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

\newaliascnt

lemmatheorem \aliascntresetthelemma \newaliascntpropositiontheorem \aliascntresettheproposition \newaliascntcorollarytheorem \aliascntresetthecorollary \newaliascntremarktheorem \aliascntresettheremark \newaliascntquestiontheorem \aliascntresetthequestion \newaliascntdefinitiontheorem \aliascntresetthedefinition \newaliascntexampletheorem \aliascntresettheexample

How strong is a Reinhardt set over extensions of CZF?

Abstract.

We investigate the lower bound of the consistency strength of 𝖢𝖹𝖥\mathsf{CZF} with Full Separation 𝖲𝖾𝗉\mathsf{Sep} and a Reinhardt set, a constructive analogue of Reinhardt cardinals. We show that 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set interprets 𝖹𝖥\mathsf{ZF^{-}} with a cofinal elementary embedding j:VVj:V\prec V. We also see that 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set interprets 𝖹𝖥\mathsf{ZF^{-}} with a model of 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}}, the Wholeness axiom for bounded formulas.

2010 Mathematics Subject Classification:
Primary 03E70; Secondary 03E55

1. Introduction

Large cardinals are one of the important topics of set theory: the linear hierarchy of large cardinals provides a scale to fathom the consistency strength of a given set-theoretic statement. Reinhardt introduced a quite strong notion of a large cardinal, now known as Reinhardt cardinal. Unfortunately, Kunen [15] showed that Reinhardt cardinals do not exist in 𝖹𝖥𝖢\mathsf{ZFC}. However, Kunen’s proof relied on the Axiom of Choice, so it remains the hope that Reinhardt cardinals are consistent if we dispose of the Axiom of Choice. Recently, there are attempts to study Reinhardt cardinals over 𝖹𝖥\mathsf{ZF} and find intrinsic evidence for choiceless large cardinals, cardinals that are incompatible with the Axiom of Choice.

We may ask choiceless large cardinals are actually consistent with subtheories of 𝖹𝖥𝖢\mathsf{ZFC}, and there are some positive answers for this question. For example, Schultzenberg [21] showed that 𝖹𝖥\mathsf{ZF} with an elementary embedding j:Vλ+2Vλ+2j:V_{\lambda+2}\prec V_{\lambda+2} is consistent if 𝖹𝖥𝖢\mathsf{ZFC} with I0I_{0} is. Matthews [17] proved that Reinhardt cardinal is compatible with 𝖹𝖥𝖢\mathsf{ZFC^{-}} if we assume the consistency of 𝖹𝖥𝖢\mathsf{ZFC} with I1I_{1}.

We may also ask the lower bound of the consistency strength of large cardinals over subtheories of 𝖹𝖥𝖢\mathsf{ZFC}, which is quite non-trivial if we remove axioms other than choice. For example, Schultzenberg proved that 𝖹𝖥\mathsf{ZF} with j:Vλ+2Vλ+2j:V_{\lambda+2}\prec V_{\lambda+2} is equiconsistent with 𝖹𝖥𝖢\mathsf{ZFC} with I0I_{0}. We do not have an actual bound for 𝖹𝖥𝖢\mathsf{ZFC^{-}} with j:VVj:V\prec V, but we can obtain the lower bound if we add the assumption VcritjVV_{\operatorname{crit}j}\in V that Matthews’ model satisfies: in that case, we can see that VλV_{\lambda} exists and it is a model of 𝖹𝖥𝖢\mathsf{ZFC} with the Wholeness axiom 𝖶𝖠\mathsf{WA}.

In this paper, we will take another ‘subtheory’ of 𝖹𝖥𝖢\mathsf{ZFC} that lacks the law of excluded middle, namely 𝖢𝖹𝖥\mathsf{CZF}. 𝖢𝖹𝖥\mathsf{CZF} is a weak theory in that its consistency strength is the same as that of Kripke-Platek set theory 𝖪𝖯\mathsf{KP} with Infinity. However, adding the law of excluded middle to 𝖢𝖹𝖥\mathsf{CZF} results in the full 𝖹𝖥\mathsf{ZF}. The aim of this paper is to measure the ‘lower bound’ of the consistency strength of 𝖢𝖹𝖥\mathsf{CZF} with Full Separation 𝖲𝖾𝗉\mathsf{Sep} and a Reinhardt set. Hence we have how hard to establish the consistency of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set. The main result of this paper is as follows, which is a consequence of Theorem 5.2 and Theorem 6.4.

Proposition \theproposition.

The theory 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} + ‘there is a Reinhardt set’ can interpret the following theory: 𝖹𝖥\mathsf{ZF^{-}} with the cofinal embedding j:VVj:V\prec V with a transitive set KK such that j(x)=xj(x)=x for all xKx\in K, Kj(K)K\in j(K) and Λ:=nωjn(K)\Lambda:=\bigcup_{n\in\omega}j^{n}(K) thinks it is a model of 𝖹𝖥\mathsf{ZF} with 𝖶𝖠0\mathsf{WA}_{0}, the Wholeness axiom for Δ0\Delta_{0}-formulas.

1.1. The structure of this paper

In Section 2 and 3, we will cover relevant preliminaries. We will review constructive set theory in Section 2 and large set axioms in Section 3, so the readers who are already familiar with these topics may skip them. In Section 4, we review and discuss Gambino’s Heyting-valued interpretation [9] over the double negation formal topology Ω\Omega, which is the main tool of the paper. It forces Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM}, the law of excluded middle for bounded formulas. By assuming the Full Separation, we may turn Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM} to the full law of excluded middle. However, forcing over Ω\Omega does not preserve every axiom of 𝖢𝖹𝖥\mathsf{CZF}: it does not preserve the Axiom of Subset Collection, a 𝖢𝖹𝖥\mathsf{CZF}-analogue of Power Set axiom. Thus the resulting theory of forcing over Ω\Omega is 𝖹𝖥\mathsf{ZF^{-}} if we start from 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep}. Unlike ‘small’ formal topologies, Ω\Omega is not absolute between transitive models, and it causes issues about the absoluteness of the Heyting-valued interpretation. We will also discuss it in this section. Section 5 is devoted to prove elementary embeddings are preserved under Ω\Omega. However, Ω\Omega does not prove the inaccessibility of a critical point of the elementary embedding, and it restricts the analysis on the consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with an elementary embedding. We will deal with this issue in Section 6 by showing that the Heyting-valued interpretation under Ω\Omega still proves the critical point enjoys a strong reflection principle, which makes the critical point a transitive model of 𝖹𝖥\mathsf{ZF} with large cardinal axioms. We briefly discuss why achieving the upper bound of the consistency strength of our object theories in Section 7, with some concluding questions.

2. Constructive set theory

In this section, we will briefly review 𝖹𝖥𝖢\mathsf{ZFC^{-}}, 𝖹𝖥𝖢\mathsf{ZFC} without power set and the constructive set theory. There are various formulations of constructive set theories, but we will focus on 𝖢𝖹𝖥\mathsf{CZF}.

2.1. 𝖹𝖥𝖢\mathsf{ZFC} without Power Set

We will frequently mention 𝖹𝖥𝖢\mathsf{ZFC} without Power Set, called 𝖹𝖥𝖢\mathsf{ZFC^{-}}. However, 𝖹𝖥𝖢\mathsf{ZFC^{-}} is not obtained by just dropping Power Set from 𝖹𝖥𝖢\mathsf{ZFC}:

Definition \thedefinition.

𝖹𝖥𝖢\mathsf{ZFC^{-}} is the theory obtained from 𝖹𝖥𝖢\mathsf{ZFC} as follows: it drops Power Set, uses Collection instead of Replacement, and the well-ordering principle instead of the usual statement of Choice. 𝖹𝖥\mathsf{ZF^{-}} is a system obtained from 𝖹𝖥𝖢\mathsf{ZFC^{-}} by dropping the well-ordering principle.

Note that using Collection instead of Replacement is necessary to avoid pathologies. See [10] for the details. It is also known by [8] that 𝖹𝖥𝖢\mathsf{ZFC^{-}} does not prove the reflection principle.

2.2. Axioms of 𝖢𝖹𝖥\mathsf{CZF}

Constructive Zermelo-Fraenkel set theory 𝖢𝖹𝖥\mathsf{CZF} is introduced by Aczel [2] with his type-theoretic interpretation of 𝖢𝖹𝖥\mathsf{CZF}. We will introduce subtheories called Basic Constructive Set Theory 𝖡𝖢𝖲𝖳\mathsf{BCST} and 𝖢𝖹𝖥\mathsf{CZF^{-}} before defining the full 𝖢𝖹𝖥\mathsf{CZF}.

Definition \thedefinition.

𝖡𝖢𝖲𝖳\mathsf{BCST} is the theory which consists of Extensionality, Pairnig, Union, Emptyset, Replacement, and Δ0\Delta_{0}-Separation. 𝖢𝖹𝖥\mathsf{CZF^{-}} is obtained by adding the following axioms to 𝖡𝖢𝖲𝖳\mathsf{BCST}: Infinity, \in-induction, and Strong Collection which states the following: if ϕ(x,y)\phi(x,y) is a formula such that xayϕ(x,y)\forall x\in a\exists y\phi(x,y) for given aa, then we can find bb such that

xabyϕ(x,y)ybxaϕ(x,y).\forall x\in a\exists\in by\phi(x,y)\land\forall y\in b\exists x\in a\phi(x,y).

We will also define some synonyms for frequently-mentioned axioms:

Definition \thedefinition.

We will use 𝖲𝖾𝗉\mathsf{Sep}, Δ𝟢-𝖲𝖾𝗉\mathsf{\Delta_{0}\text{-}Sep}, Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM} for denoting Full Separation (i.e., Separation for all formulas), Δ0\Delta_{0}-Separation and the Law of Excluded Middle for Δ0\Delta_{0}-formulas.

Full separation proves Strong Collection from Collection, but Δ0\Delta_{0}-Separation is too weak to do it. It is also known that Δ0\Delta_{0}-Separation is equivalent to the existence of the intersection of two sets. See Section 9.5 of [4] for its proof.

Proposition \theproposition.

Working over 𝖡𝖢𝖲𝖳\mathsf{BCST} without Δ0\Delta_{0}-Separation, Δ0\Delta_{0}-Separation is equivalent to the Axiom of Binary Intersection, which asserts that aba\cap b exists if aa and bb are sets. ∎

It is convenient to introduce the notion of multi-valued function to describe Strong Collection and Subset Collection that appears later. Let AA and BB be classes. A relation RA×BR\subseteq A\times B is a multi-valued function from AA to BB if domR=A\operatorname{dom}R=A. In this case, we write R:ABR:A\rightrightarrows B. We use the notation R:ABR:A\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>B if both R:ABR:A\rightrightarrows B and R:BAR:B\rightrightarrows A hold. Then we can rephrase Strong Collection as follows: for every set aa and a class R:aVR:a\rightrightarrows V, there is a set bb such that R:abR:a\rightrightarrows b.

Now we can state the Axiom of Subset Collection:

Definition \thedefinition.

The Axiom of Subset Collection states the following: let RuR_{u} be a class with a parameter uVu\in V. For each a,bVa,b\in V, we can find a set cVc\in V such that

Ru:abdc(Ru:ad).R_{u}:a\rightrightarrows b\implies\exists d\in c(R_{u}:a\rightrightarrows d).

𝖢𝖹𝖥\mathsf{CZF} is the theory by adding Subset Collection to 𝖢𝖹𝖥\mathsf{CZF^{-}}.

There is a simpler version of Subset Collection known as Fullness, which is a bit easier to understand.

Definition \thedefinition.

The Axiom of Fullness states the following: Let mv(a,b)\operatorname{mv}(a,b) the class of all multi-valued function from aa and bb. Then there is a subset cmv(a,b)c\subseteq\operatorname{mv}(a,b) such that if rmv(a,b)r\in\operatorname{mv}(a,b), then there is scs\in c such that srs\subseteq r. We call cc to be full in mv(a,b)\operatorname{mv}(a,b).

Then the following holds:

Proposition \theproposition.
  1. (1)

    (𝖢𝖹𝖥\mathsf{CZF^{-}}) Subset Collection is equivalent to Fullness.

  2. (2)

    (𝖢𝖹𝖥\mathsf{CZF^{-}}) Power Set implies Subset Collection.

  3. (3)

    (𝖢𝖹𝖥\mathsf{CZF^{-}}) Subset Collection proves the function set ba{{}^{a}}b exists for all aa and bb.

  4. (4)

    (𝖢𝖹𝖥\mathsf{CZF^{-}}) If Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM} holds, then Subset Collection implies Power Set.

We do not provide the proof for it, and the readers may consult with [5] or [4] for its proof. Note that Subset Collection does not increase the proof-theoretic strength of 𝖢𝖹𝖥\mathsf{CZF^{-}} while the Axiom of Power Set does. The following lemma is useful to establish (1) of subsection 2.2, and is also useful to treat multi-valued functions:

Lemma \thelemma.

Let R:ABR:A\rightrightarrows B be a multi-valued function. Define 𝒜(R):AA×B\mathcal{A}(R):A\rightrightarrows A\times B by

𝒜(R)={a,a,ba,bR},\mathcal{A}(R)=\{\langle a,\langle a,b\rangle\rangle\mid\langle a,b\rangle\in R\},

then the following holds:

  1. (1)

    𝒜(R):ASRS:AB\mathcal{A}(R):A\rightrightarrows S\iff R\cap S:A\rightrightarrows B,

  2. (2)

    𝒜(R):ASSR\mathcal{A}(R):A\leftleftarrows S\iff S\subseteq R.

Proof.

For the first statement, observe that 𝒜(R):AS\mathcal{A}(R):A\rightrightarrows S is equivalent to

aAsS:a,s𝒜(R).\forall a\in A\exists s\in S:\langle a,s\rangle\in\mathcal{A}(R).

By the definition of 𝒜\mathcal{A}, this is equivalent to

aAsS[bB:s=a,ba,bR].\forall a\in A\exists s\in S[\exists b\in B:s=\langle a,b\rangle\land\langle a,b\rangle\in R].

We can see that the above statement is equivalent to aAbB:a,bRS\forall a\in A\exists b\in B:\langle a,b\rangle\in R\cap S, which is the definition of RS:ABR\cap S:A\rightrightarrows B. For the second claim, observe that 𝒜(R):AS\mathcal{A}(R):A\leftleftarrows S is equivalent to

sSaA:a,s𝒜(R).\forall s\in S\exists a\in A:\langle a,s\rangle\in\mathcal{A}(R).

By rewriting 𝒜\mathcal{A} to its definition, we have

sSaA:[bB:s=a,bR].\forall s\in S\exists a\in A:[\exists b\in B:s=\langle a,b\rangle\in R].

We can see that it is equivalent to SRS\subseteq R. ∎

The following lemma provides useful applications of Strong Collection:

Lemma \thelemma.

If aAa\in A and R:aAR:a\rightrightarrows A, then there is a set bAb\in A such that bRb\subseteq R and b:aAb:a\rightrightarrows A.

Proof.

Consider 𝒜(R):aa×A\mathcal{A}(R):a\rightrightarrows a\times A. By the second-order Strong Collection over AA, there is bAb\in A such that 𝒜(R):Ab\mathcal{A}(R):A\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b. Hence by subsection 2.2, we have bRb\subseteq R and b:aAb:a\rightrightarrows A. ∎

2.3. Inductive definition

Various recursive construction on 𝖢𝖹𝖥\mathsf{CZF} is given by inductive definition. The readers might refer [5] or [4] to see general information about inductive definition, but we will review some of it for the readers who are not familiar with it.

Definition \thedefinition.

An inductive definition Φ\Phi is a class of pairs X,a\langle X,a\rangle. For an inductive definition Φ\Phi, associate ΓΦ(C)={aXCX,aΦ}\Gamma_{\Phi}(C)=\{a\mid\exists X\subseteq C\langle X,a\rangle\in\Phi\}. A class CC is Φ\Phi-closed if ΓΦ(C)C\Gamma_{\Phi}(C)\subseteq C.

We may think Φ\Phi as a generalization of a deductive system, and ΓΦ(C)\Gamma_{\Phi}(C) a class of theorems derivable from the class of axioms CC. Some authors use the notation XΦaX\vdash_{\Phi}a or X/aΦX/a\in\Phi instead of X,aΦ\langle X,a\rangle\in\Phi.

Each Inductive definitions arise the least class fixed point:

Theorem 2.1 (Class Inductive Definition Theorem).

Let Φ\Phi be an inductive definition. Then there is a smallest Φ\Phi-closed class I(Φ)I(\Phi).

The following lemma is the essential tool for the proof of Class Inductive Definition Theorem. See Lemma 12.1.2 of [4] for its proof:

Lemma \thelemma.

Every inductive definition Φ\Phi has a corresponding iteration class JJ, which satisfies Ja=Γ(xaJx)J^{a}=\Gamma\left(\bigcup_{x\in a}J^{x}\right) for all aa, where Ja={xa,xJ}J^{a}=\{x\mid\langle a,x\rangle\in J\}.

2.4. 𝖢𝖹𝖥\mathsf{CZF} versus 𝖨𝖹𝖥\mathsf{IZF}

There are two possible constructive formulations of 𝖹𝖥\mathsf{ZF}, namely 𝖨𝖹𝖥\mathsf{IZF} and 𝖢𝖹𝖥\mathsf{CZF}, although we will focus on the latter.

Definition \thedefinition.

𝖨𝖹𝖥\mathsf{IZF} is the theory that comprises the following axioms: Extensionality, Pairing, Union, Infinity, \in-induction, Separation, Collection, and Power Set.

It is known that every theorem of 𝖢𝖹𝖥\mathsf{CZF} is that of 𝖨𝖹𝖥\mathsf{IZF}. Moreover, 𝖨𝖹𝖥\mathsf{IZF} is quite strong in the sense that its proof-theoretic strength is the same as that of 𝖹𝖥\mathsf{ZF}. On the other hand, it is known that the proof-theoretic strength of 𝖢𝖹𝖥\mathsf{CZF} is equal to that of Kripke-Platek set theory 𝖪𝖯\mathsf{KP} with Infinity. 𝖨𝖹𝖥\mathsf{IZF} is deemed to be impredicative due to the presence of Full Separation and Power Set.111There is no consensus on the definition on the predicativity. The usual informal description of the predicativity is rejecting self-referencing definitions. On the other hand, 𝖢𝖹𝖥\mathsf{CZF} is viewed as predicative since it allows the type-theoretic interpretation given by Aczel [2]. However, adding the full law of excluded middle into 𝖨𝖹𝖥\mathsf{IZF} or 𝖢𝖹𝖥\mathsf{CZF} results in the same 𝖹𝖥\mathsf{ZF}.

3. Large set axioms

In this section, we will discuss large set axioms, which is an analogue of large cardinal axioms over 𝖢𝖹𝖥\mathsf{CZF}. Since ordinals over 𝖢𝖹𝖥\mathsf{CZF} could be badly behaved (for example, they need not be well-ordered), we focus on the structural properties of given sets to obtain higher infinities over 𝖢𝖹𝖥\mathsf{CZF}. We also compare the relation between large cardinal axioms over well-known theories like 𝖹𝖥\mathsf{ZF} and large set axioms.

3.1. Tiny and Small Large set axioms

The first large set notions over 𝖢𝖹𝖥\mathsf{CZF} would be regular sets. Regular sets appear first in Aczel’s paper [3] about inductive definitions over 𝖢𝖹𝖥\mathsf{CZF}. As we will see later, regular sets can ‘internalize’ most inductive constructions, which turns out to be useful in many practical cases.

Definition \thedefinition.

A transitive set AA is regular if it satisfies second-order Strong Collection:

aAR[R:aAbA(R:ab)].\forall a\in A\forall R[R:a\rightrightarrows A\to\exists b\in A(R:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b)].

A regular set AA is \bigcup-regular if aA\bigcup a\in A for all aAa\in A. A regular set AA is inaccessible if (A,)(A,\in) is a model of 𝖢𝖹𝖥\mathsf{CZF}, and furthermore, it also satisfies the second-order Subset Collection:

a,bAcAuAR(Ru:ab)dc(Ru:ad).\forall a,b\in A\exists c\in A\forall u\in A\forall R(R_{u}:a\rightrightarrows b)\to\exists d\in c(R_{u}:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>d).

The Regular Extension Axiom 𝖱𝖤𝖠\mathsf{REA} asserts that every set is contained in some regular set. The Inaccessible Extension Axiom 𝖨𝖤𝖠\mathsf{IEA} asserts that every set is contained in an inaccessible set.

There is no ‘pair-closed regular sets’ since every regular set is closed under pairings if it contains 2:

Lemma \thelemma.

If AA is regular and 2A2\in A, then a,bA\langle a,b\rangle\in A for all a,bAa,b\in A. ∎

𝖱𝖤𝖠\mathsf{REA} has various consequences: For example, 𝖢𝖹𝖥+𝖱𝖤𝖠\mathsf{CZF^{-}+REA} proves Subset Collection. Moreover, it also proves that every bounded inductive definition Φ\Phi has a set-sized fixed point I(Φ)I(\Phi).

The notion of regular sets is quite restrictive, as it does not have Separation axioms, at least for Δ0\Delta_{0}-formulas, so we have no way to do any internal construction over a regular set. The following notion is a strengthening of regular set, which resolves the issue of internal construction:

Definition \thedefinition.

A regular set AA is BCST-regular if A𝖡𝖢𝖲𝖳A\models\mathsf{BCST}. Equivalently, AA satisfies Union, Pairing, Empty set and Binary Intersection.

We do not know that 𝖢𝖹𝖥\mathsf{CZF} proves every regular set is BCST-regular, although Lubarsky and Rathjen [20] proved that the set of all hereditarily countable sets in the Feferman-Levy model is functionally regular but not \bigcup-regular. It is not even sure that the existence of a regular set implies that of BCST-regular set. However, every inaccessible set is BCST-regular, and every BCST-regular set appearing in this paper is inaccessible.

What are regular sets and inaccessible sets in the classical context? The following result illustrates how these sets look like under the well-known classical context:

Proposition \theproposition.
  1. (1)

    (𝖹𝖥\mathsf{ZF^{-}}) Every \bigcup-regular set containing 22 is a transitive model of second-order 𝖹𝖥\mathsf{ZF^{-}}, 𝖹𝖥𝟤\mathsf{ZF^{-}_{2}}.

  2. (2)

    (𝖹𝖥𝖢\mathsf{ZFC^{-}}) Every \bigcup-regular set containing 22 is of the form HκH_{\kappa} for some regular cardinal κ\kappa.

  3. (3)

    (𝖹𝖥\mathsf{ZF^{-}}) Every inaccessible set is of the form VκV_{\kappa} for some inaccessible κ\kappa.

Note that we follow the Hayut and Karagila’s definition [14] of inaccessiblity in choiceless context; that is, κ\kappa is inaccessible if Vκ𝖹𝖥𝟤V_{\kappa}\models\mathsf{ZF_{2}}.

Proof.
  1. (1)

    Let AA be a regular set containing 22. We know that AA satisfies Extensionality, \in-induction, Union and the second-order Collection. Hence it remains to show that the second-order Separation holds.

    Let XAX\subseteq A and aAa\in A. Fix cXac\in X\cap a. Now consider the function f:aAf:a\to A defined by

    f(x)={xif xX, andcotherwise.f(x)=\begin{cases}x&\text{if }x\in X,\text{ and}\\ c&\text{otherwise}.\end{cases}

    By the second-order Strong Collection over AA, we have bAb\in A such that f:abf:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b, and thus b=aXb=a\cap X.

  2. (2)

    Let AA be a regular set. Let κ\kappa be the least ordinal that is not a member of AA. Then κ\kappa must be a regular cardinal: if not, there is α<κ\alpha<\kappa and a cofinal map f:ακf:\alpha\to\kappa. By transitivity of AA and the definition of κ\kappa, we have αA\alpha\in A, so κA\kappa\in A by the second-order Replacement and Union, a contradiction.

    We can see that 𝖹𝖥𝖢\mathsf{ZFC^{-}} proves HκH_{\kappa} is a class model of 𝖹𝖥𝖢\mathsf{ZFC^{-}}, and AA satisfies the Well-ordering Principle. We can also show that AHκ={x:|TCx|<κ}A\subseteq H_{\kappa}=\{x:|\operatorname{TC}x|<\kappa\} holds: We know that AOrd=HκOrd=κA\cap\mathrm{Ord}=H_{\kappa}\cap\mathrm{Ord}=\kappa. By the second-order Separation over AA, 𝒫(Ord)A=𝒫(Ord)Hκ\mathcal{P}(\mathrm{Ord})\cap A=\mathcal{P}(\mathrm{Ord})\cap H_{\kappa}. Hence A=HκA=H_{\kappa}: for each xHκx\in H_{\kappa}, we can find θ<κ\theta<\kappa, Rθ×θR\subseteq\theta\times\theta and XθX\subseteq\theta such that (trclx,,x)(θ,R,X)(\operatorname{trcl}x,\in,x)\cong(\theta,R,X). (Here we treat xx as a unary relation.) Then (θ,R,X)A(\theta,R,X)\in A, so xAx\in A by Mostowski Collapsing Lemma.

  3. (3)

    If AA is inaccessible, then AA is closed under the true power set of its elements, since the second-order Subset Collection implies if a,bAa,b\in A then baA{{}^{a}}b\in A. Hence AA must be of the form VκV_{\kappa} for some κ\kappa. Moreover, κ\kappa is inaccessible because Vκ=A𝖹𝖥𝟤V_{\kappa}=A\models\mathsf{ZF^{-}_{2}}. ∎

Question \thequestion.

Is there a characterization of regular sets over 𝖹𝖥𝖢\mathsf{ZFC}? How about \bigcup-regular sets over 𝖹𝖥\mathsf{ZF^{-}}?

3.2. Large Large set axioms

There is no reason to stop defining large set notions up to weaker ones. Hence we define stronger large set axioms. The main tool to access strong large cardinals (up to measurable cardinals) is to use elementary embedding, so we follow the same strategy:

Definition \thedefinition (𝖢𝖹𝖥\mathsf{CZF^{-}}).

Working over the extended language \in, a unary functional symbol jj and a unary predicate symbol MM. We will extend 𝖢𝖹𝖥\mathsf{CZF^{-}} as follows: we allow jj and MM in Δ0\Delta_{0}-Separation and Strong Collection (also for Subset Collection if we start from 𝖢𝖹𝖥\mathsf{CZF}), and add the following schemes:

  1. (1)

    MM is transitive, xM(j(x))\forall xM(j(x)), and

  2. (2)

    x[ϕ(x)ϕM(j(x))]\forall\vec{x}[\phi(\vec{x})\leftrightarrow\phi^{M}(j(\vec{x}))] for every ϕ\phi which does not contain any jj or MM, where ϕM\phi^{M} is the relativization of ϕ\phi over MM.

If a transitive set KK satisfies xKj(x)=x\forall x\in Kj(x)=x and Kj(K)K\in j(K), then we call KK a critical point and we call KK a critical set if KK is also inaccessible. If M=VM=V and KK is transitive and inaccessible, then we call KK a Reinhardt set.

We will use the term critical point and critical set simultaneously, so the readers should distinguish the difference of these two terms. (For example, a critical point need not be a critical set unless it is inaccessible.) Note that the definition of critical sets is apparently stronger than that is suggested by Hayut and Karagila [14]. Finding the 𝖢𝖹𝖥\mathsf{CZF}-definition of a critical set which is classically equivalent to a critical set in the style of Hayut and Karagila would be a good future work. Also, Ziegler [24] uses the term ‘measurable sets’ to denote critical sets, but we will avoid this term for the following reasons: it does not reflect that the definition is given by an elementary embedding, and it could be confusing with measurable sets in measure theory.

We do not know every elementary embedding j:VMj:V\prec M over 𝖢𝖹𝖥\mathsf{CZF} enjoys cofinal properties. Surprisingly, the following lemma shows that jj become a cofinal map if M=VM=V, even under 𝖢𝖹𝖥\mathsf{CZF} without any assumptions. Note that the following lemma uses Subset Collection heavily. See Theorem 9.37 of [24] for its proof.

Lemma \thelemma (Ziegler [24], 𝖢𝖹𝖥\mathsf{CZF}).

Let j:VVj:V\prec V be a non-trivial elementary embedding. Then jj is cofinal, that is, we can find yy such that xj(y)x\in j(y) for each xx.

Note that Ziegler [24] uses the term set cofinality to denote our notion of cofinality. However, we will use the term cofinality to harmonize the terminology with that of Matthews [17].

4. Heyting-valued interpretation over the double negation formal topology

4.1. General Heyting-valued interpretation

We will follow Gambino’s definition [9] of Heyting-valued interpretation. We start this section by reviewing relevant facts on the Heyting-valued interpretation.

Forcing is a powerful tool to construct a model of set theory. Gambino’s definition of the Heyting-valued model (or alternatively, forcing) opens up a way to produce forcing models of 𝖢𝖹𝖥\mathsf{CZF^{-}}. His Heyting-valued model starts from formal topology, which formalizes a poset of open sets with a covering relation:

Definition \thedefinition.

A structure 𝒮=(S,,)\mathcal{S}=(S,\leq,\vartriangleleft) is formal topology is a poset (S,)(S,\leq) endowed with S×𝒫(S)\vartriangleleft\subseteq S\times\mathcal{P}(S) such that

  1. (1)

    if apa\in p, then apa\vartriangleleft p,

  2. (2)

    if aba\leq b and bpb\vartriangleleft p, then apa\vartriangleleft p,

  3. (3)

    if apa\vartriangleleft p and xp(xq)\forall x\in p(x\vartriangleleft q), then aqa\vartriangleleft q, and

  4. (4)

    if ap,qa\vartriangleleft p,q, then a(p)(q)a\vartriangleleft(\downarrow p)\cap(\downarrow q), where p={rSrp}\downarrow p=\{r\in S\mid r\leq p\}.

For each formal topology 𝒮\mathcal{S}, we have a notion of nucleus ȷp\jmath p given by

ȷp={xSxp}.\jmath p=\{x\in S\mid x\vartriangleleft p\}.

Then the class Low(𝒮)ȷ\operatorname{Low}(\mathcal{S})_{\jmath} of all lower subsets222A subset pSp\subseteq S is a lower set if p=p\downarrow p=p. that are stable under ȷ\jmath (i.e., ȷp=p\jmath p=p) form a set-generated frame:

Definition \thedefinition.

A structure 𝒜=(A,,,,,g)\mathcal{A}=(A,\leq,\bigvee,\land,\top,g) is a set-generated frame if (A,,,,)(A,\leq,\bigvee,\land,\top) is a complete distributive lattice with the generating set gAg\subseteq A, such that the class ga={xgxa}g_{a}=\{x\in g\mid x\leq a\} is a set, and a=gaa=\bigvee g_{a} for any aAa\in A.

Note that every set-generated frame has every operation of Heyting algebra: for example, we can define aba\to b by ab={xgxab}a\to b=\bigvee\{x\in g\mid x\land a\leq b\}, \bot by \varnothing, and p\bigwedge p by {xgyp(xy)}\bigvee\{x\in g\mid\forall y\in p(x\leq y)\}.

Proposition \theproposition.

For every formal topology 𝒮\mathcal{S}, the class Low(𝒮)ȷ\operatorname{Low}(\mathcal{S})_{\jmath} has a set-generated frame structure defined as follows: pq=pqp\land q=p\cap q, pq=ȷ(pq)p\lor q=\jmath(p\cup q), pq={xSxpxq}p\to q=\{x\in S\mid x\in p\to x\in q\}, p=ȷ(p)\bigvee p=\jmath(\bigcup p), p=p\bigwedge p=\bigcap p, =S\top=S, \leq as the inclusion relation, and g={{x}xS}g=\{\{x\}\mid x\in S\}.

We extend the nucleus ȷ\jmath to general classes by taking JP:={ȷppP}JP:=\bigcup\{\jmath p\mid p\subseteq P\}, and define operations on classes by PQ=PQP\land Q=P\cap Q, PQ=J(PQ)P\lor Q=J(P\cup Q) and PQ={xSxPxQ}P\to Q=\{x\in S\mid x\in P\to x\in Q\}. For a set-indexed collection of classes {PxxI}\{P_{x}\mid x\in I\}, take xIPx=xIPx\bigwedge_{x\in I}P_{x}=\bigcap_{x\in I}P_{x} and xIPx=J(xIPx)\bigvee_{x\in I}P_{x}=J\left(\bigcup_{x\in I}P_{x}\right). Note that JP=ȷPJP=\jmath P if PP is a set and we have the Full Separation.

The Heyting universe V𝒮V^{\mathcal{S}} over 𝒮\mathcal{S} is defined inductively as follows: aV𝒮a\in V^{\mathcal{S}} if and only if aa is a function from domaV𝒮\operatorname{dom}a\subseteq V^{\mathcal{S}} to Low(𝒮)ȷ\operatorname{Low}(\mathcal{S})_{\jmath}. For each set xx, we have the canonical representation xˇ\check{x} of xx defined by domxˇ={yˇyx}\operatorname{dom}\check{x}=\{\check{y}\mid y\in x\} and xˇ(yˇ)=\check{x}(\check{y})=\top. Then define the Heyting interpretation [[ϕ]][\mkern-2.0mu[\phi]\mkern-2.0mu] with parameters of V𝒮V^{\mathcal{S}} as follows:

  • [[a=b]]=(xdomaa(x)ydombb(y)[[x=y]])(ydombb(y)xdomaa(x)[[x=y]])[\mkern-2.0mu[a=b]\mkern-2.0mu]=\left(\bigwedge_{x\in\operatorname{dom}a}a(x)\to\bigvee_{y\in\operatorname{dom}b}b(y)\land[\mkern-2.0mu[x=y]\mkern-2.0mu]\right)\land\left(\bigwedge_{y\in\operatorname{dom}b}b(y)\to\bigvee_{x\in\operatorname{dom}a}a(x)\land[\mkern-2.0mu[x=y]\mkern-2.0mu]\right),

  • [[ab]]=ydombb(y)[[a=y]][\mkern-2.0mu[a\in b]\mkern-2.0mu]=\bigvee_{y\in\operatorname{dom}b}b(y)\land[\mkern-2.0mu[a=y]\mkern-2.0mu],

  • [[]]=[\mkern-2.0mu[\bot]\mkern-2.0mu]=\bot, [[ϕψ]]=[[ϕ]][[ψ]][\mkern-2.0mu[\phi\land\psi]\mkern-2.0mu]=[\mkern-2.0mu[\phi]\mkern-2.0mu]\land[\mkern-2.0mu[\psi]\mkern-2.0mu], [[ϕψ]]=[[ϕ]][[ψ]][\mkern-2.0mu[\phi\lor\psi]\mkern-2.0mu]=[\mkern-2.0mu[\phi]\mkern-2.0mu]\lor[\mkern-2.0mu[\psi]\mkern-2.0mu], and [[ϕψ]]=[[ϕ]][[ψ]][\mkern-2.0mu[\phi\to\psi]\mkern-2.0mu]=[\mkern-2.0mu[\phi]\mkern-2.0mu]\to[\mkern-2.0mu[\psi]\mkern-2.0mu],

  • [[xaϕ(x)]]=xdomaa(x)[[ϕ(x)]][\mkern-2.0mu[\forall x\in a\phi(x)]\mkern-2.0mu]=\bigwedge_{x\in\operatorname{dom}a}a(x)\to[\mkern-2.0mu[\phi(x)]\mkern-2.0mu] and [[xaϕ(x)]]=xdomaa(x)[[ϕ(x)]][\mkern-2.0mu[\exists x\in a\phi(x)]\mkern-2.0mu]=\bigvee_{x\in\operatorname{dom}a}a(x)\land[\mkern-2.0mu[\phi(x)]\mkern-2.0mu],

  • [[xϕ(x)]]=xV𝒮[[ϕ(x)]][\mkern-2.0mu[\forall x\phi(x)]\mkern-2.0mu]=\bigwedge_{x\in V^{\mathcal{S}}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu] and [[xϕ(x)]]=xV𝒮[[ϕ(x)]][\mkern-2.0mu[\exists x\phi(x)]\mkern-2.0mu]=\bigvee_{x\in V^{\mathcal{S}}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu].

Then the interpretation validates every axiom of 𝖢𝖹𝖥\mathsf{CZF^{-}} and more:

Theorem 4.1.

Working over 𝖢𝖹𝖥\mathsf{CZF^{-}}, the Heyting-valued model V𝒮V^{\mathcal{S}} also satisfies 𝖢𝖹𝖥\mathsf{CZF^{-}}. If 𝒮\mathcal{S} is set-presented and Subset Collection holds, then V𝒮𝖢𝖹𝖥V^{\mathcal{S}}\models\mathsf{CZF}. If our background theory satisfies Full Separation, then so does V𝒮V^{\mathcal{S}}.

Proof.

The first part of the theorem is shown by [9], so we will concentrate on the preservation of Full Separation. For Full Separation, it suffices to see that the proof for bounded separation over V𝒮V^{\mathcal{S}} also works for Full Separation, since Full Separation ensures [[θ]][\mkern-2.0mu[\theta]\mkern-2.0mu] is a set for every formula θ\theta. ∎

Let us finish this subsection with some constructors, which we need in a later proof.

Definition \thedefinition.

For 𝒮\mathcal{S}-names aa and bb, 𝗎𝗉(a,b)\mathsf{up}(a,b) is defined by dom(𝗎𝗉(a,b))={a,b}\operatorname{dom}(\mathsf{up}(a,b))=\{a,b\} and (𝗎𝗉(a,b))(x)=(\mathsf{up}(a,b))(x)=\top. 𝗈𝗉(a,b)\mathsf{op}(a,b) is the name defined by 𝗈𝗉(a,b)=𝗎𝗉(𝗎𝗉(a,a),𝗎𝗉(a,b))\mathsf{op}(a,b)=\mathsf{up}(\mathsf{up}(a,a),\mathsf{up}(a,b))

𝗎𝗉(a,b)\mathsf{up}(a,b) represents the unordeded pair {a,b}\{a,b\} over V𝒮V^{\mathcal{S}}. Hence the name 𝗈𝗉(a,b)\mathsf{op}(a,b) represents the ordered pair given by aa and bb over V𝒮V^{\mathcal{S}}.

4.2. Double negation formal topology

Our main tool in this paper is the Heyting-valued interpretation with the double negation formal topology. Unlike set-sized realizability or set-represented formal topology, the double negation topology and the resulting Heyting-valued interpretation need not be absolute between BCST-regular sets or transitive models of 𝖢𝖹𝖥\mathsf{CZF^{-}}. Hence we need a careful analysis of the double negation formal topology, which is the aim of this subsection.

Definition \thedefinition.

The double negation formal topology Ω\Omega is the formal topology (1,=,)(1,=,\vartriangleleft), where xpx\vartriangleleft p if and only if ¬¬(xp)\lnot\lnot(x\in p).

We can see that the class of lower sets Low(Ω)\operatorname{Low}(\Omega) is just the power set of 1, and the nucleus of 𝒮\mathcal{S} is given by the double complement

p¬¬={0¬¬(0p)}.p^{\lnot\lnot}=\{0\mid\lnot\lnot(0\in p)\}.

Hence the elements of Low(Ω)ȷ\operatorname{Low}(\Omega)_{\jmath} is the collection of all stable subsets of 1, that is, a set p1p\subseteq 1 such that p=p¬¬p=p^{\lnot\lnot}.

We will frequently mention the relativized Heyting-valued interpretation, the definition of relativized one is not different from the usual VΩV^{\Omega} and [[]][\mkern-2.0mu[\cdot]\mkern-2.0mu], thus we do not introduce its definition. Notwithstanding that, it is still worth mentioning the notational convention for relativization:

Definition \thedefinition.

Let AA be a transitive model of 𝖢𝖹𝖥\mathsf{CZF^{-}}. Then AΩ:=(VΩ)AA^{\Omega}:=(V^{\Omega})^{A} is the relativized Ω\Omega-valued universe to AA. If AA is a set, then A~\tilde{A} denotes the Ω\Omega-name defined by domA~:=AΩ\operatorname{dom}\tilde{A}:=A^{\Omega} and A~(x)=\tilde{A}(x)=\top for all xdomA~x\in\operatorname{dom}\tilde{A}.

Note that the definition of A~\tilde{A} makes sense due to subsection 4.2. We often confuse A~\tilde{A} and AΩA^{\Omega} if context is clear. It is also worth to mention that if jj is an elementary embedding, then j(K~)=j(K)~j(\tilde{K})=\widetilde{j(K)}, so we may write jn(K~)j^{n}(\tilde{K}) instead of jn(K)~\widetilde{j^{n}(K)}.

We cannot expect that Low(Ω)\operatorname{Low}(\Omega) is absolute between transitive models of 𝖢𝖹𝖥\mathsf{CZF^{-}}, and as a result, we do not know whether its Heyting-valued universe VΩV^{\Omega} and Heyting-valued interpretation [[]][\mkern-2.0mu[\cdot]\mkern-2.0mu] is absolute. Fortunately, the formula pLow(Ω)ȷp\in\operatorname{Low}(\Omega)_{\jmath} is Δ0\Delta_{0}, so it is absolute between transitive models of 𝖢𝖹𝖥\mathsf{CZF^{-}}. As a result, we have the following absoluteness result on the Heyting-valued universe:

Lemma \thelemma.

Let AA be a transitive model of 𝖢𝖹𝖥\mathsf{CZF^{-}} without Infinity. Then we have AΩ=VΩAA^{\Omega}=V^{\Omega}\cap A. Moreover, if AA is a set, then AΩA^{\Omega} is also a set.

Proof.

We will follow the proof of Lemma 6.1 of [19]. Let Φ\Phi be the inductive definition given by

X,aΦa is a function such that domaX and a(x)1a(x)¬¬=a(x) for all xdoma.\langle X,a\rangle\in\Phi\iff\text{$a$ is a function such that $\operatorname{dom}a\subseteq X$ and $a(x)\subseteq 1$, $a(x)^{\lnot\lnot}=a(x)$ for all $x\in\operatorname{dom}a$}.

We can see that the Φ\Phi defines the class VΩV^{\Omega}. Furthermore, Φ\Phi is Δ0\Delta_{0}, so it is absolute between transitive models of 𝖢𝖹𝖥\mathsf{CZF^{-}}. By Lemma 2.3, we have a class JJ such that VΩ=aVJaV^{\Omega}=\bigcup_{a\in V}J^{a}, and for each sVs\in V, Js=ΓΦ(tsJt)J^{s}=\Gamma_{\Phi}(\bigcup_{t\in s}J^{t}). Now consider the operation Υ\Upsilon given by

Υ(X):={aAYA(YXY,aΦ)}.\Upsilon(X):=\{a\in A\mid\exists Y\in A(Y\subseteq X\land\langle Y,a\rangle\in\Phi)\}.

By Lemma 2.3 again, there is a class YY such that Ys=Υ(tsYt)Y^{s}=\Upsilon(\bigcup_{t\in s}Y^{t}) for all sVs\in V. Furthermore, we can see that YsVΩY^{s}\subseteq V^{\Omega} by induction on aa.

Let Y=sAYsY=\bigcup_{s\in A}Y^{s}. We claim by induction on ss that JsAYJ^{s}\cap A\subseteq Y. Assume that JtAYJ^{t}\cap A\subseteq Y holds for all tst\in s. If aJsAa\in J^{s}\cap A, then the domain of aa is a subset of A(tsJs)A\cap\left(\bigcup_{t\in s}J^{s}\right), which is a subclass of YY by the inductive assumption and the transitivity of AA. Moreover, for each xdomax\in\operatorname{dom}a there is uu such that xYux\in Y^{u}. By Strong Collection over AA, there is vAv\in A such that for each xdomax\in\operatorname{dom}a there is uvu\in v such that xYux\in Y^{u}. Hence domauvYu\operatorname{dom}a\subseteq\bigcup_{u\in v}Y^{u}, which implies aYvYa\in Y^{v}\subseteq Y.

Hence VΩAYV^{\Omega}\cap A\subseteq Y, and we have Y=VΩAY=V^{\Omega}\cap A. We can see that the construction of YY is the relativized construction of VΩV^{\Omega} to AA, so Y=AΩY=A^{\Omega}. Hence AΩ=VΩAA^{\Omega}=V^{\Omega}\cap A. If AA is a set, then Υ(X)\Upsilon(X) is a set for each set XX, so we can see by induction on aa that YaY^{a} is also a set for each aAa\in A. Hence AΩ=Y=aAYaA^{\Omega}=Y=\bigcup_{a\in A}Y^{a} is also a set. ∎

We extended nucleus ȷ\jmath to JJ for subclasses of Low𝒮\operatorname{Low}\mathcal{S}, and use it to define the validity of formulas of the forcing language. We are working with the specific formal topology 𝒮=Ω\mathcal{S}=\Omega, and in that case, JPJP for a class P1P\subseteq 1 coincides with JP={q¬¬qP}JP=\bigcup\{q^{\lnot\lnot}\mid q\subseteq P\}. It is easy to see that PJPP¬¬P\subseteq JP\subseteq P^{\lnot\lnot}. We also define the following relativized notion for any transitive class AA such that 1A1\in A:

JAP={q¬¬qP and qA}.J^{A}P=\bigcup\{q^{\lnot\lnot}\mid q\subseteq P\text{ and }q\in A\}.

If PAP\in A, then JAP=P¬¬J^{A}P=P^{\lnot\lnot}, and in general, we have PJAPJPP¬¬P\subseteq J^{A}P\subseteq JP\subseteq P^{\lnot\lnot}. Moreover, we have

Lemma \thelemma.

Let AA and BB be transitive classes such that 1A,B1\in A,B and P1P\subseteq 1 be a class.

  1. (1)

    ABA\subseteq B implies JAPJBPJ^{A}P\subseteq J^{B}P.

  2. (2)

    If 𝒫(1)A=𝒫(1)B\mathcal{P}(1)\cap A=\mathcal{P}(1)\cap B, then JAP=JBPJ^{A}P=J^{B}P.

However, the following proposition shows that we cannot prove they are the same from 𝖢𝖹𝖥\mathsf{CZF^{-}}:

Proposition \theproposition.
  1. (1)

    If A𝒫(1)=2A\cap\mathcal{P}(1)=2 (it holds when A=2A=2 or A=VA=V and Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM} holds), then JAP=PJ^{A}P=P.

  2. (2)

    If P¬¬JPP^{\lnot\lnot}\subseteq JP for every class PP, then Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM} implies the law of excluded middle for arbitrary formulas. ∎

JAJ^{A} has a crucial role in defining Heyting-valued interpretation, but it could differ from transitive set to transitive set. It causes absoluteness problems, which is apparently impossible to emend in general. The following lemma states facts on relativized Heyting interpretations:

Lemma \thelemma.

Let ABA\subseteq B be transitive models of 𝖢𝖹𝖥\mathsf{CZF^{-}}. Assume that ϕ\phi is a formula with parameters in AΩA^{\Omega}.

  1. (1)

    If ϕ\phi is bounded, then [[ϕ]]A=[[ϕ]]B[\mkern-2.0mu[\phi]\mkern-2.0mu]^{A}=[\mkern-2.0mu[\phi]\mkern-2.0mu]^{B}.

  2. (2)

    If ϕ\phi only contains bounded quantifications, logical connectives between bounded formulas, unbounded \forall, and \land, then [[ϕ]]A=[[ϕA~]]B[\mkern-2.0mu[\phi]\mkern-2.0mu]^{A}=[\mkern-2.0mu[\phi^{\tilde{A}}]\mkern-2.0mu]^{B}.

  3. (3)

    If every conditional of \to appearing in ϕ\phi is bounded, then [[ϕ]]A[[ϕA~]]B[\mkern-2.0mu[\phi]\mkern-2.0mu]^{A}\subseteq[\mkern-2.0mu[\phi^{\tilde{A}}]\mkern-2.0mu]^{B}.

  4. (4)

    If 𝒫(1)A=𝒫(1)B\mathcal{P}(1)\cap A=\mathcal{P}(1)\cap B, then [[ϕ]]A=[[ϕA~]]B[\mkern-2.0mu[\phi]\mkern-2.0mu]^{A}=[\mkern-2.0mu[\phi^{\tilde{A}}]\mkern-2.0mu]^{B}.

Proof.

If ϕ\phi is bounded, then [[ϕ]][\mkern-2.0mu[\phi]\mkern-2.0mu] is defined in terms of double complement, Heyting connectives between subsets of 1, and set-sized union and intersection. These notions are absolute between transitive sets, so we can prove [[ϕ]][\mkern-2.0mu[\phi]\mkern-2.0mu] is also absolute by induction on ϕ\phi. (In the case of atomic formulas, we apply the induction on AΩA^{\Omega}-names.)

The remaining clauses follow from the induction on ϕ\phi: For the unbounded \forall, we have

[[xϕ(x)]]A=xAΩ[[ϕ(x)]]AxAΩ[[ϕA~(x)]]B=[[xA~ϕA~(x)]]B.[\mkern-2.0mu[\forall x\phi(x)]\mkern-2.0mu]^{A}=\bigwedge_{x\in A^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]^{A}\subseteq\bigwedge_{x\in A^{\Omega}}[\mkern-2.0mu[\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}=[\mkern-2.0mu[\forall x\in\tilde{A}\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}.

under conditions in the remaining clauses. The case for \land and \to are similar. In the case of the second or fourth clause, we can see that the above argument raises equality.

For the unbounded \exists, we have

[[xϕ(x)]]A=JA({[[ϕ(x)]]AxAΩ})JB({[[ϕA~(x)]]BxAΩ})=[[xA~ϕA~(x)]]B.[\mkern-2.0mu[\exists x\phi(x)]\mkern-2.0mu]^{A}=J^{A}\left(\bigcup\{[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]^{A}\mid x\in A^{\Omega}\}\right)\subseteq J^{B}\left(\bigcup\{[\mkern-2.0mu[\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}\mid x\in A^{\Omega}\}\right)=[\mkern-2.0mu[\exists x\in\tilde{A}\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}.

Note that if ABA\in B, then xAΩ[[ϕA~(x)]]B𝒫(1)B\bigcup_{x\in A^{\Omega}}[\mkern-2.0mu[\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}\in\mathcal{P}(1)\cap B, thus JB(xAΩ[[ϕA~(x)]]B)=(xAΩ[[ϕA~(x)]]B)¬¬J^{B}\left(\bigcup_{x\in A^{\Omega}}[\mkern-2.0mu[\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}\right)=\left(\bigcup_{x\in A^{\Omega}}[\mkern-2.0mu[\phi^{\tilde{A}}(x)]\mkern-2.0mu]^{B}\right)^{\lnot\lnot} in this case.

The case for \lor is similar to that of the unbounded \exists. For the last clause, we need JA=JBJ^{A}=J^{B} that follows from 𝒫(1)A=𝒫(1)B\mathcal{P}(1)\cap A=\mathcal{P}(1)\cap B. ∎

As a final remark, note that we may understand the forcing over Ω\Omega as a double negation translation à la forcing. See 2.3 of [11] or [6] for details.

5. Double negation translation of an elementary embedding

The following section is devoted to the following theorem:

Theorem 5.1.

(𝖢𝖹𝖥+𝖲𝖾𝗉)(\mathsf{CZF+Sep}) Let j:VVj:V\prec V be an elementary embedding and let KK be an inaccessible critical point of jj; that is, KK is inaccessible, Kj(K)K\in j(K) and j(x)=xj(x)=x for all xKx\in K. Then there is a Heyting-valued model VΩV^{\Omega} such that VΩ𝖹𝖥V^{\Omega}\models\mathsf{ZF^{-}} and it thinks j:VVj:V\prec V is cofinal and has a critical point.

We will mostly follow the proof of Ziegler [24], but we need to check his proof works on our setting since his applicative topology does not cover Heyting algebra generated by formal topologies that are not set-presentable. We will focus on Reinhardt sets, so considering the target universe MM of jj might be unnecessary. Nevertheless, we will consider MM and we will not assume M=VM=V unless it is necessary.

We need to redefine Heyting-valued interpretations to handle critical sets and Reinhardt sets, so we define the interpretation of MM and jj in the forcing language. Since jj preserves names, we can interpret jj as jj itself. We will interpret MM as the MΩM^{\Omega}. Thus, for example, [[xMϕ(x)]]=xMΩ[[ϕ(x)]][\mkern-2.0mu[\forall x\in M\phi(x)]\mkern-2.0mu]=\bigwedge_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]. We discussed that the Heyting interpretation [[ϕ]][\mkern-2.0mu[\phi]\mkern-2.0mu] need not be absolute between transitive sets. A cacophony of absoluteness issues causes technical trouble in an actual proof, so we want to assure [[ϕ]]=[[ϕ]]M[\mkern-2.0mu[\phi]\mkern-2.0mu]=[\mkern-2.0mu[\phi]\mkern-2.0mu]^{M}, which follows from 𝒫(1)=𝒫(1)M\mathcal{P}(1)=\mathcal{P}(1)\cap M. The following lemma is due to Ziegler, and see Remark 9.51 to 9.52 of [24] for its proof:

Lemma \thelemma.

Let p𝒫(1)p\in\mathcal{P}(1) and j:VMj:V\prec M be an elementary embedding. Then j(p)=pj(p)=p. Especially, we have 𝒫(1)=𝒫(1)M\mathcal{P}(1)=\mathcal{P}(1)\cap M.

Hence by subsection 4.2, [[ϕ]]M=[[ϕMΩ]][\mkern-2.0mu[\phi]\mkern-2.0mu]^{M}=[\mkern-2.0mu[\phi^{M^{\Omega}}]\mkern-2.0mu] for any formula ϕ\phi with parameters in MΩM^{\Omega}. Thus we do not need to worry about the absoluteness issue on the Heyting interpretation.

We are ready to extend our forcing language to {,j,M}\{\in,j,M\}. For each formula ϕ\phi, define

[[xMϕ(x)]]:=xMΩ[[ϕ(x)]]and[[xMϕ(x)]]:=xMΩ[[ϕ(x)]][\mkern-2.0mu[\forall x\in M\phi(x)]\mkern-2.0mu]:=\bigwedge_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]\qquad\text{and}\qquad[\mkern-2.0mu[\exists x\in M\phi(x)]\mkern-2.0mu]:=\bigvee_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]

and define the remaining as given by [9]. (We may define xMx\in M by using that this is equivalent to yM(x=y)\exists y\in M(x=y).) From this definition, we have an analogue of Lemma 4.26 of [24], which is useful to check that jj is still elementary over VΩV^{\Omega}:

Lemma \thelemma.

For any bounded formula ϕ(x)\phi(\vec{x}) with all free variables displayed in the language \in (that is, without jj and MM), we have

[[ϕ(a)]]=[[ϕMΩ(j(a))]]=[[ϕ(j(a))]][\mkern-2.0mu[\phi(\vec{a})]\mkern-2.0mu]=[\mkern-2.0mu[\phi^{M^{\Omega}}(j(\vec{a}))]\mkern-2.0mu]=[\mkern-2.0mu[\phi(j(\vec{a}))]\mkern-2.0mu]

for every aVΩ\vec{a}\in V^{\Omega}.

Proof.

For the first equality, we have

(1) [[ϕ(a)]]=j([[ϕ(a)]])=[[ϕMΩ(j(a))]].[\mkern-2.0mu[\phi(\vec{a})]\mkern-2.0mu]=j([\mkern-2.0mu[\phi(\vec{a})]\mkern-2.0mu])=[\mkern-2.0mu[\phi^{M^{\Omega}}(j(\vec{a}))]\mkern-2.0mu].

by Section 5. Note that the last equality of (1) follows from the induction on ϕ\phi. The second equality also follows from the induction on ϕ\phi. ∎

Moreover, we can check the following equalities easily:

Proposition \theproposition.
  1. (1)

    [[x,y(x=yj(x)=j(y))]]=[\mkern-2.0mu[\forall x,y(x=y\to j(x)=j(y))]\mkern-2.0mu]=\top,

  2. (2)

    [[xj(x)M]]=[\mkern-2.0mu[\forall xj(x)\in M]\mkern-2.0mu]=\top,

  3. (3)

    [[x(xMyx(yM))]]=[\mkern-2.0mu[\forall x(x\in M\to\forall y\in x(y\in M))]\mkern-2.0mu]=\top.

Proof.

The first equality follows from [[x=y]]=[[j(x)=j(y)]][\mkern-2.0mu[x=y]\mkern-2.0mu]=[\mkern-2.0mu[j(x)=j(y)]\mkern-2.0mu], and the remaining two follow from the direct calculation. ∎

Lemma \thelemma.

For every aVΩ\vec{a}\in V^{\Omega} and a formula ϕ\phi that does not contain jj or MM, we have [[ϕ(a)ϕM(j(a))]]=[\mkern-2.0mu[\phi(\vec{a})\leftrightarrow\phi^{M}(j(\vec{a}))]\mkern-2.0mu]=\top.

Proof.

Section 5 shows that this lemma holds for bounded formulas ϕ\phi. We will use full induction on ϕ\phi to prove [[ϕ(a)]]=[[ϕM(j(a))]][\mkern-2.0mu[\phi(\vec{a})]\mkern-2.0mu]=[\mkern-2.0mu[\phi^{M}(j(\vec{a}))]\mkern-2.0mu] for all aVΩ\vec{a}\in V^{\Omega}. If ϕ\phi is xψ(x,a)\forall x\psi(x,\vec{a}), we have [[xϕ(x,a)]]=xVΩ[[ϕ(x,a)]][\mkern-2.0mu[\forall x\phi(x,\vec{a})]\mkern-2.0mu]=\bigwedge_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu]. Since

0xVΩ[[ϕ(x,a)]]\displaystyle 0\in\bigwedge_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu] xVΩ(0[[ϕ(x,a)]])\displaystyle\iff\forall x\in V^{\Omega}(0\in[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu])
x(VΩ)M(0[[ϕ(x,j(a))]]),\displaystyle\iff\forall x\in(V^{\Omega})^{M}(0\in[\mkern-2.0mu[\phi(x,j(\vec{a}))]\mkern-2.0mu]),

where the last equivalence follows from applying jj to the above formula. Since the last formula is equivalent to 0xMΩ[[ϕ(x)]]0\in\bigwedge_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu], we have xVΩ[[ϕ(x)]]=xMΩ[[ϕ(x)]]=[[xMϕM(x)]]\bigwedge_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]=\bigwedge_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x)]\mkern-2.0mu]=[\mkern-2.0mu[\forall x\in M\phi^{M}(x)]\mkern-2.0mu].

If ϕ\phi is xψ(x,a)\exists x\psi(x,a), we have [[xϕ(x,a)]]=xVΩ[[ϕ(x,a)]][\mkern-2.0mu[\exists x\phi(x,a)]\mkern-2.0mu]=\bigvee_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu]. Moreover,

0xVΩ[[ϕ(x,a)]]\displaystyle 0\in\bigvee_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu] p1[pxVΩ[[ϕ(x,a)]] and 0p¬¬]\displaystyle\iff\exists p\subseteq 1\left[p\subseteq\bigcup_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu]\text{ and }0\in p^{\lnot\lnot}\right]
p1[px(VΩ)M[[ϕM(x,j(a))]] and 0p¬¬]\displaystyle\iff\exists p\subseteq 1\left[p\subseteq\bigcup_{x\in(V^{\Omega})^{M}}[\mkern-2.0mu[\phi^{M}(x,j(\vec{a}))]\mkern-2.0mu]\text{ and }0\in p^{\lnot\lnot}\right]

Hence 0xVΩ[[ϕ(x,a)]]0\in\bigvee_{x\in V^{\Omega}}[\mkern-2.0mu[\phi(x,\vec{a})]\mkern-2.0mu] if and only if 0xMΩ[[ϕ(x,j(a))]]0\in\bigvee_{x\in M^{\Omega}}[\mkern-2.0mu[\phi(x,j(\vec{a}))]\mkern-2.0mu]

We extended the language of set theory to treat elementary embedding jj and the target universe MM. We also expand the Full Separation 𝖲𝖾𝗉\mathsf{Sep} and Strong Collection to the extended language. Thus we need to check that Full Separation and Strong Collection under the extended language are also persistent under the double negation interpretation. We can see that the proof given by [9] and Theorem 4.1 carries over, so we have the following:

Proposition \theproposition.
  1. (1)

    If we assume Full Separation for the extended language, then VΩV^{\Omega} also thinks Full Separation for the extended language holds.

  2. (2)

    VΩV^{\Omega} thinks Strong Collection for the extended language holds.

The essential property of a critical set is that it is inaccessible. Unfortunately, there is no hope to preserve the inaccessibility of a critical set. The main reason is that an inaccessible set must satisfy second-order Set Collection, which is not preserved by forcing under Ω\Omega. Fortunately, being a critical point is preserved provided if it is regular:

Lemma \thelemma.

Let KK be a regular set such that Kj(K)K\in j(K) and j(x)=xj(x)=x for all xKx\in K. Then [[Kˇj(K~)xK~(j(x)=x)]]=[\mkern-2.0mu[\check{K}\in j(\tilde{K})\land\forall x\in\tilde{K}(j(x)=x)]\mkern-2.0mu]=\top.

Proof.

Since (j(K) is inaccessible)M(\text{$j(K)$ is inaccessible})^{M}, j(K)Ω=(j(K)Ω)M=j(K)VΩj(K)^{\Omega}=(j(K)^{\Omega})^{M}=j(K)\cap V^{\Omega} is a set by subsection 4.2. Also, Kj(K)K\in j(K) implies K~j(K)\tilde{K}\in j(K). Since the domain of j(K~)=j(K)~j(\tilde{K})=\widetilde{j(K)} is j(K)VΩj(K)\cap V^{\Omega}, we have K~domj(K~)\tilde{K}\in\operatorname{dom}j(\tilde{K}), which implies [[K~j(K~)]]=[\mkern-2.0mu[\tilde{K}\in j(\tilde{K})]\mkern-2.0mu]=\top. For the second assertion, observe that if xdomK~x\in\operatorname{dom}\tilde{K} then j(x)=xj(x)=x, so we have the desired conclusion. ∎

Note that the canonical name Kˇ\check{K}, is also a critical point in VΩV^{\Omega}. However, we stick to use K~\tilde{K}, whose reason becomes apparent in Section 6.

We do not make use of Full Separation or Subset Collection until now. However, the following proof requires Full Separation (or at least, Separation for Σ\Sigma-formulas) or 𝖱𝖤𝖠\mathsf{REA}.

Lemma \thelemma.

If j:VVj:V\prec V is a cofinal elementary embedding, then VΩV^{\Omega} thinks jj is cofinal.

Proof.

Let aVΩa\in V^{\Omega}. Then there is a set XX such that aj(X)a\in j(X). If we assume Full Separation, then XVΩX\cap V^{\Omega} is a set and j(XVΩ)=j(X)VΩj(X\cap V^{\Omega})=j(X)\cap V^{\Omega}. Let bb be a name such that domb=XVΩ\operatorname{dom}b=X\cap V^{\Omega} and b(y)=1b(y)=1 for all ydomby\in\operatorname{dom}b. Then [[aj(b)]]=[\mkern-2.0mu[a\in j(b)]\mkern-2.0mu]=\top.

We need some work if we assume 𝖱𝖤𝖠\mathsf{REA} instead: Take a set XX such that aj(X)a\in j(X). By 𝖱𝖤𝖠\mathsf{REA}, we can find a regular YY such that XYX\in Y. By subsection 4.2, YΩ=YVΩY^{\Omega}=Y\cap V^{\Omega} is a set. The remaining argument is identical to the previous one. ∎

Combining the above lemmas shows our main theorem and more:

Theorem 5.2.

The consistency of the former implies the consistency of the latter. Here the former always assume that the critical point of jj is regular (but not for the latter), MM can be equal to VV, and we always assume that jj is non-trivial:

  1. (1)

    𝖢𝖹𝖥+j:VM\mathsf{CZF^{-}}+j:V\prec M and 𝖢𝖹𝖥+Δ𝟢-𝖫𝖤𝖬+j:VM\mathsf{CZF^{-}+\Delta_{0}\text{-}LEM}+j:V\prec M,

  2. (2)

    𝖢𝖹𝖥+𝖱𝖤𝖠+j:VV\mathsf{CZF+REA}+j:V\prec V and 𝖢𝖹𝖥+Δ𝟢-𝖫𝖤𝖬+j:VV\mathsf{CZF^{-}+\Delta_{0}\text{-}LEM}+j:V\prec V is cofinal,

  3. (3)

    𝖢𝖹𝖥+𝖲𝖾𝗉+j:VV (is cofinal)\mathsf{CZF^{-}+Sep}+j:V\prec V\text{ (is cofinal)} and 𝖹𝖥+j:VV (is cofinal)\mathsf{ZF^{-}}+j:V\prec V\text{ (is cofinal)},

  4. (4)

    𝖢𝖹𝖥+𝖲𝖾𝗉+j:VV\mathsf{CZF+Sep}+j:V\prec V and 𝖹𝖥+j:VV\mathsf{ZF^{-}}+j:V\prec V is cofinal,

  5. (5)

    𝖨𝖹𝖥+j:VM\mathsf{IZF}+j:V\prec M and 𝖹𝖥+j:VM\mathsf{ZF}+j:V\prec M,

Proof.

It follows from our previous lemmas in this section, subsection 3.2, and that Heyting-valued interpretation preserves 𝖲𝖾𝗉\mathsf{Sep} and 𝖯𝗈𝗐\mathsf{Pow}. ∎

Unfortunately, the above result answers little about the consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set, since we know little about the consistency strength of 𝖹𝖥\mathsf{ZF^{-}} with a cofinal embedding j:VVj:V\prec V. The author thought that its consistency strength is very high from the following argument: consider the principle known as the Relation Reflection Scheme (𝖱𝖱𝖲\mathsf{RRS}) defined by Aczel [1]. It is known that 𝖱𝖱𝖲\mathsf{RRS} is persistent under Heyting-valued interpretation, and 𝖹𝖥𝖢\mathsf{ZFC^{-}} proves the equivalence between 𝖱𝖱𝖲\mathsf{RRS} and the reflection principle. Moreover, in an earlier version of [17], Matthews claimed that 𝖹𝖥𝖢\mathsf{ZFC^{-}} with the reflection principle and the existence of a cofinal elementary embedding is sufficient to derive the contradiction. Hence Reinhardt cardinals are incompatible with choice over 𝖹𝖥\mathsf{ZF^{-}} with the reflection principle. However, Matthews pointed out to the author that the claim as stated above is incorrect: the problem is that the reflection principle over 𝖹𝖥𝖢\mathsf{ZFC^{-}} does not ensure that VcritjV_{\operatorname{crit}j} is a set, which seems necessary to derive the contradiction. In spite of that, we can still see that assuming 𝖣𝖢μ\mathsf{DC}_{\mu}-scheme for every cardinal μ\mu proves there is no cofinal elementary embedding over 𝖹𝖥𝖢\mathsf{ZFC^{-}}.

Can we see the revised result as a form of Kunen’s inconsistency phenomenon, so that it is evidence of the consistency strength of 𝖹𝖥\mathsf{ZF^{-}} with a cofinal elementary embedding? We can still prove that 𝖹𝖥𝖢\mathsf{ZFC^{-}} and 𝖹𝖥𝖢\mathsf{ZFC^{-}} with 𝖣𝖢μ\mathsf{DC}_{\mu}-scheme for all μ\mu are equiconsistent since LL satisfies global choice. We know that LL is compatible with small large cardinals over 𝖹𝖥𝖢\mathsf{ZFC}, so we may guess the same holds for 𝖹𝖥𝖢\mathsf{ZFC^{-}}. However, we do not aware well about large cardinals over 𝖹𝖥𝖢\mathsf{ZFC^{-}} to conclude that the consistency strength of 𝖹𝖥\mathsf{ZF^{-}} with a cofinal elementary is high. Even worse, LL is not compatible with large large cardinals. Thus we cannot extend the same argument further. Analyzing the notion of large cardinals over 𝖹𝖥𝖢\mathsf{ZFC^{-}} and their consistency with 𝖣𝖢μ\mathsf{DC}_{\mu}-scheme and reflection principles would be an interesting topic, but beyond the scope of this paper.

Question \thequestion.

What is the exact consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set?

Despite that, we will see in the next section that 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set is still quite strong.

6. An analysis on the critical point

In the previous section, we saw that non-trivial elementary embeddings over 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} have a strong consistency strength. However, we observed nothing about how much large cardinal properties of the critical point of an jj are preserved by the double negation translation. In this section, we will extract the large cardinal properties of the critical point of an elementary embedding. Especially, we will focus on the critical point of a Reinhardt embedding.

The main result of this section is as follows:

Theorem 6.1.

Let j:VMj:V\prec M be a elementary embedding with an inaccessible critical point KK. Then VΩV^{\Omega} thinks n,mω(n<mjn(K~) is BCST-regular)jm(K~)\forall n,m\in\omega(n<m\to\text{$j^{n}(\tilde{K})$ is BCST-regular})^{j^{m}(\tilde{K})}.

Furthermore, if j:VVj:V\prec V, then VΩV^{\Omega} also thinks jjn(K~)jn+1(K~)j\mathrel{\upharpoonright}j^{n}(\tilde{K})\in j^{n+1}(\tilde{K}).

The main strategy of this theorem is to internalize the proof of VΩ𝖢𝖹𝖥V^{\Omega}\models\mathsf{CZF^{-}} into a BCST-regular set. We mostly follow the proof of [9], but for the sake of verification, we will provide most of the detail of relevant lemmas and their proof. Throughout this section, ABA\in B sets of the same power set of 1 (i.e., 𝒫(1)A=𝒫(1)B\mathcal{P}(1)\cap A=\mathcal{P}(1)\cap B) such that AA is BCST-regular and BB is a transitive model of 𝖢𝖹𝖥\mathsf{CZF^{-}}, and RBR\in B is a multi-valued function unless specified.

Lemma \thelemma.

Let aAa\in A, R:aAR:a\rightrightarrows A, Qa×AQ\subseteq a\times A, and QBQ\in B. Moreover, assume that

  1. (1)

    x,yRyQx={yx,yQ}\langle x,y\rangle\in R\to y\subseteq Q_{x}=\{y\mid\langle x,y\rangle\in Q\}, and

  2. (2)

    (Monotone Closeness) x,yR\langle x,y\rangle\in R and yzQxy\subseteq z\subseteq Q_{x} implies x,zR\langle x,z\rangle\in R,

then there is fAAaf\in A\cap{{}^{a}}A such that x,f(x)R\langle x,f(x)\rangle\in R for all xax\in a.

Proof.

By subsection 2.2, there is bAb\in A such that bRb\subseteq R and b:aAb:a\rightrightarrows A. Take ff as xbx={yx,yb}x\mapsto\bigcup b_{x}=\bigcup\{y\mid\langle x,y\rangle\in b\}. Then fAf\in A, since AA satisfies Union and the second-order Replacement. Moreover, by monotone-closedness of RR, we have x,f(x)R\langle x,f(x)\rangle\in R for all xax\in a. ∎

Lemma \thelemma.

Let PAP\subseteq A, PBP\in B, aAa\in A and R:aA𝒫(P)R:a\rightrightarrows A\cap\mathcal{P}(P). Furthermore, assume that RR is monotone closed, that is, y,zAy,z\in A, yzPy\subseteq z\subseteq P, and x,yR\langle x,y\rangle\in R implies x,zR\langle x,z\rangle\in R. Then there is bAb\in A such that bPb\subseteq P and x,bR\langle x,b\rangle\in R for all xax\in a.

Proof.

Applying Section 6 to RR provides a function fAAaf\in A\cap{{}^{a}}A such that x,f(x)R\langle x,f(x)\rangle\in R for all xAx\in A. Now take b={f(x)xa}Ab=\bigcup\{f(x)\mid x\in a\}\in A, then we have x,bR\langle x,b\rangle\in R by monotone closeness of RR. ∎

The following lemma has a critical role in the proof of our theorem. Moreover, this lemma requires 𝒫(1)A=𝒫(1)B\mathcal{P}(1)\cap A=\mathcal{P}(1)\cap B:

Lemma \thelemma.

Let aAΩa\in A^{\Omega}, R:aAR:a\rightrightarrows A, RBR\in B and pAp\in A be such that p1p\subseteq 1 and p=p¬¬p=p^{\lnot\lnot}. Define

P={x,y,tdoma×AΩ×1t(pa(x)[[𝗈𝗉(x,y)R]]B)}.P=\{\langle x,y,t\rangle\in\operatorname{dom}a\times A^{\Omega}\times 1\mid t\in(p\land a(x)\land[\mkern-2.0mu[\mathsf{op}(x,y)\in R]\mkern-2.0mu]^{B})\}.

Furthermore, assume that we have p[[R:aA~]]Bp\subseteq[\mkern-2.0mu[R:a\rightrightarrows\tilde{A}]\mkern-2.0mu]^{B}. Then there is rAr\in A such that rPr\subseteq P and pa(x){tyAΩx,y,tr}¬¬p\land a(x)\subseteq\{t\mid\exists y\in A^{\Omega}\langle x,y,t\rangle\in r\}^{\lnot\lnot}.

Proof.

Let Q={x,tyAΩ(x,y,tP)}Q=\{\langle x,t\rangle\mid\exists y\in A^{\Omega}(\langle x,y,t\rangle\in P)\} and Qx={tx,tQ}1Q_{x}=\{t\mid\langle x,t\rangle\in Q\}\subseteq 1. Then Qx𝒫(1)B=𝒫(1)AQ_{x}\in\mathcal{P}(1)\cap B=\mathcal{P}(1)\cap A. From p[[R:aA~]]Bp\subseteq[\mkern-2.0mu[R:a\rightrightarrows\tilde{A}]\mkern-2.0mu]^{B}, we can deduce pa(x)Qx¬¬p\land a(x)\subseteq Q_{x}^{\lnot\lnot}. By applying Section 6 to the relation

{x,vdoma×(𝒫(1)A)pa(x)v¬¬ and vQx}\{\langle x,v\rangle\in\operatorname{dom}a\times(\mathcal{P}(1)\cap A)\mid p\land a(x)\subseteq v^{\lnot\lnot}\text{ and }v\subseteq Q_{x}\}

we have a function fAdomaAf\in{{}^{\operatorname{dom}a}}A\cap A such that pa(x)f(x)¬¬p\land a(x)\subseteq f(x)^{\lnot\lnot} and f(x)Qxf(x)\subseteq Q_{x} for all xdomax\in\operatorname{dom}a. (The condition 𝒫(1)B=𝒫(1)A\mathcal{P}(1)\cap B=\mathcal{P}(1)\cap A is necessary to ensure the above relation is a multi-valued function of domain doma\operatorname{dom}a.) Now let

q={x,txdoma and tf(x)}.q=\{\langle x,t\rangle\mid x\in\operatorname{dom}a\text{ and }t\in f(x)\}.

Then x,tqyAΩ(x,y,tP)\forall\langle x,t\rangle\in q\exists y\in A^{\Omega}(\langle x,y,t\rangle\in P) holds. That is, P:qAΩP:q\rightrightarrows A^{\Omega}. By subsection 2.2, there is rAr\in A such that rPr\subseteq P and r:qAΩr:q\rightrightarrows A^{\Omega}. It is easy to see that rr satisfies the desired property. ∎

There is some technical note for the proof: there is no need that PP, QQ, and QxQ_{x} are definable over AA in general. The reason is that we do not know either RR or [[]]B[\mkern-2.0mu[\cdot]\mkern-2.0mu]^{B} is accessible from AA. However, we do not need to care about it since we are relying on the second-order Strong Collection over AA.

Theorem 6.2.

Let AA be a BCST-regular set. Then BΩB^{\Omega} thinks A~\tilde{A} is BCST-regular.

Proof.

It is easy to see that BΩB^{\Omega} thinks A~\tilde{A} is transitive and closed under Pairing, Union, and Binary Intersection. Hence it remains to show that BΩB^{\Omega} thinks A~\tilde{A} satisfies second-order Strong Collection, that is,

[[aA~R[R:aA~bA~(R:ab)]]]B=.[\mkern-2.0mu[\forall a\in\tilde{A}\forall R[R:a\rightrightarrows\tilde{A}\to\exists b\in\tilde{A}(R:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b)]]\mkern-2.0mu]^{B}=\top.

Take adomA~a\in\operatorname{dom}\tilde{A}, RBΩR\in B^{\Omega} and pAp\in A such that p1p\subseteq 1 and p=p¬¬p=p^{\lnot\lnot}. We claim that if p[[R:aA~]]Bp\subseteq[\mkern-2.0mu[R:a\rightrightarrows\tilde{A}]\mkern-2.0mu]^{B}, then there is bdomA~b\in\operatorname{dom}\tilde{A} such that p[[R:ab]]Bp\subseteq[\mkern-2.0mu[R:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b]\mkern-2.0mu]^{B}. By Section 6, we have rAr\in A such that rPr\subseteq P and pa(x){tyx,y,zr}¬¬p\land a(x)\subseteq\{t\mid\exists y\langle x,y,z\rangle\in r\}^{\lnot\lnot}. Define bb such that domb={yx,t(x,y,tr)}\operatorname{dom}b=\{y\mid\exists x,t(\langle x,y,t\rangle\in r)\} and

b(y)={0xx,y,0r}¬¬.b(y)=\{0\mid\exists x\langle x,y,0\rangle\in r\}^{\lnot\lnot}.

for ydomby\in\operatorname{dom}b. (Note that b(y)Ab(y)\in A.) Then we have p[[R:ab]]Bp\subseteq[\mkern-2.0mu[R:a\>\mathrlap{\leftleftarrows}{\rightrightarrows}\>b]\mkern-2.0mu]^{B}. ∎

Hence we have

Corollary \thecorollary.

[[jn(K~) is BCST-regular]]jm(K)=[\mkern-2.0mu[j^{n}(\tilde{K})\text{\text{ is BCST-regular}}]\mkern-2.0mu]^{j^{m}(K)}=\top. Furthermore, [[(jn(K~) is BCST-regular)jm(Kˇ)]]=[\mkern-2.0mu[(j^{n}(\tilde{K})\text{\text{ is BCST-regular}})^{j^{m}(\check{K})}]\mkern-2.0mu]=\top.

Proof.

The first statement follows from Theorem 6.2 by taking A=KA=K and B=jm(K)B=j^{m}(K) for n=1n=1, and applying jj n1n-1 times for general cases. For the last statement, we can apply the third clause of subsection 4.2. ∎

However, it does not directly result in our desired theorem, since we do not know Ω\Omega forces jn(K~)j^{n}(\tilde{K}) is uniformly BCST-regular. We need some work to see this.

There are two possible meanings of jn(K)j^{n}(K): the first is applying jj nn times to KK. Here nn must be a natural number over the metatheory, and this description lacks a way to describe the sequence jn(K)nω\langle j^{n}(K)\mid n\in\omega\rangle. The second way to see jn(K)j^{n}(K) is to understand it as it is given by the following recursion:

j0(K)=K and jn+1(K)=j(jn(K)).j^{0}(K)=K\text{ and }j^{n+1}(K)=j(j^{n}(K)).

Thus the formal statement of ϕ(jn(K))\phi(j^{n}(K)) is

f[domf=ωf(0)=Kmω(f(m+1)=j(f(m)))]ϕ(f(n)).\exists f[\operatorname{dom}f=\omega\land f(0)=K\land\forall m\in\omega(f(m+1)=j(f(m)))]\land\phi(f(n)).
Theorem 6.3.

VΩV^{\Omega} thinks the following statement is valid:

(2) f[f is a functiondomf=ωf(0)=Kmω(f(m+1)=j(f(m)))]n,mω[n<m(f(n) is BCST-regular)f(m)].\exists f[\text{$f$ is a function}\land\operatorname{dom}f=\omega\land f(0)=K\land\forall m\in\omega(f(m+1)=j(f(m)))]\\ \land\forall n,m\in\omega[n<m\to(f(n)\text{ is BCST-regular})^{f(m)}].
Proof.

Let domf={𝗈𝗉(nˇ,jn(K~))nω}\operatorname{dom}f=\{\mathsf{op}(\check{n},j^{n}(\tilde{K}))\mid n\in\omega\}. We claim that ff witnesses our theorem. The first three conditions are easy to prove. To see the last condition, let n,mωn,m\in\omega. We can show the following facts by induction on mm:

  1. (1)

    If 0[[nˇ<mˇ]]0\in[\mkern-2.0mu[\check{n}<\check{m}]\mkern-2.0mu] then n<mn<m.

  2. (2)

    [[f(m)=jm(K~)]]=[\mkern-2.0mu[f(m)=j^{m}(\tilde{K})]\mkern-2.0mu]=\top.

By combining these facts with Section 6, we have [[nˇ<mˇ]][[(f(n) is BCST-regular)f(m)]][\mkern-2.0mu[\check{n}<\check{m}]\mkern-2.0mu]\subseteq[\mkern-2.0mu[(f(n)\text{ is BCST-regular})^{f(m)}]\mkern-2.0mu]. Hence the result follows. ∎

Hence K~\tilde{K} in VΩV^{\Omega} has the following reflection property: For every n<mn<m, jm(K~)(jn(K~)𝖢𝖹𝖥𝟤)j^{m}(\tilde{K})\models(j^{n}(\tilde{K})\models\mathsf{CZF^{-}_{2}}). How much is this reflection principle strong? To see this, assume that we started from 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with an elementary embedding j:VVj:V\prec V and produced the Heyting interpretation under Ω\Omega. Then the Heyting interpretation interprets 𝖹𝖥\mathsf{ZF^{-}} with a cofinal elementary embedding j:VVj:V\prec V. By Theorem 6.1, a critical point K~\tilde{K} of jj satisfies j(K~)(K~𝖢𝖹𝖥𝟤)j(\tilde{K})\models(\tilde{K}\models\mathsf{CZF^{-}_{2}}). Due to the help of the classical logic, we have j(K~)(K~𝖹𝖥𝟤)j(\tilde{K})\models(\tilde{K}\models\mathsf{ZF_{2}^{-}}). For each xK~x\in\tilde{K}, we have j(K~)X(xXX𝖹𝖥𝟤)j(\tilde{K})\models\exists X(x\in X\land X\models\mathsf{ZF_{2}^{-}}). By the property of jj, we have K~xX(xXX𝖹𝖥𝟤)\tilde{K}\models\forall x\exists X(x\in X\land X\models\mathsf{ZF^{-}_{2}}). That is, K~\tilde{K} satisfies 𝖱𝖤𝖠\mathsf{REA}. Since 𝖱𝖤𝖠\mathsf{REA} implies the Axiom of Subset Collection, which is equivalent to Power Set in the classical context, we have K~𝖹𝖥\tilde{K}\models\mathsf{ZF} and j(K~)(K~𝖹𝖥𝟤)j(\tilde{K})\models(\tilde{K}\models\mathsf{ZF_{2}})! We may extend it further to stronger notions of large cardinal properties, like inaccessibility, Mahloness, or indescribability. However, the following example describes there is a limit of large cardinal property we can achieve from the reflection property of K~\tilde{K}:

Example \theexample.

Work over 𝖹𝖥\mathsf{ZF} with a Reinhardt cardinal. Let j:VVj:V\prec V be an elementary embedding and κ=critj\kappa=\operatorname{crit}j. Take K=LκK=L_{\kappa}. Since κ\kappa is a critical cardinal, it is strongly inaccessible by Proposition 3.3 of [14]. Hence Lj(κ)L_{j(\kappa)} also thinks κ\kappa is inaccessible. Especially, Lj(κ)L_{j(\kappa)} thinks Vκ=LκV_{\kappa}=L_{\kappa} is a model of 𝖹𝖥𝖢𝟤\mathsf{ZFC_{2}}.

Since LL is incompatible with large cardinals stronger than the existence of 00^{\sharp}, the above example shows the previous argument with KK does not yield large cardinal properties stronger than the existence of 00^{\sharp}. However, it does not mean there is no room for stronger properties of KK if jj has a stroger property. The following result shows we can extract more large cardinal properties from KK if j:VVj:V\prec V, by bringing the elementary embedding jj over 𝖢𝖹𝖥\mathsf{CZF} to an elementary embedding of jω(K~)j^{\omega}(\tilde{K}) in the Heyting interpretation:

Lemma \thelemma (𝖢𝖹𝖥\mathsf{CZF^{-}}).

Let j:VVj:V\prec V. Then VΩV^{\Omega} thinks jjn(K~)jn+1(K~)j\mathrel{\upharpoonright}j^{n}(\tilde{K})\in j^{n+1}(\tilde{K}) for all nωn\in\omega.

Proof.

Observe that jn+1(K)j^{n+1}(K) is regular and jjn(K):jn(K)jn+1(K)j\mathrel{\upharpoonright}j^{n}(K):j^{n}(K)\to j^{n+1}(K) is a multi-valued function. By subsection 2.2, there is bjn(K)b\in j^{n}(K) such that bjjn(K)b\subseteq j\mathrel{\upharpoonright}j^{n}(K) and b:jn(K)jn+1(K)b:j^{n}(K)\rightrightarrows j^{n+1}(K). Since jjn(K)j\mathrel{\upharpoonright}j^{n}(K) is a function, bb is also a function of domain jn(K)j^{n}(K). Hence jjn(K)=bjn+1(K)j\mathrel{\upharpoonright}j^{n}(K)=b\in j^{n+1}(K).

From the previous argument, we also have jjn(K~)=(jjn(K))jn(K~)jn+1(K)j\mathrel{\upharpoonright}j^{n}(\tilde{K})=(j\mathrel{\upharpoonright}j^{n}(K))\mathrel{\upharpoonright}j^{n}(\tilde{K})\in j^{n+1}(K). Now let cc be a name such that

domc={𝗈𝗉(x,y)x,yjjn(K~)}and\operatorname{dom}c=\{\mathsf{op}(x,y)\mid\langle x,y\rangle\in j\mathrel{\upharpoonright}j^{n}(\tilde{K})\}\quad\text{and}\quad

and c(x)=c(x)=\top for all xdomcx\in\operatorname{dom}c. By definition of cc, cjn+1(K~)c\in j^{n+1}(\tilde{K}). Moreover, it is easy to see that VΩV^{\Omega} thinks cc is a function of the domain jn(K~)j^{n}(\tilde{K}), and direct calculation shows [[xy𝗈𝗉(x,y)cj(x)=y]]=[\mkern-2.0mu[\forall x\forall y\mathsf{op}(x,y)\in c\to j(x)=y]\mkern-2.0mu]=\top. Hence [[c=jjn(K~)]]=[\mkern-2.0mu[c=j\mathrel{\upharpoonright}j^{n}(\tilde{K})]\mkern-2.0mu]=\top. ∎

Note that this lemma does not work for general j:VMj:V\prec M, since we do not know jn(K)j^{n}(K) is regular in VV.

Theorem 6.4 (𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep}).

VΩV^{\Omega} thinks jω(K~):=nωjn(K~)j^{\omega}(\tilde{K}):=\bigcup_{n\in\omega}j^{n}(\tilde{K}) satisfies 𝖹𝖥\mathsf{ZF} with a cofinal elementary embedding jj from itself to itself. Moreover, jω(K~)j^{\omega}(\tilde{K}) satisfies Δ0\Delta_{0}-Separation with jj be allowed to appear.

Proof.

Let K~=j(K~)\tilde{K}=j(\tilde{K}) and Λ:=jω(K~)\Lambda:=j^{\omega}(\tilde{K}). We will rely on a completely internal argument to VΩV^{\Omega}, which is a model of 𝖹𝖥\mathsf{ZF^{-}}.

Assume that Λϕ(a)\Lambda\models\phi(\vec{a}), where ϕ(x)\phi(\vec{x}) is a Δ0\Delta_{0}-formula with all free variables displayed in the language \in (i.e., without jj.) Since j(Λ)=Λj(\Lambda)=\Lambda, we have Λϕ(j(a))\Lambda\models\phi(j(\vec{a})). Hence j:VVj:V\prec V.

It remains to show that Λ\Lambda satisfies Δ0\Delta_{0}-separation for formulas with jj be allowed to appear. Let aΛa\in\Lambda, then there is nn such that ajn(K~)a\in j^{n}(\tilde{K}). For a bounded formula ϕ\phi with parameters in jn(K~)j^{n}(\tilde{K}), let ϕ\phi^{\prime} be the formula obtained by every occurrence of jj to jjn(K~)j\mathrel{\upharpoonright}j^{n}(\tilde{K}). Then {xaϕ(x)}={xaϕ(x)}jn+1(K~)\{x\in a\mid\phi(x)\}=\{x\in a\mid\phi^{\prime}(x)\}\in j^{n+1}(\tilde{K}). Thus Λ\Lambda satisfies Δ0\Delta_{0}-separation for formulas with jj. ∎

How much is the resulting theory strong? We can see that Λ\Lambda is a model of 𝖹𝖥\mathsf{ZF} with the Wholeness axiom for Δ0\Delta_{0}-formulas 𝖶𝖠𝟢\mathsf{WA_{0}} proposed by Hamkins [13]. The author does not know the exact consistency strength of 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}} in 𝖹𝖥𝖢\mathsf{ZFC}-context. However, we can still find a lower bound of it: we can see that Λ\Lambda also thinks κ=rankK\kappa=\operatorname{rank}K is a critical point of jj, and the critical sequence defined by κ0=κ\kappa_{0}=\kappa and κn+1=j(κn)\kappa_{n+1}=j(\kappa_{n}) is cofinal over Ord\mathrm{Ord}. (Note that the cofinal sequence is still definable, although it may not a set. See Proposition 3.2 of [7] for details.) From this, we have

Lemma \thelemma (𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}}).

If the critical sequence is cofinal, then κ0\kappa_{0} is extendible.

Proof.

Let η\eta be an ordinal. Take nn such that η<jn(κ)\eta<j^{n}(\kappa), then jn:Vκ+ηVjn(κ+η)j^{n}:V_{\kappa+\eta}\prec V_{j^{n}(\kappa+\eta)} and critjn=κ\operatorname{crit}j^{n}=\kappa. Hence κ\kappa satisfies η\eta-extendibility. ∎

By an easy reflection argument, we can see also see that 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}} with the cofinal critical sequence proves not only there is an extendible cardinal, but also the consistency of 𝖹𝖥\mathsf{ZF} with a proper class of extendible cardinals, an extendible limit of extendible cardinals, and many more. Since extendible cardinals are preserved by Woodin’s forcing (Theorem 226 of [23]), we have a lower bound of the consistency strength of 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}}, e.g., 𝖹𝖥𝖢\mathsf{ZFC} with there is a proper class of extendible cardinals.

7. Concluding Questions

We may wonder how to find the upper bound of the consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a non-trivial elementary embedding, in terms of extensions of 𝖹𝖥𝖢\mathsf{ZFC} or 𝖹𝖥𝖢\mathsf{ZFC^{-}}. Most construction of interpretations of 𝖢𝖹𝖥\mathsf{CZF} from classical theories rely on realizability, and employ type-theoretic interpretations (like [18] or [12]) or set-as-tree interpretation (like [16] or functional realizability model over 𝖹𝖥𝖢\mathsf{ZFC^{-}} given by Swan [22].333Swan worked over 𝖹𝖥𝖢\mathsf{ZFC}, but his proof for soundness of functional realizability still works over 𝖹𝖥𝖢\mathsf{ZFC^{-}}. Also, note that his functional realizability is a part of his two-stage Kripke model.) These interpretations usually satisfy the Axiom of Subcountability, which states every set is an image of a subset of ω\omega. However, Ziegler [24] proved that the existence of non-trivial elementary embedding j:VMj:V\prec M contradicts with the Axiom of Subcountability. The author thinks that the Axiom of Subcountability comes from that we are using countable pca to construct interpretations, but delimiting the size of pca does not guarantee that we can reach the upper bound of 𝖢𝖹𝖥\mathsf{CZF} with a non-trivial elementary embedding. We do not know about the type-theoretic analogue of 𝖢𝖹𝖥\mathsf{CZF} with an elementary embedding, so we do not know we can employ type-theoretic interpretations. Functional realizability or set-as-tree interpretations also have a problem. We need a pca of size greater than not only the critical point but also its successive application to the elementary embedding to ensure the critical set exists under the interpretation since the size of the pca delimits the size of every set under the interpretation. However, nothing much is known about realizability under large pcas. Especially, these large pcas are not fixed under the elementary embedding, which makes them hard to handle. One possible way to control the large pcas under the elementary embedding jj is to add jj into the pca. However, finding the realizer of ϕM(j(x))ϕ(x)\phi^{M}(j(x))\to\phi(x) seems not easy.

Question \thequestion.

What is the upper bound of the consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a critical set or a Reinhardt set? Can we find the bound in terms of large cardinals compatible with the Axiom of Choice?

Our method is also restricted to the analysis on 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep}, mainly because the forcing over the double negation topology only provides Δ𝟢-𝖫𝖤𝖬\mathsf{\Delta_{0}\text{-}LEM}. Full Separation is necessary to turn it to the full excluded middle. We may ask we can analyze the strength of 𝖢𝖹𝖥\mathsf{CZF} with large large set axioms without any help of full Separation.

Question \thequestion.

Is there any non-trivial result for the consistency strength of 𝖢𝖹𝖥\mathsf{CZF} with a critical set or a Reinhardt set?

Improving the lower bound of the consistency strength of 𝖢𝖹𝖥+𝖲𝖾𝗉\mathsf{CZF+Sep} with a Reinhardt set is also an issue. We have not made use of the full strength of the resulting theory stated in Section 1 to get the lower bound of the consistency strength. For example, we never used the cofinality of jj. Even worse, we did not extract the full consistency strength of 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}}: extendibility (or a proper class of extendibles) is far below 𝖶𝖠𝟢\mathsf{WA_{0}}. One may try to force the Axiom of Choice over 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}}, by using Woodin’s forcing (see Theorem 226 of [23]) and a method given by Hamkins [13]. However, the quotient of the limit stages of Woodin’s forcing is not sufficiently γ\gamma-closed, so Hamkins’ argument is not applied well.

Question \thequestion.

Can we obtain a better lower bound of the consistency strength of 𝖹𝖥\mathsf{ZF^{-}} with a cofinal embedding j:VVj:V\prec V, with a critical point K𝖹𝖥K\models\mathsf{ZF} such that jω(K)(K𝖹𝖥𝟤)j^{\omega}(K)\models(K\models\mathsf{ZF_{2}^{-}})? Especially, are 𝖹𝖥+𝖶𝖠𝟢\mathsf{ZF+WA_{0}} and 𝖹𝖥𝖢+𝖶𝖠𝟢\mathsf{ZFC+WA_{0}} equiconsistent?

Acknowledgements

The author wants to thank Richard Matthews for pointing out an error and providing comments on earlier versions of this manuscript.

References

  • [1] Peter Aczel “The Relation Reflection Scheme” In Mathematical Logic Quarterly 54.1 Wiley Online Library, 2008, pp. 5–11
  • [2] Peter Aczel “The type theoretic interpretation of constructive set theory” In Logic colloquium 77, 1978, pp. 55–66
  • [3] Peter Aczel “The type theoretic interpretation of constructive set theory: inductive definitions” In Studies in Logic and the Foundations of Mathematics 114 Elsevier, 1986, pp. 17–49
  • [4] Peter Aczel and Michael Rathjen “CST Book draft”, 2010 URL: https://www1.maths.leeds.ac.uk/~rathjen/book.pdf
  • [5] Peter Aczel and Michael Rathjen “Notes on Constructive Set Theory”, 2001
  • [6] John L. Bell “Intuitionistic set theory” College Publications, 2014
  • [7] Paul Corazza “The wholeness axiom and Laver sequences” In Annals of Pure and Applied Logic 105.1-3 Elsevier, 2000, pp. 157–260
  • [8] Sy-David Friedman, Victoria Gitman and Vladimir Kanovei “A model of second-order arithmetic satisfying AC but not DC” In Journal of Mathematical Logic 19.01 World Scientific, 2019, pp. 1850013
  • [9] Nicola Gambino “Heyting-valued interpretations for constructive set theory” In Annals of Pure and Applied Logic 137.1-3, 2006, pp. 164–188
  • [10] Victoria Gitman, Joel David Hamkins and Thomas A Johnstone “What is the theory 𝖹𝖥𝖢\mathsf{ZFC} without power set?” In Mathematical Logic Quarterly 62.4-5 Wiley Online Library, 2016, pp. 391–406
  • [11] Robin J. Grayson “Heyting-valued models for intuitionistic set theory” In Applications of Sheaves Springer, 1979, pp. 402–414
  • [12] Edward Griffor and Michael Rathjen “The strength of some Martin-Löf type theories” In Archive for Mathematical Logic 33.5 Springer, 1994, pp. 347–385
  • [13] Joel David Hamkins “The wholeness axioms and V= HOD” In Archive for Mathematical Logic 40.1 Springer, 2001, pp. 1–8
  • [14] Yair Hayut and Asaf Karagila “Critical cardinals” In Israel Journal of Mathematics 236.1 Springer, 2020, pp. 449–472
  • [15] Kenneth Kunen “Elementary embeddings and infinitary combinatorics” In The Journal of Symbolic Logic 36.3 Cambridge University Press, 1971, pp. 407–413
  • [16] Robert S. Lubarsky “CZF and Second Order Arithmetic” In Annals of Pure and Applied Logic 141, 2006, pp. 29–34
  • [17] Richard Matthews “Taking Reinhardt’s Power Away”, 2020 arXiv:2009.01127
  • [18] Michael Rathjen “Constructive Set Theory and Brouwerian Principles” In Journal of Universal Computer Science 11.12, 2005, pp. 2008–2033
  • [19] Michael Rathjen “Realizability for constructive Zermelo-Fraenkel set theory” In Logic Colloquium ’03 24, Lecture Notes in Logic La Jolla, CA: Association for Symbolic Logic, 2006, pp. 282–314
  • [20] Michael Rathjen and Robert S. Lubarsky “On the regular extension axiom and its variants” In Mathematical Logic Quarterly: Mathematical Logic Quarterly 49.5 Wiley Online Library, 2003, pp. 511–518
  • [21] Farmer Schlutzenberg “On the consistency of ZF with an elementary embedding from Vλ+2V_{\lambda+2} into Vλ+2V_{\lambda+2}”, 2020 arXiv:2006.01077
  • [22] Andrew W. Swan “CZF does not have the existence property” In Annals of Pure and Applied Logic 165.5 Elsevier, 2014, pp. 1115–1147
  • [23] W. Woodin “Suitable extender models I” In Journal of Mathematical Logic 10.01n02 World Scientific, 2010, pp. 101–339
  • [24] Albert Ziegler “Large sets in constructive set theory”, 2014