subseteq X_i$ .

  3. (3) For all $i\ge k_{l+1}$ , $p_i\Vdash ` `\dot {h}\upharpoonright k_l= h_0\upharpoonright k_l".$

  4. (4) For all $i\leq k_l$ , there is $u_i^{l+1}\leq u_i\leq p_i$ such that $u_i^{l+1}\Vdash ` `\varphi (\dot {h}[X_i\setminus k_{l+1}])<m-\max \{\varphi (h_0[F]):F\subseteq k_l\land \varphi (h_0[F])<m\}"$ .

Note that 4) can be achieved since $\dot {h}$ is forced to be finite to one, and $u_i$ forces $\dot {h}[X_i]\in \mathscr {I}=Exh(\varphi )$ . Conditions 1)–3) are easy to arrange.

Assume the sequence $\langle k_j:j\in \omega \rangle $ has been constructed. Without loss of generality, we can assume that $Y=\bigcup _{j\in \omega }[k_{3j+1},k_{3j+2})\in \mathcal U$ . Let $j_0$ be such that $Z\setminus k_{3j_0}\subseteq X\cap Y$ (recall that Z is a pseudointersection of $\mathcal M\cap \mathcal U$ , and $X\cap Y\in \mathcal U\cap \mathcal M$ ). Note that for any $l> j_0$ we have $[k_{3l-1},k_{3l+1})\cap Z=\emptyset $ . Fix one of such l. We claim that $u_{k_{3l}}^{3l+1}\Vdash ` `\varphi (\dot {h}[Z])< m"$ . Since $u_{k_{3l}}^{3l+1}\leq p_{k_{3l}}$ , by 3) above we have

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash ` `\dot{h}\upharpoonright k_{3l-1}= h_0\upharpoonright k_{3l-1}", \end{align*} $$

which in turn implies

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash ` `\dot{h}[Z\cap k_{3l-1}]= h_0[Z\cap k_{3l-1}]" \end{align*} $$

and so

(1) $$ \begin{align} u_{k_{3l}}^{3l+1}\Vdash ` `\varphi(\dot{h}[Z\cap k_{3l-1}])\leq\max\{\varphi(h_0[F]):F\subseteq k_{3l}\land\varphi(h_0[F])<m\}". \end{align} $$

Also, by 2) and the choice of l we have

(2) $$ \begin{align} Z\setminus k_{3l+1}\subseteq X\cap Y\setminus k_{3l+1}\subseteq X_{k_{3l}}\setminus k_{3l+1.} \end{align} $$

Note that since $Z\cap [k_{3l-1},k_{3l+1})=\emptyset $ , the following holds:

(3) $$ \begin{align} u_{k_{3l}}^{3l+1}\Vdash ` `\varphi(\dot{h}[Z])\leq\varphi(\dot{h}[Z\setminus k_{3l}])+\varphi &(\dot{h}[Z\cap k_{3l}])= \\ &\qquad \varphi(\dot{h}[Z\setminus k_{3l+1}])+\varphi(\dot{h}[Z\cap k_{3l-1}])". \nonumber \end{align} $$

Note that (1) implies the following:

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash` `m-\max\{\varphi(h_0[F]):F\subseteq k_{3l}\land\varphi(h_0[F])<m\}\leq m-\varphi(\dot{h}[Z\cap k_{3l-1}])". \end{align*} $$

This inequality, joint with 4), gives the following:

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash` `\varphi(\dot{h}[X_{k_{3l}}\setminus k_{3l+1}])< m-\varphi(\dot{h}[Z\cap k_{3l-1}])" \end{align*} $$

or equivalently

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash ` `\varphi(\dot{h}[X_{k_{3l}}\setminus k_{3l+1}])+\varphi(\dot{h}[Z\cap k_{3l-1}])<m". \end{align*} $$

By (2) and the monotonicity of $\varphi $ , we have that

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash ` `\varphi(\dot{h}[Z\setminus k_{3l+1}])+\varphi(\dot{h}[Z\cap k_{3l-1}])<m". \end{align*} $$

Finally by (3) we get

$$ \begin{align*} u_{k_{3l}}^{3l+1}\Vdash ` `\varphi(\dot{h}[Z])<m". \end{align*} $$

Now let $l>j_0$ be a natural number big enough so that $n<k_{3l-1}$ , and let $q\leq u_{k_{3l}}^{3l+1}$ be an $(\mathcal {M},{\mathbb {P}})$ -generic condition. Define the following:

  1. (a) In $\mathcal {M}$ , let $\tilde {s}_n$ be the $\mathbb {P}$ -name $\dot {q}_{k_{3l}}$ .

  2. (b) In $\mathcal {M}$ , let be a $\mathbb {P}$ -name for the fine derived sequence $\tilde {\delta }(\langle r^{\prime }_i:i\in \omega \rangle ,\dot {f})$ , where $r^{\prime }_i=r_{i+k_{3l}}$ , for all $i\in \omega $ .

  3. (c) In $\mathcal {M}$ , let $\tilde {h}_{n}$ be the $\mathbb {P}$ -name $\dot {h}$ .

  4. (d) Define $p_{n,\mathcal {M}}=q$ .

We claim that $p_{n,\mathcal {M}}$ , $\tilde {s}_n$ , $\tilde {h}_n$ and are as required.

Conditions (1)–(4) are easily seen to be hold by the construction. Condition $i)$ , follows from the fact that $p_{n,\mathcal {M}}\leq p_{k_{3l}}\leq p_0$ . The first part of $ii)$ follows from a) and $(p_{k_{3l}},\dot {q}_{k_{3l}})\leq (p,\dot {q})$ and the choice of $\tilde {s}_n$ . The second part of condition $ii)$ follows from the fact that $\langle (p_k,\dot {q}_k):k\in \omega \rangle \in \mathcal {M}$ , $p_{n,\mathcal {M}}\leq p_{k_{3l}}$ and $\tilde {s}_n=\dot {q}_{k_{3l}}$ .

Let us see condition $(iii)$ . By definition $\tilde {\delta }(\langle r_i':i\in \omega \rangle ,\dot {f})$ , is a fine interpretation of $\dot {f}$ . To see that $\tilde {\delta }(\langle r_i':i\in \omega \rangle ,\dot {f})$ is forced to be below $\tilde {s}_n$ by $p_{n,\mathcal {M}}$ , let $G\subseteq \mathbb {P}$ be any generic filter such that $p_{n,\mathcal {M}}\in G$ . By definition $r_0'=(p_{k_{3l}},\dot {q}_{k_{3l}})$ . Since $p_{n,\mathcal {M}}\leq p_{k_{3l}}$ and $p_{n,\mathcal {M}}\in G$ , it follows that the first term of $\tilde {\delta }(\langle r_i':i\in \omega \rangle )$ is $\dot {q}_{k_{3l}}[G]=\tilde {s}_n[G]$ , and all the remaining terms are below the first one since $\tilde {\delta }(\langle r_i':i\in \omega \rangle ,\dot {f})$ is a decreasing sequence. To see that $p_{n,\mathcal {M}}\Vdash ` `\varphi (\tilde {h}_n[Z])<m"$ , recall that $l>j_0$ , $p_{n,\mathcal {M}}\leq u_{k_{3l}}^{3l+1}$ and $u_{k_{kl}}^{3l+1}\Vdash ` `\varphi (\dot {h}[Z])<m"$ .

To see that condition $(iv)$ holds, just note that $n < k_{3l-1}$ , $(p_{n,\mathcal {M}},\tilde {s}_n)\leq (p_{k_{3l}},\dot {q}_{k_{3l}})$ and

$$ \begin{align*} (p_{k_{3l}},\dot{q}_{k_{3l}})\Vdash` `\dot{f}\upharpoonright k_{3l}=h_0\upharpoonright k_{3l}", \end{align*} $$

which implies

$$ \begin{align*} p_{n,\mathcal{M}}\Vdash` `\tilde{s}_n\Vdash` `\dot{f}\upharpoonright k_{3l}=h_0\upharpoonright k_{3l}"" \end{align*} $$

and also

$$ \begin{align*} p_{k_{3l}}\Vdash` `\tilde{h}_n\upharpoonright k_{3l}=\dot{h}\upharpoonright k_{3l}=h_0\upharpoonright k_{3l}", \end{align*} $$

so it follows that

$$ \begin{align*} p_{n,\mathcal{M}}\Vdash` `\tilde{s}_n\Vdash` `\dot{f}\upharpoonright k_{3l}=\tilde{h}_n\upharpoonright k_{3l}"". \end{align*} $$

Since $n<k_{3l-1}$ , the required property follows.

Before proving Theorem 5.1 we introduce some notation and conventions to be used in the proof. Given a countable support iteration $\mathbb {P}=\langle \mathbb {P}_n,\dot {\mathbb {Q}}_n:n\in \omega \rangle $ which is factorized as $\mathbb {P}=\mathbb {P}_n*\tilde {\mathbb {P}}_{[n+1,\omega )}$ , and a condition $p\in \mathbb {P}$ , we denote by $(p_0,\dot {p}_{[1,\omega )})$ the factorization of the condition p relative to $\mathbb {P}_n*\tilde {\mathbb {P}}_{[n+1,\omega )}$ . In a similar way, if $\langle r_n:n\in \omega \rangle $ is a sequence of conditions in $\mathbb {P}$ , we denote by $(r_0^k,\dot {r}_{[1,\omega )}^k)$ the factorization of the condition $r_k$ relative to $\mathbb {P}_n*\tilde {\mathbb {P}}_{[n+1,\omega )}$ . On the other hand, consider $\mathbb {P}$ factorized as $\mathbb {P}_n*\dot {\mathbb {Q}}_{n+1}*\tilde {\mathbb {P}}_{[n+2,\omega )}$ and let $G_{n}\subseteq \mathbb {P}_{n}$ be a generic filter over V. We will be translating $\dot {\mathbb {Q}}_{n+1}[G_{n}]$ -names in $V[G_n]$ to $\mathbb {P}_{n+1}$ -names names in V, and vice versa. In this setting, when working in $V[G_n]$ , we consider $\dot {\mathbb {Q}}_{n+1}[G_n]$ as the evaluation of the $\mathbb {P}_n$ -name $\dot {\mathbb {Q}}_{n+1}$ , and $\tilde {\mathbb {P}}_{[n+2,\omega )}[G_n]$ as a suitable $\dot {\mathbb {Q}}_{n+1}[G_n]$ -name for the residual iteration after $\mathbb {P}_{n+1}=\mathbb {P}_n*\dot {\mathbb {Q}}_{n+1}$ , or more exactly, for the quotient $\mathbb {P}_{\omega }/G_{n+1}$ , where $G_{n+1}\subseteq \mathbb {P}_{n+1}$ is a generic filter over V. So, when working in $V[G_n]$ , we will write $\dot {\mathbb {Q}}_{n+1}[G_n]*\tilde {\mathbb {P}}_{n+1}[G_n]$ for the quotient forcing $\mathbb {P}_{\omega }/G_n$ in $V[G_n]$ . The point of this is the two step factorization of $\mathbb {P}_{\omega }/G_n$ , so we can apply the previous lemma in $V[G_n]$ , and then make a translation of elements from $V[G_n]$ to $\mathbb {P}_n$ -names in V. Finally, we will be making an abuse of notation when referring to the $\mathbb {P}_{\omega }$ -name $\dot {f}$ : when talking about $\dot {f}$ in the extensions given by $\mathbb {P}_n$ , we implicitly assume that we are taking about an appropriated $\mathbb {P}_{[n+1,\omega )}$ -name which is equivalent to $\dot {f}$ after forcing with $\mathbb {P}_{[n+1,\omega )}$ on the generic extension by $\mathbb {P}_n$ .

Proof of Theorem 5.1

Let us first note that the successor step is easy to handle: let $\mathbb {P}*\dot {\mathbb {Q}}$ be a two step iteration of proper forcings such that $\mathbb {P}$ preserves $(\mathscr {I},p)$ -points and $\mathbb {P}\Vdash ` `\dot {\mathbb {Q}}\text { preserves } (\mathscr {I},p)\text {-points}"$ . Let $\mathcal {U}$ be an $(\mathscr {I},p)$ -point in the ground model. Let $G\subseteq \mathbb {P}$ be V-generic. Then, in $V[G]$ , $\mathcal {U}$ generates a $(\mathscr {I},p)$ -point, and $\dot {\mathbb {Q}}[G]$ preserves $(\mathscr {I},p)$ -points. If $H\subseteq \dot {\mathbb {Q}}$ is $V[G]$ -generic, then $\mathcal {U}$ generates an $(\mathscr {I},p)$ -point in $V[G][H]=V[G*H]$ . So $\mathbb {P}*\dot {\mathbb {Q}}$ preserves $\mathcal {U}$ as an $(\mathscr {I},p)$ -point.

We are left with the limit step. Let us assume all the hypothesis in the statement of Theorem 5.1. Note that we can assume that the iteration has length $\omega $ , since every real appears in a successor stage or in a countable cofinality stage, so we assume $\mathbb {P}_{\omega }=\langle \mathbb {P}_n,\dot {\mathbb {Q}}_n:n\in \omega \rangle $ . Let $\mathcal U$ be an $(\mathscr {I},p)$ -point, $\dot f$ a $\mathbb P_{\omega }$ -name for a function from $\omega $ to $\omega $ and $p\in \mathbb P_{\omega }$ a condition. Note that by Theorem 2.8, we have that $\mathcal {U}$ generates a p-point after forcing with $\mathbb {P}_{\omega }$ , so we can assume that $p\Vdash ` `\mathcal {U}\text { generates a } p\text {-point}"$ . If there is a condition $q\leq p$ and $X\in \mathcal U$ such that $q\Vdash ` `\dot {f}\upharpoonright X \textit {is constant}"$ , then we are done. Otherwise, there is $q\leq p$ and there exists $A\in \mathcal U$ such that $\dot f\upharpoonright A$ is forced to be finite to one by q, so we can assume that $\dot f$ is finite to one. Let $\mathcal M\prec H(\theta )$ be a countable elementary submodel such that $\dot {f},\mathcal U,p,\mathbb P_{\omega }\in \mathcal M$ . Fix $Z\in \mathcal U$ a pseudointersection of $\mathcal U\cap \mathcal M$ . In $\mathcal M$ , let $(\langle r_n:n\in \omega \rangle ,h_0)$ be a fine interpretation for $\dot {f}$ below p, and assume $\varphi (h_0[Z])< m$ for some fixed $m\in \omega $ . For each $n\in \omega $ , we assume $r_n$ has the form $\langle r_n^k:k\in \omega \rangle $ . We will construct by recursion four sequences $\langle q_n:n\in \omega \rangle $ , $\langle \dot {s}_n:n\in \omega \rangle $ , $\langle \dot {h}_n:n\in \omega \rangle $ and such that the following holds:

  1. (a) For $n=0$ , $q_0$ is a $(\mathcal {M},\mathbb {P}_0)$ -generic condition.

  2. (b) For all $n\in \omega $ , $q_n$ is $(\mathcal {M},\mathbb {P}_n)$ -generic and $q_{n}=(q_0,\ldots ,\dot {q}_n)$ , where $\dot {q}_n$ is a $\mathbb {P}_{n-1}$ -name for $n>0$ .

  3. (c) For all positive $n\in \omega $ , $\dot {h}_n$ is a $\mathbb {P}_{n-1}$ -name and $q_{n-1}\Vdash ` `\dot {h}_n\in \mathcal {M}[\dot {G}_{\mathbb {P}_{n-1}}]"$ .

  4. (d) For all positive $n\in \omega $ , is a $\mathbb {P}_{n-1}$ -name for a decreasing sequence of conditions in $\tilde {\mathbb {P}}_{[n,\omega )}[\dot {G}_{\mathbb {P}_{n-1}}]$ .

  5. (e) For all positive $n\in \omega $ , $\dot {s}_n$ is a $\mathbb {P}_{n-1}$ -name and $q_{n-1}$ forces that $\dot {s}_n$ is a condition in $\tilde {\mathbb {P}}_{[n,\omega )}[\dot {G}_{\mathbb {P}_{n-1}}]\cap \mathcal {M}[\dot {G}_{\mathbb {P}_{n-1}}]$ .

  6. (f) For all positive $n\in \omega $ , $q_{n-1}$ forces that is a fine interpretation of $\dot {f}$ below $\dot {s}_n$ .

  7. (g) For all positive $n\in \omega $ , $q_{n-1}$ forces that $\varphi (\dot {h}_n[Z])<m$ .

  8. (h) For all positive $n\in \omega $ , $q_{n-1}\Vdash ` `\dot {s}_n\Vdash ` `\dot {f}\upharpoonright n=\dot {h}_n\upharpoonright n""$ .

  9. (i) For all positive $n\in \omega $ , $q_{n-1}\Vdash ` `\dot {q}_n\leq \dot {s}_{0}^n"$ and $q_0\leq r_0^0$ .

  10. (j) For all $n\in \omega $ , $q_n\Vdash ` `\dot {s}_{n+1}\leq \dot {s}_{[1,\omega )}^{n}"$ .

Suppose first we have succeeded in the above construction and consider the condition $q=\bigcup _{n\in \omega }q_n$ . Note that by construction of q we have $q\leq q_0$ . We claim that $q\Vdash ` `\varphi (\dot {f}[Z])\leq m"$ . It is enough to prove $q\Vdash "(\forall n\in \omega )(\dot {f}\upharpoonright n=\dot {h}_n\upharpoonright n)"$ , since by g), this would imply that q forces the measure of the initial segments of $\dot {f}[Z]$ to be smaller than m, and by the lower semicontinuity of $\varphi $ , this implies that $\varphi (\dot {f}[Z])$ is at most m. First note that for any positive $n\in \omega $ , it holds that $q_{n}^{\frown } \dot {s}_{n+1}\leq q_{n-1}^{\frown } \dot {s}_{n}$ . To see this fix a positive $n\in \omega $ . Clause j) is equivalent to write $q_{n-1}^{\frown } \dot {q}_n\Vdash ` `\dot {s}_{n+1}\leq \dot {s}_{[1,\omega )}^n"$ . This, together with $q_{n-1}\Vdash ` `\dot {q}_n\leq \dot {s}_0^n"$ (see clause i)) and the factorization $\dot {s}_{n}=(\dot {s}_0^n,\dot {s}_{[1,\omega )}^n)$ gives the conclusion that $q_n^{\frown }\dot {s}_{n+1}\leq q_{n-1}^{\frown } \dot {s}_n$ . Then, note that this implies that for all $k\ge n$ , it holds that $q_k^{\frown } \dot {s}_{k+1}\leq q_{n-1}^{\frown } \dot {s}_{n}$ . This implies that for all $k\in \omega $ , $q\upharpoonright k=q_k\leq q_{n-1}^{\frown } \dot {s}_{n}\upharpoonright k$ , which means that $q\leq q_{n-1}^{\frown } \dot {s}_{n}$ . Since clause h) is equivalent to $q_{n-1}^{\frown } \dot {s}_n\Vdash ` `\dot {f}\upharpoonright n=\dot {h}_n\upharpoonright n"$ , we conclude that $q\Vdash ` `\dot {f}\upharpoonright n=\dot {h}_n\upharpoonright n"$ . Finally, since n was an arbitrary positive natural number, we conclude that $q\Vdash "(\forall n\in \omega )(\dot {f}\upharpoonright n=\dot {h}_n\upharpoonright n)"$ .

Let us construct the sequences. We start defining $\dot {s}_0=p_0$ , $\dot {h}_0=h_0$ and . Consider the factorization $\mathbb {P}=\mathbb {P}_0*\tilde {\mathbb {P}}_{[1,\omega )}$ . Relative to this factorization, any condition r in $\mathbb {P}_{\omega }$ is factored as $r=(r_0,\dot {r}_{[1,\omega )})$ . For the sequence $\langle r_n:n\in \omega \rangle $ we will write $r_n=(r_0^n,\dot {r}_{[1,\omega )}^{n})$ . Then note that all conditions from Lemma 5.5 are satisfied for $\mathscr {I}$ , $\varphi $ , $\mathbb {P}_0*\tilde {\mathbb {P}}_{[1,\omega )}$ , $\mathcal {U}$ , $\dot {f}$ , $p=(p_0,\dot {p}_{[1,\omega )})$ , $(\langle r_n:n\in \omega \rangle ,h_0),\ Z$ , $m,$ and $\mathcal {M}$ . So for $n=1$ , by an application of Lemma 5.5 we get $q_{1,\mathcal {M}}$ , $\tilde {s}_1$ , $\tilde {h}_1$ , that satisfy conditions 1) $-$ 4) from Lemma 5.5, and such that conditions $i)$ $iv)$ are adapted as follows:

  1. (1) $q_{1,\mathcal {M}}\leq r_0^0$ is a $(\mathcal {M},\mathbb {P}_0)$ -generic condition.

  2. (2) $q_{1,\mathcal {M}}\Vdash ` `\tilde {s}_{1}\leq \dot {p}_{[1,\omega )}"$ and $q_{\mathcal {M}}\Vdash ` `\tilde {s}_{1}\in \tilde {\mathbb {P}}_{[1,\omega )}[\dot {G}_{\mathbb {P}_0}]\cap \mathcal {M}[\dot {G}_{\mathbb {P}_0}]"$ .

  3. (3) $q_{1,\mathcal {M}}$ forces that is a fine interpretation of $\dot {f}$ below $\tilde {s}_1$ , as well as $\varphi (\tilde {h}_1[Z])<m$ .

  4. (4) $q_{1,\mathcal {M}}\Vdash ` `\tilde {s}_1\Vdash ` `\dot {f}\upharpoonright 1=\tilde {h}_1\upharpoonright 1""$ .

Now define $\dot {q}_0=q_{1,\mathcal {M}}$ , $\dot {s}_1=\tilde {s}_1$ , $\dot {h}_1=\tilde {h}_1$ , and just as was obtained. Let us see that conditions a)–j) are satisfied. Conditions a) and b) follow trivially from 1) above. Condition c) follows from 3) above. Condition d) follows from 3) above. Condition e) follows from 2) above. Conditions f) and g) follow from 3) above. Condition h) is condition 4) above. Condition i) follows from 1) and 2). Condition j) follows from 2).

Now assume that we have constructed $q_{n-1}$ , $\dot {s}_n$ , $\dot {h}_n,$ and with the required properties. Consider the factorization $\mathbb {P}_{\omega }=\mathbb {P}_{n-1}*\dot {\mathbb {Q}}_{n}*\tilde {\mathbb {P}}_{[n+1,\omega )}$ . Let $G_{n-1}\subseteq \mathbb {P}_{n-1}$ be a generic filter over V such that $q_{n-1}\in G_{n-1}$ . Then let $s_n=\dot {s}_n[G_{n-1}]$ , $h_{n}=\dot {h}_{n}[G_{n-1}],$ and $\vec {t}_n$ be the evaluations of $\dot {s}_{n}$ , $\dot {h}_n,$ and , respectively, by the filter $G_{n-1}$ . We work in $\mathcal {M}[G_{n-1}]$ . Then we have that $\dot {\mathbb {Q}}_{n}[G_{n-1}]*\tilde {\mathbb {P}}_{[n+1,\omega )}[G_{n-1}]\in \mathcal {M}[G_{n-1}]$ . Moreover,

$$ \begin{align*} \mathcal{M}[G_{n-1}]\vDash (\vec{t}_n,h_{n})\text{ is a fine interpretation of } \dot{f} \text{ below } s_{n} \end{align*} $$

and

$$ \begin{align*} V[G_{n-1}]\vDash\varphi(h_{n}[Z])< m. \end{align*} $$

Summarizing, we have the following:

  1. (1) $s_n=(s_0^n,\dot {s}_{[1,\omega )}^n),\dot {f},\mathscr {I},\varphi ,\mathcal {U},\dot {\mathbb {Q}}_{n}[G_{n-1}]*\tilde {\mathbb {P}}_{[n+1,\omega )}[G_{n-1}]\in \mathcal {M}[G_{n-1}]$ .

  2. (2) $(\vec {t}_n,h_{n})\in \mathcal {M}[G_{n-1}]$ is a fine interpretation of $\dot {f}$ below $s_{n}$ .

  3. (3) Z is a pseudointersection of $\mathcal {M}[G_{n-1}]\cap \mathcal {U}$ and $\varphi (h_n[Z])<m.$

Then we are in conditions to apply Lemma 5.5 for $n+1$ and $s_n$ taking the place of the condition p in the lemma. We get the following:

  1. (1) A $(\mathcal {M}[G_{n-1}],\dot {\mathbb {Q}}_{n}[G_{n-1}])$ -generic condition $q_{\mathcal {M}[G_{n-1}]}$ .

  2. (2) A $\dot {\mathbb {Q}}_{n}[G_{n-1}]$ -name $\tilde {s}_{n+1}$ for a condition in

    $\tilde {\mathbb {P}}_{[n+1,\omega )}[G_{n-1}][\dot {G}_{\dot {\mathbb {Q}}_n[G_{n-1}]}]\cap \mathcal {M}[G_{n-1}][\dot {G}_{\dot {\mathbb {Q}}_n[G_{n-1}]}]$ .

  3. (3) A $\dot {\mathbb {Q}}_{n}[G_{n-1}]$ -name $\tilde {h}_{n+1}$ for a finite to one function from $\omega $ to $\omega $ .

  4. (4) A $\dot {\mathbb {Q}}_{n}[G_{n-1}]$ -name for a sequence of conditions in $\tilde {\mathbb {P}}_{[n+1,\omega )}[G_{n-1}][\dot {G}_{\dot {\mathbb {Q}}_n[G_{n-1}]}]$ .

And such that the following hold:

  1. (i) $q_{\mathcal {M}[G_{n-1}]}\leq s_n^0$ is a $(\mathcal {M}[G_{n-1}],\dot {\mathbb {Q}}_n[G_{n-1}])$ -generic condition.

  2. (ii) $q_{\mathcal {M}[G_{n-1}]}\Vdash ` `\tilde {s}_{n+1}\leq \dot {s}_{[1,\omega )}^{n}"$ and $q_{\mathcal {M}}\Vdash ` `\tilde {s}_{n+1}\in \tilde {\mathbb {P}}_{[n+1,\omega )}[G_{n-1}] [\dot {G}_{\mathbb {Q}_n[G_{n-1}]}]\cap \mathcal {M}[\dot {G}_{n-1}][\dot {G}_{\mathbb {Q}_n[G_{n-1}]}]"$ .

  3. (iii) $q_{\mathcal {M}[G_{n-1}]}$ forces that is a fine interpretation of $\dot {f}$ below $\tilde {s}_{n+1}$ , as well as $\varphi (\tilde {h}_{n+1}[Z])<m$ .

  4. (iv) $q_{\mathcal {M}[G_{n-1}]}\Vdash ` `\tilde {s}_{n+1}\Vdash ` `\dot {f}\upharpoonright (n+1)=\tilde {h}_{n+1}\upharpoonright (n+1)""$ .

All of these happen regarding $V[G_{n-1}]$ and $\mathcal {M}[G_{n-1}]$ , and the construction was achieved by only assuming that $q_{n-1}\in G_{n-1}$ , and so everything above is forced by $q_{n-1}$ . Now, going back to V, we define the corresponding $\mathbb {P}_{n}$ -names:

  1. $\alpha $ ) $\dot {q}_n$ as a $\mathbb {P}_{n-1}$ -name for $\dot {q}_{\mathcal {M}[G_{n-1}]}$ .

  2. $\beta $ ) $\dot {s}_{n+1}$ as a $\mathbb {P}_{n-1}*\dot {\mathbb {Q}}_{n}$ -name for $\tilde {s}_{n+1}$ .

  3. $\gamma $ ) $\dot {h}_{n+1}$ as a $\mathbb {P}_{n-1}*\dot {\mathbb {Q}}_{n}$ -name for $\tilde {h}_{n+1}$ .

  4. $\delta $ ) as a $\mathbb {P}_{n-1}*\dot {\mathbb {Q}}_{n}$ -name for .

Finally, define $q_n=q_{n-1}\ ^{\frown } \dot {q}_n$ . Checking that these names are as required is routing as for case $n=1$ . This finishes the construction of the sequences, and therefore, the proof of Theorem 5.1.

6 Answer to Question 1.4

Now we are ready to answer Question 1.4 stated in Section 1. We only state a remark that help us to prove a slightly more general result.

Lemma 6.1. Let $\mathscr {I}$ an analytic p-ideal. There is a summable ideal $\mathscr {J}$ such that $\mathscr {J}\subseteq \mathscr {I}$ .

Proof Let $\varphi $ be a lscsm which defines $\mathscr {I}$ . Define $Z=\{n\in \omega :\varphi (\{n\})=0\}$ . Note that since $\varphi $ is a lscsm, the set Z is coinfinite. Now, for $n\notin Z$ , define $g(n)=\varphi (\{n\})$ , and for $n\in Z$ , define $g(n)=1/2^n$ . The lower semicontinuity of $\varphi $ implies that $\mathscr {I}_g\subseteq \mathscr {I}$ .

It is clear that if $\mathscr {J}\subseteq \mathscr {I}$ are ideals, and $\mathcal {U}$ is an $\mathscr {J}$ -ultrafilter, then $\mathcal {U}$ is an $\mathscr {I}$ -ultrafilter as well.

Theorem 6.2. It is consistent that there is no rapid ultrafilter but given any family $\mathcal D$ of analytic p-ideals such that $\lvert \mathcal D\rvert <\mathfrak {d}$ , there is an ultrafilter $\mathcal U$ which is an $\mathscr {I}$ -ultrafilter for all $\mathscr {I}\in \mathcal D$ .

Proof First note that for all summable ideal $\mathscr {I}_g$ , it holds that $\mathscr {I}_g=Fin(\varphi _g)=Exh(\varphi _g)$ , where $\varphi _g$ is the lscsm induced by $g:\omega \to \mathbb {R}^+$ . The forcing $\mathbb {P}_{\omega _2}$ is an $\omega _2$ -length countable support iteration of the Rational Perfect set forcing over a model of $\mathsf {ZFC}+\mathsf {CH}$ . Theorem 4.8 states that Rational Perfect forcing preserves $(\mathscr {I},p)$ -points whenever $\mathscr {I}$ in an analytic p-ideal, and Theorem 5.1 makes sure that $(\mathscr {I},p)$ -points are preserved along the iteration whenever $\mathscr {I}$ is an $F_{\sigma } \ p$ -ideal. Recall that in the Rational Perfect set model the dominating number is equal to $\omega _2$ . So let $\mathcal {D}$ be a family of analytic p-ideals such that $\vert \mathcal {D}\vert =\omega _1$ . Since every real appears before a stage of cofinality $\omega _1$ , it can be assumed that each ideal in $\mathcal {D}$ belongs to the ground model. By Lemma 6.1, for every $\mathscr {I}\in \mathcal {D}$ there is a summable ideal $\mathscr {J}(\mathscr {I})$ such that $\mathscr {J}(\mathscr {I})\subseteq \mathscr {I}$ . Define $\mathcal {D}'=\{\mathscr {J}(\mathscr {I}):\mathscr {I}\in \mathcal {D}\}$ . Since these ideals are from the ground model, there is an ultrafilter $\mathcal {U}$ that is an $(\mathscr {I},p)$ -point for every ideal $\mathscr {I}\in \mathcal {D}'$ (for example, $\mathcal {U}$ may be a Ramsey ultrafilter). This ultrafilter remains as an $(\mathscr {I},p)$ -point for every $\mathscr {I}\in \mathcal {D}'$ in the forcing extension by $\mathbb {P}_{\omega _2}$ , which implies that $\mathcal {U}$ is an $(\mathscr {I},p)$ -point for any $\mathscr {I}\in \mathcal {D}$ .

7 Final remarks

Let us recall that T. Bartoszyński and S. Shelah have proved that in the Rational Perfect set model, the Hausdorff ultrafilters are dense in the Rudin–Blass ordering (see [Reference Bartoszyński and Shelah2]).

Theorem 7.1 (Bartoszyński and Shelah, see [Reference Bartoszyński and Shelah2]).

In the Rational Perfect set model, Hausdorff ultrafilters are dense in the Rudin–Blass ordering.

Theorem 6.2 implies the same fact for a broader class of ideals. First let us recall the Near Coherence of Filters principle:

Definition 7.2 (Near Coherence of Filters principle (NCF); Blass, see [Reference Blass4]).

Let $\mathcal {U}$ and $\mathcal {V}$ be two ultrafilters on $\omega $ . Then there is a finite to one function $f:\omega \to \omega $ such that $f(\mathcal {U})=f(\mathcal {V})$ .

Theorem 7.3 (Blass and Shelah, see [Reference Blass and Shelah6]).

The Near Coherence of Filters holds in the Rational Perfect set model.

Let us work in the Rational Perfect set model. Let $\mathscr {I}$ be any analytic p-ideal, and apply Theorem 6.2 to the family $\mathcal {D}=\{\mathscr {I}\}$ to obtain an $\mathscr {I}$ -ultrafilter $\mathcal {U}$ . Now let $\mathcal {V}$ be an arbitrary ultrafilter on $\omega $ . By Theorem 7.3, there is a finite to one function $f\in \omega ^{\omega }$ such that $f(\mathcal {U})=f(\mathcal {V})$ . Since $f(\mathcal {U})$ is a $\mathscr {I}$ -ultrafilter and $f(\mathcal {U})=f(\mathcal {V})\leq _{RB}\mathcal {V}$ , we have found an $\mathscr {I}$ -ultrafilter below $\mathcal {V}$ . We have proved the following:

Corollary 7.4. It is relatively consistent with $\mathsf {ZFC}$ that for any analytic p-ideal, $\mathscr {I}$ -ultrafilters are dense in the Rudin–Blass ordering. Actually, this holds in the Rational Perfect set model.

Also, recall that C. Laflamme and J. Zhu, in their article The Rudin Blass ordering of ultrafilters from 1998 (see [Reference Laflamme and Zhu11]), have proved in $\mathsf {ZFC}$ that there is an ultrafilter $\mathcal {U}$ with no Rudin–Blass predecessors which are rapid. In particular, no predecessor of $\mathcal {U}$ is a q-point, equivalently, a weak $\mathcal {ED}_{fin}$ -ultrafilter. Corollary 7.4 implies that it is not possible to prove the same for $\mathscr {I}$ -ultrafilters whenever $\mathscr {I}$ is an analytic p-ideal.

Recall that given two ideals $\mathscr {I}$ and $\mathscr {J}$ on $\omega $ , we say that $\mathscr {I}$ is Katětov below $\mathscr {J}$ , denoted by $\mathscr {I}\leq _K\mathscr {J}$ , if there is a function $f:\omega \to \omega $ such that for all $A\in \mathscr {I}$ , $f^{-1}[A]\in \mathscr {J}$ . If the function f is required to be finite to one, then we say that $\mathscr {I}$ is Katětov–Blass below $\mathscr {J}$ , and it is denoted by $\mathscr {I}\leq _{KB}\mathscr {J}$ . For a given ideal $\mathscr {I}$ and an ultrafilter $\mathcal {U}$ , it is easily seen that $\mathcal {U}$ is an $\mathscr {I}$ -ultrafilter if and only if $\mathscr {I}\nleq _K\mathcal {U}^*$ . Similarly, $\mathcal {U}$ is a weak $\mathscr {I}$ -ultrafilter if and only if $\mathscr {I}\nleq _{KB}\mathcal {U}^*$ . Using this terminology, Laflamme and Zhu’s theorem can be restated as follows:

Theorem 7.5 (Laflamme and Zhu, see [Reference Laflamme and Zhu11]).

There exists an ultrafilter $\mathcal {U}$ such that for all finite to one $f\in \omega ^{\omega }$ , $\mathcal {ED}_{fin}\leq _{KB}f(\mathcal {U})^*$ .

Thus, it follows trivially that for any Borel ideal $\mathscr {I}\leq _{KB}\mathcal {ED}_{fin}$ , there is an ultrafilter for which any Rudin–Blass predecessor $\mathcal {V}\leq _{RB}\mathcal {U}$ , its dual ideal is Katětov–Blass above $\mathscr {I}$ , that is $\mathscr {I}\leq _{KB}\mathcal {V}^*$ . However, we do not know of a Borel ideal $\mathscr {I}\nleq _{KB}\mathcal {ED}_{fin}$ having this property.

Definition 7.6. Let $\mathscr {I}$ be a tall ideal on $\omega $ . We say that $\mathscr {I}$ is Laflamme–Zhu if there is an ultrafilter $\mathcal {U}$ all of whose Rudin–Blass predecessors are such that its dual ideal is Katětov–Blass above of $\mathscr {I}$ . We say that $\mathscr {I}$ is trivially Laflamme–Zhu if $\mathscr {I}<_{KB}\mathcal {ED}_{fin}$ .

The previous remarks lead us to ask the following:

Question 7.7. Does there exist a Borel or analytic tall ideal other than $\mathcal {ED}_{fin}$ which is non-trivially Laflamme–Zhu?

Question 7.8. Is there a critical ideal $\mathscr {I}$ for the Borel (analytic) Laflamme–Zhu ideals, i.e., such that any ideal $\mathscr {J}$ is Laflamme–Zhu if and only if $\mathscr {J}\leq _{KB}\mathscr {I}$ ?

Question 7.9. In case there is an analytic ideal $\mathscr {I}\nleq _{KB}\mathcal {ED}_{fin}$ which is Laflamme–Zhu, let $\mathbf {LZ}_{Borel}$ and $\mathbf {LZ}_{Analytic}$ be the families of all Borel and analytic ideals which are Laflamme–Zhu, respectively. What is the structure of the Katětov–Blass order restricted to such classes?

In the case of an affirmative answer to Question 7.7 , such ideal can not be Katětov–Blass above the ideal $\mathsf {conv}$ generated by convergent sequences of rationals in $\mathbb {Q}\cap [0,1]$ , since p-points are characterized as the $\mathsf {conv}$ -ultrafilters. Also, by Corollary 7.4, such an ideal can not be an analytic p-ideal. In particular it can not be a summable ideal. In [Reference Hrušák and Meza-Alcántara9] it has been proved that Hausdorff ultrafilters are exactly the $\mathcal {G}_{fc}$ -ultrafilters, which together with Theorem 7.1 imply that any Lafflame–Zhu ideal can not be Katětov–Blass above the ideal $\mathcal {G}_{fc}$ , the ideal on $[\omega ]^2$ of graphs with finite chromatic number.

Footnotes

1 The cardinal invariant $\mathfrak {d}$ is the dominating number. This cardinal invariant is defined in the next section.

References

Abraham, U., Proper forcing , Handbook of Set Theory (Foreman, M. and Kanamori, A., editors), Springer, Dordrecht, 2009, pp. 333394.Google Scholar
Bartoszyński, T. and Shelah, S., On the density of Hausdorff ultrafilters , Logic Colloquium 2004 (A. Andretta, K. Kearnes, and D. Zambella, editors), Lecture Notes in Logic, vol. 29, Association of Symbolic Logic, Chicago, 2008, pp. 1832.Google Scholar
Baumgartner, J. E., Ultrafilters on $\omega$ , this Journal, vol. 60 (1995), no. 2, pp. 624–639.Google Scholar
Blass, A., Near coherence of filters, I: Cofinal equivalence of models o arithmetic . Notre Dame Journal of Formal Logic , vol. 27 (1986), no. 4, pp. 579591.CrossRefGoogle Scholar
Blass, A. and Shelah, S., There may be simple ${P}_{\aleph_1}$ - and ${P}_{\aleph_2}$ -points and the Rudin–Keisler ordering may be downward directed . Annals of Pure and Applied Logic , vol. 33 (1987), no. 3, pp. 213243.CrossRefGoogle Scholar
Blass, A. and Shelah, S., Near coherence of filters. III. A simplified consistency proof . Notre Dame Journal of Formal Logic , vol. 30 (1989), no. (4), pp. 530538.CrossRefGoogle Scholar
Cancino-Manríquez, J., Every maximal ideal may be Katětov above of all ${F}_{\sigma }$ ideals . Transactions of the American Mathematical Society , vol. 375 (2022), 18611881.CrossRefGoogle Scholar
Flašková, J., Rapid ultrafilters and summable ideals, preprint, 2012. Available at https://arxiv.org/abs/1208.3690, http://home.zcu.cz/~blobner/research/IgnotRapid-2012.pdf.Google Scholar
Hrušák, M. and Meza-Alcántara, D., Katětov order, Fubini property and Hausdorff ultrafilters . Rendiconti dell’Istituto di Matematica dell’Università di Trieste , vol. 44 (2012), pp. 19.Google Scholar
Laflamme, C., Filter games and combinatorial properties of winning strategies . Contemporary Mathematics , vol. 192 (1996), pp. 5167.CrossRefGoogle Scholar
Laflamme, C. and Zhu, J.-P., The Rudin–Blass ordering of ultrafilters, this Journal, vol. 63 (1998), no. 2, pp. 584–592.Google Scholar
Mazur, K., ${F}_{\sigma }$ -ideals and ${\omega}_1{\omega}_1^{\ast }$ -gaps in the Boolean algebras $\mathbf{\mathcal{P}}\left(\omega \right)/ \mathcal{I}$ . Fundamenta Mathematicae , vol. 138 (1991), no. 2, pp. 103111. Available at http://eudml.org/doc/211857.CrossRefGoogle Scholar
Miller, A. W., Rational perfect set forcing . Contemporary Mathematics , vol. 31 (1984), pp. 143159.CrossRefGoogle Scholar
Solecki, S., Analytic ideals and their applications . Annals of Pure and Applied Logic , vol. 99 (1999), pp. 5172.CrossRefGoogle Scholar
Vojtáš, P., On ${\omega}^{\ast }$ and absolutely divergent series . Topology Proceedings , vol. 19 (1994), pp. 335348.Google Scholar
Wohofsky, W., On the existence of $\boldsymbol{p}$ -points and other ultrafilters in the Stone–Čech-compactification of $\mathbb{N}$ , Master’s thesis, Institute of Discrete Mathematics and Geometry, Vienna University of Technology, 2008.Google Scholar