This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On Semiproperness of Namba forcings and Ideals in Prikry extensions

Kenta Tsukuura
Abstract.

In this paper, we study some variations of Namba forcing Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) over 𝒫κλ\mathcal{P}_{\kappa}\lambda and show that its semiproperness implies SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},<\kappa).

In particular, Prikry forcing at μ\mu forces Namba forcing Nm(μ+)\mathrm{Nm}(\mu^{+}) is not semiproper. This shows that there is no semiproper saturated ideal in the extensions by Prikry-type forcings. We also show that [λ+]μ+[\lambda^{+}]^{\mu^{+}} cannot carry a λ+\lambda^{+}-saturated semiproper ideal if μ\mu is singular with countable cofinality.

Key words and phrases:
Namba forcing, Semistationary reflection principles, Prikry-type forcing, saturated ideal, huge cardinal
2020 Mathematics Subject Classification:
03E35, 03E40, 03E55

1. Introduction

The notion of a saturated ideal is one of generic large cardinal axioms. The first model of a saturated ideal over 1\aleph_{1} is due to Kunen. He established

Theorem 1.1 (Kunen [15] for μ=0\mu=\aleph_{0}).

If κ\kappa is a huge cardinal then for every regular cardinal μ<κ\mu<\kappa, there is a poset which forces that μ+=κ\mu^{+}=\kappa carries a saturated ideal.

Kunen’s proof is useful to construct a model with a saturated ideal over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda. Many saturated ideals, which have strengthnengths of a saturation property, were found by Kunen’s method like in [16] and  [9]. In [26], the author proved that Kunen’s original ideal of Theorem 1.1 is not so strong in the sense of saturation properties.

But, if μ1\mu\geq\aleph_{1} then the saturated ideals obtained by collapsing a huge cardinal are proper* (For example, see Proposition 5.1). It seems that properness* of ideals is very strong by the results of [17]. Here, by proper*, we mean the properness in the sense of forcing. For an ideal II over ZZ, the properness of II is ZIZ\not\in I in the usual sense. To distinguish properness of ideals and that of forcings, we write “proper*” according to [22].

In the above story, μ\mu is a regular cardinal. We are interested in the case of singular cardinal μ\mu. A model with a saturated ideal over μ+\mu^{+} was given by Foreman [6]. Then, the ideal is not proper*. Moreover, it is known that 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda cannot carry a proper* ideal if μ\mu singular by Matsubara–Shelah [18].

On the other hand, Sakai obtained a model with a semiproper ideal over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda in [22]. In Sakai’s proof, μ\mu can be taken as a singular cardinal. Notice that the ideal of Sakai’s model is not saturated. We ask

Question 1.2.

For a singular cardinal μ\mu, can μ+\mu^{+} carry a saturated and semiproper ideal?

All known models with a saturated ideal over a successor of a singular cardinal were given by using Prikry-type forcing, as far as the author knows. This paper aims to show all known saturated ideals are not semiproper by proving the following theorem.

Theorem 1.3.

Suppose that VWV\subseteq W are inner models and μ<λ\mu<\lambda are cardinals such that:

  1. (1)

    μ\mu and λ\lambda are regular cardinals in VV.

  2. (2)

    μ\mu is a singular cardinal in WW of cofinality ω\omega.

  3. (3)

    λ\lambda is a regular cardinal in WW.

  4. (4)

    (Eμλ)V(E^{\lambda}_{\mu})^{V} is a stationary subset in WW.

Then there is no semiproper precipitous ideals over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda in WW. In particular, Prikry-type forcing over μ\mu (like Prikry forcing and Woodin’s modification) forces there is no semiproper precipitous ideals over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda for all regular λμ+\lambda\geq\mu^{+}.

Moreover, if we consider an ideal over [λ+]μ+[\lambda^{+}]^{\mu^{+}} instead of 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda, the answer of Question 1.2 is NO.

Theorem 1.4.

Suppose that μ\mu is a singular cardinal with cofinality ω\omega and λ>μ\lambda>\mu is a regular cardinal. If [λ+]μ+[\lambda^{+}]^{\mu^{+}} carries a normal, fine, μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal II then II is not semiproper.

The key of the proof of Theorems 1.3 and 1.4 is the study of semiproperness of Namba forcing. Namba forcing was introduced by Namba [20] as an example of Boolean algebras that is (ω,ω1)(\omega,\omega_{1})-distributive but not (ω,ω2)(\omega,\omega_{2})-distributive.

In recent research, Namba forcing is studied in the context of semiproper forcing sometimes. For example, Namba forcing appears as one of the characterizations of dagger principle ()(\dagger) that was introduced by Foreman–Magidor–Shelah [5].

Theorem 1.5 (Shelah [24] for (1) \leftrightarrow (2), Doebler–Schindler [3] for (3) \to (2)).

The following are equivalent:

  1. (1)

    ()(\dagger) holds. That is, every ω1\omega_{1}-stationary preserving poset is semiproper.

  2. (2)

    SSR([λ]ω)\mathrm{SSR}([\lambda]^{\omega}) holds for all λ2\lambda\geq\aleph_{2}.

  3. (3)

    Nm(λ)\mathrm{Nm}(\lambda) is semiproper for all λ2\lambda\geq\aleph_{2}.

SSR([λ]ω)\mathrm{SSR}([\lambda]^{\omega}) and Nm(λ)\mathrm{Nm}(\lambda) are semistationary reflection principles and Namba forcing over λ\lambda, respectively. For the definition of them, we refer to Section 2. There is also a local relationship between (2) and (3) above as follows.

  • For λ\lambda-strongly compact cardinal κ\kappa, the Levy collapse Coll(1,<κ)\mathrm{Coll}(\aleph_{1},<\kappa) forces both SSR([λ]ω)\mathrm{SSR}([\lambda]^{\omega}) and the semiproperness of Nm(λ)\mathrm{Nm}(\lambda).

  • (Todorčević [25]) If Nm(2)\mathrm{Nm}(\aleph_{2}) is semiproper then SR([2]ω)\mathrm{SR}([\aleph_{2}]^{\omega}) holds. In particular, SSR([2]ω)\mathrm{SSR}([\aleph_{2}]^{\omega}) holds.

SR([2]ω)\mathrm{SR}([\aleph_{2}]^{\omega}) is the stationary reflection principle. We are interested in the tolerance between SSR([λ]ω)\mathrm{SSR}([\lambda]^{\omega}) and the semiproperness of Nm(λ)\mathrm{Nm}(\lambda). In this paper, we introduce Namba forcing Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) over 𝒫κλ\mathcal{P}_{\kappa}\lambda and show the following.

Theorem 1.6.

Suppose that μω<κ\mu^{\omega}<\kappa for all μκ2\mu\in\kappa\setminus\aleph_{2} and 2κλ\aleph_{2}\leq\kappa\leq\lambda are regular cardinals. The following are equivalent:

  1. (1)

    Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) preserves the semistationarity of any subset of [λ]ω[\lambda]^{\omega}.

  2. (2)

    SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa).

SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa) is one of variations of semistationary reflection principles introduced by Sakai [23]. This is equivalent with SSR([λ]ω)\mathrm{SSR}([\lambda]^{\omega}) if κ=2\kappa=\aleph_{2}. The semiproperness of Nm(κ,κ)\mathrm{Nm}(\kappa,\kappa) is also equivalent with that of Nm(κ)\mathrm{Nm}(\kappa) (See Lemma 3.10). Theorem 1.6 generalizes Todorčević’s result. We can regard the semiproperness of Namba forcing as one of the reflection principles.

Theorem 1.6 brings us the following observation. The proof of Theorem 1.3 is almost the same as that of the following theorem.

Theorem 1.7.

Suppose that VWV\subseteq W are inner models and μ<λ\mu<\lambda are cardinals such that:

  1. (1)

    μ\mu and λ\lambda are regular cardinals in VV.

  2. (2)

    μ\mu is a singular cardinal in WW of cofinality ω\omega.

  3. (3)

    λ\lambda is a regular cardinal in WW.

  4. (4)

    (Eμλ)V(E^{\lambda}_{\mu})^{V} is a stationary subset in WW.

Then Nm(μ+,λ)\mathrm{Nm}(\mu^{+},\lambda) is not semiproper for all regular λμ+\lambda\geq\mu^{+} in WW. In particular, Prikry-type forcing over μ\mu (like Prikry forcing and Woodin’s modification) forces Nm(μ+,λ)\mathrm{Nm}(\mu^{+},\lambda) is not semiproper for all regular λμ+\lambda\geq\mu^{+}. In particular, Nm(μ+)\mathrm{Nm}(\mu^{+}) is forced to be non semiproper111This also follows from the proof of [3, Theorem 5.7]..

The structure of this paper is as follows. In Section 2, we recall the basic facts of semistationary subsets, Namba forcings, and saturated ideals. In Section 3, we introduce Namba forcing Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) over 𝒫κλ\mathcal{P}_{\kappa}\lambda. We study the basic properties of Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda). In Section 4, we prove Theorems 1.6 and 1.7. In Section 5, we study the semiproperness of ideals in Prikry-type extensions. The proofs of Theorems 1.3 and 1.4 are given in this section. Section 6 is not related to saturated ideals. However, our results can be positioned in the context of strong compactness of cardinals. We will call these principles ()(\dagger)-aspects of strong compactness and conclude this paper with some observations.

2. Preliminaries

In this section, we recall basic facts of semistationary subsets, Namba forcings, and saturated ideals. We use [14] as a reference for set theory in general. For the topics of Namba forcings and saturated ideals, we refer to [24, Section XII] and  [7], respectively.

Our notation is standard. In this paper, by κ\kappa and λ\lambda, we mean regular cardinals greater than 2\aleph_{2} unless otherwise stated. We use μ\mu to denote an infinite cardinal. For κ<λ\kappa<\lambda, EκλE^{\lambda}_{\kappa} and E<κλE^{\lambda}_{<\kappa} denote the set of all ordinals below λ\lambda of cofinality κ\kappa and <κ<\kappa, respectively. We also write [κ,λ]={ξκξλ}[\kappa,\lambda]=\{\xi\mid\kappa\leq\xi\leq\lambda\}. By Reg\mathrm{Reg}, we mean the class of regular cardinals. For \subseteq-increasing finite sequences s,t[𝒫(X)]<ωs,t\in[\mathcal{P}(X)]^{<\omega}, by sts\sqsubseteq t, we means tt end-extends ss that is, atsbs(ba)\forall a\in t\setminus s\forall b\in s(b\subseteq a). λ\lambda-strongly compact cardinal κ\kappa is a regular cardinal in which 𝒫κλ\mathcal{P}_{\kappa}\lambda carries a fine ultrafilter.

In the proof of Theorem 1.3 and 1.7, we use the properties of Prikry forcing and Woodin’s modifications. The former was introduced in [21] and the letter appeared in [11].

Theorem 2.1 (Prikry [21]).

If μ\mu is a measurable cardinal then there is a poset PP with the following conditions:

  1. (1)

    PP is μ\mu-centered and PP preserves all cardinal.

  2. (2)

    Pcf˙(μ)=ωP\Vdash\dot{\mathrm{cf}}(\mu)=\omega.

  3. (3)

    If Pω<cf˙(δ)<μP\Vdash\omega<\dot{\mathrm{cf}}(\delta)<\mu then Pcf˙(δ)=cfˇ(δ)P\Vdash\dot{\mathrm{cf}}(\delta)=\check{cf}(\delta).

Proof.

For (1) and (2), we refer to [12]. For (3), see [2, Section 11]

Theorem 2.2 (Woodin).

If μ\mu is a measurable cardinal and 2μ=μ+2^{\mu}=\mu^{+} then there is a poset PP with the following conditions:

  1. (1)

    PP is μ\mu-centered.

  2. (2)

    Pμ=ωP\Vdash\mu=\aleph_{\omega}.

  3. (3)

    If Pω<cf˙(δ)<μP\Vdash\omega<\dot{\mathrm{cf}}(\delta)<\mu then ω<cf(δ)<μ\omega<{cf}(\delta)<\mu.

Proof.

See [11]. ∎

Lemma 2.3.

If WW is a generic extension by Prikry forcing over μ\mu or Woodin’s modification over μ\mu. Then, for every regular λμ+\lambda\geq\mu^{+}, the following holds.

  1. (1)

    μ\mu is a singular cardinal in WW of cofinality ω\omega.

  2. (2)

    λ\lambda is a regular cardinal in WW.

  3. (3)

    (Eμλ)V(E^{\lambda}_{\mu})^{V} is a stationary subset in WW.

Proof.

Easy. ∎

2.1. Semistationary subsets and Namba forcing

For a set Wω1W\supseteq\omega_{1}, a semistationary subset is an S[W]ωS\subseteq[W]^{\omega} such that S𝐜𝐥={x[W]ωyS(yω1x)}{S}^{\mathbf{cl}}=\{x\in[W]^{\omega}\mid\exists y\in S(y\sqsubseteq_{\omega_{1}}x)\} is stationary in [λ]ω[\lambda]^{\omega}. By yω1xy\sqsubseteq_{\omega_{1}}x, we mean yxyω1=xω1y\subseteq x\land y\cap\omega_{1}=x\cap\omega_{1}.

For a poset PP, we say that PP is semiproper if and only if {M[θ]ωpMPqp(q\{M\in[\mathcal{H}_{\theta}]^{\omega}\mid\forall p\in M\cap P\exists q\leq p(q is (M,P)(M,P)-semigeneric)})\} contains a club for all sufficiently large regular θ\theta. (M,P)(M,P)-semigeneric is a condition that forces M[G˙]ω1=Mω1M[\dot{G}]\cap\omega_{1}=M\cap\omega_{1}. The following lemma is well-known.

Lemma 2.4 (Shelah [24]).

The following are equivalent:

  1. (1)

    PP is semiproper.

  2. (2)

    PP preserves the semistationarity of any sets.

We also introduce a useful lemma due to Menas to study semistationary subsets.

Lemma 2.5 (Menas [19]).

Let WW and W¯\overline{W} be sets with ω1WW¯\omega_{1}\subseteq W\subseteq\overline{W}.

  1. (1)

    If C[W]ωC\subseteq[W]^{\omega} is a club then the set {x¯[W¯]ωx¯WC}\{\overline{x}\in[\overline{W}]^{\omega}\mid\overline{x}\cap W\in C\} is a club in [W¯]ω[\overline{W}]^{\omega}.

  2. (2)

    If C¯[W¯]ω\overline{C}\subseteq[\overline{W}]^{\omega} is a club then the set {x¯W[W]ωx¯C¯}\{\overline{x}\cap W\in[{W}]^{\omega}\mid\overline{x}\in\overline{C}\} contains a club in [W]ω[W]^{\omega}.

Lemma 2.6.

If S[W]ωS\subseteq[W]^{\omega} is semistationary in [W]ω[W]^{\omega} then SS is semistationary in [W¯]ω[\overline{W}]^{\omega}.

Proof.

This directly follows from Lemma 2.5. ∎

SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},<\kappa) is the statement that claims, for every semistationary subset S[λ]ωS\subseteq[\lambda]^{\omega}, there is an R𝒫κλR\in\mathcal{P}_{\kappa}\lambda with the following properties:

  1. (1)

    ω1Rκκ\omega_{1}\subseteq R\cap\kappa\in\kappa

  2. (2)

    S[R]ωS\cap[R]^{\omega} is semistationary.

We say that SS reflects to RR if (2)(2) above holds. To give an witness R𝒫κλR\in\mathcal{P}_{\kappa}\lambda of SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa), it is enough to give an R𝒫κλR\in\mathcal{P}_{\kappa}\lambda in which some semistatoinary S[λ]ωS\subseteq[\lambda]^{\omega} reflects to RR. So (1)(1) can be removed as follows.

Lemma 2.7.

If a semistationary subset S[λ]ωS\subseteq[\lambda]^{\omega} reflects to R𝒫κλR\in\mathcal{P}_{\kappa}\lambda then there is an R𝒫κλR^{\prime}\in\mathcal{P}_{\kappa}\lambda such that SS reflects to RR^{\prime} and ω1Rκκ\omega_{1}\subseteq R^{\prime}\cap\kappa\in\kappa.

Proof.

See [22]. ∎

Lemma 2.8.

If a semistationary subset S[λ]ωS\subseteq[\lambda]^{\omega} reflects to R𝒫κλR\in\mathcal{P}_{\kappa}\lambda then {a𝒫κλS[a]ω\{a\in\mathcal{P}_{\kappa}\lambda\mid S\cap[a]^{\omega} is semistationary}\} is co-bounded.

Proof.

This follows from Lemma 2.6. ∎

In the proof of Theorem 1.7, we use the following theorem.

Theorem 2.9 (Sakai [23]).

SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa) implies Refl(Eωλ,<κ)\mathrm{Refl}(E_{\omega}^{\lambda},<\kappa).

Here, Refl(Eωλ,<κ)\mathrm{Refl}(E_{\omega}^{\lambda},<\kappa) is the statement that, for every stationary subset SEωλS\subseteq E_{\omega}^{\lambda}, there is an αE<κλ\alpha\in E_{<\kappa}^{\lambda} such that SαS\cap\alpha is stationary in α\alpha.

Lastly, we conclude this subsection with the definition of Namba forcing. Namba forcing Nm(λ)\mathrm{Nm}(\lambda) is the set of all Namba trees. Namba tree is a tree p[λ]<ωp\subseteq[\lambda]^{<\omega} with the following conditions:

  1. (1)

    pp has a trunk tr(p)\mathrm{tr}(p), that is, tr(p)\mathrm{tr}(p) is the maximal tpt\in p such that sp(stts)\forall s\in p(s\subseteq t\lor t\subseteq s).

  2. (2)

    For all sps\in p, if str(p)s\supseteq\mathrm{tr}(p) then Suc(s)={ξ<λsξp}\mathrm{Suc}(s)=\{\xi<\lambda\mid s{{}^{\frown}}\langle\xi\rangle\in p\} is unbounded in λ\lambda.

Nm(λ)\mathrm{Nm}(\lambda) is ordered by inclusion. Note that we wrote sξs{{}^{\frown}}\langle\xi\rangle to denote s{ξ}s\cup\{\xi\}. But if we write sξs{{}^{\frown}}\langle\xi\rangle then we always assume maxs<ξ\max{s}<\xi. Therefore, sξs{{}^{\frown}}\langle\xi\rangle end-extends aa. We use a similar notation for finite subsets of an ordered set later.

Nm(λ)\mathrm{Nm}(\lambda) is ω1\omega_{1}-stationary preserving and that forces cf˙(λ)=ω\dot{\mathrm{cf}}(\lambda)=\omega. Therefore Nm(λ)\mathrm{Nm}(\lambda) is semiproper under ()(\dagger). Namba forcing is a representative of semiproper posets that change uncountable cofinalities in the following sense.

Theorem 2.10 (Shelah [24]).

The following are equivalent:

  1. (1)

    Nm(λ)\mathrm{Nm}(\lambda) is semiproper.

  2. (2)

    There is a poset that is semiproper and forces cf˙(λ)=ω\dot{\mathrm{cf}}(\lambda)=\omega.

  3. (3)

    Φλ\Phi_{\lambda}.

For the definition of Φλ\Phi_{\lambda}, see Section 3.

2.2. Saturated ideals

For an ideal II over ZZ, we write I+=𝒫(Z)II^{+}=\mathcal{P}(Z)\setminus I for the set of II-positive sets. 𝒫(Z)/I\mathcal{P}(Z)/I is a poset I+,\langle I^{+},\subseteq\rangle. We say that II is semiproper if 𝒫(Z)/I\mathcal{P}(Z)/I is semiproper. We say that II is λ\lambda-saturated if 𝒫(Z)I\mathcal{P}(Z)\setminus I has the λ\lambda-c.c. For an ideal over Z=𝒫κλZ=\mathcal{P}_{\kappa}\lambda, we simply say that II is saturated if II is λ+\lambda^{+}-saturated. Note that we can see an ideal over κ\kappa as an ideal over 𝒫κκ\mathcal{P}_{\kappa}\kappa.

For a precipitous ideal II over ZZ, by comp(I)\mathrm{comp}(I), we mean the least κ\kappa such that II is not κ+\kappa^{+}-complete. Then, the critical point of the generic ultrapower mapping j˙:VM˙V[G˙]\dot{j}:V\to\dot{M}\subseteq V[\dot{G}] is forced to be comp(I)\mathrm{comp}(I). We say that an ideal II is exactly and uniformly κ\kappa-complete if comp(IA)=κ\mathrm{comp}(I\upharpoonright A)=\kappa for all AI+A\in I^{+}. If Z=𝒫κXZ=\mathcal{P}_{\kappa}X or Z=[X]κZ=[X]^{\kappa} then every κ\kappa-complete ideal over ZZ is exactly and uniformly κ\kappa-complete.

Lemma 2.11.

Suppose that II is a normal, fine, exactly and uniformly μ+\mu^{+}-complete precipitous ideal over Z𝒫(X)Z\subseteq\mathcal{P}(X). Let j˙\dot{j} be a 𝒫(Z)/I\mathcal{P}(Z)/I-name for the generic ultrapower mapping j˙:VM˙V[G˙]\dot{j}:V\to\dot{M}\subseteq V[\dot{G}]. Then the following are forced by 𝒫(Z)/I\mathcal{P}(Z)/I:

  1. (1)

    M˙μ+M˙V[G˙]{{}^{\mu^{+}}\dot{M}}\subseteq\dot{M}\cap V[\dot{G}].

  2. (2)

    j˙``XM˙\dot{j}``X\in\dot{M}. If II is |X|+|X|^{+}-saturated then M˙|X|V[G˙]M˙{{}^{|X|}}\dot{M}\cap V[\dot{G}]\subseteq\dot{M}.

Proof.

See [7, Section 2]

For a later purpose, we see the cofinalities in the generic ultrapower. Let us introduce Shelah’s famous result.

Lemma 2.12 (Shelah).

Suppose that VWV\subseteq W are inner models. Let λ\lambda be a regular cardinal in VV. If (λ+)V(\lambda^{+})^{V} is a cardinal in WW then cfW(λ)=cfW(|λ|)\mathrm{cf}^{W}(\lambda)=\mathrm{cf}^{W}(|\lambda|).

Proof.

See [4, Theorem 4.73]

We will use Lemma 2.13 in Section 5. An example of use is computations of cofinalities. For a given saturated ideal over μ+\mu^{+}, if μ\mu is singular with the countable cofinality then II forces cf˙(μ+)=ω\dot{\mathrm{cf}}(\mu^{+})=\omega by Lemma 2.13.

Lemma 2.13.

Suppose that II is a normal, fine, exactly and uniformly μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal over Z𝒫(X)Z\subseteq\mathcal{P}(X). Suppose that λ=|X|\lambda=|X| is a regular cardinal. Then the following holds:

  1. (1)

    If Z𝒫μ+(X)Z\subseteq\mathcal{P}_{\mu^{+}}(X) then 𝒫(Z)/I\mathcal{P}(Z)/I forces that cf˙(λ)=cf˙(μ)\dot{\mathrm{cf}}(\lambda)=\dot{\mathrm{cf}}(\mu).

  2. (2)

    If Z[X]μ+Z\subseteq[X]^{\mu^{+}} and λ\lambda is a successor cardinal then 𝒫(Z)/I\mathcal{P}(Z)/I forces that cf˙(λ)=cf˙(μ)\dot{\mathrm{cf}}(\lambda^{-})=\dot{\mathrm{cf}}(\mu).

Proof.

By (2) of Lemma 2.11 and the assumption, λ+=ot(j˙``λ+)=j˙(μ+)\lambda^{+}=\mathrm{ot}(\dot{j}``\lambda^{+})=\dot{j}(\mu^{+}) in the extension. Therefore |λ+|=μ|\lambda^{+}|=\mu. Note that λ+\lambda^{+} is a cardinal by the λ+\lambda^{+}-saturation of II. So Lemma 2.12 shows the required result. ∎

In this paper, we often consider an ideal in some extension. We introduce the preservation theorem of saturated ideals. Foreman proved the conclusion of Lemma 2.14 in the case of Z=μ+Z=\mu^{+} in [6] at first but the same thing holds for a loud class.

Lemma 2.14 (Foreman [6]).

Suppose that II is a normal, fine, exactly and uniformly μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal over Z𝒫(X)Z\subseteq\mathcal{P}(X). Suppose one of the following.

  1. (1)

    Z𝒫μ+(X)Z\subseteq\mathcal{P}_{\mu^{+}}(X) and |X|=λ|X|=\lambda.

  2. (2)

    Z[X]μ+Z\subseteq[X]^{\mu^{+}} and |X|=λ+|X|=\lambda^{+}.

Then, for every μ\mu-centered poset PP forces I¯\overline{I} is a normal, fine, exactly and uniformly μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal over ZZ. I¯\overline{I} is PP-name for the ideal generated by II.

Proof.

This follows from Foreman’s duality theorem [8]. ∎

Lastly, we recall forcing projections for Proposition 5.1. For posets PP and QQ, a projection π:QP\pi:Q\to P is an order-preserving mapping with the property that qPπ(p)q\leq_{P}\pi(p) implies rQp(π(r)Pq)\exists r\leq_{Q}p(\pi(r)\leq_{P}q) and π(1Q)=1P\pi(1_{Q})=1_{P}. Whenever a projection π:QP\pi:Q\to P is given, for every dense DD in PP, π1D\pi^{-1}D is also dense in QQ. It follows that Qπ``H˙Q\Vdash\pi``\dot{H} generates a (V,P)(V,P)-generic filter, where H˙\dot{H} is the canonical name of (V,Q)(V,Q)-generic filter. The quotient forcing is defined by PQ/G˙={qQπ(q)G˙}P\Vdash Q/\dot{G}=\{q\in Q\mid\pi(q)\in\dot{G}\}, ordered by P\leq_{P}. Then PQ/G˙P\ast Q/\dot{G} is equivalent with QQ in the sense of Boolean completion.

We say that a projection π:QP\pi:Q\to P between complete Boolean algebras is continuous if π``X=π(X)\prod\pi``X=\pi(\prod X) for all XQX\subseteq Q with X0\prod X\not=0. The continuity of projection is useful when we try to analyze quotient forcing Q/G˙Q/\dot{G}. The following lemma will be used in a proof of Proposition 5.1.

Lemma 2.15.

If π:QP\pi:Q\to P is a continuous projection between complete Boolean algebras and QQ is κ\kappa-closed then PP forces Q/G˙Q/\dot{G} is κ\kappa-closed.

Proof.

Note that PP has a κ\kappa-closed dense subset. Let p{q˙ii<ν}[Q/G˙]<κp\Vdash\{\dot{q}_{i}\mid i<\nu\}\in[Q/\dot{G}]^{<\kappa} be a descending sequence. By induction, let us construct pip_{i} and qiq_{i} such that

  1. (1)

    pp0p1p\geq p_{0}\geq p_{1}\geq\cdots

  2. (2)

    ipi0\prod_{i}p_{i}\not=0.

  3. (3)

    piq˙i=qip_{i}\Vdash\dot{q}_{i}=q_{i}.

We note that piπ(qi)p_{i}\leq\pi(q_{i}) for each ii by (3). Since QQ is κ\kappa-closed, iqiQ\prod_{i}q_{i}\in Q. Then the continuity of π\pi shows

ipiiπ(qi)=π(iqi).\textstyle\prod_{i}p_{i}\leq\prod_{i}\pi(q_{i})=\pi(\prod_{i}q_{i}).

By the definition of Q/G˙Q/\dot{G}, ipi\prod_{i}p_{i} forces that iqiQ/G˙\prod_{i}q_{i}\in Q/\dot{G} and this is a lower bound of {q˙ii<ν}\{\dot{q}_{i}\mid i<\nu\}, as desired. ∎

3. Properties of Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)

From here, throughout this paper, we fix regular cardinals ω2κλ\omega_{2}\leq\kappa\leq\lambda. In this section, for a fine ideal II over 𝒫κλ\mathcal{P}_{\kappa}\lambda, we introduce Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) and study this. Most things in this section are analogies of original Nm(λ)\mathrm{Nm}(\lambda). An importance is that Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is ω1\omega_{1}-stationary preserving, this is shown in Lemma 3.8. Our proof of this is essentially due to Shelah (See Lemma 3.7).

For a fine ideal II over 𝒫κλ\mathcal{P}_{\kappa}\lambda, a (II-)Namba tree pp is a set p[𝒫κλ]<ωp\subseteq[\mathcal{P}_{\kappa}\lambda]^{<\omega} with the following conditions:

  1. (1)

    pp is a tree. That is, each sps\in p is \subseteq-increasing and pp closed under the initial segment.

  2. (2)

    pp has a trunk tr(p)\mathrm{tr}(p).

  3. (3)

    For each sps\in p, if str(p)s\sqsupseteq\mathrm{tr}(p) then Suc(s)={a𝒫κλsa}I+\mathrm{Suc}(s)=\{a\in\mathcal{P}_{\kappa}\lambda\mid s{{}^{\frown}}\langle a\rangle\}\in I^{+}.

Let Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) be the set of all II-Namba trees. Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is ordered by inclusion. we denote Nm(κ,λ,Jκλbd)\mathrm{Nm}(\kappa,\lambda,J_{\kappa\lambda}^{bd}) by Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda). JκλbdJ^{bd}_{\kappa\lambda} is the bounded ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda.

For qNm(κ,λ,I)q\in\mathrm{Nm}(\kappa,\lambda,I) and n<ωn<\omega, we write Levq(n)=q[𝒫κλ]n\mathrm{Lev}_{q}(n)=q\cap[\mathcal{P}_{\kappa}\lambda]^{n}. This set is the nn-th levels of qq. For p,qNm(κ,λ,I)p,q\in\mathrm{Nm}(\kappa,\lambda,I), by qpq\leq^{\ast}p, we mean qpq\leq p and tr(p)=tr(q)\mathrm{tr}(p)=\mathrm{tr}(q).

Lemma 3.1.

If II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda then, for all n<ωn<\omega, Dn={pNm(κ,λ,I)|tr(p)|n}D_{n}=\{p\in\mathrm{Nm}(\kappa,\lambda,I)\mid|\mathrm{tr}(p)|\geq n\} is a dense subset of Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I).

Proof.

Easy. ∎

Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) changes the cofinalities of regular cardinals lying between κ\kappa and λ\lambda.

Lemma 3.2.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. For all δ[κ,λ]Reg\delta\in[\kappa,\lambda]\cap\mathrm{Reg}, Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) forces cf(δ)=ω\mathrm{cf}(\delta)=\omega.

Proof.

Consider a Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name g˙\dot{g} for the set {tr(p)pG˙}\bigcup\{\mathrm{tr}(p)\mid p\in\dot{G}\}. g˙\dot{g} is forced to be a countable \subseteq-increasing sequence. Let g˙n\dot{g}_{n} be Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for the nn-th element of g˙\dot{g}. It is easy to see

  1. (1)

    ng˙n=λ\Vdash\bigcup_{n}\dot{g}_{n}=\lambda and

  2. (2)

    p{g˙0,,g˙n1}=tr(p)p\Vdash\{\dot{g}_{0},...,\dot{g}_{n-1}\}=\mathrm{tr}(p). Here, n=|tr(p)|n=|\mathrm{tr}(p)|.

For each δ[κ,λ]Reg\delta\in[\kappa,\lambda]\cap\mathrm{Reg}, by (1), ng˙nδ=δ\Vdash\bigcup_{n}\dot{g}_{n}\cap\delta=\delta. By (1), g˙nδ(𝒫κδ)V\Vdash\dot{g}_{n}\cap\delta\in(\mathcal{P}_{\kappa}\delta)^{V} for each n<ωn<\omega. Since cf(δ)=δκ\mathrm{cf}(\delta)=\delta\geq\kappa, ξ˙n=supg˙nδ\dot{\xi}_{n}=\sup\dot{g}_{n}\cap\delta is forced to be <δ<\delta. Of course, supnξn=δ\Vdash\sup_{n}\xi_{n}=\delta, as desired. ∎

We need a fusion sequence argument sometime.

Lemma 3.3.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. For a \leq-descending sequence qnn<ω\langle q_{n}\mid n<\omega\rangle, suppose that there is a increasing l:ωωl:\omega\to\omega such that

  • tr(qn)=tr(qn+1)\mathrm{tr}(q_{n})=\mathrm{tr(q_{n+1})} for every n<ωn<\omega.

  • For every n<ωn<\omega, Levqn(l(n))=Levqn+1(l(n))\mathrm{Lev}_{q_{n}}(l(n))=\mathrm{Lev}_{q_{n+1}}(l(n)) for every n<ωn<\omega.

Then, nqnNm(κ,λ,I)\prod_{n}q_{n}\in\mathrm{Nm}(\kappa,\lambda,I).

Proof.

Let q=nqnq=\bigcap_{n}q_{n} then qq is a tree with trunk tr(q0)\mathrm{tr}(q_{0}). By the assumption, for every n<ωn<\omega, let mm be a natural number such that n<l(m)n<l(m) then Levq(n)=Levqm(n)\mathrm{Lev}_{q}(n)=\mathrm{Lev}_{q_{m}}(n). Therefore qq is II-Namba tree, as desired. ∎

We often deal with an II-Namba tree that has a form of psp\upharpoonright s. Lemma 3.4 describes what this tree is and its properties.

Lemma 3.4.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. For pNm(κ,λ,I)p\in\mathrm{Nm}(\kappa,\lambda,I) and sps\in p, ps={tpstts}Nm(κ,λ,I)p\upharpoonright s=\{t\in p\mid s\sqsubseteq t\lor t\sqsubseteq s\}\in\mathrm{Nm}(\kappa,\lambda,I) is an II-Namba tree such that

  1. (1)

    tr(ps)=s\mathrm{tr}(p\upharpoonright s)=s.

  2. (2)

    For every qpq\leq p, if str(q)s\sqsubseteq\mathrm{tr}(q) then qpspq\leq p\upharpoonright s\leq p.

Proof.

Easy. ∎

We fix a Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for a ordinal α˙\dot{\alpha} such that α˙w\Vdash\dot{\alpha}\in w for some set |w|<κ|w|<\kappa. For an II-Namba tree pp, we say pp is α˙\dot{\alpha}-bad, if there is no qpq\leq^{\ast}p that decides the value of α˙\dot{\alpha}.

Lemma 3.5.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. For every Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for an ordinal α˙\dot{\alpha} and a set |w|<κ|w|<\kappa with α˙w\Vdash\dot{\alpha}\in w, if pNm(κ,λ,I)p\in\mathrm{Nm}(\kappa,\lambda,I) is α˙\dot{\alpha}-bad then {a𝒫κλp(tr(p)a)\{a\in\mathcal{P}_{\kappa}\lambda\mid p\upharpoonright(\mathrm{tr}(p){{}^{\frown}}\langle a\rangle) is α˙\dot{\alpha}-bad }I+\}\in I^{+}.

Proof.

We show the contraposition. Since A={a𝒫κλp(tr(p)a)A=\{a\in\mathcal{P}_{\kappa}\lambda\mid p\upharpoonright(\mathrm{tr}(p){{}^{\frown}}\langle a\rangle) is not α˙\dot{\alpha}-bad }I\}\in I^{\ast}, for each aAa\in A, there are pap(tr(p)a)p_{a}\leq^{\ast}p\upharpoonright(\mathrm{tr}(p){{}^{\frown}}\langle a\rangle) and αa\alpha_{a} such that paα˙=αap_{a}\Vdash\dot{\alpha}=\alpha_{a}. Since II is κ\kappa-complete, there is a βw\beta\in w such that {aAαa=β}I+\{a\in A\mid\alpha_{a}=\beta\}\in I^{+}. Consider a tree q={paaAαa=β}q=\bigcup\{p_{a}\mid a\in A\land\alpha_{a}=\beta\}. Then qpq\leq^{\ast}p forces α˙=β\dot{\alpha}=\beta, as desired. ∎

Lemma 3.6.

If II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda then, for every Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for an ordinal α˙\dot{\alpha} and a set |w|<κ|w|<\kappa with α˙w\Vdash\dot{\alpha}\in w, there is no α˙\dot{\alpha}-bad condition.

Proof.

Suppose otherwise. Let qPq\in P be an α˙\dot{\alpha}- bad condition. By Lemmas 3.3 and  3.5, We can define qpq\leq^{\ast}p such that, for all sqs\in q, Suc(s)={a𝒫κλq(sa)\mathrm{Suc}(s)=\{a\in\mathcal{P}_{\kappa}\lambda\mid q\upharpoonright(s^{\frown}\langle a\rangle) is α˙\dot{\alpha}-bad }\}. Then every extension of qq does not decide the value of α˙\dot{\alpha}. This is a contradiction. ∎

To prove Lemmas 3.8 and 3.9, we need the following lemma.

Lemma 3.7.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. For a set aa, if [a]ω[a]^{\omega} has a club subset DD of size <κ<\kappa then, for every pNm(κ,λ,I)p\in\mathrm{Nm}(\kappa,\lambda,I), if pC˙[a]ωp\Vdash\dot{C}\subseteq[a]^{\omega} is a club then {z[a]ωqp(qzC˙)}\{z\in[a]^{\omega}\mid\exists q\leq p(q\Vdash z\in\dot{C})\} contains a club.

Proof.

For each zDz\in D, let us define a two player game GzG_{z} of length ω\omega as follows:

Player I X0X_{0}, ξ0\xi_{0} \cdots Xi,ξiX_{i},\xi_{i} \cdots
Player II a0,y0a_{0},y_{0}, q0q_{0} \cdots ai,yi,qia_{i},y_{i},q_{i} \cdots

Players I and II must choose to satisfy the following:

  • XiIX_{i}\in I.

  • ξiz\xi_{i}\in z, yiDy_{i}\in D, and yiyi+1y_{i}\subseteq y_{i+1}.

  • pq0q1p\geq q_{0}\geq q_{1}\geq\cdots.

  • aiSucqi(tr(qi))Xia_{i}\in\mathrm{Suc}_{q_{i}}(\mathrm{tr}(q_{i}))\setminus X_{i}.

  • tr(qi)=|tr(p)|+i\mathrm{tr}(q_{i})=|\mathrm{tr}(p)|+i and max(tr(qi))=ai\mathrm{max}(\mathrm{tr}(q_{i}))=a_{i}.

  • qiξiyiC˙q_{i}\Vdash\xi_{i}\in y_{i}\in\dot{C}.

Note that Player II can choose qiq_{i} anytime by Lemma 3.6. Player II wins if iyiz\bigcup_{i}y_{i}\subseteq z. It is easy to see that GzG_{z} is an open game, and thus, GzG_{z} is determined.

First, we claim that C0={z[a]ωC_{0}=\{z\in[a]^{\omega}\mid Player II has a winning strategy in Gz}G_{z}\} contains a club subset. Suppose otherwise. Then, there is a stationary subset TDT\subseteq D such that, for every zTz\in T, Player I has a winning strategy in GzG_{z} by the determinacy of games. Let wszzT\langle\mathrm{ws}_{z}\mid z\in T\rangle be I’s winning strategies. Let θ\theta be a sufficiently large regular cardinal. Consider an elementary substructure MΨM\prec\mathcal{H}_{\Psi} such that

  1. (1)

    MM is countable.

  2. (2)

    κ,λ,I,p,C˙,D,wszzTM\kappa,\lambda,I,p,\dot{C},D,\langle\mathrm{ws}_{z}\mid z\in T\rangle\in M.

  3. (3)

    Ma=z¯TM\cap a=\overline{z}\in T.

To prove the contradiction, let us find a sequence of Player II’s move ai,yi,qii<ω\langle a_{i},y_{i},q_{i}\mid i<\omega\rangle in which ai,yi,qiM\langle a_{i},y_{i},q_{i}\rangle\in M and this is a regal move after Player I taked wsz¯(aj,yj,qjj<i)\mathrm{ws}_{\overline{z}}(a_{j},y_{j},q_{j}\mid j<i). In this play, Player II wins while Player I uses her winning strategy, as we see later.

Suppose that aj,yj,qjj<iM\langle a_{j},y_{j},q_{j}\mid j<i\rangle\in M has been obtained. For each zTz\in T, let Xiz,xiz=wsz(aj,yj,qjj<i)\langle X_{i}^{z},x_{i}^{z}\rangle=\mathrm{ws}_{z}(\langle a_{j},y_{j},q_{j}\mid j<i\rangle) if aj,yj,qjj<i\langle a_{j},y_{j},q_{j}\mid j<i\rangle is a partial play of GzG_{z} along wsz\mathrm{ws}_{z}. Note that Xiz,xizzTM\langle X_{i}^{z},x_{i}^{z}\mid z\in T\rangle\in M by the induction hypothesis and wszzTM\langle\mathrm{ws}_{z}\mid z\in T\rangle\in M. By |D|<κ|D|<\kappa, X=zTXizIX=\bigcup_{z\in T}X_{i}^{z}\in I. So, in MM, we can pick aiSucqi1(tr(qi1))XSucqi1(tr(qi1))Xiz¯a_{i}\in\mathrm{Suc}_{q_{i-1}}(\mathrm{tr}(q_{i-1}))\setminus X\subseteq\mathrm{Suc}_{q_{i-1}}(\mathrm{tr}(q_{i-1}))\setminus X_{i}^{\overline{z}}. Then qi1tr(qi1)aiMq_{i-1}\upharpoonright\mathrm{tr}(q_{i-1}){{}^{\frown}}\langle a_{i}\rangle\in M. By the rule of Gz¯G_{\overline{z}}, ξiz¯z¯=Ma\xi_{i}^{\overline{z}}\in\overline{z}=M\cap a. By Lemma 3.6, we can choose qiq_{i} and yiy_{i} such that qiξiz¯yiG˙q_{i}\Vdash\xi_{i}^{\overline{z}}\in y_{i}\in\dot{G} and yi1yiy_{i-1}\subseteq y_{i}. The induction is completed.

Player I used a winning strategy wsz¯\mathrm{ws}_{\overline{z}} in this play but Player II wins. Indeed, iyiMa=z¯\bigcup_{i}y_{i}\subseteq M\cap a=\overline{z} by each yiy_{i} is in MM. This is a contradiction.

Fix zC0z\in C_{0}. Lastly, we claim that there is a qpq\leq p that forces zC˙z\in\dot{C}. Let {ξ0,ξ1,,}\{\xi_{0},\xi_{1},...,\} enumerates zz. By induction, let us define a descending sequence q0q1q_{0}\geq^{\ast}q_{1}\geq^{\ast}\cdots with the following conditions:

  1. (1)

    q0=pq_{0}=p.

  2. (2)

    Levj(qi)=Levj(qi+1)\mathrm{Lev}_{j}(q_{i})=\mathrm{Lev}_{j}(q_{i+1}) for all j<|tr(p)|+ij<|\mathrm{tr}(p)|+i.

  3. (3)

    For every sLev|tr(p)|+i(qi)s\in\mathrm{Lev}_{|\mathrm{tr}(p)|+i}(q_{i}), if we write str(p)={a0s,,ais}s\setminus\mathrm{tr}(p)=\{a_{0}^{s},...,a_{i}^{s}\} then there is a yisy_{i}^{s} such that ais,yis,qis=ws(Xjs,ξjji)\langle a^{s}_{i},y^{s}_{i},q^{s}_{i}\rangle=\mathrm{ws}(\langle X^{s}_{j},\xi_{j}\mid j\leq i) for some X0s,,XisIX^{s}_{0},...,X^{s}_{i}\in I.

Suppose that qiq_{i} has been defined. Let us define qi+1q_{i+1}. For each sLev|tr(p)|+i(qi)s\in\mathrm{Lev}_{|\mathrm{tr}(p)|+i}(q_{i}), first we consider q=qisq^{\prime}=q_{i}\upharpoonright s. Let XsX_{s} be the set of all aSucq(s)a\in\mathrm{Suc}_{q^{\prime}}(s) such that aa appears in ws(Xjs,ξjj<iX,ξi))\mathrm{ws}(\langle X^{s}_{j},\xi_{j}\mid j<i\rangle{{}^{\frown}}\langle X,\xi_{i}\rangle)) for some XIX\in I. XsX_{s} need to be an II-positive set, that is XsI+X_{s}\in I^{+}.

Suppose otherwise, let a,y,q=ws(Xjs,ξjj<iXs,ξi)\langle a,y,q\rangle=\mathrm{ws}(\langle X^{s}_{j},\xi_{j}\mid j<i\rangle{{}^{\frown}}\langle X_{s},\xi_{i}\rangle). By the definition of XsX_{s}, aXsa\in X_{s}. On the other hand, by the rule of GzG_{z}, aXsa\not\in X_{s}. This is a contradiction. So XsI+X_{s}\in I^{+}. For each aXsa\in X_{s}, there are yay_{a} and qaq_{a} such that a,ya,qa=ws(Xjs,ξjj<iX,ξi)\langle a,y_{a},q_{a}\rangle=\mathrm{ws}(\langle X^{s}_{j},\xi_{j}\mid j<i\rangle{{}^{\frown}}\langle X,\xi_{i}\rangle) for some XIX\in I. Define qi+1q_{i+1} by

{qasLev|tr(p)|+i(qi)\bigcup\{q_{a}\mid s\in\mathrm{Lev}_{|\mathrm{tr}(p)|+i}(q_{i}) and aXs}a\in X_{s}\}

Note that tr(qa)=tr(q)sa\mathrm{tr}(q_{a})=\mathrm{tr}(q){{}^{\frown}}s{{}^{\frown}}\langle a\rangle for each aXsa\in X_{s}, Sucqi+1(s)=Xs\mathrm{Suc}_{q_{i+1}}(s)=X_{s} and Levqi+1(|tr(p)|+i)=Levqi(|tr(q)|+i)\mathrm{Lev}_{q_{i+1}}(|\mathrm{tr}(p)|+i)=\mathrm{Lev}_{q_{i}}(|\mathrm{tr}(q)|+i).

By the definition of {qii<ω}\{q_{i}\mid i<\omega\} and Lemma 3.3, a lower bound q=iqipq=\prod_{i}q_{i}\leq p exists. Let y˙i\dot{y}_{i} be a Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name such that

qy˙i=yisq\Vdash\dot{y}_{i}=y_{i}^{s} if and only if sg˙s\sqsubseteq\dot{g}.

This is well-defined. It is easy to see that qξiy˙iy˙i+1C˙q\Vdash\xi_{i}\in\dot{y}_{i}\subseteq\dot{y}_{i+1}\in\dot{C} for each ii. Then qiy˙i=zC˙q\Vdash\bigcup_{i}{\dot{y}_{i}}=z\in\dot{C}, as desired. ∎

Lemma 3.8.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. If SS is semistationary subset of [a]ω[a]^{\omega}. If [a]ω[a]^{\omega} has a club subset DD of size <κ<\kappa then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) preserves the semistationarity of SS. In particular, Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is ω1\omega_{1}-stationary preserving.

Proof.

Let pNm(κ,λ,I)p\in\mathrm{Nm}(\kappa,\lambda,I) and C˙\dot{C} be such that pC˙[a]ωp\Vdash\dot{C}\subseteq[a]^{\omega} is a club. By Lemma 3.7, we can take qpq\leq p and ySy\in S such that yω1zy\sqsubseteq_{\omega_{1}}z and qzC˙q\Vdash z\in\dot{C}, as desired. ∎

Proposition 3.9.

If SS is a stationary subset of [a]ω[a]^{\omega}. If [a]ω[a]^{\omega} has a club subset DD of size <κ<\kappa then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) preserves the stationarity of SS. In particular, if cf(ν)ω<κ\mathrm{cf}(\nu)^{\omega}<\kappa and cf(μ)>ω\mathrm{cf}(\mu)>\omega then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) forces cf(ν)>ω{\mathrm{cf}}(\nu)>\omega.

Proof.

For the preservation of stationary subsets, Lemma 3.7 works as well as the proof of Lemma 3.8. We only check about uncountable cofinalities. Let X={νξξ<cf(ν)}X=\{\nu_{\xi}\mid\xi<\mathrm{cf}(\nu)\} be a cofinal subset of ν\nu. Then [X]ω[X]^{\omega} has a club subset of size <κ<\kappa. Therefore, [X]ω[X]^{\omega} is forced to be a stationary subset. So any countable subset of XX in the extension is covered by some element of ([X]ω)V([X]^{\omega})^{V}, as desired. ∎

In the proof of Theorem 1.7, we will use the following lemma:

Lemma 3.10.

The following are equivalent:

  1. (1)

    Nm(κ,κ)\mathrm{Nm}(\kappa,\kappa) is semiproper.

  2. (2)

    Nm(κ)\mathrm{Nm}(\kappa) is semiproper.

To show Lemma 3.10, we need to characterize the semiproperness of Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) in terms of game theory. Let us introduce the principle Φκ,λ,I\Phi_{\kappa,\lambda,I}. Φκ,λ,I\Phi_{\kappa,\lambda,I} is the statement that the player II has a winning strategy for the Galvin game (I,A)\Game(I,A) for all AI+A\in I^{+}. (I,A)\Game(I,A) is a game of length ω\omega with two players as follows:

Player I F0:Aω1F_{0}:A\to\omega_{1} \cdots Fi:Aω1F_{i}:A\to\omega_{1} \cdots
Player I ξ0<ω1\xi_{0}<\omega_{1} \cdots ξi<ω1\xi_{i}<\omega_{1} \cdots

Let ξ=supnξn\xi=\sup_{n}\xi_{n}. II wins if n<ωFn1ξI+\bigcap_{n<\omega}F_{n}^{-1}\xi\in I^{+}. We write Φκ,λ\Phi_{\kappa,\lambda} for Φκ,λ,Jκλbd\Phi_{\kappa,\lambda,J_{\kappa\lambda}^{\mathrm{bd}}}. Note that, the Galvin game can be defined for any ideal II over any set ZZ. Φκ\Phi_{\kappa} is the statement that Player II has a winning strategy for (μ+,Jκbd)\Game(\mu^{+},J_{\kappa}^{\mathrm{bd}}). Φκ\Phi_{\kappa} is equivalent with the semiproperness of Nm(κ)\mathrm{Nm}(\kappa) as we saw in Theorem 2.10. Lemma 3.11 is an analog of this. Lemma 3.10 follows by Φκ,κΦκ\Phi_{\kappa,\kappa}\leftrightarrow\Phi_{\kappa} (See Lemmas 3.12 and 3.13).

Note that Φκ\Phi_{\kappa} is a game theoretical variation of Chang’s conjecture. For a detail, we refer to  [24, Theorem 2.5 of Section XII]. We see Φκ,λ\Phi_{\kappa,\lambda} as a variation of Chang’s conjectures.

Lemma 3.11.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. The following are equivalent.

  1. (1)

    Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper.

  2. (2)

    Φκ,λ,I\Phi_{\kappa,\lambda,I} holds.

Proof.

First, we show the forward direction. Let C[θ]ωC\subseteq[\mathcal{H}_{\theta}]^{\omega} be a club such that, for every MCM\in C,

  • MθM\prec\mathcal{H}_{\theta},

  • κ,λ,IM\kappa,\lambda,I\in M, and,

  • for every pMNm(κ,λ,I)p\in M\cap\mathrm{Nm}(\kappa,\lambda,I), there is a qpq\leq p that forces M[G˙]ω1=Mω1M[\dot{G}]\cap\omega_{1}=M\cap\omega_{1}.

Consider an expansion 𝒜=θ,,C,κ,λ,I\mathcal{A}=\langle\mathcal{H}_{\theta},\in,C,\kappa,\lambda,I\rangle. Let us describe a winning strategy for Player II in (A,I)\Game(A,I).

When a function F:Aω1F:A\to\omega_{1} is given, let α˙F\dot{\alpha}_{F} be a Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for supF``g˙\sup F``\dot{g}. Let pAp_{A} be a condition such that tr(pA)=\mathrm{tr}(p_{A})=\emptyset and SucpA(s)A\mathrm{Suc}_{p_{A}}(s)\subseteq A for all spAs\in p_{A}. Note that pAg˙Ap_{A}\Vdash\dot{g}\subseteq A. By Lemma 3.8, pAα˙F<ω1p_{A}\Vdash\dot{\alpha}_{F}<\omega_{1}.

Suppose that Player I played F0,,Fi\langle F_{0},...,F_{i}\rangle. Let MiM_{i} be a Skolem hull Sk𝒜({pA}{F0,,Fi})\mathrm{Sk}_{\mathcal{A}}(\{p_{A}\}\cup\{F_{0},...,F_{i}\}). Player II choose ξi=Miω1\xi_{i}=M_{i}\cap\omega_{1}. Then Player II wins. Note that M=iMiCM=\bigcup_{i}M_{i}\in C by the definition of MiM_{i}’s. Let ξ=supiξi\xi=\sup_{i}\xi_{i}. We have ξ=Mω1\xi=M\cap\omega_{1}. By the choice of CC, we can take a qpAq\leq p_{A} that forces ξ=Mω1=M[G˙]ω1\xi=M\cap\omega_{1}=M[\dot{G}]\cap\omega_{1}.

Let n=|tr(q)|n=|\mathrm{tr}(q)| and let g˙n\dot{g}_{n} be the Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for the nn-th element of g˙\dot{g}. We have

Fi(g˙n)α˙FiM[G˙]ω1=ξ\Vdash F_{i}(\dot{g}_{n})\leq\dot{\alpha}_{F_{i}}\in M[\dot{G}]\cap\omega_{1}=\xi.

For every aSucq(tr(q))a\in\mathrm{Suc}_{q}(\mathrm{tr}(q)), q(tr(q)a)q\upharpoonright(\mathrm{tr}(q){{}^{\frown}}\langle a\rangle) forces g˙n=a\dot{g}_{n}=a, and thus, Fi(a)α˙Fi<ξF_{i}(a)\leq\dot{\alpha}_{F_{i}}<\xi. So Sucq(tr(q))iFi1ξ\mathrm{Suc}_{q}(\mathrm{tr}(q))\subseteq\bigcap_{i}F^{-1}_{i}\xi is II-positive, as desired.

Let us show the inverse direction. We fix a sequence of winning strategies wsAAI+\langle\mathrm{ws}_{A}\mid A\in I^{+}\rangle. For a countable MθM\prec\mathcal{H}_{\theta} with wsAAI+M\langle\mathrm{ws}_{A}\mid A\in I^{+}\rangle\in M. For pPp\in P, let us find qpq\leq^{\ast}p which forces Mω1=M[G˙]ω1M\cap\omega_{1}=M[\dot{G}]\cap\omega_{1}. Let α˙0,α˙1,\dot{\alpha}_{0},\dot{\alpha}_{1},... be an enumeration of Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-names for a countable ordinal belonging to MM.

We put t=tr(p)t=\mathrm{tr}(p). Let F0t:Suc(t)ω1F^{t}_{0}:\mathrm{Suc}(t)\to\omega_{1} be a function such that:

  • p(ta)p\upharpoonright(t^{\frown}\langle a\rangle) has a direct extension pa0p^{0}_{a} that forces α˙0=F0t(a)\dot{\alpha}_{0}=F^{t}_{0}(a).

Let ξ0\xi_{0} be II’s first play using wsSuc(t)\mathrm{ws}_{\mathrm{Suc}(t)} and A0t=(F0t)1{ξ0}A^{t}_{0}=({F^{t}_{0}})^{-1}\{\xi_{0}\}. Define p0=aA0pa0p_{0}=\bigcup_{a\in A^{0}}p_{a}^{0}.

Next, let us define p1p_{1}. First, for each sLev|t|+1(p0)s\in\mathrm{Lev}_{|t|+1}(p_{0}), let F0s:Suc(s)ω1F^{s}_{0}:\mathrm{Suc}(s)\to\omega_{1} be a function such that

  • p0(ta)p_{0}\upharpoonright(t^{\frown}\langle a\rangle) has a direct extension pasp^{s}_{a} that forces α˙1=F0t(a)\dot{\alpha}_{1}=F^{t}_{0}(a).

Let ξ0s\xi_{0}^{s} be II’s first play using wsSuc(s)\mathrm{ws}_{\mathrm{Suc}(s)} and A0s=(F0s)1{ξ0s}A^{s}_{0}=(F^{s}_{0})^{-1}\{\xi_{0}^{s}\}. Let us define F1t:Suc(t)ω1F^{t}_{1}:\mathrm{Suc}(t)\to\omega_{1} by

  • If aA0ta\in A_{0}^{t} then F1t(a)=ξ0taF^{t}_{1}(a)=\xi_{0}^{t^{\frown}\langle a\rangle}.

  • If aA0ta\not\in A_{0}^{t} then F1t(a)=0F^{t}_{1}(a)=0.

Let ξ1t\xi_{1}^{t} be II’s second play using wsSuc(s)\mathrm{ws}_{\mathrm{Suc}(s)} and A1t=(F1t)1{ξ1t}A^{t}_{1}=(F^{t}_{1})^{-1}\{\xi_{1}^{t}\}. Define p1=aA1tbA0taptabp_{1}=\bigcup_{a\in A^{t}_{1}}\bigcup_{b\in A_{0}^{t^{\frown}\langle a\rangle}}p_{t^{\frown}\langle a\rangle}^{b}. Of course, p1p0p_{1}\leq^{\ast}p_{0}.

Similarly, we continue this process for all α˙n\dot{\alpha}_{n}. Then q=npnq=\bigcap_{n}p_{n} is a tree such that, for all sqs\in q, if sts\sqsupseteq t then Suq(s)=nAns\mathrm{Suq}(s)=\bigcap_{n}A_{n}^{s}. Each of the components AnsA_{n}^{s} was a response by II using her winning strategy. Therefore nAnsI+\bigcap_{n}A_{n}^{s}\in I^{+}. So qNm(κ,λ,I)q\in\mathrm{Nm}(\kappa,\lambda,I) and that forces α˙n\dot{\alpha}_{n} is bounded by ξnt\xi_{n}^{t} and it is in MM. In particular, qMω1=M[G˙]ω1q\Vdash M\cap\omega_{1}=M[\dot{G}]\cap\omega_{1}, as desired. ∎

Lemma 3.12.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. If Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper then Nm(δ)\mathrm{Nm}(\delta) is semiproper for all δ[κ,λ]Reg\delta\in[\kappa,\lambda]\cap\mathrm{Reg}.

Proof.

By the assumption and Lemma 3.2, Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is a semiproper poset that changes the cofinality of δ\delta to ω\omega. By Theorem 2.10, Nm(δ)\mathrm{Nm}(\delta) is semiproper. ∎

Lemma 3.13.

Φκ\Phi_{\kappa} implies Φκ,κ\Phi_{\kappa,\kappa}.

Proof.

Let ws\mathrm{ws}^{\prime} be a II’s winning strategy of (κ,Jκbd)\Game(\kappa,J_{\kappa}^{\mathrm{bd}}). Since AA has a subset {aξξ<κ}A\{a_{\xi}\mid\xi<\kappa\}\subseteq A such that ξ<ζot(aξ)<ot(aζ)\xi<\zeta\to\mathrm{ot}(a_{\xi})<\mathrm{ot}(a_{\zeta}) and supξot(aξ)=κ\sup_{\xi}\mathrm{ot}(a_{\xi})=\kappa. Define an ordinal αξ=ot(aξ)\alpha_{\xi}=\mathrm{ot}(a_{\xi}). For F:Aω1F:A\to\omega_{1}, define F¯:κω1\overline{F}:\kappa\to\omega_{1} by F¯(ξ)=F(aξ)\overline{F}(\xi)=F(a_{\xi}). Define II’s strategy ws\mathrm{ws} for (A,Jκλbd)\Game(A,J_{\kappa\lambda}^{\mathrm{bd}}) by ws(F0,,Fn)=ws(F¯0,,F¯n)\mathrm{ws}(F_{0},...,F_{n})=\mathrm{ws}^{\prime}(\overline{F}_{0},...,\overline{F}_{n}). It is easy to see that ws\mathrm{ws} is a winning strategy. ∎

The rest of this section is not related to the main theorems, but we introduce these. The consistency of Φκ,λ,I\Phi_{\kappa,\lambda,I} was shown in [10]. Recall that PP is ω+1\omega+1-strategically closed if Player II has a winning strategy for the game in which Player I and II alternatively choose pnp^{n} and pn+1p^{n+1} such that p0p1pnpn+1p_{0}\geq p_{1}\geq\cdots\geq p_{n}\geq p_{n+1}\geq\cdots. Player II wins if npn0\prod_{n}p_{n}\not=0. The following lemma is essentially due to Galvin–Jech–Magidor [10].

Lemma 3.14.

If II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda such that 𝒫(𝒫κλ))/I\mathcal{P}(\mathcal{P}_{\kappa}\lambda))/I is ω+1\omega+1-strategically closed then Φκ,λ,I\Phi_{\kappa,\lambda,I} holds

Proof.

Let ws\mathrm{ws}^{\prime} be II’s winning strategy witnessing ω+1\omega+1-strategically closedness. We let to describe II’s winning strategy ws\mathrm{ws} for (A,I)\Game(A,I) for all AI+A\in I^{+}.

When Player I chooses Fi:Aω1F_{i}:A\to\omega_{1}, since II is ω1\omega_{1}-complete, we define AiI+A_{i}\in I^{+} and ξi\xi_{i} such that Fi1{ξi}j<iAjI+F^{-1}_{i}\{\xi_{i}\}\cap\bigcap_{j<i}A_{j}\in I^{+}. Let Ai=ws(Fk1{ξk}j<kAjki)A_{i}=\mathrm{ws}^{\prime}(F^{-1}_{k}\{\xi_{k}\}\cap\bigcap_{j<k}A_{j}\mid k\leq i). Define ws(F0,,Fi)=ξi\mathrm{ws}(F_{0},...,F_{i})=\xi_{i}.

Then Player II wins. In fact, iAiF1(supiξi)\bigcap_{i}A_{i}\subseteq F^{-1}(\sup_{i}\xi_{i}) is in I+I^{+} since AiA_{i} is taken by the II’s strategy ws\mathrm{ws}^{\prime}, as desired. ∎

Galvin–Jech–Magidor proved that 2\aleph_{2} carries an ideal II in which 𝒫(2)/I\mathcal{P}(\aleph_{2})/I is ω+1\omega+1-strategically closed if a measurable cardinal is collapsed to 2\aleph_{2}.

4. Proof of Theorems 1.6 and 1.7

The first half of this section is devoted to Theorem 1.6. The rest is about Theorem 1.7.

Lemma 4.1.

Suppose that a semistationary subset S[λ]ωS\subseteq[\lambda]^{\omega} does not reflect to any R𝒫κλR\in\mathcal{P}_{\kappa}\lambda. Then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) destroys the semistationarity of SS.

Proof.

First, we fix a bijection φ:λ<ωλ\varphi:{{}^{<\omega}}\lambda\to\lambda. For each a𝒫κλa\in\mathcal{P}_{\kappa}\lambda, by the assumption, there is a function Fa:[a]<ωaF_{a}:[a]^{<\omega}\to a such that there is no xS𝐜𝐥[a]<ωx\in{S}^{\mathbf{cl}}\cap[a]^{<\omega} which closed under FaF_{a}.

Let g˙\dot{g} be a Nm(κ,λ){\mathrm{Nm}(\kappa,\lambda)}-name for {tr(p)pG˙}\bigcup\{\mathrm{tr}(p)\mid p\in\dot{G}\}. Let g˙i\dot{g}_{i} be the ii-th element of g˙\dot{g}. Again, we list the properties of g˙i\dot{g}_{i}.

  1. (1)

    g˙=λ\Vdash\bigcup\dot{g}=\lambda

  2. (2)

    ptr(p)g˙p\Vdash\mathrm{tr}(p)\sqsubseteq\dot{g}.

  3. (3)

    g˙ig˙i+1\dot{g}_{i}\subseteq\dot{g}_{i+1}.

We may assume that ω1g˙0\Vdash\omega_{1}\subseteq\dot{g}_{0} by shrinking Sucp()\mathrm{Suc}_{p}(\emptyset).

Define a Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda)-name for a function F˙:[λ]<ωλ\dot{F}:[\lambda]^{<\omega}\to\lambda by

F˙(ξ0,,ξn)={φ(Fg˙0(ξ0,,ξi),,Fg˙j(ξ0,,ξi))n=2i3j0 o.w. \dot{F}(\xi_{0},...,\xi_{n})=\begin{cases}\varphi(F_{\dot{g}_{0}}(\xi_{0},...,\xi_{i}),...,F_{\dot{g}_{j}}(\xi_{0},...,\xi_{i}))&n=2^{i}3^{j}\\ 0&\text{ o.w. }\end{cases}.

Here, Fg˙k(ξ0,,ξi)F_{\dot{g}_{k}}(\xi_{0},...,\xi_{i}) denotes Fg˙k({ξ0,,ξi}g˙k)F_{\dot{g}_{k}}(\{\xi_{0},...,\xi_{i}\}\cap\dot{g}_{k}) for each kjk\leq j.

Suppose that F˙\Vdash\dot{F} closed under some x˙S𝐜𝐥\dot{x}\in{S}^{\mathbf{cl}}. By Lemma 2.5, we may assume that x˙\dot{x} can decode φ\varphi. That is, if φ(α0,,αn)x˙\Vdash\varphi(\alpha_{0},...,\alpha_{n})\in\dot{x} then αix˙\Vdash\alpha_{i}\in\dot{x}.

By the choice of x˙\dot{x}, we can choose pTp\in T and ySy\in{S} such that pyω1x˙p\Vdash y\sqsubseteq_{\omega_{1}}\dot{x}. We may assume that yay\subseteq a for some atr(p)a\in\mathrm{tr}(p). Then p|x˙a|=ωp\Vdash|\dot{x}\cap a|=\omega. We claim that x˙\dot{x} closed under FaF_{a}. For every ξ0,,ξnx˙\xi_{0},...,\xi_{n}\in\dot{x}, we can choose an expansion ζ0,,ζ2n3in1x(ξn+1)\zeta_{0},...,\zeta_{2^{n}3^{i}-n-1}\in x\setminus(\xi_{n}+1) by p|x˙a|=ωp\Vdash|\dot{x}\cap a|=\omega. pp forces

F˙(ξ0,,ξn,ζ0,,ζ2n3in1)=φ(Fa0(ξ0,,ξn),,Fai(ξ0,,ξn))x˙\dot{F}(\xi_{0},...,\xi_{n},\zeta_{0},...,\zeta_{2^{n}3^{i}-n-1})=\varphi(F_{a_{0}}(\xi_{0},...,\xi_{n}),...,F_{a_{i}}(\xi_{0},...,\xi_{n}))\in\dot{x}.

And thus, Fi(ξ0,,ξn)xF_{i}(\xi_{0},...,\xi_{n})\in x. Therefore it is forced by pp that x˙S𝐜𝐥[ai]ω\dot{x}\in{S}^{\mathbf{cl}}\cap[a_{i}]^{\omega} closed under FaiF_{a_{i}}. Let Hull(y,Fa)\mathrm{Hull}(y,F_{a}) denote the least set that contains yy as a subset and closed under FaF_{a}. We have

pHull(y,Fa)x˙ap\Vdash\mathrm{Hull}(y,F_{a})\subseteq\dot{x}\cap a.

Since pyω1=x˙ω1p\Vdash y\cap\omega_{1}=\dot{x}\cap\omega_{1}, we also have

pHull(y,Fa)ω1=yω1p\Vdash\mathrm{Hull}(y,F_{a})\cap\omega_{1}=y\cap\omega_{1}.

In particular, Hull(y,Fa)S𝐜𝐥[a]ω\mathrm{Hull}(y,F_{a})\in{S}^{\mathbf{cl}}\cap[a]^{\omega} closed under FaF_{a}. This is contradicting to the choice of FaF_{a}. So, the semistationarity of SS is destroyed by pp, as desired. ∎

Lemma 4.2.

If a semistationary subset S[λ]ωS\subseteq[\lambda]^{\omega} reflects to some R𝒫κλR\in\mathcal{P}_{\kappa}\lambda then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) preserves the semistationarity of SS.

Proof.

By Lemma 3.8, Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) preserves the semistationarity of S[R]ωS\cap[R]^{\omega}. By Lemma 2.6, the semistationarity of SS is also preserved. ∎

Proof of Theorem 1.6.

The forward direction follows from Lemma 4.1. The inverse direction follows from 4.2. ∎

We conclude this section with Theorem 1.7.

Proof of Theorem 1.7.

It is easy to see that Refl(Eωλ,<μ+)\mathrm{Refl}(E_{\omega}^{\lambda},{<}\mu^{+}) fails in WW. Indeed, EμλE_{\mu}^{\lambda} does not reflect to any ordinals in E<μ+λE_{<\mu^{+}}^{\lambda}. Since (Eμλ)V(Eωλ)W(E_{\mu}^{\lambda})^{V}\subseteq(E_{\omega}^{\lambda})^{W} is a stationary subset, this witnesses ¬Refl(Eωλ,<κ)\lnot\mathrm{Refl}(E_{\omega}^{\lambda},{<}\kappa) in WW. By Theorem 2.9, SSR([λ]ω,<μ+)\mathrm{SSR}([\lambda]^{\omega},{<}\mu^{+}) also fails in WW. By Lemma 3.10, Nm(μ+,λ)\mathrm{Nm}(\mu^{+},\lambda) is not semiproper, as desired. In particular, by Lemma 3.11, Nm(μ+)\mathrm{Nm}(\mu^{+}) is non semiproper in WW. For Prikry-type forcings, Lemma 2.3 works. ∎

5. Semiproperness of saturated ideals

In this section, we discuss about ideals. First, we note that successors of regular cardinals can carry an ideal that is both saturated and semiproper by Proposition 5.1. The rest is devoted to Theorems 1.3 and 1.4. We also study their corollaries.

Proposition 5.1.

If j:VMj:V\to M is an almost huge embedding with critical point κ\kappa. For regular cardinals 1μ<κλ<j(κ)\aleph_{1}\leq\mu<\kappa\leq\lambda<j(\kappa), there is a poset which forces the following:

  1. (1)

    𝒫κλ\mathcal{P}_{\kappa}\lambda carries a saturated and proper* ideal.

  2. (2)

    μ+=κ\mu^{+}=\kappa and λ=j(κ)\lambda=j(\kappa).

Proof.

Let P=P(μ,κ)Coll˙(λ,<j(κ))P=P(\mu,\kappa)\ast\dot{\mathrm{Coll}}(\lambda,<j(\kappa)). Here, P(μ,κ)P(\mu,\kappa) is the μ\mu-support diagonal product of Levy collapses α[μ,κ)Reg<μColl(μ,<κ)\prod_{\alpha\in[\mu,\kappa)\cap\mathrm{Reg}}^{<\mu}\mathrm{Coll}(\mu,<\kappa). For a detail, we refer to [27, Theorem 1.2]. This paper gave a continuous projection π:P(μ,j(κ))P(μ,κ)Coll˙(λ,<j(κ))\pi:P(\mu,j(\kappa))\to P(\mu,\kappa)\ast\dot{\mathrm{Coll}}(\lambda,<j(\kappa)) and a P(μ,κ)Coll˙(λ,<j(κ))P(\mu,\kappa)\ast\dot{\mathrm{Coll}}(\lambda,<j(\kappa))-name I˙\dot{I} such that P(μ,κ)Coll˙(λ,<j(κ))P(\mu,\kappa)\ast\dot{\mathrm{Coll}}(\lambda,<j(\kappa)) forces the following properties:

  1. (1)

    I˙\dot{I} is a saturated ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda.

  2. (2)

    𝒫(𝒫κλ)/I˙P(μ,j(κ))/G˙H˙\mathcal{P}(\mathcal{P}_{\kappa}\lambda)/\dot{I}\simeq P(\mu,j(\kappa))/\dot{G}\ast\dot{H}.

It is easy to see that P(μ,j(κ))P(\mu,j(\kappa)) is μ\mu-closed. By Lemma 2.15, P(μ,j(κ))/G˙H˙P(\mu,j(\kappa))/\dot{G}\ast\dot{H} is forced to be μ\mu-closed. By (2), 𝒫(𝒫κλ)/I˙\mathcal{P}(\mathcal{P}_{\kappa}\lambda)/\dot{I} has a μ\mu-closed dense subset. Since μ1\mu\geq\aleph_{1}, 𝒫(𝒫κλ)/I˙\mathcal{P}(\mathcal{P}_{\kappa}\lambda)/\dot{I} is proper*, as desired. ∎

A conclusion of Proposition 5.1 fails if μ\mu is singular as follows.

Theorem 5.2 (Matsubara–Shelah [18]).

If μ\mu is a singular cardinal then 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda cannot carries a proper* ideal.

Proof.

Matsubara–Shelah proved non-properness* for a more loud class of ideals, that contains all ideals over successors of singular cardinals. We refer to [18].

Here, to see a common point between saturated ideals and Namba forcings, we show non-properness* assuming the cofinality of μ\mu is countable and the ideal is saturated. Let II be a saturated ideal over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda. By (1) of Lemma 2.13, 𝒫(𝒫μ+λ)/I\mathcal{P}(\mathcal{P}_{\mu^{+}}\lambda)/I forces cf˙(λ)=ω\mathrm{\dot{cf}}(\lambda)=\omega. Therefore this is not proper* by the same reason of non-properness* of Namba forcing. ∎

A model in which 𝒫ω+1λ\mathcal{P}_{\aleph_{\omega+1}}\lambda carries a semiproper ideal II such that 𝒫(𝒫ω+1λ)/I\mathcal{P}(\mathcal{P}_{\aleph_{\omega+1}}\lambda)/I forces cf(λ)=ω\mathrm{cf}(\lambda)=\omega was given by Sakai [22]. Note that II is not saturated in Sakai’s model.

We study the semiproperness of ideals in terms of the semistationary reflection principle. The following lemma is an analog of Theorem 1.7 for precipitous ideals.

Lemma 5.3.

Suppose that II is a normal, fine, exactly and uniformly μ+\mu^{+}-complete precipitous ideal over Z𝒫(X)Z\subseteq\mathcal{P}(X). If |x|μ+|x|\leq\mu^{+} for all xZx\in Z then the following are equivalent:

  1. (1)

    SSR([μ+]ω,<μ+)\mathrm{SSR}([\mu^{+}]^{\omega},{<}\mu^{+}).

  2. (2)

    𝒫(Z)/I\mathcal{P}(Z)/I preserves all semistationary subset of [μ+]ω[\mu^{+}]^{\omega}.

Proof.

Let λ=|X|\lambda=|X|. We may assume X=λX=\lambda. First, we need the following claim:

Claim 5.4.

There is a CIC\in I^{\ast} such that |xμ+|<μ+|x\cap\mu^{+}|<\mu^{+} for all xCx\in C.

Proof of Claim.

Let GG be an arbitrary (V,𝒫(Z)/I)(V,\mathcal{P}(Z)/I)-generic and j:VMV[G]j:V\to M\subseteq V[G] be the generic ultrapower mapping by GG. Then the following holds:

  • j``λ=[id]Mj``\lambda=[\mathrm{id}]\in M and Mμ+V[G]M{{}^{\mu^{+}}}M\cap V[G]\subseteq M.

  • j(μ+)λj(\mu^{+})\geq\lambda.

The first item is well-known. The second item follows from the assumption.

Therefore, we have M|μ+|=|j``λj(μ+)|<j(μ+)M\models|\mu^{+}|=|j``\lambda\cap j(\mu^{+})|<j(\mu^{+}). By Łós’s theorem, {xZ|xμ+|<μ+}G\{x\in Z\mid|x\cap\mu^{+}|<\mu^{+}\}\in G. Since GG is arbitrary, {xZ|xμ+|<μ+}I\{x\in Z\mid|x\cap\mu^{+}|<\mu^{+}\}\in I^{\ast}, as desired. ∎

First, we show the forward direction. Let SS be a semistationary subset of [μ+]ω[\mu^{+}]^{\omega}. By Lemma 2.8, D={a𝒫μ+μ+S[a]ωD=\{a\in\mathcal{P}_{\mu^{+}}\mu^{+}\mid S\cap[a]^{\omega} is semistationary}\} is co-bounded. By the claim and Lemma 2.5, A={xCxμ+D}IA=\{x\in C\mid x\cap\mu^{+}\in D\}\in I^{\ast}. It is easy to see that A={xZS[xμ+]ωA=\{x\in Z\mid S\cap[x\cap\mu^{+}]^{\omega} is semistationary}\}. By Łós’s theorem, AA forces that (j(S)[j``λj(μ+)]ω(j(S)\cap[j``\lambda\cap j(\mu^{+})]^{\omega} is semistationary)M)^{M}. Since MM is forced to be closed under μ+\mu^{+}-sequence, S=j(S)[j``λj(μ+)]ωS=j(S)\cap[j``\lambda\cap j(\mu^{+})]^{\omega} is semistationary, as desired.

Let SS be semistationary subset of [μ+]ω[\mu^{+}]^{\omega}. By the previous argument, we have A={xZ|xμ+|<μ+A=\{x\in Z\mid|x\cap\mu^{+}|<\mu^{+} and S[xμ+]ωS\cap[x\cap\mu^{+}]^{\omega} is semistationary}I\}\in I^{\ast}. Since II is fine and ω1\omega_{1}-complete, we may assume that ω1xμ+\omega_{1}\subseteq x\cap\mu^{+} for all xAx\in A. Therefore SS reflect to xμ+x\cap\mu^{+} for all xAx\in A, as desired.∎

From this, let us show Theorems 1.3 and 1.4 .

Proof of Theorem 1.3.

By Lemma 5.3, the same proof of Theorem 1.7 works as well. ∎

Therefore there is no semiproper saturated ideals over 𝒫μ+λ\mathcal{P}_{\mu^{+}}\lambda in the extension by Prikry forcing or Woodin’s modification over μ\mu. Moreover, there is no semiproper precipitous ideals222It is unknown whether every semiproper ideal is precipitous or not. But, under the GCH, the semiproperness shows the precipitousness, and thus, we can omit the condition of precipitousness from Theorem 1.3. For detail, we refer to [22].

Proof of Theorem 1.4.

Since λ\lambda is a regular cardinal, S=Eλλ+S=E_{\lambda}^{\lambda^{+}} is a non-reflecting stationary subset. That is, for every α<λ+\alpha<\lambda^{+}, SαS\cap\alpha is not stationary in α\alpha. Let GG be (V,𝒫([λ+]μ+))(V,\mathcal{P}([\lambda^{+}]^{\mu^{+}}))-generic and j:VMV[G]j:V\to M\subseteq V[G] be the generic ultrapower mapping induced by GG. By Lemmas 2.11 and 2.13, MM has the following properties:

  1. (1)

    MλV[G]M{{}^{\lambda}M}\cap V[G]\subseteq M.

  2. (2)

    λ+=j(μ+)\lambda^{+}=j(\mu^{+}).

  3. (3)

    cfV[G](λ)=cfM(λ)=ω\mathrm{cf}^{V[G]}(\lambda)=\mathrm{cf}^{M}(\lambda)=\omega.

Note that SS remains a stationary subset since II is λ+\lambda^{+}-saturated. (1) shows that SMS\in M and SS is non-reflection stationary subset in MM. By (3), SS is forced to be a subset of Eωλ+=(Eωλ+)M=j(Eωμ+)E_{\omega}^{\lambda^{+}}=(E_{\omega}^{\lambda^{+}})^{M}=j(E_{\omega}^{\mu^{+}}). Therefore, by (2),

MM\models there is a non-reflecting stationary subset of j(Eωμ+)j(E^{\mu^{+}}_{\omega}).

By the elementarity of jj, we have a non-reflecting stationary subset of Eωμ+E_{\omega}^{\mu^{+}} in the ground. In particular, SSR([μ+]ω,<μ+)\mathrm{SSR}([\mu^{+}]^{\omega},{<}\mu^{+}) fails by Theorem 2.9. By Lemma 5.3, II is not semiproper, as desired. ∎

We have a new mutual inconsistency.

Corollary 5.5.

If μ\mu is a singular cardinal with cofinality ω\omega then the following are mutually inconsistent:

  1. (1)

    μ+\mu^{+} carries a semiproper ideal.

  2. (2)

    [λ+]μ+[\lambda^{+}]^{\mu^{+}} carries a normal, fine, μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal for some regular λ\lambda.

Proof.

If (1) holds then SSR([μ+]ω,<μ+)\mathrm{SSR}([\mu^{+}]^{\omega},{<}\mu^{+}) holds. By the proof of Theorem 1.4, (2) cannot hold. ∎

It seems that the “huge-type” saturated ideal is an anti-compactness principle around singular cardinals in the following sense.

Corollary 5.6.

If [λ+]μ+[\lambda^{+}]^{\mu^{+}} carries a normal, fine, μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal for some regular λ\lambda and singular μ\mu of cofinality δ\delta then there is no supercompact cardinal between δ\delta and μ\mu.

Proof.

Suppose otherwise. Let κ\kappa be a supercompact such that δ<κ<μ\delta<\kappa<\mu. Let I¯\overline{I} be a normal, fine, μ+\mu^{+}-complete λ+\lambda^{+}-saturated ideal. If δ=κ\delta=\kappa then Coll(ω1,<κ)\mathrm{Coll}(\omega_{1},<\kappa) forces both ()(\dagger) and the ideal JJ generated by II is a λ+\lambda^{+}-saturated. The former follows from [5, Theorem 14]. The latter follows from Lemma 2.14

Note that a saturated ideal is ω1\omega_{1}-stationary preserving. Therefore the ideal is semiproper and λ+\lambda^{+}-saturated ideal over [λ+]μ+[\lambda^{+}]^{\mu^{+}} in the extension. This contradicts to Theorem 1.4.

We assume δ\delta is uncountable then δ<ω{{}^{<\omega}}\delta forces that κ\kappa is supercompact and the generated ideal KK by II is a λ+\lambda^{+}-saturated ideal. The latter also follows from Lemma 2.14. In the extension, [λ+]μ+[\lambda^{+}]^{\mu^{+}} carries a λ+\lambda^{+}-saturated ideal and cf˙(μ)=|δ|=ω\dot{\mathrm{cf}}(\mu)=|\delta|=\omega. This is impossible.∎

Lastly, we point out that the semiproperness of saturated ideals can characterized by the following form:

Proposition 5.7.

For a fine ideal II over 𝒫κλ\mathcal{P}_{\kappa}\lambda with the disjointing property, the following are equivalent:

  1. (1)

    II is semiproper.

  2. (2)

    Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper.

  3. (3)

    Φκ,λ,I\Phi_{\kappa,\lambda,I} holds.

Proof.

We only check (3) \to (1). By the disjointing property of II, for every countable MθM\prec\mathcal{H}_{\theta} with IMI\in M 𝒫(𝒫κλ)\mathcal{P}(\mathcal{P}_{\kappa}\lambda)-name α˙M\dot{\alpha}\in M for a countable ordinals, and AI+A\in I^{+}, there is an F:𝒫κλω1F:\mathcal{P}_{\kappa}\lambda\to\omega_{1} and BI+B\in I^{+} such that B[F]G˙=α˙B\Vdash[F]_{\dot{G}}=\dot{\alpha}. So, if MM has Player II’s winning strategy of Galvin game then this strategy shows C=FMF1Mω1AC=\bigcap_{F\in M}F^{-1}M\cap\omega_{1}\subseteq A is II-positive. CC is a (M,𝒫(𝒫κλ))(M,\mathcal{P}(\mathcal{P}_{\kappa}\lambda))-semigeneric condition. ∎

6. ()(\dagger)-aspects of Strong compactness

In this section, we introduce the ()(\dagger)-aspects of strong compactness. We begin with Proposition 6.1.

Proposition 6.1.

The following are equivalent

  1. (1)

    ()(\dagger).

  2. (2)

    SSR([λ]ω,<2)\mathrm{SSR}([\lambda]^{\omega},{<}\aleph_{2}) for all λ2\lambda\geq\aleph_{2}.

  3. (3)

    Nm(2,λ)\mathrm{Nm}(\aleph_{2},\lambda) is semiproper for all λ2\lambda\geq\aleph_{2}.

Proof.

This follows from Theorems 1.5 and 1.6. ∎

Proposition 6.2.

If κ\kappa is λ\lambda-strongly compact then Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper.

Proof.

Note that 𝒫κλ\mathcal{P}_{\kappa}\lambda carries a fine ultrafilter UU. Then 𝒫(𝒫κλ)/U\mathcal{P}(\mathcal{P}_{\kappa}\lambda)/U is ω+1\omega+1-strategically closed by obvious reason. By Lemma 3.14, Φκ,λ,U\Phi_{\kappa,\lambda,U^{\ast}} holds. In particular, Φκ,λ\Phi_{\kappa,\lambda} holds too. ∎

Recall that our definition of κ\kappa is λ\lambda-strongly compact is that 𝒫κλ\mathcal{P}_{\kappa}\lambda carries a fine ultrafilter. On the other hand, the original definition of strong compactness is due to Tarski. He introduced a logic κκ\mathcal{L}_{\kappa\kappa} that admits disjunctions of <κ{<}\kappa-many formulas and <κ{<}\kappa-many quantifiers. κ\kappa is strongly compact if, for every κκ\mathcal{L}_{\kappa\kappa}-theory TT, if TT is <κ{<}\kappa-consistent then TT is consistent. Here, by <κ{<}\kappa-consistent, we mean that SS is consistent for all S[T]<κS\in[T]^{<\kappa}. For a detail, we refer to [14, Section 4].

There is another formulation of strong compactness. We say that κ\kappa is λ\lambda-compact333This phrase is due to Gitik. if every κ\kappa-complete filter over λ\lambda can be extended to κ\kappa-complete ultrafilter. It was shown that κ\kappa is strongly compact if κ\kappa is λ\lambda-compact for all λκ\lambda\geq\kappa.

Hayut studied these compactnesses.

Theorem 6.3 (Hayut [13]).

If λ=λ<κ\lambda={\lambda^{<\kappa}} then the following are equivalent.

  1. (1)

    κ\kappa is λ\lambda-compact.

  2. (2)

    For every κκ\mathcal{L}_{\kappa\kappa}-theory TT with 2λ2^{\lambda}-many symbols, if TT is <κ{<}\kappa-consistent then TT is consistent.

(2) is called κ,κ\mathcal{L}_{\kappa,\kappa}-compactness for languages of size λ\lambda. Then, Hayut claimed that λ\lambda-compactness affects to cardinals up to 2λ2^{\lambda}, contrary to λ\lambda-strong compactness only effects for λ\lambda. In [13], he also drew a picture like

2λ2^{\lambda}-strong compactness \Rightarrow λ\lambda-compact
\Leftrightarrow κ,κ\mathcal{L}_{\kappa,\kappa}-compactness for language of size 2λ2^{\lambda} \Rightarrow λ\lambda-strong compactness.

We see a similar phenomenon in the view of semistationary reflection principles and semiproperness of Namba forcings.

Proposition 6.4.

If αω<κ\alpha^{\omega}<\kappa for all α<κ\alpha<\kappa then the following are equivalent.

  1. (1)

    SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa) for all λκ\lambda\geq\kappa.

  2. (2)

    Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper for all λκ\lambda\geq\kappa.

Proof.

(2) to (1) follows from Theorem 1.6. (1) to (2) follows from Lemma 3.8. Indeed, for every semistationary subset S[δ]ωS\subseteq[\delta]^{\omega} for δ\delta, let us check the semistationarity of SS is preserved by Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda).

If δ<κ\delta<\kappa then Lemma 3.8 works.

If δκ\delta\geq\kappa then SS reflect to some a𝒫κδa\in\mathcal{P}_{\kappa}\delta. By Lemma 3.8, the semistationarity of S[a]ωS\cap[a]^{\omega} is preserved by Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) for all λ\lambda. By Lemma 2.6, that of SS is also preserved. ∎

We let to denote ()κ(\dagger)^{\kappa} for convenience by (1) or (2) above holds. ()2(\dagger)^{\aleph_{2}} is equivalent with ()(\dagger). Of course, ()κ(\dagger)^{\kappa} is equivalent to (1) and (2), respectively.

Proposition 6.5.

If κ\kappa is strongly compact. For an uncountable regular cardinal μ<κ\mu<\kappa, Coll(μ,<κ)\mathrm{Coll}(\mu,<\kappa) forces ()κ(\dagger)^{\kappa}.

Proof.

The standard generic elementary embedding argument shows SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa) for all λκ\lambda\geq\kappa. ∎

The following lemma is essentially due to Todorčević [25].

Lemma 6.6.

If II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda then the following are equivalent:

  1. (1)

    Φκ,λ,I\Phi_{\kappa,\lambda,I} holds. This is equivalent with Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper.

  2. (2)

    WIA={M[ω1ω1A]ω{f1Mω1f:Aω1fM}I}W_{I}^{A}=\{M\in[\omega_{1}\cup{{}^{A}}\omega_{1}]^{\omega}\mid\bigcap\{f^{-1}M\cap\omega_{1}\mid f:A\to\omega_{1}\land f\in M\}\in I\} is non-stationary for all AIA\in I.

Proof.

The forward direction is clear. Let us check the inverse direction. We define a winning strategy of Player II for the Galvin game (I,A)\Game(I,A). Let θ\theta be a sufficiently large regular cardinal. Let CθC\subseteq\mathcal{H}_{\theta} be a club subset such that {M(ω1)𝒫κλMC}WIA=\{M\cap({\omega_{1}\cup{{}^{\mathcal{P}_{\kappa}\lambda}}})\mid M\in C\}\cap W_{I}^{A}=\emptyset. Consider an expansion 𝒜=θ,,A,C,κ,λ,I,\mathcal{A}=\langle\mathcal{H}_{\theta},\in,A,C,\kappa,\lambda,I,...\rangle. For given Player I’s move F0,,Fn\langle F_{0},...,F_{n}\rangle, Let ws(F0,,Fn)=Sk𝒜({F0,,Fn})ω1\mathrm{ws}(F_{0},...,F_{n})=\mathrm{Sk}_{\mathcal{A}}(\{F_{0},...,F_{n}\})\cap\omega_{1}. Sk𝒜\mathrm{Sk}_{\mathcal{A}} denotes the Skolem hull. It is easy to see that ws\mathrm{ws} is a winning strategy. ∎

Importance is WIAW_{I}^{A} is forced to be non-semiproper by Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) even if WIAW^{A}_{I} is stationary as follows.

Lemma 6.7.

Suppose that II is a fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda and AI+A\in I^{+}. Then Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) forces WIAW_{I}^{A} is non-semistationary.

Proof.

Let pAp_{A} be II-Namba tree such that

  • tr(pA)=\mathrm{tr}(p_{A})=\emptyset.

  • Suc(s)A\mathrm{Suc}(s)\subseteq A for all spAs\in p_{A}.

For every F:Aω1F:A\to\omega_{1}, let us define α˙Fn\dot{\alpha}_{F}^{n} by pAα˙Fn=F(g˙n)p_{A}\Vdash\dot{\alpha}_{F}^{n}=F(\dot{g}_{n}). Here, g˙n\dot{g}_{n} is the nn-th element of g˙\dot{g}. Note that pAp_{A} force g˙nA\dot{g}_{n}\in A by the choise of pAp_{A}. So α˙Fn\dot{\alpha}_{F}^{n} is well-defined.

Let θ\theta be a sufficiently large regular cardinal. We claim that if M(ω1ω1𝒫κλ)WIM\cap(\omega_{1}\cup{{}^{\mathcal{P}_{\kappa}\lambda}\omega_{1}})\in W_{I} then pAM[G˙]ω1Mω1p_{A}\Vdash M[\dot{G}]\cap\omega_{1}\not=M\cap\omega_{1} for any M[θ]ωM\in[\mathcal{H}_{\theta}]^{\omega}. Suppose otherwise, let us fix qpAq\leq p_{A} such that qq forces Mω1=M[G˙]ω1M\cap\omega_{1}=M[\dot{G}]\cap\omega_{1}. Putting s=tr(q)s=\mathrm{tr}(q), for every F:Aω1MF:A\to\omega_{1}\in M, we have

qF(g˙|s|)=α˙F|s|M[G˙]ω1=Mω1q\Vdash F(\dot{g}_{|s|})=\dot{\alpha}_{F}^{|s|}\in M[\dot{G}]\cap\omega_{1}=M\cap\omega_{1}.

So Suc(s){f1Mω1f:Aω1fM}I+\mathrm{Suc}(s)\subseteq\bigcap\{f^{-1}M\cap\omega_{1}\mid f:A\to\omega_{1}\land f\in M\}\in I^{+}. This is a contradiction.

Let 𝒜˙\dot{\mathcal{A}} be a Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I)-name for an expansion ˙θ,ˇθ,,κ,λ,A,pA,G˙,\langle\dot{\mathcal{H}}_{\theta},\check{\mathcal{H}}_{\theta},\in,\kappa,\lambda,A,p_{A},\dot{G},...\rangle. For a contradiction, we assume that there is a qpAq\leq p_{A} that forces NWIAN\in W_{I}^{A} and Nω1M˙𝒜˙N\sqsubseteq_{\omega_{1}}\dot{M}\prec\dot{\mathcal{A}}. By the claim above, we have qq forces

Nω1N[G˙]ω1M˙ω1=Nω1N\cap\omega_{1}\subsetneq N[\dot{G}]\cap\omega_{1}\subseteq\dot{M}\cap\omega_{1}=N\cap\omega_{1}.

This is a contradiction. Therefore WIAW_{I}^{A} is forced to be non-semistationary, as desired. ∎

Theorem 6.8.

If αω<κ\alpha^{\omega}<\kappa for all α<κ\alpha<\kappa then the following implication holds:

SSR([2λ<κ]ω,<κ)\mathrm{SSR}([2^{\lambda^{<\kappa}}]^{\omega},{<}\kappa) \to Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper for all fine ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda
\to Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper
\to SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa).
Proof.

The third implication follows from Theorem 1.6. The second implication is trivial.

Let us see the first implication. First, we note that SSR([2λ<κ]ω,<κ)\mathrm{SSR}([2^{{\lambda}^{<\kappa}}]^{\omega},{<}\kappa) implies that every semistationary subset of size 2λ<κ2^{\lambda^{<\kappa}} is preserved by Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) by Lemmas  2.6 and 3.8. By Lemma 6.7, WIAW_{I}^{A} needs to be non-semistationary for all AI+A\in I^{+}. By Lemma 6.6, Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is semiproper, as desired. ∎

Theorem 6.9.

Assume the same assumption of Theorem 6.8. The inverse direction of the first and third line of Theorem 6.8 cannot be proven in ZFC\mathrm{ZFC} modulo the existence of large cardinals.

Proof.

For the first line, let us give a model in which SSR([2]ω,<2)\mathrm{SSR}([\aleph_{2}]^{\omega},{<}{\aleph_{2}}) but Nm(2,2)\mathrm{Nm}(\aleph_{2},\aleph_{2}) is not semiproper. By the result of Baumgartner [1], Velic̆ković [29], and Sakai [23], it is known that SSR([2]ω,<2)\mathrm{SSR}([\aleph_{2}]^{\omega},{<}\aleph_{2}) is equiconsistent with the existence of a weakly compact cardinal.

On the other hand, If Nm(2,2)\mathrm{Nm}(\aleph_{2},\aleph_{2}) is semiproper then so Nm(2)\mathrm{Nm}(\aleph_{2}) is (Lemma 3.10). Shelah proved that the semiproperness of Nm(2)\mathrm{Nm}(\aleph_{2}) implies the Chang’s conjecture. So the consistency strength of the semiproperness of Nm(2)\mathrm{Nm}(\aleph_{2}) is stronger than that of an ω1\omega_{1}-Erdős cardinal.

Therefore, let us assume weakly compact cardinal exists in LL. In LL, by the above observations, we can get a model in which SSR([2]ω,<2)\mathrm{SSR}([\aleph_{2}]^{\omega},{<}\aleph_{2}) but Nm(2,2)\mathrm{Nm}(\aleph_{2},\aleph_{2}) is not semiproper, as desired.

For the third line, let us give a model in which Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper but SSR([2λ<κ]ω,<κ)\mathrm{SSR}([2^{\lambda^{<\kappa}}]^{\omega},{<}\kappa) fails assuming κ\kappa is a λ\lambda-strongly compact cardinal. We may assume 2λκ=λ+2^{\lambda^{\kappa}}=\lambda^{+}. Then λ<κ=λ\lambda^{<\kappa}=\lambda. Note that SSR([λ+]ω,<κ)\mathrm{SSR}([\lambda^{+}]^{\omega},{<}\kappa) implies ¬λ\lnot\square_{\lambda} by Theorem 2.9. Let PP be the standard poset, that adds λ\square_{\lambda}-sequence, of [2]. Note that PP does not changes (𝒫(𝒫κλ))<κ{{}^{<\kappa}}(\mathcal{P}(\mathcal{P}_{\kappa}\lambda)) and thus κ\kappa remains λ\lambda-strongly compact. By Proposition 6.2, the extension is a required model. ∎

As we have seen in Proposition 6.5, ()κ(\dagger)^{\kappa} holds if κ\kappa is strongly compact or κ\kappa is Levy collapsed to some successor cardinal. Note that another reflection principle holds from the effect of strong compactness. For example, one of them is Rado’s conjecture.

For a tree TT of height ω1\omega_{1}, we say that TT is a special if there is an F:TωF:T\to\omega such that F1{n}TF^{-1}\{n\}\subseteq T is an anti-chain for each n<ωn<\omega. RC(κ,λ)\mathrm{RC}(\kappa,\lambda) is the following statement: For every non-special tree TT of size λ\lambda has a non-special subtree of size <κ<\kappa.

Theorem 6.10.

RC(κ,λ)\mathrm{RC}(\kappa,\lambda) implies SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},{<}\kappa) holds.

Proof.

The same proof of [30, Theorem 5.2] works as well. ∎

It is also known that the inverse direction cannot be proven in ZFC\mathrm{ZFC}444For example, see [28, Theorem 1.7].. By the contents of this section, under the GCH\mathrm{GCH}, we have the following diagram of strong compactness of κ\kappa.

λκ(RC(κ,λ))\forall\lambda\geq\kappa(\mathrm{RC}(\kappa,\lambda)) \rightarrow ()κ(\dagger)^{\kappa}
\downarrow \swarrow         \searrow
\vdots \vdots \vdots
\searrow
RC(κ,λ+)\mathrm{RC}(\kappa,\lambda^{+}) \rightarrow SSR([λ+]ω,<κ)\mathrm{SSR}([\lambda^{+}]^{\omega},<{\kappa}) \leftarrow Nm(κ,λ+)\mathrm{Nm}(\kappa,\lambda^{+}) is semiproper
\searrow
RC(κ,λ)\mathrm{RC}(\kappa,\lambda) \rightarrow SSR([λ]ω,<κ)\mathrm{SSR}([\lambda]^{\omega},<{\kappa}) \leftarrow Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper
\searrow
\vdots \vdots \vdots
\searrow
RC(κ,κ)\mathrm{RC}(\kappa,\kappa) \rightarrow SSR([κ]ω,<κ)\mathrm{SSR}([\kappa]^{\omega},<{\kappa}) \leftarrow Nm(κ,κ)\mathrm{Nm}(\kappa,\kappa) is semiproper

All of the inverse directions of the arrows above cannot be proven in ZFC\mathrm{ZFC} modulo the existence of large cardinals. By this diagram, we can observe the strong compactness in more detail. We conclude this paper with the following question.

Question 6.11.

Is it consistent that Nm(κ,λ)\mathrm{Nm}(\kappa,\lambda) is semiproper but Nm(κ,λ,I)\mathrm{Nm}(\kappa,\lambda,I) is not semiproper for some fine ideal II?

References

  • [1] James E. Baumgartner. A new class of order types. Ann. Math. Logic, 9(3):187–222, 1976.
  • [2] James Cummings, Matthew Foreman, and Menachem Magidor. Squares, scales and stationary reflection. J. Math. Log., 1(1):35–98, 2001.
  • [3] Philipp Doebler and Ralf Schindler. Π2\Pi_{2} consequences of BMM+NSω1\mathrm{BMM}+\mathrm{NS}_{\omega_{1}} is precipitous and the semiproperness of stationary set preserving forcings. Math. Res. Lett., 16(5):797–815, 2009.
  • [4] Todd Eisworth. Successors of singular cardinals. In Handbook of set theory. Vols. 1, 2, 3, pages 1229–1350. Springer, Dordrecht, 2010.
  • [5] M. Foreman, M. Magidor, and S. Shelah. Martin’s maximum, saturated ideals, and nonregular ultrafilters. I. Ann. of Math. (2), 127(1):1–47, 1988.
  • [6] Matthew Foreman. More saturated ideals. In Cabal seminar 79–81, volume 1019 of Lecture Notes in Math., pages 1–27. Springer, Berlin, 1983.
  • [7] Matthew Foreman. Ideals and generic elementary embeddings. In Handbook of set theory. Vols. 1, 2, 3, pages 885–1147. Springer, Dordrecht, 2010.
  • [8] Matthew Foreman. Calculating quotient algebras of generic embeddings. Israel J. Math., 193(1):309–341, 2013.
  • [9] Matthew Foreman and Richard Laver. Some downwards transfer properties for 2\aleph_{2}. Adv. in Math., 67(2):230–238, 1988.
  • [10] F. Galvin, T. Jech, and M. Magidor. An ideal game. J. Symbolic Logic, 43(2):284–292, 1978.
  • [11] Moti Gitik. The negation of the singular cardinal hypothesis from o(κ)=κ++o(\kappa)=\kappa^{++}. Ann. Pure Appl. Logic, 43(3):209–234, 1989.
  • [12] Gitik, M. Prikry-type forcings. In Handbook of set theory. Vols. 1, 2, 3, pages 1351–1447. Springer, Dordrecht, 2010.
  • [13] Yair Hayut. Partial strong compactness and squares. Fund. Math., 246(2):193–204, 2019.
  • [14] Akihiro Kanamori. The higher infinite. Springer Monographs in Mathematics. Springer-Verlag, Berlin, second edition, 2003. Large cardinals in set theory from their beginnings.
  • [15] Kenneth Kunen. Saturated ideals. J. Symbolic Logic, 43(1):65–76, 1978.
  • [16] Richard Laver. An (2,2,0)(\aleph_{2},\,\aleph_{2},\,\aleph_{0})-saturated ideal on ω1\omega_{1}. In Logic Colloquium ’80 (Prague, 1980), volume 108 of Stud. Logic Foundations Math., pages 173–180. North-Holland, Amsterdam-New York, 1982.
  • [17] Yo Matsubara. Stationary preserving ideals over 𝒫κλ\mathcal{P}_{\kappa}\lambda. J. Math. Soc. Japan, 55(3):827–835, 2003.
  • [18] Yo Matsubara and Saharon Shelah. Nowhere precipitousness of the non-stationary ideal over 𝒫κλ\mathcal{P}_{\kappa}\lambda. J. Math. Log., 2(1):81–89, 2002.
  • [19] Telis K. Menas. On strong compactness and supercompactness. Ann. Math. Logic, 7:327–359, 1974/75.
  • [20] Kanji Namba. Independence proof of (ω,ωα)(\omega,\,\omega_{\alpha})-distributive law in complete Boolean algebras. Comment. Math. Univ. St. Paul., 19:1–12, 1971.
  • [21] Prikry, K. L. Changing measurable into accessible cardinals. Dissertationes Math. (Rozprawy Mat.), 68:55, 1970.
  • [22] Hiroshi Sakai. Semiproper ideals. Fund. Math., 186(3):251–267, 2005.
  • [23] Hiroshi Sakai. Semistationary and stationary reflection. J. Symbolic Logic, 73(1):181–192, 2008.
  • [24] Saharon Shelah. Proper and improper forcing. Perspectives in Mathematical Logic. Springer-Verlag, Berlin, second edition, 1998.
  • [25] Stevo Todorčević. Conjectures of Rado and Chang and cardinal arithmetic. In Finite and infinite combinatorics in sets and logic (Banff, AB, 1991), volume 411 of NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., pages 385–398. Kluwer Acad. Publ., Dordrecht, 1993.
  • [26] Kenta Tsukuura. ON THE CENTEREDNESS OF SATURATED IDEALS. Tsukuba J. Math., 47(1):29–39, 2023.
  • [27] Kenta Tsukuura. The extent of saturation of induced ideals. submitted.
  • [28] Toshimichi Usuba. Reflection principles for ω2\omega_{2} and the semi-stationary reflection principle. J. Math. Soc. Japan, 68(3):1081–1098, 2016.
  • [29] Boban Velic̆ković. Forcing axioms and stationary sets. Adv. Math., 94(2):256–284, 1992.
  • [30] Jing Zhang. Rado’s conjecture and its Baire version. J. Math. Log., 20(1):1950015, 35, 2020.