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

Can you take Akemann–Weaver’s 1\diamondsuit_{\aleph_{1}} away?

Daniel Calderón Department of Mathematics and Statistics, York University, 4700 Keele Street, Toronto, Ontario, Canada, M3J 1P3 [email protected]  and  Ilijas Farah Department of Mathematics and Statistics, York University, 4700 Keele Street, Toronto, Ontario, Canada, M3J 1P3 Matematički Institut SANU, Kneza Mihaila 36, 11000 Beograd, p.p. 367, Serbia [email protected]
(Date: April 3, 2025)
Abstract.

By Glimm’s dichotomy, a separable, simple C\textrm{C}^{*}-algebra has continuum many unitarily inequivalent irreducible representations if, and only if, it is non-type I while all of its irreducible representations are unitarily equivalent if, and only if, it is type I. Naimark asked whether the latter equivalence holds for all C\textrm{C}^{*}-algebras.

In 2004, Akemann and Weaver gave a negative answer to Naimark’s problem using Jensen’s Diamond Principle 1\diamondsuit_{\aleph_{1}}, a powerful diagonalization principle that implies the Continuum Hypothesis (𝖢𝖧\operatorname{\mathsf{CH}}). By a result of Rosenberg, a separably represented, simple C\textrm{C}^{*}-algebra with a unique irreducible representation is necessarily of type I. We show that this result is sharp by constructing an example of a separably represented, simple C\textrm{C}^{*}-algebra that has exactly two inequivalent irreducible representations, and therefore does not satisfy the conclusion of Glimm’s dichotomy. Our construction uses a weakening of Jensen’s 1\diamondsuit_{\aleph_{1}}, denoted 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}, that holds in the original Cohen’s model for the negation of 𝖢𝖧\operatorname{\mathsf{CH}}. We also prove that 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} suffices to give a negative answer to Naimark’s problem. Our main technical tool is a forcing notion that generically adds an automorphism of a given C\textrm{C}^{*}-algebra with a prescribed action on its space of pure states.

Key words and phrases:
Representations of C\textrm{C}^{*}-algebras, forcing, Jensen’s Diamond, Naimark’s problem
2010 Mathematics Subject Classification:
Primary: 03E35, 46L30.
Partially supported by NSERC
Corresponding author: Ilijas Farah. ORCID: 0000-0001-7703-6931

A Carlos Di Prisco en su cumpleaños número 70.

1. Introduction

A major early result in the theory of operator algebras (and, at the time, possibly the deepest result in the theory; see [2, §IV.1.5]) was Glimm’s 1960 dichotomy theorem. It states (among other things) that a separable and simple C\textrm{C}^{*}-algebra AA either has a unique irreducible representation up to unitary equivalence, or it has 202^{\aleph_{0}} inequivalent irreducible representations. The former condition is equivalent to AA being isomorphic to the algebra of compact operators on a separable Hilbert space, while the latter is equivalent to AA not being of type I (see [2, Theorem IV.1.5.1] for the full statement).

Parts of Glimm’s theorem were extended to non-separable C\textrm{C}^{*}-algebras by Sakai (see [2, IV.1.5.8] for a discussion). In the 1970s further progress on extending Glimm’s theorem to all simple C\textrm{C}^{*}-algebras slowed down to a halt. The most obvious question, asked by Naimark already in the 1950s, was whether a C\textrm{C}^{*}-algebra with a unique irreducible representation up to unitary equivalence is necessarily isomorphic to an algebra of compact operators. A counterexample to Naimark’s problem is a C\textrm{C}^{*}-algebra that is not isomorphic to an algebra of compact operators, yet still has only one irreducible representation up to unitary equivalence.

In a seminal paper [1], Akemann and Weaver constructed a counterexample to Naimark’s problem using Jensen’s 1\diamondsuit_{\aleph_{1}} principle111It is not known whether κ\diamondsuit_{\kappa} has any bearing on Naimark’s problem for any κ2\kappa\geq\aleph_{2}.. By related proofs, also conditioned on Jensen’s 1\diamondsuit_{\aleph_{1}} on 1\aleph_{1}, several counterexamples with additional properties (e.g., a prescribed tracial simplex [30], not isomorphic to its opposite algebra [12]) were obtained. N.C. Phillips observed that the Akemann–Weaver construction provides a nuclear C\textrm{C}^{*}-algebra. The range of applications of this construction was extended to other problems: In [12] it was shown that Glimm’s dichotomy can fail: Assuming 1\diamondsuit_{\aleph_{1}}, there exists a simple C\textrm{C}^{*}-algebra with exactly mm inequivalent irreducible representations for all m0m\leq\aleph_{0} (the latter was announced in [9, §8.2]) and a hyperfinite II1 factor not isomorphic to its opposite was constructed in [13].

The following is a special case of our Theorem 5.4.

Theorem A.

It is relatively consistent with 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} that there exists a counterexample to Naimark’s problem while 1\diamondsuit_{\aleph_{1}} fails.

More specifically, we isolate a combinatorial principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} that, together with the Continuum Hypothesis, implies the existence of a counterexample to Naimark’s problem. Then we show that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} is consistent with the failure of 1\diamondsuit_{\aleph_{1}} (see Appendix B).

A simple C\textrm{C}^{*}-algebra with a unique irreducible representation up to unitary equivalence that is represented on a Hilbert space of density character strictly smaller than 202^{\aleph_{0}} is necessarily isomorphic to an algebra of compact operators (see e.g., [10, Corollary 5.5.6]). We prove the ‘next best thing’ by constructing a separably represented counterexample to Glimm’s dichotomy. More precisely, we obtain the following:

Theorem B.

For any m2m\geq 2, 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} implies that there exists a separably represented, simple, unital C\textrm{C}^{*}-algebra with exactly mm irreducible representations up to unitary equivalence.

This is a special case of Theorem 7.1 proved below.

Our principal technical contribution is the introduction of a forcing notion that generically adds an automorphism of a given C\textrm{C}^{*}-algebra with a prescribed action on its space of pure states (see §3, §4, and §6), which constitute a generalization of a celebrated theorem of Kishimoto, Ozawa, and Sakai (see [22]).

Acknowledgments

Some of the results of this paper come from the first author’s masters thesis written under the second author’s supervision. We are indebted to Ryszard Nest and Chris Schafhauser for enlightening discussions. We also wish to thank Andrea Vaccaro and Alessandro Vignati for helpful comments on the early versions of this paper, and to Assaf Rinot for precious comments on precious stones222Diamonds. that considerably improved the presentation of the transfinite constructions in this paper.

2. Preliminaries and notation

The reader is assumed to be familiar with the basics of forcing and the basics of C\textrm{C}^{*}-algebras. Our notation and terminology follow [2] for operator algebras, [19] and [23] for set theory (in particular, forcing), and [10] for both (except forcing). It is understood that [10] is used as a reference only for the reader’s (as well as the author’s) convenience; none of the results referred to are claimed to be due to the author of [10].

2.1. C\textrm{C}^{*}-algebras and their representations

C\textrm{C}^{*}-algebras are complex Banach algebras with an involution that satisfy the C\textrm{C}^{*}-equality, aa=a2\|aa^{*}\|=\|a\|^{2}. By a result of Gelfand and Naimark, these are exactly the algebras isomorphic to a norm-closed, -closed, subalgebra of the algebra (H)\mathscr{B}(H) of bounded linear operators on a complex Hilbert space HH. A homomorphism between C\textrm{C}^{*}-algebras that preserves the involution is called a -homomorphism, and a -homomorphism into (H)\mathscr{B}(H) is a representation. A representation π:A(H)\pi\colon A\to\mathscr{B}(H) is irreducible if HH has no nontrivial closed subspaces invariant under the image of AA. A representation is faithful if it is injective. Every faithful representation is necessarily isometric. More generally, all -homomorphisms are contractive (take note that in the theory of operator algebras ‘contractive’ is synonymous with ‘1-Lipshitz.’) When AA is unital, 𝖴(A)\mathsf{U}(A) is the set of its unitary elements, i.e., those uAu\in A such that uu=uu=1Auu^{*}=u^{*}u=1_{A}.

2.2. States

A bounded linear functional φ\varphi on a C\textrm{C}^{*}-algebra AA is a state if it is positive, i.e., φ(aa)0\varphi(a^{*}a)\geq 0 for all aAa\in A, and φ=1\|\varphi\|=1. The space of all states on AA is denoted 𝖲(A)\mathsf{S}(A). Via the Gelfand–Naimark–Segal (GNS) construction (see [10, §1.10]), every state φ\varphi on AA is associated with a representation πφ:A(Hφ)\pi_{\varphi}\colon A\to\mathscr{B}(H_{\varphi}) such that a unique (up to multiplication by a scalar of modulus 1) unit vector, ξφ\xi_{\varphi} in HφH_{\varphi} satisfies that φ(a)=(πφ(a)ξφ|ξφ)\varphi(a)=(\pi_{\varphi}(a)\xi_{\varphi}|\xi_{\varphi}) for all aAa\in A. The triplet (πφ,Hφ,ξφ)(\pi_{\varphi},H_{\varphi},\xi_{\varphi}) is the GNS triplet associated with φ\varphi. Conversely, every representation (π,H)(\pi,H) of AA with a cyclic vector (i.e., some ξH\xi\in H such that the π[A]\pi[A]-orbit of ξ\xi is dense in HH) is of the form πφ\pi_{\varphi} for some state φ\varphi on 𝖲(A)\mathsf{S}(A). When AA is unital, the states space of AA is a weak-compact and convex set. The extreme points of 𝖲(A)\mathsf{S}(A) are called pure states and a state is pure if, and only if, the corresponding GNS representation is irreducible, if, and only if, πφ[A]\pi_{\varphi}[A] is dense in (Hφ)\mathscr{B}(H_{\varphi}) with respect to the weak operator topology (see [10, §3.6]). The space of all pure states on AA is denoted 𝖯(A)\mathsf{P}(A).

2.3. Automorphisms and crossed products

We will say that an automorphism Φ\Phi of a C\textrm{C}^{*}-algebra AA is inner if it is of the form Adu(a):=uau\operatorname{Ad}u(a):=uau^{*} for some unitary uu in the unitization of AA. It is approximately inner if there exists a net of unitaries (up:p𝔊)(u_{p}:p\in\mathfrak{G}) in the unitization of AA such that Φ(a)=lim𝔊Adup(a)\Phi(a)=\lim_{\mathfrak{G}}\operatorname{Ad}u_{p}(a) for each aAa\in A. The automorphism group of AA is denoted Aut(A)\operatorname{Aut}(A).

An automorphism Φ\Phi of AA determines a continuous action of \mathbb{Z} on AA given by n.a:=Φn(a)n.a:=\Phi^{n}(a). To such a non-commutative dynamical system one can associate a reduced crossed product, AΦA\rtimes_{\Phi}\mathbb{Z}. This C\textrm{C}^{*}-algebra is generated by an isomorphic copy of AA (routinely identified with AA) and a unitary uu that implements Φ\Phi on AA, in the sense that Adu(a)=Φ(a)\operatorname{Ad}u(a)=\Phi(a) for all aAa\in A. For more details see e.g., [3, §4.1] or [10, §2.4.2].

2.4. Equivalences of states and representations

We say that two representations (π0,H0)(\pi_{0},H_{0}) and (π1,H1)(\pi_{1},H_{1}) of a C\textrm{C}^{*}-algebra are spatially equivalent, and we write π0π1\pi_{0}\sim\pi_{1}, if there exists a *-isomorphism Φ:(H0)(H1)\Phi\colon\mathscr{B}(H_{0})\to\mathscr{B}(H_{1}) such that Φπ0=π1\Phi\circ\pi_{0}=\pi_{1}. Two pure states φ\varphi and ψ\psi of AA are called conjugate if there exists an automorphism Φ\Phi of AA such that φΦ=ψ\varphi\circ\Phi=\psi. If Φ\Phi can be chosen to be inner, we say that φ\varphi and ψ\psi are unitarily equivalent and write φψ\varphi\sim\psi. Using a GNS argument plus the Kadison transitivity theorem, it can be shown (see [10, Lemma 3.8.1]) that φψ\varphi\sim\psi if, and only if, πφπψ\pi_{\varphi}\sim\pi_{\psi}, if, and only if, there is a unitary uu in the unitization of AA such that φAduψ<2\|\varphi\circ\operatorname{Ad}u-\psi\|<2. This implies that Naimark’s problem as stated in the previous section is equivalent to asking whether every C\textrm{C}^{*}-algebra with a unique pure state up to unitary equivalence is isomorphic to an algebra of compact operators.

2.5. The space 𝖯m(A)\mathsf{P}_{m}(A)

Following [10, §5.6], for mm\in\mathbb{N}, the space of mm-tuples of pairwise inequivalent pure states of AA is denoted 𝖯m(A)\mathsf{P}_{m}(A). A typical element of 𝖯m(A)\mathsf{P}_{m}(A) is of the form φ¯=(φi:i<m)\bar{\varphi}=(\varphi_{i}:i<m). The automorphism group of AA acts naturally on 𝖯m(A)\mathsf{P}_{m}(A) as follows: If φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A) and ΦAut(A)\Phi\in\text{Aut}(A) then φ¯Φ:=(φiΦ:i<m)\bar{\varphi}\circ\Phi:=(\varphi_{i}\circ\Phi:i<m).

We write GAG\Subset A if GG is a finite subset of AA. If GAG\Subset A and δ>0\delta>0, we write

φ¯G,δψ¯if, and only if,maxbG(maxi<m|φi(b)ψi(b)|)<δ.\bar{\varphi}\approx_{G,\delta}\bar{\psi}\quad\text{if, and only if,}\quad\max_{b\in G}\left(\max_{i<m}|\varphi_{i}(b)-\psi_{i}(b)|\right)<\delta.

Thus,

UG,δ(φ¯):={ψ¯:ψ¯G,δφ¯}U_{G,\delta}(\bar{\varphi}):=\{\bar{\psi}:\bar{\psi}\approx_{G,\delta}\bar{\varphi}\}

is a typical weak-open neighbourhood of φ¯\bar{\varphi} in 𝖯m(A)\mathsf{P}_{m}(A). These sets range over a weak-neighbourhood basis of φ¯\bar{\varphi} in 𝖯m(A)\mathsf{P}_{m}(A) as δ\delta ranges over positive reals and GG ranges over finite subsets of AA (or finite subsets of a fixed dense subset of AA, for this use the fact that states have norm 11).

It is worth mentioning that the notation φ¯ψ¯\bar{\varphi}\sim\bar{\psi} is reserved for the existence of a unitary u𝖴(A)u\in\mathsf{U}(A) such that φ¯Adu=ψ¯\bar{\varphi}\circ\operatorname{Ad}u=\bar{\psi} (cf. Definition 4.1).

Given two tuples of pairwise inequivalent pure states φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A) and ψ¯𝖯l(A)\bar{\psi}\in\mathsf{P}_{l}(A), we denote by φ¯ψ¯\bar{\varphi}^{\frown}\bar{\psi} their concatenation (φ0,,φm1,ψ0,,ψl1)𝖯(A)m+l(\varphi_{0},\dots,\varphi_{m-1},\psi_{0},\dots,\psi_{l-1})\in\mathsf{P}(A)^{m+l}.

2.6. Type I C\textrm{C}^{*}-algebras

A C\textrm{C}^{*}-algebra AA is type I if the ideal of compact operators on HH is included in π[A]\pi[A] for every irreducible representation (π,H)(\pi,H) of AA, and non-type I if it is not type I. An example of a non-type I C\textrm{C}^{*}-algebra is the CAR algebra, M2:=M2()M_{2^{\infty}}:=\bigotimes_{\mathbb{N}}M_{2}(\mathbb{C}), and by a result due to Glimm (see [10, Theorem 3.7.2]), a C\textrm{C}^{*}-algebra is non-type I if, and only if, it has a C\textrm{C}^{*}-subalgebra which has a quotient isomorphic to M2M_{2^{\infty}}.

2.7. Transitive models of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}

Our ambient theory is 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}, the Zermelo–Fraenkel set theory with the Axiom of Choice (see e.g., [10, §A.1]). Because of meta-mathematical obstructions of no direct relevance to the present paper, while working in 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} one cannot prove the existence of a model of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}. Fortunately, for any uncountable regular cardinal κ\kappa the set H(κ)H(\kappa) of all sets whose hereditary closure has cardinality smaller than κ\kappa (see e.g., [10, §A.7]) is a model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}, the theory obtained by removing the Power Set axiom from 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}.333Purists may prefer working with transitive structures closed under the rudimentary functions, see e.g., [19, Definition 27.2]. This fragment of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} suffices for our purposes.444It may be worth mentioning that the Power Set axiom is far from being useless. By a result of Harvey Friedman, Borel Determinacy cannot be proved in 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}–its proof even requires uncountably many iterations of the power set operation–and yet it is a theorem of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}; see [24].

Borel subsets of a Polish space with a fixed countable basis can be coded by elements of \mathbb{N}^{\mathbb{N}} (see e.g., [19, p. 504]). This coding is sufficiently absolute, so that a transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}} that does not include the set of all real numbers can still contains codes for some Borel sets and be correct about their properties (the proof of [19, Lemma 25.46] applies to show this).

2.8. Forcing

A forcing notion is a partially ordered set \mathbb{P}. The elements of \mathbb{P} are also called conditions, and if pqp\leq q then pp is said to extend qq. Two conditions are compatible if a single condition extends both of them. A subset DD of \mathbb{P} is called open if it contains all extensions of all of its elements. A subset DD of \mathbb{P} is called dense if it contains some extension of every condition in \mathbb{P}. A subset 𝔊\mathfrak{G} of \mathbb{P} is a filter if it satisfies the following two conditions: (i) p𝔊p\in\mathfrak{G} and pqp\leq q implies q𝔊q\in\mathfrak{G}, and (ii) every two elements of 𝔊\mathfrak{G} have a common extension in 𝔊\mathfrak{G}. If 𝒟\mathcal{D} is a family of dense open subsets of \mathbb{P}, then a filter 𝔊\mathfrak{G} is called 𝒟\mathcal{D}-generic if it intersects every element of 𝒟\mathcal{D} non-trivially.

If MM is a transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}, \mathbb{P} is a forcing notion in MM, and a filter 𝔊\mathfrak{G}\subseteq\mathbb{P} intersects all dense open subsets of \mathbb{P} that belong to MM, then 𝔊\mathfrak{G} is said to be MM-generic. In this situation, one can define the forcing (also called generic) extension M[𝔊]M[\mathfrak{G}] which is a transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}} that includes MM and contains 𝔊\mathfrak{G}. The model MM is usually referred to as the ground model.

3. Forcing an approximately inner automorphism

Let AA be a simple and unital C\textrm{C}^{*}-algebra. Given two elements φ¯\bar{\varphi} and ψ¯\bar{\psi} of 𝖯m(A)\mathsf{P}_{m}(A), we will define a forcing notion 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), depending on a +i\mathbb{Q}+i\mathbb{Q}-subalgebra AA^{\circ} of AA, whose generic object codes an approximately inner automorphism Φ𝔊\Phi_{\mathfrak{G}} of AA such that φ¯Φ𝔊=ψ¯\bar{\varphi}\circ\Phi_{\mathfrak{G}}=\bar{\psi}.

Besides 1\diamondsuit_{\aleph_{1}}, the Akemann–Weaver construction uses a refinement of a deep 2001 result due to Kishimoto, Ozawa, and Sakai (see [22], also [10, §5.6]) that implies that all pure states of a separable, simple, and unital C\textrm{C}^{*}-algebra are conjugate by an approximately (and even asymptotically) inner automorphism. A crucial lemma in the proof of the Kishimoto–Ozawa–Sakai theorem (see [22, Lemma 2.2], also [10, Lemma 5.6.7]) together with [16, Property 7] serves as a motivation for the following lemma.

Lemma 3.1.

Let AA be a simple, unital, and infinite-dimensional C\textrm{C}^{*}-algebra. For all m1m\geq 1, φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), FAF\Subset A, and ε>0\varepsilon>0 the following holds: There exist a GAG\Subset A and δ>0\delta>0 such that for all θ¯𝖯m(A)\bar{\theta}\in\mathsf{P}_{m}(A), if φ¯G,δθ¯\bar{\varphi}\approx_{G,\delta}\bar{\theta} then for all KAK\Subset A and every γ>0\gamma>0 there exists a unitary v𝖴(A)v\in\mathsf{U}(A) such that φ¯AdvK,γθ¯\bar{\varphi}\circ\operatorname{Ad}v\approx_{K,\gamma}\bar{\theta} and bAdv(b)<ε\|b-\operatorname{Ad}v(b)\|<\varepsilon for every bFb\in F.

Proof.

We commence the proof by restating it in the language introduced in §2.5. It asserts that for every m1m\geq 1, all FAF\Subset A and ε>0\varepsilon>0, and all φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), there is a basic open neighbourhood UG,δ(φ¯)U_{G,\delta}(\bar{\varphi}) of φ¯\bar{\varphi} such that for every θ¯\bar{\theta} in this neighbourhood and every basic open neighbourhood UK,γ(θ¯)U_{K,\gamma}(\bar{\theta}) of θ¯\bar{\theta}, some v𝖴(A)v\in\mathsf{U}(A) satisfies maxbF[v,b]<ε\max_{b\in F}\|[v,b]\|<\varepsilon and φ¯AdvUK,γ(θ¯)\bar{\varphi}\circ\operatorname{Ad}v\in U_{K,\gamma}(\bar{\theta}).

Equivalently, for every m1m\geq 1, all FAF\Subset A and ε>0\varepsilon>0, and all φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), there is a basic open neighbourhood UG,δ(φ¯)U_{G,\delta}(\bar{\varphi}) of φ¯\bar{\varphi} such that the set

{φ¯Advv𝖴(A),maxbF[v,b]<ε}\{\bar{\varphi}\circ\operatorname{Ad}v\mid v\in\mathsf{U}(A),\max_{b\in F}\|[v,b]\|<\varepsilon\}

is weak-dense in UG,δ(φ¯)U_{G,\delta}(\bar{\varphi}).

[10, Lemma 5.6.7] falls just a little short of proving this, under the same assumptions on AA. In the case when AA is unital, hence A=A~A=\tilde{A}, it asserts that for every m1m\geq 1, all FAF\Subset A and ε>0\varepsilon>0, and all φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), there is a basic open neighbourhood UG,δ(φ¯)U_{G,\delta^{\prime}}(\bar{\varphi}) of φ¯\bar{\varphi} such that for every ψ¯UG,δ(φ¯)\bar{\psi}\in U_{G,\delta^{\prime}}(\bar{\varphi}) with ψ¯φ¯\bar{\psi}\sim\bar{\varphi} (see §2.4), there exists u1𝖴(A)u_{1}\in\mathsf{U}(A) such that maxbF[b,u1]<24ε\max_{b\in F}\|[b,u_{1}]\|<24\varepsilon and φ¯Adu1=ψ¯\bar{\varphi}\circ\operatorname{Ad}u_{1}=\bar{\psi}.555The unitary u1u_{1} obtained in [10, Lemma 5.6.7] is homotopic to 1A1_{A}, and all the unitaries in the homotopy path satisfy maxbF[b,ut]<24ε\max_{b\in F}\|[b,u_{t}]\|<24\varepsilon, but we don’t need this.

Hence, in order to complete the proof, it will suffice to replace ε\varepsilon with ε/24\varepsilon/24 and show that the set {ψ¯ψ¯φ¯}\{\bar{\psi}\mid\bar{\psi}\sim\bar{\varphi}\} is dense in UG,δU_{G,\delta^{\prime}}. Since AA is simple and infinite-dimensional, every representation π:A(H)\pi\colon A\to\mathscr{B}(H) satisfies π[A]𝒦(H)={0}\pi[A]\cap\mathscr{K}(H)=\{0\}. It is a standard consequence of this condition and Glimm’s Lemma that for every ψ¯\bar{\psi} in 𝖯m(A)\mathsf{P}_{m}(A), the unitary orbit {ψ¯Adu:u𝖴(A)}\{\bar{\psi}\circ\operatorname{Ad}u:u\in\mathsf{U}(A)\} is dense in 𝖯m(A)\mathsf{P}_{m}(A) (see e.g., [10, Proposition 5.2.9]). This completes the proof. ∎

The previous lemma motivates the following definition.

Definition 3.2.

Given a C\textrm{C}^{*}-algebra AA and m1m\geq 1, let φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), FAF\Subset A, and ε>0\varepsilon>0. We will say that a pair (G,δ)(G,\delta) with GAG\Subset A and δ>0\delta>0 is (φ¯,F,ε)(\bar{\varphi},F,\varepsilon)-good if for all θ¯𝖯m(A)\bar{\theta}\in\mathsf{P}_{m}(A), if φ¯G,δθ¯\bar{\varphi}\approx_{G,\delta}\bar{\theta} then for all KAK\Subset A and every γ>0\gamma>0 there exists a unitary u𝖴(A)u\in\mathsf{U}(A) such that φ¯AduK,γθ¯\bar{\varphi}\circ\operatorname{Ad}u\approx_{K,\gamma}\bar{\theta} and bAdu(b)<ε\|b-\operatorname{Ad}u(b)\|<\varepsilon for every bFb\in F.

Analogously to the restatement of Lemma 3.1 given in the first paragraph of its proof, one obtains the following.

Lemma 3.3.

For m1m\geq 1, FAF\Subset A, ε>0\varepsilon>0, and φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), a pair (G,δ)(G,\delta) with GAG\Subset A and δ>0\delta>0 is (φ¯,F,ε)(\bar{\varphi},F,\varepsilon)-good if and only if the set

{φ¯Advv𝖴(A),maxbF[v,b]<ε}\{\bar{\varphi}\circ\operatorname{Ad}v\mid v\in\mathsf{U}(A),\max_{b\in F}\|[v,b]\|<\varepsilon\}

is weak-dense in UG,δ(φ¯)U_{G,\delta}(\bar{\varphi}). ∎

Thus we have the following equivalent reformulation of Lemma 3.1.

Lemma 3.4.

If AA is a simple, unital, infinite-dimensional C\textrm{C}^{*}-algebra, then for all m1m\geq 1, φ¯𝖯m(A)\bar{\varphi}\in\mathsf{P}_{m}(A), FAF\Subset A, and ε>0\varepsilon>0, there exists a (φ¯,F,ε)(\bar{\varphi},F,\varepsilon)-good pair (G,δ)(G,\delta). ∎

Fix now a simple and unital C\textrm{C}^{*}-algebra AA, and tuples φ¯\bar{\varphi} and ψ¯\bar{\psi} in 𝖯m(A)\mathsf{P}_{m}(A). We also fix a norm-dense +i\mathbb{Q}+i\mathbb{Q}-subalgebra AA^{\circ} of AA with minimal cardinality such that 𝖴(A)A\mathsf{U}(A)\cap A^{\circ} is norm-dense in 𝖴(A)\mathsf{U}(A). In particular, when AA is separable, AA^{\circ} will be countable.

In Definition 3.5, we will introduce a forcing notion 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) which generically adds an automorphism Φ𝔊\Phi_{\mathfrak{G}} of AA such that φ¯Φ𝔊=ψ¯\bar{\varphi}\circ\Phi_{\mathfrak{G}}=\bar{\psi}. More precisely, it adds two nets of unitaries, vpv_{p} and wpw_{p}, for p𝔊p\in\mathfrak{G}, such that each of the nets of Advp\operatorname{Ad}v_{p} and Adwp\operatorname{Ad}w_{p}^{*} indexed by p𝔊p\in\mathfrak{G} converges pointwise to an automorphism of AA (this is assured by condition (c)). The automorphism Φ𝔊\Phi_{\mathfrak{G}} is the composition of these two automorphisms.

Definition 3.5.

Let 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) be the set of tuples

q=(Fq,Gq,εq,δq,vq,wq)q=(F_{q},G_{q},\varepsilon_{q},\delta_{q},v_{q},w_{q})

such that:

  1. (1)

    FqF_{q} and GqG_{q} are finite subsets of AA^{\circ},

  2. (2)

    εq\varepsilon_{q} and δq\delta_{q} are positive real numbers,

  3. (3)

    vqv_{q} and wqw_{q} are unitaries of AA in AA^{\circ},

  4. (4)

    (Gq,δq)(G_{q},\delta_{q}) is a (φ¯Advq,FqAdvq[Fq],εq/3)(\bar{\varphi}\circ\operatorname{Ad}v_{q},F_{q}\cup\operatorname{Ad}v_{q}^{*}[F_{q}],\varepsilon_{q}/3)-good pair, and

  5. (5)

    φ¯AdvqGq,δqψ¯Adwq\bar{\varphi}\circ\operatorname{Ad}v_{q}\approx_{G_{q},\delta_{q}}\bar{\psi}\circ\operatorname{Ad}w_{q}.

We order 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) by pqp\leq q if:

  1. (a)

    FpFqF_{p}\supseteq F_{q}, GpGqG_{p}\supseteq G_{q},

  2. (b)

    εpεq\varepsilon_{p}\leq\varepsilon_{q}, δpδq\delta_{p}\leq\delta_{q}, and

  3. (c)

    for all bFqb\in F_{q}

    max{Advp(b)Advq(b),Advp(b)Advq(b)}\displaystyle\max\left\{\left\|\operatorname{Ad}v_{p}(b)-\operatorname{Ad}v_{q}(b)\right\|,\left\|\operatorname{Ad}v_{p}^{*}(b)-\operatorname{Ad}v_{q}^{*}(b)\right\|\right\} εqεp, and\displaystyle\leq\varepsilon_{q}-\varepsilon_{p},\text{ and}
    max{Adwp(b)Adwq(b),Adwp(b)Adwq(b)}\displaystyle\max\left\{\left\|\operatorname{Ad}w_{p}(b)-\operatorname{Ad}w_{q}(b)\right\|,\left\|\operatorname{Ad}w_{p}^{*}(b)-\operatorname{Ad}w_{q}^{*}(b)\right\|\right\} εqεp.\displaystyle\leq\varepsilon_{q}-\varepsilon_{p}.
Remark 3.6.

The bound εqεp\varepsilon_{q}-\varepsilon_{p} in (c) of Definition 3.5 is used to assure that the relation \leq is transitive on 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}). This idea is taken from [14].

Lemma 3.7.

For all finite subsets FF and GG of AA^{\circ}, and all positive real numbers ε\varepsilon and δ\delta, the set D(F,G,ε,δ)D(F,G,\varepsilon,\delta) of conditions pp such that FFpF\subseteq F_{p}, GGpG\subseteq G_{p}, εpε\varepsilon_{p}\leq\varepsilon, and δpδ\delta_{p}\leq\delta is a dense and open666See §2.8 for definitions. subset of 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}).

Proof.

It is obvious that an extension of every condition in D(F,G,ε,δ)D(F,G,\varepsilon,\delta) belongs to D(F,G,ε,δ)D(F,G,\varepsilon,\delta).

Second, we need to prove that every condition has an extension that belongs to D(F,G,ε,δ)D(F,G,\varepsilon,\delta). Let q𝔼A(φ¯,ψ¯)q\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) be fixed but arbitrary. By Lemma 3.1, there exists a (ψ¯Adwq,FqAdwq[Fq],εq)(\bar{\psi}\circ\operatorname{Ad}w_{q},F_{q}\cup\operatorname{Ad}w_{q}^{*}[F_{q}],\varepsilon_{q})-good pair; denote it (G1,δ1)(G_{1},\delta_{1}). Since qq is a condition, (Gq,δq)(G_{q},\delta_{q}) is (φ¯Advq,FqAdvq[Fq],εq)(\bar{\varphi}\circ\operatorname{Ad}v_{q},F_{q}\cup\operatorname{Ad}v_{q}^{*}[F_{q}],\varepsilon_{q})-good and φ¯AdvqGq,δqψ¯Adwq\bar{\varphi}\circ\operatorname{Ad}v_{q}\approx_{G_{q},\delta_{q}}\bar{\psi}\circ\operatorname{Ad}w_{q}.

Using the goodness of (Gq,δq)(G_{q},\delta_{q}), choose v𝖴(A)Av\in\mathsf{U}(A)\cap A^{\circ} to be some unitary such that φ¯AdvqvG1,δ1ψ¯Adwq\bar{\varphi}\circ\operatorname{Ad}v_{q}v\approx_{G_{1},\delta_{1}}\bar{\psi}\circ\operatorname{Ad}w_{q} and bAdv(b)<εq/3\|b-\operatorname{Ad}v(b)\|<\varepsilon_{q}/3 for all bFqAdvq[Fq]b\in F_{q}\cup\operatorname{Ad}v_{q}^{*}[F_{q}]. Set εp:=min{ε,εq/3}\varepsilon_{p}:=\min\{\varepsilon,\varepsilon_{q}/3\}, Fp:=FqFF_{p}:=F_{q}\cup F, and vp:=vqvv_{p}:=v_{q}v.

By Lemma 3.1, there is a (φ¯Advp,FpAdvp[Fp],εp)(\bar{\varphi}\circ\operatorname{Ad}v_{p},F_{p}\cup\operatorname{Ad}v_{p}^{*}[F_{p}],\varepsilon_{p})-good pair, denoted (G2,δ2)(G_{2},\delta_{2}). Let Gp:=GGqG2G_{p}:=G\cup G_{q}\cup G_{2} and δp:=min{δ,δq,δ2}\delta_{p}:=\min\{\delta,\delta_{q},\delta_{2}\}.

Using now the fact that the pair (G1,δ1)(G_{1},\delta_{1}) is a good pair, let w𝖴(A)Aw\in\mathsf{U}(A)\cap A^{\circ} be such that φ¯AdvpGp,δpψ¯Adwqw\bar{\varphi}\circ\operatorname{Ad}v_{p}\approx_{G_{p},\delta_{p}}\bar{\psi}\circ\operatorname{Ad}w_{q}w and bAdw(b)<εq/3\|b-\operatorname{Ad}w(b)\|<\varepsilon_{q}/3 for every bFqAdwq[Fq]b\in F_{q}\cup\operatorname{Ad}w_{q}^{*}[F_{q}]. Define wpw_{p} as wqww_{q}w and set pp to be (Fp,Gp,εp,δp,vp,wp)(F_{p},G_{p},\varepsilon_{p},\delta_{p},v_{p},w_{p}).

Clearly FFpF\subseteq F_{p}, GGpG\subseteq G_{p}, εpε\varepsilon_{p}\leq\varepsilon and δpδ\delta_{p}\leq\delta. Also, since the pair (Gp,δp)(G_{p},\delta_{p}) is (φ¯Advp,FpAdvp[Fp],εp)(\bar{\varphi}\circ\operatorname{Ad}v_{p},F_{p}\cup\operatorname{Ad}v_{p}^{*}[F_{p}],\varepsilon_{p})-good, p𝔼A(φ¯,ψ¯)p\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}). Finally, if bFqb\in F_{q} then

Advq(b)Advp(b)\displaystyle\|\operatorname{Ad}v_{q}^{*}(b)-\operatorname{Ad}v_{p}^{*}(b)\| =Adv(Advq(b))Advq(b)\displaystyle=\|\operatorname{Ad}v(\operatorname{Ad}v_{q}^{*}(b))-\operatorname{Ad}v_{q}^{*}(b)\|
<εq/3<εqεq/3εqεp.\displaystyle<\varepsilon_{q}/3<\varepsilon_{q}-\varepsilon_{q}/3\leq\varepsilon_{q}-\varepsilon_{p}.

Also, Advq(b)Advp(b)=bAdv(b)<εqεp\|\operatorname{Ad}v_{q}(b)-\operatorname{Ad}v_{p}(b)\|=\|b-\operatorname{Ad}v(b)\|<\varepsilon_{q}-\varepsilon_{p}. The calculations for wpw_{p} and wqw_{q} are analogous and therefore pqp\leq q. ∎

Two forcing notions 0\mathbb{P}_{0} and 1\mathbb{P}_{1} are said to be forcing-equivalent if they give rise to the same generic extensions, i.e., every generic extension M[G]M[G] by 0\mathbb{P}_{0} can be realized as a generic extension M[H]M[H] for some MM-generic filter HH on 1\mathbb{P}_{1}, and vice versa. The simplest mechanism for assuring forcing equivalence is by finding an isomorphic copy of 0\mathbb{P}_{0} in 1\mathbb{P}_{1} that is also dense (in the forcing sense, see §2.8) in it. (A proof of this fact is contained in [23, Lemma III.3.68 (β\beta)]. The only point is that if \mathbb{P} is a dense subordering of 1\mathbb{P}_{1} then a filter 𝔊1\mathfrak{G}\subseteq\mathbb{P}_{1} is generic over MM if and only if 𝔊\mathfrak{G}\cap\mathbb{P} is generic over MM.) It is straightforward to use this observation to prove that every two countable atomless (i.e., with no minimal elements) forcing notions are equivalent (see e.g., the hint for [23, Exercise III.3.70]).

Theorem 3.8.

Let AA be a simple and unital C\textrm{C}^{*}-algebra, m1m\geq 1 and let φ¯\bar{\varphi} and ψ¯\bar{\psi} be elements of 𝖯m(A)\mathsf{P}_{m}(A). Then

  1. (1)

    Forcing with 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) adds an approximately inner automorphism Φ𝔊\Phi_{\mathfrak{G}} of AA such that φ¯Φ𝔊=ψ¯\bar{\varphi}\circ\Phi_{\mathfrak{G}}=\bar{\psi}.

  2. (2)

    If AA is separable, then 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) is equivalent to the Cohen forcing.

Proof.

(1) Let MM be a countable transitive model of a large enough fragment of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} such that 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) is an element of MM, and let 𝔊\mathfrak{G} be an MM-generic filter on 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}). By Lemma 3.7, for any FAF\Subset A, and for all ε>0\varepsilon>0, there exists some q𝔼A(φ¯,ψ¯)q\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) such that if pqp\leq q then for all bFb\in F, Advp(b)Advq(b)<ε\|\operatorname{Ad}v_{p}(b)-\operatorname{Ad}v_{q}(b)\|<\varepsilon and Advp(b)Advq(b)<ε\|\operatorname{Ad}v_{p}^{*}(b)-\operatorname{Ad}v_{q}^{*}(b)\|<\varepsilon. Therefore, the nets Advp\operatorname{Ad}v_{p} and Advp\operatorname{Ad}v_{p}^{*}, for p𝔊p\in\mathfrak{G}, are Cauchy with respect to the point-norm topology in Aut(A)\text{Aut}(A). By [10, Lemma 2.6.3], ΦLAut(A)\Phi_{L}\in\text{Aut}(A) defined pointwise as ΦL(a):=lim𝔊Advp(a)\Phi_{L}(a):=\lim_{\mathfrak{G}}\operatorname{Ad}v_{p}(a) is an endomorphism of AA, and its inverse is given by ΦL1(a)=lim𝔊Advp(a)\Phi_{L}^{-1}(a)=\lim_{\mathfrak{G}}\operatorname{Ad}v_{p}^{*}(a) for each aAa\in A. An analogous argument shows that ΦR(a):=lim𝔊Adwp(a)\Phi_{R}(a):=\lim_{\mathfrak{G}}\operatorname{Ad}w_{p}(a) for each aAa\in A is an approximately inner automorphism of AA.

Let now aAa\in A be arbitrary and let ε>0\varepsilon>0. Again using Lemma 3.7, choose p𝔊p\in\mathfrak{G} such that some bFpGpb\in F_{p}\cap G_{p} satisfies ab<ε/3\|a-b\|<\varepsilon/3 and that max{εp,δp}<ε/3\max\{\varepsilon_{p},\delta_{p}\}<\varepsilon/3. Then |φ¯ΦL(a)ψ¯ΦR(a)|<ε|\bar{\varphi}\circ\Phi_{L}(a)-\bar{\psi}\circ\Phi_{R}(a)|<\varepsilon. Set Φ𝔊:=ΦLΦR1\Phi_{\mathfrak{G}}:=\Phi_{L}\circ\Phi_{R}^{-1}. Since ε\varepsilon was arbitrary, φ¯Φ𝔊=ψ¯\bar{\varphi}\circ\Phi_{\mathfrak{G}}=\bar{\psi}.

(2) Since AA is separable, AA^{\circ} is countable. Also, the conditions such that both εp\varepsilon_{p} and δp\delta_{p} are rational, comprise a dense subset of 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}). This set is countable, and by [20, Proposition 10.20], 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) is equivalent to the Cohen forcing. ∎

We should point out that the C\textrm{C}^{*}-algebra AA in Theorem 3.8 is not required to be separable. If AA is (for example) the C\textrm{C}^{*}-algebra with pure states of characters 0\aleph_{0} and an uncountable κ\kappa (such AA exists by [10, Proposition 2.9.4]) then the existence of an automorphism as guaranteed by Theorem 3.8 in a generic extension implies that κ\kappa is collapsed to 0\aleph_{0}. Thus for some nonseparable C\textrm{C}^{*}-algebras the forcing 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) may collapse cardinals.

The idea of restricting to a countable dense set in order to assure the countable chain condition of a poset used in the proof of Theorem 3.8 was first used in the context of operator algebras in [32].

4. The Unique Extension Property of pure states

The most remarkable property of the generic automorphism introduced by the forcing notion 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) is the genericity of its action on 𝖯(A)\mathsf{P}(A) (see Theorem 4.4).

Let AA be a simple, unital, and non-type I C\textrm{C}^{*}-algebra. We will see later on (see Proposition A.4) that after forcing with 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), new equivalence classes of pure states of AA will appear. The aim of this section is to study the ground-model pure states of AA that have a unique pure state extension to AΦ𝔊A\rtimes_{\Phi_{\mathfrak{G}}}\mathbb{Z} in M[𝔊]M[\mathfrak{G}]. In order to assure that pure states of AA in a prescribed set have unique pure extensions to the crossed product mentioned above, we use the tools introduced in [1] and presented in a gory detail in [10, §5.4].

By Theorem 3.1 in [21], if AA is simple and Φ\Phi is outer, then AΦA\rtimes_{\Phi}\mathbb{Z} is simple as well. By Theorem 2 in [1] (see also [10, Proposition 5.4.7]), a pure state φ\varphi on AA has a unique extension to a pure state on AΦA\rtimes_{\Phi}\mathbb{Z} if, and only if, φ\varphi is not equivalent to φΦn\varphi\circ\Phi^{n} for all n1n\geq 1. Since the set of all extensions of φ\varphi is a face in 𝖲(AΦ)\mathsf{S}(A\rtimes_{\Phi}\mathbb{Z}) (see [10, Lemma 5.4.1]), φ\varphi has a unique extension if, and only if, it has a unique pure state extension. If two pure states φ\varphi and ψ\psi on AA have unique extensions φ~\tilde{\varphi} and ψ~\tilde{\psi} to AΦA\rtimes_{\Phi}\mathbb{Z}, then these extensions are equivalent if, and only if, φ\varphi is equivalent to ψΦn\psi\circ\Phi^{n} for some nn\in\mathbb{Z} (this is the case Γ=\Gamma=\mathbb{Z} of [10, Theorem 5.4.8]).

In the following, we use the notation established in Theorem 3.8. Note that for a fixed AA and AA^{\circ}, the conditions in all forcings 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) have the same format and that φ¯\bar{\varphi} and ψ¯\bar{\psi} behave as side-conditions. We will now relate these forcing notions.

Definition 4.1.

Let AA be a unital C\textrm{C}^{*}-algebra, m1m\geq 1, and let φ¯\bar{\varphi} and ψ¯\bar{\psi} be elements of 𝖯m(A)\mathsf{P}_{m}(A). We will say that φ¯\bar{\varphi} is pointwise unitarily equivalent to ψ¯\bar{\psi}, in symbols φ¯pψ¯\bar{\varphi}\sim_{p}\bar{\psi}, if there exists a tuple (u0,,um1)𝖴(A)m(u_{0},\dots,u_{m-1})\in\mathsf{U}(A)^{m} such that for each i<mi<m, we have that φ¯iAdui=ψ¯i\bar{\varphi}_{i}\circ\operatorname{Ad}u_{i}=\bar{\psi}_{i}.

Lemma 4.2.

Suppose AA is a simple and unital C\textrm{C}^{*}-algebra, m1m\geq 1, and φ¯\bar{\varphi} and ψ¯\bar{\psi} are elements of 𝖯m(A)\mathsf{P}_{m}(A). Also suppose that l0l\geq 0, ρ¯\bar{\rho} and σ¯\bar{\sigma} belong to 𝖯l(A)\mathsf{P}_{l}(A), and moreover φ¯ρ¯\bar{\varphi}^{\frown}\bar{\rho} and ψ¯σ¯\bar{\psi}^{\frown}\bar{\sigma} belong to 𝖯m+l(A)\mathsf{P}_{m+l}(A). Let us write 0:=𝔼A(φ¯,ψ¯)\mathbb{P}_{0}:=\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), 1:=𝔼A(φ¯ρ¯,ψ¯σ¯)\mathbb{P}_{1}:=\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\bar{\rho},\bar{\psi}^{\frown}\bar{\sigma}), and j\leq_{j} for the ordering on j\mathbb{P}_{j} for j<2j<2. Then

  1. (1)

    Every condition in 1\mathbb{P}_{1} is a condition in 0\mathbb{P}_{0}. Moreover, if pp and qq are in 1\mathbb{P}_{1} then p0qp\leq_{0}q if, and only if, p1qp\leq_{1}q.

  2. (2)

    For every q0q\in\mathbb{P}_{0} there exists some ρ¯pρ¯\bar{\rho}^{\prime}\sim_{p}\bar{\rho} such that some p0qp\leq_{0}q belongs to 1:=𝔼A(φ¯ρ¯,ψ¯σ¯)\mathbb{P}_{1}^{\prime}:=\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\bar{\rho}^{\prime},\bar{\psi}^{\frown}\bar{\sigma}).777In short, 1\mathbb{P}_{1} is a subordering of 0\mathbb{P}_{0} and the union of all 1\mathbb{P}_{1}^{\prime} as in (2) is dense in 0\mathbb{P}_{0}. This formulation is dangerously misleading, since 1\mathbb{P}_{1} is typically not a regular subordering of 0\mathbb{P}_{0}.

Proof.

To see that the first part of (1) holds, fix q1q\in\mathbb{P}_{1}. Conditions (1)–(3) of Definition 3.5 do not depend on the tuples of pure states, while (4) and (5) are weakened as one passes to sub-tuples of pure states. The second part of (1) follows because conditions (a)–(c) of Definition 3.5 do not refer to the pure states.

(2) Fix q0q\in\mathbb{P}_{0}. We will assume that each of ρ¯\bar{\rho} and σ¯\bar{\sigma} consists of a single pure state, ρ\rho and σ\sigma respectively. Once this is proved, the general case will follow by induction. By Lemma 3.1, there exists a ((ψ¯σ)Adwq,FqAdwq[Fq],εq)((\bar{\psi}^{\frown}\sigma)\circ\operatorname{Ad}w_{q},F_{q}\cup\operatorname{Ad}w_{q}^{*}[F_{q}],\varepsilon_{q})-good pair; denote it (G1,δ1)(G_{1},\delta_{1}). Since qq is a condition, (Gq,δq)(G_{q},\delta_{q}) is (φ¯Advq,FqAdvq[Fq],εq)(\bar{\varphi}\circ\operatorname{Ad}v_{q},F_{q}\cup\operatorname{Ad}v_{q}^{*}[F_{q}],\varepsilon_{q})-good, and

φ¯AdvqGq,δqψ¯Adwq.\bar{\varphi}\circ\operatorname{Ad}v_{q}\approx_{G_{q},\delta_{q}}\bar{\psi}\circ\operatorname{Ad}w_{q}.

Using the goodness of (Gq,δq)(G_{q},\delta_{q}), choose v𝖴(A)Av\in\mathsf{U}(A)\cap A^{\circ} to be some unitary such that φ¯AdvqvG1,δ1ψ¯Adwq\bar{\varphi}\circ\operatorname{Ad}v_{q}v\approx_{G_{1},\delta_{1}}\bar{\psi}\circ\operatorname{Ad}w_{q} and bAdv(b)<εq/3\|b-\operatorname{Ad}v(b)\|<\varepsilon_{q}/3 for all bFqAdvq[Fq]b\in F_{q}\cup\operatorname{Ad}v_{q}^{*}[F_{q}]. Set εp:=εq/3\varepsilon_{p}:=\varepsilon_{q}/3, Fp:=FqF_{p}:=F_{q}, and vp:=vqvv_{p}:=v_{q}v. Apply now [10, Proposition 5.2.9] to find some unitary u𝖴(A)Au\in\mathsf{U}(A)\cap A^{\circ} such that if ρ:=ρAdu\rho^{\prime}:=\rho\circ\operatorname{Ad}u, then

(φ¯ρ)AdvpG1,δ1(ψ¯σ)Adwq.(\bar{\varphi}^{\frown}\rho^{\prime})\circ\operatorname{Ad}v_{p}\approx_{G_{1},\delta_{1}}(\bar{\psi}^{\frown}\sigma)\circ\operatorname{Ad}w_{q}.

By Lemma 3.1, there is a ((φ¯ρ)Advp,FpAdvp[Fp],εp)((\bar{\varphi}^{\frown}\rho^{\prime})\circ\operatorname{Ad}v_{p},F_{p}\cup\operatorname{Ad}v_{p}^{*}[F_{p}],\varepsilon_{p})-good pair, denoted (G2,δ2)(G_{2},\delta_{2}). Also, let Gp:=GqG2G_{p}:=G_{q}\cup G_{2} and δp:=min{δq,δ2}\delta_{p}:=\min\{\delta_{q},\delta_{2}\}.

Using now the fact that the pair (G1,δ1)(G_{1},\delta_{1}) is a good pair, let w𝖴(A)Aw\in\mathsf{U}(A)\cap A^{\circ} be such that (φ¯ρ)AdvpGp,δp(ψ¯σ)Adwqw(\bar{\varphi}^{\frown}\rho^{\prime})\circ\operatorname{Ad}v_{p}\approx_{G_{p},\delta_{p}}(\bar{\psi}^{\frown}\sigma)\circ\operatorname{Ad}w_{q}w and bAdw(b)<εq/3\|b-\operatorname{Ad}w(b)\|<\varepsilon_{q}/3 for every bFqAdwq[Fq]b\in F_{q}\cup\operatorname{Ad}w_{q}^{*}[F_{q}]. Define wpw_{p} as wqww_{q}w and set pp to be (Fp,Gp,εp,δp,vp,wp)(F_{p},G_{p},\varepsilon_{p},\delta_{p},v_{p},w_{p}).

Since the pair (Gp,δp)(G_{p},\delta_{p}) is ((φ¯ρ)Advp,FpAdvp[Fp],εp)((\bar{\varphi}^{\frown}\rho^{\prime})\circ\operatorname{Ad}v_{p},F_{p}\cup\operatorname{Ad}v_{p}^{*}[F_{p}],\varepsilon_{p})-good, p1p\in\mathbb{P}_{1}. Also, note that if bFqb\in F_{q} then

Advq(b)Advp(b)\displaystyle\|\operatorname{Ad}v_{q}^{*}(b)-\operatorname{Ad}v_{p}^{*}(b)\| =Adv(Advq(b))Advq(b)\displaystyle=\|\operatorname{Ad}v(\operatorname{Ad}v_{q}^{*}(b))-\operatorname{Ad}v_{q}^{*}(b)\|
<εq/3<εqεq/3εqεp.\displaystyle<\varepsilon_{q}/3<\varepsilon_{q}-\varepsilon_{q}/3\leq\varepsilon_{q}-\varepsilon_{p}.

Also, Advq(b)Advp(b)=bAdv(b)<εqεp\|\operatorname{Ad}v_{q}(b)-\operatorname{Ad}v_{p}(b)\|=\|b-\operatorname{Ad}v(b)\|<\varepsilon_{q}-\varepsilon_{p}. The calculations for wpw_{p} and wqw_{q} are analogous and therefore p0qp\leq_{0}q. ∎

Remark 4.3.

We will follow an abuse of terminology common in set theory and refer to ‘reals’ as elements of any uncountable Polish space fixed in advance that is definable in the sense that it belongs to every model of (a sufficiently large fragment of) 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}. This is justified by a classical result of Kuratowski, asserting that any two uncountable Polish spaces are Borel-isomorphic. Using properties of the standard coding for Borel sets (see e.g., [19, p. 504]) one sees that the property of being a Borel-isomorphism between two Polish spaces is absolute between transitive models of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}. This has as a consequence the fact that a forcing notion adds a new element to \mathbb{R} if, and only if, it adds a new element to some (every) uncountable Polish space in the ground model. Thus the phrase ‘\mathbb{P} does not add new reals’ is unambiguous, even with the gratuitous use of the phrase ‘the reals’.

A forcing notion \mathbb{P} adds a new real to the model MM if, and only if, it adds a new element to some (equivalently, every) non-trivial C\textrm{C}^{*}-algebra. Because of this, in a forcing extension M[𝔊]M[\mathfrak{G}] we identify AA with its completion, and pure states of AA with their unique continuous extensions to the completion of AA. As common in set theory, by AMA^{M} we denote the original C\textrm{C}^{*}-algebra AA in MM and by AM[𝔊]A^{M[\mathfrak{G}]} we denote its completion in M[𝔊]M[\mathfrak{G}] (see Appendix A). Note that AMA^{M} is an element of M[𝔊]M[\mathfrak{G}] which is an algebra over the field M\mathbb{C}^{M}. In the extension, the latter is a proper subfield of M[𝔊]\mathbb{C}^{M[\mathfrak{G}]}, hence the former is not a complex algebra, hence not a C\textrm{C}^{*}-algebra. It is however dense in AM[𝔊]A^{M[\mathfrak{G}]}, which suffices for our purposes.

Theorem 4.4.

Suppose AA is a simple, unital, separable, and non-type I C\textrm{C}^{*}-algebra, and let φ¯\bar{\varphi} and ψ¯\bar{\psi} be elements of 𝖯m(A)\mathsf{P}_{m}(A). If ρ\rho is a pure state on AA, then some condition in 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) forces that ρΦ𝔊n\rho\circ\Phi^{n}_{\mathfrak{G}} is equivalent to a ground-model pure state σ\sigma for some n1n\geq 1 if, and only if, n=1n=1 and there exists i<mi<m such that ρφi\rho\sim\varphi_{i} and σψi\sigma\sim\psi_{i}.

Proof of Theorem 4.4.

The converse implication is the conclusion of Theorem 3.8.

The direct implication uses the well-known fact that pure states η\eta and σ\sigma on AA that are inequivalent in a model of a large enough fragment of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} remain inequivalent in every forcing extension (and even in every larger transitive model of a large enough fragment of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}). A proof of this fact is included below.

Suppose that the direct implication is false for some nn, ρ\rho, and σ\sigma. Choose the minimal such nn. Note that if ρφi\rho\sim\varphi_{i} for some ii then necessarily σψi\sigma\sim\psi_{i}, and similarly if σψi\sigma\sim\psi_{i} then ρφi\rho\sim\varphi_{i}. Since AA is separable, by Glimm’s theorem ([10, Corollary 5.5.5]) it has 202^{\aleph_{0}} inequivalent pure states, if n>1n>1 then we can choose mutually inequivalent pure states ρ1,,ρn1\rho_{1},\dots,\rho_{n-1} each of which is inequivalent to all φi\varphi_{i}, ψi\psi_{i}, ρ\rho, and σ\sigma. Therefore, by our assumption, (writing ρ0:=ρ\rho_{0}:=\rho and ρ¯:=(ρ0,ρ1,,ρn1)\bar{\rho}:=(\rho_{0},\rho_{1},\dots,\rho_{n-1})) both φ¯ρ¯\bar{\varphi}^{\frown}\bar{\rho} and ψ(ρ1,,ρn1,σ)\psi^{\frown}(\rho_{1},\dots,\rho_{n-1},\sigma) belong to 𝖯m+n(A)\mathsf{P}_{m+n}(A).

Fix q𝔼A(φ¯,ψ¯)q\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) which forces that ρΦ𝔊nσ\rho\circ\Phi^{n}_{\mathfrak{G}}\sim\sigma. By the minimality of nn, we can extend qq so that it forces ρΦ𝔊j≁σ\rho\circ\Phi^{j}_{\mathfrak{G}}\not\sim\sigma for all 1j<n1\leq j<n. Since the ground-model 𝖴(A)\mathsf{U}(A) is dense in 𝖴(A)\mathsf{U}(A) of the generic extension, by further extending qq, we may assume that there exists u𝖴(A)Au\in\mathsf{U}(A)\cap A^{\circ} such that qρΦ𝔊nσAdu<1/2q\Vdash\|\rho\circ\Phi^{n}_{\mathfrak{G}}-\sigma\circ\operatorname{Ad}u\|<1/2. Using the abundance of inequivalent pure states of AA again, let ησ\eta\nsim\sigma be such that if η¯:=(ρ1,,ρn1,η)\bar{\eta}:=(\rho_{1},\dots,\rho_{n-1},\eta), then ψ¯η¯\bar{\psi}^{\frown}\bar{\eta} belongs to 𝖯m+n(A)\mathsf{P}_{m+n}(A). By Lemma 4.2 there exists ρ¯=(ρ0,ρ1,ρn1)\bar{\rho}^{\prime}=(\rho^{\prime}_{0},\rho^{\prime}_{1},\dots\rho^{\prime}_{n-1}) in 𝖯n(A)\mathsf{P}_{n}(A) such that ρ¯pρ¯\bar{\rho}^{\prime}\sim_{p}\bar{\rho} and some p𝔼A(φ¯ρ¯,ψ¯η¯)p\in\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\bar{\rho}^{\prime},\bar{\psi}^{\frown}\bar{\eta}) extends qq.

Let \mathfrak{H} be an MM-generic filter on 𝔼A(φ¯ρ¯,ψ¯η¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\bar{\rho}^{\prime},\bar{\psi}^{\frown}\bar{\eta}) containing pp. By Theorem 3.8, in M[]M[\mathfrak{H}] we have that ρ¯Φ=η¯\bar{\rho}^{\prime}\circ\Phi_{\mathfrak{H}}=\bar{\eta}. Let v𝖴(A)v\in\mathsf{U}(A) be such that ρ=ρ0Adv\rho=\rho^{\prime}_{0}\circ\operatorname{Ad}v and set v:=Φn(v)v_{\mathfrak{H}}:=\Phi_{\mathfrak{H}}^{-n}(v). Since AdvΦn=ΦnAdv\operatorname{Ad}v\circ\Phi_{\mathfrak{H}}^{n}=\Phi_{\mathfrak{H}}^{n}\circ\operatorname{Ad}v_{\mathfrak{H}}, we have

ρΦn=ρ0AdvΦn=ρ0ΦnAdv.\rho\circ\Phi_{\mathfrak{H}}^{n}=\rho^{\prime}_{0}\circ\operatorname{Ad}v\circ\Phi_{\mathfrak{H}}^{n}=\rho^{\prime}_{0}\circ\Phi_{\mathfrak{H}}^{n}\circ\operatorname{Ad}v_{\mathfrak{H}}.

Since σ\sigma is not equivalent to ηAdv\eta\circ\operatorname{Ad}v_{\mathfrak{H}}, by [10, Proposition 3.8.1] some aA1Ma\in A^{M}_{\leq 1} satisfies

|(ηAdv)(a)(σAdu)(a)|2.\left|\left(\eta\circ\operatorname{Ad}v_{\mathfrak{H}}\right)(a)-\left(\sigma\circ\operatorname{Ad}u\right)(a)\right|\geq 2.

Since AMA^{M} is norm-dense in AM[]A^{M[\mathfrak{H}]}, let r𝔼A(φ¯ρ¯,ψ¯η¯)r\in\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\bar{\rho}^{\prime},\bar{\psi}^{\frown}\bar{\eta}), with rpr\leq p, be such that some bGrb\in G_{r} satisfies bAdv(a)<1/6\|b-\operatorname{Ad}v_{\mathfrak{H}}(a)\|<1/6, and δr<1/6\delta_{r}<1/6. The (easy) first part of Lemma 4.2 implies that r𝔼A(φ¯,ψ¯)r\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) and that rr extends qq as an element of 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}). However, rr forces (in either of the posets) that

|(ρΦn)(a)(ηAdv)(a)|\displaystyle\left|\left(\rho\circ\Phi_{\mathfrak{H}}^{n}\right)(a)-\left(\eta\circ\operatorname{Ad}v_{\mathfrak{H}}\right)(a)\right| =|(ρ0Φn)(Adv(a))η(Adv(a))|\displaystyle=\left|(\rho^{\prime}_{0}\circ\Phi_{\mathfrak{H}}^{n})(\operatorname{Ad}v_{\mathfrak{H}}(a))-\eta(\operatorname{Ad}v_{\mathfrak{H}}(a))\right|
|(ρ0Φn)(b)η(b)|+2bAdv(a)<1/2.\displaystyle\leq\left|(\rho^{\prime}_{0}\circ\Phi_{\mathfrak{H}}^{n})(b)-\eta(b)\right|+2\|b-\operatorname{Ad}v_{\mathfrak{H}}(a)\|<1/2.

By [10, Proposition 3.8.1], ρΦn\rho\circ\Phi_{\mathfrak{H}}^{n} is equivalent to η\eta, and hence η\eta and σ\sigma are equivalent in the extension. But η\eta and σ\sigma are inequivalent ground model pure states. Since 𝖴(A)M\mathsf{U}(A)^{M} is dense in 𝖴(A)M[]\mathsf{U}(A)^{M[\mathfrak{H}]}, we can choose v𝖴(A)Mv\in\mathsf{U}(A)^{M} such that ηAdvσ<2\|\eta\circ\operatorname{Ad}v-\sigma\|<2. By using [10, Proposition 3.8.1] again we conclude that η\eta and σ\sigma are equivalent in MM; contradiction. ∎

In the following corollary there is no need to explicitly refer to the ground model.

Corollary 4.5.

Suppose that AA is a simple, unital, and non-type I C\textrm{C}^{*}-algebra. Fix elements φ¯\bar{\varphi} and ψ¯\bar{\psi} of 𝖯m(A)\mathsf{P}_{m}(A), and fix ρ\rho and σ\sigma in 𝖯(A)\mathsf{P}(A). If 𝔊\mathfrak{G} is a generic filter in 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), then the following statements hold in the forcing extension:

  1. (1)

    ρ\rho has multiple pure state extensions to AΦ𝔊A\rtimes_{\Phi_{\mathfrak{G}}}\mathbb{Z} if, and only if, there exists i<mi<m such that ρφiψi\rho\sim\varphi_{i}\sim\psi_{i}.

  2. (2)

    If ρ\rho and σ\sigma are inequivalent and they both have unique pure state extensions to AΦ𝔊A\rtimes_{\Phi_{\mathfrak{G}}}\mathbb{Z} then these extensions are equivalent if, and only if, some i<mi<m satisfies ρφi\rho\sim\varphi_{i} and σψi\sigma\sim\psi_{i} or ρψi\rho\sim\psi_{i} and σφi\sigma\sim\varphi_{i}.

Proof.

By a result from [1] (or see [10, Proposition 5.4.7]), a pure state ζ\zeta has a unique pure state extension to the crossed product if, and only if, ζζΦ𝔊n\zeta\nsim\zeta\circ\Phi_{\mathfrak{G}}^{n} for all n0n\neq 0. By Theorem 3.8 and Theorem 4.4, this happens if, and only if, ρφiψi\rho\sim\varphi_{i}\sim\psi_{i} for some i<mi<m. This proves (1).

In (2) only the direct implication requires a proof. It uses the case when Γ=\Gamma=\mathbb{Z} of [10, Theorem 5.4.8]. This theorem asserts that with an action α\alpha of a discrete group Γ\Gamma on AA, two pure states ρ\rho and σ\sigma on AA have equivalent extensions to the reduced crossed product if and only if some gg in Γ\Gamma satisfies ραgσ\rho\circ\alpha_{g}\sim\sigma. Therefore if ρ\rho and σ\sigma have unique and equivalent extensions to the reduced crossed product, then 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), forces that ρσΦn\rho\sim\sigma\circ\Phi^{n} for some nn\in\mathbb{Z}. Since ρ\rho and σ\sigma are inequivalent, they remain inequivalent in the forcing extension and therefore n0n\neq 0. By Theorem 4.4, if n>0n>0 then n=1n=1, ρφi\rho\sim\varphi_{i}, and σψi\sigma\sim\psi_{i} for some ii. By the same theorem, if n<0n<0 then n=1n=-1, ρψi\rho\sim\psi_{i}, and σφi\sigma\sim\varphi_{i}. This exhausts the possibilities and concludes the proof. ∎

5. A proof of Theorem A from 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}

In this section, we introduce our weakening of Jensen’s 1\diamondsuit_{\aleph_{1}} principle, 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}, and use it to construct the C\textrm{C}^{*}-algebra as required in Theorem A.

A subset of 1\aleph_{1} is closed and unbounded (club) if it is unbounded and contains the supremum of each of its bounded subsets.888In other words, it is unbounded and closed in the ordinal topology. A subset of 1\aleph_{1} is stationary if it intersects every club non-trivially.

Let us first recall Jensen’s 1\diamondsuit_{\aleph_{1}} principle.

Definition 5.1.

A 1\diamondsuit_{\aleph_{1}} sequence is an indexed family of sets SααS_{\alpha}\subseteq\alpha, for α<1\alpha<\aleph_{1}, such that for every X1X\subseteq\aleph_{1} the set {αXα=Sα}\{\alpha\mid X\cap\alpha=S_{\alpha}\} is stationary. The 1\diamondsuit_{\aleph_{1}} principle asserts that a 1\diamondsuit_{\aleph_{1}} sequence exists.

It is well known that 1\diamondsuit_{\aleph_{1}} is relatively consistent with 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}, that it implies 𝖢𝖧\operatorname{\mathsf{CH}}, and that it is not a consequence of 𝖢𝖧\operatorname{\mathsf{CH}} (see [23]).

Definition 5.2.

A chain (Mα:α<1)(M_{\alpha}:\alpha<\aleph_{1}) is a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain if:

  1. 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(a)

    Each MαM_{\alpha} is a (not necessarily countable) transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}.

  2. 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(b)

    For every X1X\subseteq\aleph_{1}, the set {α<1:XαMα}\{\alpha<\aleph_{1}:X\cap\alpha\in M_{\alpha}\} is stationary.

  3. 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(c)

    For every α<1\alpha<\aleph_{1}, some real in Mα+1M_{\alpha+1} is Cohen-generic over MαM_{\alpha}.

We say that the principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} holds if there exists a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain.

Clearly, 1\diamondsuit_{\aleph_{1}} implies 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}. (To see this, fix a 1\diamondsuit_{\aleph_{1}} sequence SαS_{\alpha}, for α<1\alpha<\aleph_{1}, and choose an increasing chain of countable elementary submodels (Nα)(N_{\alpha}) of H(2)H(\aleph_{2}) such that SαNαS_{\alpha}\in N_{\alpha} for all α\alpha, and assure that Nα+1N_{\alpha+1} contains a real which is Cohen over NαN_{\alpha}; this is possible because NαN_{\alpha} is countable. Take MαM_{\alpha} to be the transitive collapse of NαN_{\alpha}.) For a partial converse, note that if each model MαM_{\alpha} in a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain is countable, or if its intersection with 2α2^{\alpha} is countable, then 1\diamondsuit_{\aleph_{1}} holds. This is a consequence of [23, Theorem III.7.8]. We will discuss the relative consistency of 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} in Appendix B.

Let us now discuss the relevance of the condition 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(b).

First, observe that it implies that α<1Mα\bigcup_{\alpha<\aleph_{1}}M_{\alpha} contains all subsets of ω\omega, and therefore all elements of any Polish space (affectionately known as ‘reals’—see Remark 4.3 for both the terminology and the proof).

Second, this condition applies when XX is replaced with any object of cardinality 1\aleph_{1} or with a complete metric space of density character 1\aleph_{1}. More specifically, every C\textrm{C}^{*}-algebra of density character κ\kappa can be coded by a subset of κ\kappa (see the introduction to §3 in [12], or [10, §7.1-2]). Similarly, if ψ¯\bar{\psi} is a tuple of states of a C\textrm{C}^{*}-algebra BB of density character κ\kappa, then the structure (B,ψ¯)(B,\bar{\psi}) can be coded by a subset of κ\kappa. Moreover, the version of Löwenheim–Skolem theorem for logic of metric structures stated in [10, Theorem 7.1.4] implies the following.

Lemma 5.3.

Suppose that κ\kappa is a regular and uncountable cardinal, A=limα<κAαA=\varinjlim_{\alpha<\kappa}A_{\alpha} is a C\textrm{C}^{*}-algebra such that the density character of each AαA_{\alpha} is strictly smaller than κ\kappa, Aβ=limα<βAαA_{\beta}=\varinjlim_{\alpha<\beta}A_{\alpha} for every limit ordinal β\beta, φ¯\bar{\varphi} is a tuple of states of AA, and XκX\subseteq\kappa is a code for the structure (A,φ¯)(A,\bar{\varphi}). Then the set

{α<κ:Xα is a code for (Aα,φ¯Aα)}\left\{\alpha<\kappa:X\cap\alpha\text{ is a code for }(A_{\alpha},\bar{\varphi}\upharpoonright A_{\alpha})\right\}

includes a club.∎

If MM is a transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}} then we slighty abuse the terminology and say that a C\textrm{C}^{*}-algebra BB belongs to MM if some code for BB belongs to MM. The analogous remark applies to states of BB.

Glimm’s dichotomy (see e.g. [10, Corollary 5.5.8]) implies that every simple C\textrm{C}^{*}-algebra AA of density character κ<20\kappa<2^{\aleph_{0}} either has a unique pure state up to unitary equivalence, in which case A𝒦(2(κ))A\cong\mathscr{K}(\ell_{2}(\kappa)), or has 202^{\aleph_{0}} unitary equivalence classes of pure states. The conclusion of the following theorem was deduced from 1\diamondsuit_{\aleph_{1}} in [12, Theorem 1.2], as announced in [9, §8.2] (see also [10, Theorem 11.2.2]). The special case when m=1m=1 (using the full 1\diamondsuit_{\aleph_{1}}) is the Akemann–Weaver result.

The case when m=1m=1 of Theorem 5.4 below is Theorem A. In its proof we adopt the approach to 1\diamondsuit_{\aleph_{1}} constructions introduced in [26].

Theorem 5.4.

If 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} holds, then for all m1m\geq 1 there is a simple C\textrm{C}^{*}-algebra of density character 1\aleph_{1} with exactly mm pure states up to unitary equivalence that is not isomorphic to any algebra of compact operators on a complex Hilbert space.

Proof.

Let (Mα:α<1)(M_{\alpha}:\alpha<\aleph_{1}) be a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain. Using the Continuum Hypothesis, fix a surjection f:1H(1)f\colon\aleph_{1}\to H(\aleph_{1}) such that every element of H(1)H(\aleph_{1}) is listed cofinally often. By recursion on β<1\beta<\aleph_{1}, we will define an inductive system of separable, simple, unital, and non-type I C\textrm{C}^{*}-algebras AβA_{\beta}.

Let A0A_{0} be a separable, simple, unital, non-type I C\textrm{C}^{*}-algebra and let φi\varphi_{i}, for i<mi<m, be inequivalent pure states of A0A_{0}. At the latter stages of the construction we will assure that the following conditions hold for all α<1\alpha<\aleph_{1}.

  1. (1)

    If ξ<α\xi<\alpha then AξA_{\xi} is a unital C\textrm{C}^{*}-subalgebra of AαA_{\alpha}.

  2. (2)

    With γ(α):=min{γ:AαMγ}\gamma(\alpha):=\min\{\gamma:A_{\alpha}\in M_{\gamma}\}, Aα+1A_{\alpha+1} belongs to Mγ(α)+1M_{\gamma(\alpha)+1}.999This function is well-defined: Since AαA_{\alpha} is separable, it is coded by a real and therefore belongs to α<1Mα\bigcup_{\alpha<\aleph_{1}}M_{\alpha}.

  3. (3)

    Every pure state of AαA_{\alpha} that belongs to Mγ(α)M_{\gamma(\alpha)} has a unique pure state extension to Aα+1A_{\alpha+1}.

  4. (4)

    If f(α)f(\alpha) is a code for a pair (Aξ,ψ)(A_{\xi},\psi), where ξ<α\xi<\alpha and ψ\psi is a pure state of AξA_{\xi} which has a unique pure state extension to AαA_{\alpha}, then ψ\psi has a unique pure state extension to Aα+1A_{\alpha+1}, and this extension is equivalent to (the unique pure state extension of) some φi\varphi_{i}, for i<mi<m.

To describe the recursive construction, suppose that β\beta is a countable ordinal such that AαA_{\alpha} as required has been defined for all α<β\alpha<\beta.

Consider first the case when β\beta is a successor ordinal, β=α+1\beta=\alpha+1. Suppose for a moment that f(α)f(\alpha) is a code for a pair (Aξ,ψ)(A_{\xi},\psi) with the following properties:

  1. (a)

    ξ<α\xi<\alpha.

  2. (b)

    ψ\psi is a pure state of AξA_{\xi} that has a unique extension ψ~\tilde{\psi} to a pure state of AαA_{\alpha}.

  3. (c)

    For all i<mi<m, ψ~\tilde{\psi} is inequivalent to the unique extension of φi\varphi_{i} to AαA_{\alpha} (still denoted φi\varphi_{i}).

By the second part of Theorem 3.8, 𝔼Aα(φ0,ψ~)\mathbb{E}_{A^{\circ}_{\alpha}}(\varphi_{0},\tilde{\psi}) is forcing-equivalent to the poset for adding a single Cohen real. Since Mγ(α)+1M_{\gamma(\alpha)+1} contains a real that is Cohen-generic over Mγ(α)M_{\gamma(\alpha)}, by [23, Lemma IV.4.7], it contains an Mγ(α)M_{\gamma(\alpha)}-generic filter 𝔊\mathfrak{G} on 𝔼Aα(φ0,ψ~)\mathbb{E}_{A^{\circ}_{\alpha}}(\varphi_{0},\tilde{\psi}). By the first part of Theorem 3.8, Φ𝔊\Phi_{\mathfrak{G}} is an approximately inner automorphism of AαA_{\alpha} such that φ0Φ𝔊=ψ~\varphi_{0}\circ\Phi_{\mathfrak{G}}=\tilde{\psi}. By Corollary 4.5, the C\textrm{C}^{*}-algebra Aα+1:=AαΦ𝔊A_{\alpha+1}:=A_{\alpha}\rtimes_{\Phi_{\mathfrak{G}}}\mathbb{Z} has the property that every pure state of AαA_{\alpha} that belongs to Mγ(α)M_{\gamma(\alpha)} has a unique pure state extension to Aα+1A_{\alpha+1}. By the second part of Corollary 4.5, the unique pure state extensions of φi\varphi_{i}, for i<mi<m, to Aα+1A_{\alpha+1} are inequivalent. Also, Aα+1A_{\alpha+1} is separable, simple, unital, and non-type I by Corollary A.2.

If f(α)f(\alpha) does not satisfy the conditions (a)–(c), let Aα+1:=AαA_{\alpha+1}:=A_{\alpha}.

If β\beta is a limit ordinal, take Aβ:=limα<βAαA_{\beta}:=\varinjlim_{\alpha<\beta}A_{\alpha}.

Finally, let A1:=limα<1AαA_{\aleph_{1}}:=\varinjlim_{\alpha<\aleph_{1}}A_{\alpha}.

By the construction described above, each one of the the pure states φi\varphi_{i}, for i<mi<m, of A0A_{0} has a unique pure state extension to A1A_{\aleph_{1}}, and these pure state extensions are inequivalent.

Suppose that ψ\psi is a pure state of A1A_{\aleph_{1}}. In order to prove that it is equivalent to φi\varphi_{i} for some i<mi<m, fix a code X1X\subseteq\aleph_{1} for the pair (A1,ψ)(A_{\aleph_{1}},\psi). By [10, Proposition 7.3.10], the set {α<1:ψAα is pure}\{\alpha<\aleph_{1}:\psi\upharpoonright A_{\alpha}\text{ is pure}\} is a club, and by Lemma 5.3, the set

{α<1:Xα is a code for (Aα,ψAα)}\{\alpha<\aleph_{1}:X\cap\alpha\text{ is a code for }(A_{\alpha},\psi\upharpoonright A_{\alpha})\}

is a club as well. By 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(b), there exists some α\alpha in the intersection of these two clubs such that XαMαX\cap\alpha\in M_{\alpha}. In particular, both AαA_{\alpha} and ψAα\psi\upharpoonright A_{\alpha} belong to MαM_{\alpha}, i.e., γ(α)=α\gamma(\alpha)=\alpha.

Claim.

The pure state ψAα\psi\upharpoonright A_{\alpha} of AαA_{\alpha} has a unique pure state extension to AβA_{\beta} for every β>α\beta>\alpha.

Proof.

This is proved by induction on β\beta: At the successor stages, it follows from properties (2) and (3). At the limits, note that the unique pure state extension of ψAα\psi\upharpoonright A_{\alpha} is definable from ψAξ\psi\upharpoonright A_{\xi}, for α<ξ<β\alpha<\xi<\beta, and therefore belongs to the relevant model. ∎

By the choice of the function ff, there exists some β<1\beta<\aleph_{1} such that f(β)f(\beta) codes the pair (Aα,ψAα(A_{\alpha},\psi\upharpoonright{A_{\alpha}}). By the definition of Aβ+1A_{\beta+1}, the restrictions of ψ\psi and some φi\varphi_{i}, for i<mi<m, to Aβ+1A_{\beta+1} are equivalent. This proves that A1A_{\aleph_{1}} has exactly mm inequivalent pure states. Since A1A_{\aleph_{1}} is infinite-dimensional and unital, it is not isomorphic to any algebra of compact operators. ∎

The proof of Theorem A will be completed in Appendix B. In this section, we will prove that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} is relatively consistent with the negation of 1\diamondsuit_{\aleph_{1}}. Once proven, this will provide a model of 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} in which both 1\diamondsuit_{\aleph_{1}} and Glimm’s dichotomy fail.

6. A proof of Theorem B, part I: Extending a GNS representation gently

This section contains finer analysis of the forcing notion 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), culminating in Lemma 6.5. The following is an analog of Theorem 4.4.

Theorem 6.1.

Suppose that Θ\Theta is an outer automorphism of a separable, simple, unital, non-type I C\textrm{C}^{*}-algebra AA, m1m\geq 1, φ¯\bar{\varphi} and ψ¯\bar{\psi} belong to 𝖯m(A)\mathsf{P}_{m}(A), and ρ\rho is a pure state of AA inequivalent to all φi\varphi_{i}, for i<mi<m. If Θ𝔊\Theta_{\mathfrak{G}} is defined as Φ𝔊ΘΦ𝔊1\Phi_{\mathfrak{G}}\circ\Theta\circ\Phi_{\mathfrak{G}}^{-1}, then 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) forces that ρΘ𝔊\rho\circ\Theta_{\mathfrak{G}} is inequivalent to any ground-model pure state of AA.

Proof.

Towards obtaining a contradiction, assume that in M[𝔊]M[\mathfrak{G}] we have ρΘ𝔊σ\rho\circ\Theta_{\mathfrak{G}}\sim\sigma for a ground-model pure state σ\sigma. Then fix u𝖴(A)Au\in\mathsf{U}(A)\cap A^{\circ} and q𝔼A(φ¯,ψ¯)q\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) such that qq forces that ρΘ𝔊σAdu<1/2\|\rho\circ\Theta_{\mathfrak{G}}-\sigma\circ\operatorname{Ad}u\|<1/2. This implies that

ρΦ𝔊ΘσAduΦ𝔊<1/2.\|\rho\circ\Phi_{\mathfrak{G}}\circ\Theta-\sigma\circ\operatorname{Ad}u\circ\Phi_{\mathfrak{G}}\|<1/2.

We first consider the most difficult case, when ρ\rho is equivalent to σ\sigma.

Since Θ\Theta is outer, by the slight extension of [21, Theorem 2.1] proved in [12, Theorem 2.4], there exists an uncountable set of pure states η\eta of AA each of which satisfies ηΘη\eta\circ\Theta\nsim\eta. We can therefore choose a pure state η\eta such that ηηΘ\eta\nsim\eta\circ\Theta and each one of η\eta and ηΘ\eta\circ\Theta is inequivalent to all ψi\psi_{i}, for i<mi<m. By Lemma 4.2, there are a pure state ρρ\rho^{\prime}\sim\rho and a condition pqp\leq q in 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) such that pp also belongs to the forcing notion 𝔼A(φ¯ρ,ψ¯η)\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\rho^{\prime},\bar{\psi}^{\frown}\eta).

The remainder is analogous to the corresponding part of the proof of Theorem 4.4: if \mathfrak{H} is an MM-generic filter on 𝔼A(φ¯ρ,ψ¯η)\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\rho^{\prime},\bar{\psi}^{\frown}\eta) containing pp then, in M[]M[\mathfrak{H}], ρΦ=η\rho^{\prime}\circ\Phi_{\mathfrak{H}}=\eta. Let v𝖴(A)v\in\mathsf{U}(A) be such that ρ=ρAdv\rho=\rho^{\prime}\circ\operatorname{Ad}v and set v:=Φ1(v)v_{\mathfrak{H}}:=\Phi_{\mathfrak{H}}^{-1}(v), so that ρΦ=ρAdvΦ=ρΦAdv\rho\circ\Phi_{\mathfrak{H}}=\rho^{\prime}\circ\operatorname{Ad}v\circ\Phi_{\mathfrak{H}}=\rho^{\prime}\circ\Phi_{\mathfrak{H}}\circ\operatorname{Ad}v_{\mathfrak{H}}. Since ηηΘ\eta\nsim\eta\circ\Theta, we have that ηAdvΘηAdv\eta\circ\operatorname{Ad}v_{\mathfrak{H}}\circ\Theta\nsim\eta\circ\operatorname{Ad}v_{\mathfrak{H}}. Because of this, we can find aA1Ma\in A_{\leq 1}^{M} such that

|(ηAdvAdu)(a)(ηAdvΘ)(a)|3/2.\left|\left(\eta\circ\operatorname{Ad}v_{\mathfrak{H}}\circ\operatorname{Ad}u\right)(a)-\left(\eta\circ\operatorname{Ad}v_{\mathfrak{H}}\circ\Theta\right)(a)\right|\geq 3/2.

Since AMA^{M} is norm-dense in AM[]A^{M[\mathfrak{H}]}, there is a condition r𝔼A(φ¯ρ,ψ¯η)r\in\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\rho^{\prime},\bar{\psi}^{\frown}\eta) extending pp, with δr<1/6\delta_{r}<1/6, and such that some bb and cc in GrG_{r} satisfy

max{bAdvu(a),cAdv(Θ(a))}<1/6.\max\left\{\|b-\operatorname{Ad}v_{\mathfrak{H}}u(a)\|,\|c-\operatorname{Ad}v_{\mathfrak{H}}(\Theta(a))\|\right\}<1/6.

By the easy part of Lemma 4.2, r𝔼A(φ¯,ψ¯)r\in\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) and it extends qq in this poset. From the choice of rr, we can conclude that, in 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), rr forces both

|(ρΦAdu)(a)(η\displaystyle|(\rho\circ\Phi_{\mathfrak{H}}\circ\operatorname{Ad}u)(a)-(\eta\circ AdvAdu)(a)|\displaystyle\operatorname{Ad}v_{\mathfrak{H}}\circ\operatorname{Ad}u)(a)|
=|(ρΦ)(Advu(a))η(Advu(a))|\displaystyle=|(\rho^{\prime}\circ\Phi_{\mathfrak{H}})(\operatorname{Ad}v_{\mathfrak{H}}u(a))-\eta(\operatorname{Ad}v_{\mathfrak{H}}u(a))|
|(ρΦ)(b)η(b)|+2bAdvu(a)<1/2\displaystyle\leq|(\rho^{\prime}\circ\Phi_{\mathfrak{H}})(b)-\eta(b)|+2\|b-\operatorname{Ad}v_{\mathfrak{H}}u(a)\|<1/2

and

|(ρΦΘ)(a)(η\displaystyle|(\rho\circ\Phi_{\mathfrak{H}}\circ\Theta)(a)-(\eta AdvΘ)(a)|\displaystyle\circ\operatorname{Ad}v_{\mathfrak{H}}\circ\Theta)(a)|
=|(ρΦ)(Adv(Θ(a)))η(Adv(Θ(a)))|\displaystyle=|(\rho^{\prime}\circ\Phi_{\mathfrak{H}})(\operatorname{Ad}v_{\mathfrak{H}}(\Theta(a)))-\eta(\operatorname{Ad}v_{\mathfrak{H}}(\Theta(a)))|
|(ρΦ)(c)η(c)|+2cAdv(Θ(a))<1/2.\displaystyle\leq|(\rho^{\prime}\circ\Phi_{\mathfrak{H}})(c)-\eta(c)|+2\|c-\operatorname{Ad}v_{\mathfrak{H}}(\Theta(a))\|<1/2.

By the triangle inequality and our choice of aa, we obtain 1/2+1/2>3/21/2+1/2>3/2; contradiction. This concludes the discussion of the case when ρσ\rho\sim\sigma.

Suppose now that ρσ\rho\nsim\sigma. As in the first case, in each of the two subcases of this case (corresponding to (1) and (2) below), we will use Lemma 4.2 to define a forcing notion \mathbb{P} and a condition pqp\leq q in 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) that also belongs to \mathbb{P}:

  1. (1)

    If σφi\sigma\sim\varphi_{i} for some i<mi<m, choose a pure state ζ\zeta that is not equivalent to any of the ψj\psi_{j} and such that in addition ζΘψi\zeta\circ\Theta\nsim\psi_{i}. This is possible because AA has 202^{\aleph_{0}} inequivalent pure states. By Lemma 4.2, there are is a ρρ\rho^{\prime}\sim\rho and a condition pqp\leq q in the poset :=𝔼A(φ¯ρ,ψ¯ζ)\mathbb{P}:=\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\rho^{\prime},\bar{\psi}^{\frown}\zeta).

  2. (2)

    If σ\sigma is not equivalent to any of the φi\varphi_{i}, choose two pure states, ζ\zeta and η\eta, that are not equivalent to any of the ψj\psi_{j} and such that in addition ζΘη\zeta\circ\Theta\nsim\eta. By Lemma 4.2, there are pure states ρρ\rho^{\prime}\sim\rho, σσ\sigma^{\prime}\sim\sigma, and a condition pqp\leq q in the poset :=𝔼A(φ¯ρσ,ψ¯ζη)\mathbb{P}:=\mathbb{E}_{A^{\circ}}(\bar{\varphi}^{\frown}\rho^{\prime\frown}\sigma^{\prime},\bar{\psi}^{\frown}\zeta^{\frown}\eta).

In each of these two cases the proof that the assumptions lead to a contradiction is analogous to the proof in the case when ρσ\rho\sim\sigma and is therefore omitted. ∎

Corollary 6.2.

Suppose that AA is a separable, simple, unital C\textrm{C}^{*}-algebra, Θ\Theta is an outer automorphism of AA of order two, m1m\geq 1, φ¯\bar{\varphi} and ψ¯\bar{\psi} belong to 𝖯m(A)\mathsf{P}_{m}(A), and ρ\rho is a pure state of AA inequivalent to all the φi\varphi_{i}, for i<mi<m. If Θ𝔊:=Φ𝔊ΘΦ𝔊1\Theta_{\mathfrak{G}}:=\Phi_{\mathfrak{G}}\circ\Theta\circ\Phi_{\mathfrak{G}}^{-1}, then ρ\rho has multiple pure state extensions to AΘ𝔊/2A\rtimes_{\Theta_{\mathfrak{G}}}\mathbb{Z}/2\mathbb{Z} if, and only if, there exists some i<mi<m such that ρφiψi\rho\sim\varphi_{i}\sim\psi_{i}.

Proof.

As in the proof of Corollary 4.5, by Theorem 3.8 only the direct implication requires a proof. We prove its contrapositive, whose proof is analogous to the proof of Corollary 4.5. Assume that there is no ii such that ρφiψi\rho\sim\varphi_{i}\sim\psi_{i}. Then Theorem 6.1 (with σ=ρ\sigma=\rho) implies that ρΘ𝔊\rho\circ\Theta_{\mathfrak{G}} is inequivalent to ρ\rho. Since Θ𝔊\Theta_{\mathfrak{G}} has order two, [10, Theorem 5.4.8] implies that ρ\rho has a unique pure state extension to the reduced crossed product. ∎

In order to prove Theorem B, we need to take a closer look at the inner workings of the GNS construction (see [10, §1.10]). If φ\varphi is a state on a C\textrm{C}^{*}-algebra AA, then it defines a sesquilinear form on AA by (a|b)φ:=φ(ba)(a|b)_{\varphi}:=\varphi(b^{*}a). The completion of AA with respect to this norm is a Hilbert space 2(A,φ)\ell_{2}(A,\varphi) (denoted by HφH_{\varphi} in [10] and §2.2), and the representation πφ\pi_{\varphi} is defined by the left multiplication. If AA is a C\textrm{C}^{*}-subalgebra of BB and φ~\tilde{\varphi} is a state on BB that extends φ\varphi, then 2(A,φ)\ell_{2}(A,\varphi) is naturally identified with a closed subspace of 2(B,φ~)\ell_{2}(B,\tilde{\varphi}).

Lemma 6.3.

Let AA be a unital C\textrm{C}^{*}-algebra with an outer automorphism Φ\Phi such that Φn=idA\Phi^{n}=\operatorname{id}_{A} for some n2n\geq 2. Let B:=AΦ/nB:=A\rtimes_{\Phi}\mathbb{Z}/n\mathbb{Z}, and let uBu\in B be the unitary such that Φ(a)=Adu(a)\Phi(a)=\operatorname{Ad}u(a) for every aAa\in A. Suppose that φ\varphi is a state (not necessarily pure) on AA such that φ=φΦ\varphi=\varphi\circ\Phi and that ψ𝖲(B)\psi\in\mathsf{S}(B) is an extension of φ\varphi satisfying ψ(u)n=1\psi(u)^{n}=1. Then the following conclusions hold.

  1. (1)

    For every a0,,an1Aa_{0},\dots,a_{n-1}\in A, we have that

    ψ(j<najuj)=j<nψ(u)jφ(aj),\psi\left(\sum_{j<n}a_{j}u^{j}\right)=\sum_{j<n}\psi(u)^{j}\varphi(a_{j}),

    hence ψ\psi is uniquely determined by φ\varphi and ψ(u)\psi(u).

  2. (2)

    2(A,φ)=2(B,ψ)\ell_{2}(A,\varphi)=\ell_{2}(B,\psi).

  3. (3)

    If φ\varphi is pure, so is ψ\psi.

Proof.

We first prove that if ψ(u)\psi(u) belongs to the unit circle then φΦ=φ\varphi\circ\Phi=\varphi. For every aAa\in A, we have Φ(a)=uau\Phi(a)=uau^{*}. Since ψ\psi is a state, it is self-adjoint, hence ψ(u)=ψ(u)¯\psi(u^{*})=\overline{\psi(u)}. Finally, since the spectrum of uu is included in 𝕋\mathbb{T}, ψ(u)\psi(u) is an extreme point of this spectrum, and therefore [10, Propositon 1.7.8] applied twice implies

φ(Φ(a))=ψ(uau)=ψ(u)ψ(au)=ψ(u)ψ(a)ψ(u)=ψ(a)ψ(u)ψ(u)¯=φ(a).\varphi(\Phi(a))=\psi(uau^{*})=\psi(u)\psi(au^{*})=\psi(u)\psi(a)\psi(u^{*})=\psi(a)\psi(u)\overline{\psi(u)}=\varphi(a).

(1) Since ψ(u)\psi(u) is an element of the unit circle and uu is a unitary, by [10, Proposition 1.7.8], for every continuous, complex-valued function ff defined on the unit circle and every aAa\in A we have that ψ(af(u))=f(ψ(u))ψ(a)=f(ψ(u))φ(a)\psi(af(u))=f(\psi(u))\psi(a)=f(\psi(u))\varphi(a). Therefore ψ(uj)=ψ(u)j\psi(u^{j})=\psi(u)^{j} and ψ(auj)=ψ(u)jφ(a)\psi(au^{j})=\psi(u)^{j}\varphi(a) for all jj and all aAa\in A. Thus (1) follows by the additivity of ψ\psi. For use in the proof of 2, we note that, since Φn=idA\Phi^{n}=\operatorname{id}_{A}, every element of the crossed product BB is of the form j<najuj\sum_{j<n}a_{j}u^{j},

(2) By the GNS construction (see [10, §1.10]), to prove that 2(A,φ)=2(B,ψ)\ell_{2}(A,\varphi)=\ell_{2}(B,\psi), it is enough to prove that for every aAa\in A and every k<nk<n, there exists some aAa^{\prime}\in A such that aaukψ=0\|a^{\prime}-au^{k}\|_{\psi}=0. Once proven, this will imply that a dense subspace of 2(B,ψ)\ell_{2}(B,\psi) is included in 2(A,φ)\ell_{2}(A,\varphi), and therefore the spaces coincide. We claim that a:=ψ(u)kaa^{\prime}:=\psi(u)^{k}a is as desired:

ψ(u)kaaukψ2\displaystyle\left\|\psi(u)^{k}a-au^{k}\right\|^{2}_{\psi} =ψ((ψ(u)kaauk)(ψ(u)kaauk))\displaystyle=\psi\left(\left(\psi(u)^{k}a-au^{k}\right)^{*}\left(\psi(u)^{k}a-au^{k}\right)\right)
=ψ((ψ(u)kuk)aa(ψ(u)kuk)).\displaystyle=\psi\left(\left(\psi(u)^{-k}-u^{-k}\right)a^{*}a\left(\psi(u)^{k}-u^{k}\right)\right).

Now, (ψ(u)kuk)aa(ψ(u)kuk)aa(ψ(u)kuk)(ψ(u)kuk)\left(\psi(u)^{-k}-u^{-k}\right)a^{*}a\left(\psi(u)^{k}-u^{k}\right)\leq\|a^{*}a\|\left(\psi(u)^{-k}-u^{-k}\right)\left(\psi(u)^{k}-u^{k}\right), and a simple calculation using ψ(uk)=ψ(u)k\psi(u^{k})=\psi(u)^{k} proven in (1) gives

ψ((ψ(u)kuk)(ψ(u)kuk))=0.\psi(\left(\psi(u)^{-k}-u^{-k}\right)\left(\psi(u)^{k}-u^{k}\right))=0.

Since ψ\psi is positive, by the earlier calculation this implies ψ(u)kaaukψ2=0\left\|\psi(u)^{k}a-au^{k}\right\|^{2}_{\psi}=0 as required.

(3) If φ\varphi is pure, then an extension ψ\psi of φ\varphi is pure if, and only if, it is an extreme point of the set :={θ𝖲(B):θ extends φ}\mathcal{E}:=\{\theta\in\mathsf{S}(B):\theta\text{ extends }\varphi\} by [10, Lemma 5.4.1]. Let g:g\colon\mathcal{E}\to\mathbb{C} be given by g(θ):=θ(u)g(\theta):=\theta(u). Then g[]g[\mathcal{E}] is included in the convex closure of the spectrum of uu, co(sp(u))\textrm{co}(\textrm{sp}(u)). Since gg is an affine and weak-continuous function on a closed face of 𝖲(B)\mathsf{S}(B), g[]g[\mathcal{E}] is a compact, convex subset of co(sp(u))\textrm{co}(\textrm{sp}(u)), and the gg-preimage of a face of g[]g[\mathcal{E}] is a face of \mathcal{E}. But, being on the unit circle, ψ(u)\psi(u) is an extreme point of g[]g[\mathcal{E}], and (1) implies that ψ\psi is uniquely determined by ψ(u)\psi(u). Therefore g1({ψ(u)})g^{-1}(\{\psi(u)\}) is a singleton. Its unique element ψ\psi is an extreme point of \mathcal{E}, and therefore a pure state by [10, Lemma 5.4.1]. ∎

Lemma 6.5 below is based on [12, Lemma 2.7]. The key property of the CAR algebra M2M_{2^{\infty}} used in it is extracted in the following lemma implicit in [12].

Lemma 6.4.

There are inequivalent pure states ρj\rho_{j}, σj\sigma_{j}, ηj\eta_{j} for jj\in\mathbb{N} on M2M_{2^{\infty}} and an automorphism Θ\Theta of M2M_{2^{\infty}} of order two such that the following conditions hold.

  1. (1)

    σj=ρjΘ\sigma_{j}=\rho_{j}\circ\Theta and ηj=ηjΘ\eta_{j}=\eta_{j}\circ\Theta for all jj.

  2. (2)

    M2Θ/2M_{2^{\infty}}\rtimes_{\Theta}\mathbb{Z}/2\mathbb{Z} is isomorphic to M2M_{2^{\infty}}.

Proof.

Identify M2M_{2^{\infty}} with An\bigotimes_{\mathbb{N}}A_{n}, where AnM2n()A_{n}\cong M_{2^{n}}(\mathbb{C}). Let φj\varphi_{j}, for jj\in\mathbb{N}, be a family of separated product states of M2M_{2^{\infty}} (see [12, Definition 2.5]). The existence of such family is guaranteed by [12, Lemma 2.6, (1) implies (2)]. Let unu_{n} be a self-adjoint unitary in AnA_{n} as defined in the proof of [12, Lemma 2.7], so that for every nn the projections in AnA_{n} separating the pure states satisfy the analogues of conditions (6)–(8). Let Θ:=Adun\Theta:=\bigotimes_{\mathbb{N}}\operatorname{Ad}u_{n}. Then the action of Θ\Theta on the distinguished pure states is as required. To complete the proof, note that as in [12, Lemma 2.7], the classification of AF algebras implies that M2Θ/2M_{2^{\infty}}\rtimes_{\Theta}\mathbb{Z}/2\mathbb{Z} is isomorphic to M2M_{2^{\infty}}. ∎

Lemma 6.5.

Suppose that XX, YY, and ZZ are disjoint finite sets of pure states of M2M_{2^{\infty}} and 𝖥:XY\mathsf{F}\colon X\to Y is a bijection. Then there are φ¯\bar{\varphi} and ψ¯\bar{\psi} in 𝖯m+l(M2)\mathsf{P}_{m+l}(M_{2^{\infty}}) (where m=|X|m=|X| and l=|Z|l=|Z|), and ΘAut(M2)\Theta\in\operatorname{Aut}(M_{2^{\infty}}) such that 𝔼M2(φ¯,ψ¯)\mathbb{E}_{M_{2^{\infty}}^{\circ}}(\bar{\varphi},\bar{\psi}) forces the following:

  1. (1)

    With Θ𝔊:=Φ𝔊ΘΦ𝔊1\Theta_{\mathfrak{G}}:=\Phi_{\mathfrak{G}}\circ\Theta\circ\Phi_{\mathfrak{G}}^{-1}, we have that B:=M2Θ𝔊/2B:=M_{2^{\infty}}\rtimes_{\Theta_{\mathfrak{G}}}\mathbb{Z}/2\mathbb{Z} is isomorphic to M2M_{2^{\infty}}.

  2. (2)

    Every ηZ\eta\in Z has exactly two pure state extensions, denoted η+1\eta_{+1} and η1\eta_{-1}, to BB, and 2(A,η)=2(B,η±1)\ell_{2}(A,\eta)=\ell_{2}(B,\eta_{\pm 1}).

  3. (3)

    For every ηX\eta\in X, η\eta and F(η)F(\eta) have unique pure state extensions to BB, and these extensions are equivalent.

Proof.

For convenience, we write A:=M2A:=M_{2^{\infty}}. The plan is to match the pure states in XX, YY, and ZZ to those provided by Lemma 6.4 and import Θ\Theta from there. More specifically, fix ρj\rho_{j}, σj\sigma_{j}, ηj\eta_{j}, and Θ\Theta as guaranteed by Lemma 6.4. Enumerate XX as φj\varphi_{j}, for j<mj<m, and let φm+j:=𝖥(φj)\varphi_{m+j}:=\mathsf{F}(\varphi_{j}) for j<mj<m. Enumerate YY as φ2m+j\varphi_{2m+j}, for j<lj<l. Now let ψj:=ρj\psi_{j}:=\rho_{j} and ψm+j:=σj\psi_{m+j}:=\sigma_{j} if j<mj<m, and let ψ2m+j:=ηj\psi_{2m+j}:=\eta_{j} if j<kj<k.

  1. (1)

    Since Θ𝔊\Theta_{\mathfrak{G}} is conjugate to Θ\Theta, we have BM2B\cong M_{2^{\infty}}.

  2. (2)

    Fix ηZ\eta\in Z. Then η=φi=ψi\eta=\varphi_{i}=\psi_{i} for some ii, and therefore Lemma 6.3 implies that η\eta has exactly two pure state extensions, η±1\eta_{\pm 1}, to BB and that 2(A,η)=2(B,η±1)\ell_{2}(A,\eta)=\ell_{2}(B,\eta_{\pm 1}).

  3. (3)

    If ηX\eta\in X and ζ:=𝖥(η)\zeta:=\mathsf{F}(\eta), then φi=η\varphi_{i}=\eta and ψi=ζ\psi_{i}=\zeta for some ii. Theorem 4.4 implies that η\eta and ζ\zeta have unique pure state extensions to BB, and Theorem 3.8 implies that they are equivalent.

This concludes the proof. ∎

7. A proof of Theorem B, part II: The diagonalization argument

In this section, we prove that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} implies that the conclusion of Glimm’s dichotomy fails even for separably represented C\textrm{C}^{*}-algebras. More precisely, we prove the following:

Theorem 7.1.

Assume 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}}. For every m0m\geq 0 and every n1n\geq 1 there exists a simple, unital C\textrm{C}^{*}-algebra of density character 1\aleph_{1} with exactly m+nm+n unitary equivalence classes of irreducible representations such that mm of these representations are on a separable Hilbert space and nn of these representations are on a non-separable Hilbert space.

A simple C\textrm{C}^{*}-algebra with irreducible representations on both separable and non-separable Hilbert spaces can be constructed in 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} (see [10, Theorem 10.4.3]). Both this C\textrm{C}^{*}-algebra and the one in Theorem 7.1 are inductive limits of inductive systems of C\textrm{C}^{*}-algebras all of which are isomorphic to the CAR algebra.

Proof.

With Lemma 6.5 at our disposal, this proof is analogous to that of Theorem 5.4. Fix a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain (Mα:α<1)(M_{\alpha}:\alpha<\aleph_{1}). Using the Continuum Hypothesis, fix a surjection f:1H(1)f\colon\aleph_{1}\to H(\aleph_{1}) such that every element of H(1)H(\aleph_{1}) is listed cofinally often. By recursion on α<1\alpha<\aleph_{1}, we will define an inductive system of separable, simple, unital, non-type I C\textrm{C}^{*}-algebras, AαA_{\alpha}. For each AαA_{\alpha} we will have a distinguished (m+n)(m+n)-tuple of inequivalent pure states, (φiα:i<m+n)(\varphi^{\alpha}_{i}:i<m+n).

Let A0A_{0} be the CAR algebra with inequivalent pure states φi0\varphi^{0}_{i}, for i<m+ni<m+n. At the latter steps of the construction, we will assure that for all α<1\alpha<\aleph_{1} the following conditions hold:

  1. (1)

    If ξ<α\xi<\alpha then AξA_{\xi} is a unital C\textrm{C}^{*}-subalgebra of AαA_{\alpha}.

  2. (2)

    With γ(α):=min{γ:AαMγ}\gamma(\alpha):=\min\{\gamma:A_{\alpha}\in M_{\gamma}\}, Aα+1A_{\alpha+1} belongs to Mγ(α)+1M_{\gamma(\alpha)+1}.

  3. (3)

    Every pure state ψ\psi of AαA_{\alpha} that belongs to Mγ(α)M_{\gamma(\alpha)}, except φiα\varphi^{\alpha}_{i}, for i<mi<m, has a unique pure state extension to Aα+1A_{\alpha+1}.

  4. (4)

    If f(α)f(\alpha) is a code for a pair (Aξ,ψ)(A_{\xi},\psi), where ξ<α\xi<\alpha, ψ\psi is a pure state of AξA_{\xi} which has a unique pure state extension ψ~\tilde{\psi} to AαA_{\alpha}, and ψ~\tilde{\psi} is inequivalent to φiα\varphi^{\alpha}_{i} for all i<m+ni<m+n, then ψ\psi has a unique pure state extension to Aα+1A_{\alpha+1}, and this extension is equivalent to φmα+1\varphi^{\alpha+1}_{m}.

  5. (5)

    For all i<mi<m, φiα+1\varphi_{i}^{\alpha+1} extends φiα\varphi_{i}^{\alpha} and 2(Aα+1,φiα+1)=2(Aα,φiα)\ell_{2}(A_{\alpha+1},\varphi^{\alpha+1}_{i})=\ell_{2}(A_{\alpha},\varphi^{\alpha}_{i}).101010For the notation, see the discussion preceding Lemma 6.3.

In order to describe the recursive construction, suppose that β\beta is a countable ordinal such that AαA_{\alpha} as required has been defined for all α<β\alpha<\beta. As in the proof of Theorem 5.4, the interesting case is when β=α+1\beta=\alpha+1 for some α\alpha and f(α)f(\alpha) is a code for a pair (Aξ,ψ)(A_{\xi},\psi) that satisfies the following conditions:

  1. (a)

    ξ<α\xi<\alpha.

  2. (b)

    ψ\psi is a pure state of AξA_{\xi} that has a unique extension ψ~\tilde{\psi} to a pure state of AαA_{\alpha}.

  3. (c)

    For all i<m+ni<m+n, ψ~\tilde{\psi} is inequivalent to φiα\varphi^{\alpha}_{i}.

By the second part of Theorem 3.8, any forcing notion of the form 𝔼Aα(ρ¯,σ¯)\mathbb{E}_{A^{\circ}_{\alpha}}(\bar{\rho},\bar{\sigma}) is forcing-equivalent to the forcing notion for adding a single Cohen real. Since Mγ(α)+1M_{\gamma(\alpha)+1} contains a real that is Cohen-generic over Mγ(α)M_{\gamma(\alpha)}, by [23, Lemma IV.4.7], it contains an Mγ(α)M_{\gamma(\alpha)}-generic filter for any forcing notion of this form. Therefore, Lemma 6.5 implies that in Mγ(α)+1M_{\gamma(\alpha)+1} there exists an automorphism Θ𝔊\Theta_{\mathfrak{G}} of AαA_{\alpha} of order two and such that the C\textrm{C}^{*}-algebra Aα+1:=AαΘ𝔊/2A_{\alpha+1}:=A_{\alpha}\rtimes_{\Theta_{\mathfrak{G}}}\mathbb{Z}/2\mathbb{Z} is isomorphic to the CAR algebra, each φiα\varphi^{\alpha}_{i} for mi<m+nm\leq i<m+n has a unique pure state extension to Aα+1A_{\alpha+1}, ψ\psi has a unique pure state extension to Aα+1A_{\alpha+1} equivalent to111111Note that n1n\geq 1, hence φmα\varphi^{\alpha}_{m} is well-defined for every α<1\alpha<\aleph_{1}. φmα+1\varphi^{\alpha+1}_{m}, and φiαΘ𝔊=φiα\varphi^{\alpha}_{i}\circ\Theta_{\mathfrak{G}}=\varphi^{\alpha}_{i} for i<mi<m. Lemma 6.3 implies that, for i<mi<m, φiα\varphi^{\alpha}_{i} has a pure state extension φiα+1\varphi^{\alpha+1}_{i} to Aα+1A_{\alpha+1} that satisfies 2(Aα,φiα)=2(Aα+1,φiα+1)\ell_{2}(A_{\alpha},\varphi^{\alpha}_{i})=\ell_{2}(A_{\alpha+1},\varphi_{i}^{\alpha+1}). By Corollary 6.2, any other pure states of AαA_{\alpha} that belongs to Mγ(α)M_{\gamma(\alpha)} has a unique pure state extension to Aα+1A_{\alpha+1}. Also, Aα+1A_{\alpha+1} is separable, simple, unital, and non-type I by Corollary A.2.

If f(α)f(\alpha) does not satisfy the conditions (a)–(c), let Aα+1:=AαA_{\alpha+1}:=A_{\alpha} and, for each i<m+ni<m+n, let φiα+1:=φiα\varphi^{\alpha+1}_{i}:=\varphi^{\alpha}_{i}.

If β\beta is a limit ordinal, take Aβ:=limα<βAαA_{\beta}:=\varinjlim_{\alpha<\beta}A_{\alpha} and, for each i<m+ni<m+n, define φiβ\varphi_{i}^{\beta} as the unique pure state of AβA_{\beta} that extends φiα\varphi^{\alpha}_{i} for all α<β\alpha<\beta. Since φiβ\varphi^{\beta}_{i} is definable from its restrictions, it belongs to the relevant model.

This describes the recursive construction.

Let A1:=limα<1AαA_{\aleph_{1}}:=\varinjlim_{\alpha<\aleph_{1}}A_{\alpha}, and for i<m+ni<m+n let φi\varphi_{i} be the unique pure state of A1A_{\aleph_{1}} that extends φiα\varphi^{\alpha}_{i} for all α<1\alpha<\aleph_{1}.

By the construction, the pure states φi\varphi_{i}, for i<m+ni<m+n, are inequivalent. By (5) and induction, for i<mi<m, we have 2(A1,φi)=2(A0,φi0)\ell_{2}(A_{\aleph_{1}},\varphi_{i})=\ell_{2}(A_{0},\varphi_{i}^{0}), and therefore the GNS Hilbert space associated with φi\varphi_{i} is separable. If mi<m+nm\leq i<m+n, then φiα+1\varphi^{\alpha+1}_{i} is the unique extension of φiα\varphi^{\alpha}_{i} and therefore Lemma 6.5 implies that 2(Aα,φiα)\ell_{2}(A_{\alpha},\varphi^{\alpha}_{i}) is a proper subspace of 2(Aα+1,φiα+1)\ell_{2}(A_{\alpha+1},\varphi^{\alpha+1}_{i}) for all α<1\alpha<\aleph_{1}. Therefore, the GNS Hilbert space associated with φi\varphi_{i} is non-separable.

It only remains to prove that every pure state of A1A_{\aleph_{1}} is equivalent to some φi\varphi_{i}, but this proof is analogous to the corresponding proof in Theorem 5.4 and therefore omitted. ∎

The reader may wonder whether it is possible to sharpen the conclusion of Theorem 7.1 and obtain a simple, unital, infinite-dimensional C\textrm{C}^{*}-algebra AA with at most m0m\leq\aleph_{0} irreducible representations up to unitary equivalence such that every irreducible representation of AA is on a separable Hilbert space. The answer is well-known to be negative in the case when m=1m=1 (it is Rosenberg’s result, see e.g., [10, Corollary 5.5.6], that a counterexample to Naimark’s problem cannot be separably represented). A proof analogous to that of Rosenberg’s result provides a negative answer in the general case.

Proposition 7.2.

Suppose that AA is a non-type I C\textrm{C}^{*}-algebra all of whose irreducible representations are on a separable Hilbert space. Then AA has at least 202^{\aleph_{0}} spatially inequivalent irreducible representations.

Proof.

The assumption on AA is used only to prove that it has a self-adjoint element aa whose spectrum is a perfect set. In M2M_{2^{\infty}}, there exists a positive contraction a0a_{0} with this property. To see this, note that the diagonal masa (i.e., the maximal abelian subalgebra that is the inductive limit of the algebras of diagonal matrices) is isomorphic to the algebra of continuous functions on the Cantor space, and let a0a_{0} correspond to the identity map on the Cantor space via the continuous functional calculus. By Glimm’s theorem ([10, Theorem 1.7.2], but see also [7]), AA has a subalgebra whose quotient is isomorphic to M2M_{2^{\infty}}. Let aa be a self-adjoint lift of a0a_{0} to AA (it exists by [10, §2.5]). Then the spectrum of aa includes the spectrum of a0a_{0}, hence aa is as required.

For every element xx of the spectrum of aa, fix a pure state φx\varphi_{x} on AA such that φx(a)=x\varphi_{x}(a)=x. We can take φx\varphi_{x} to be a pure state extension of the point-evaluation at xx. Then the cyclic vector is an xx-eigenvector of πx(a)\pi_{x}(a), where πx\pi_{x} is the representation associated to φx\varphi_{x}. Since the eigenvectors corresponding to distinct eigenvalues are orthogonal, in every irreducible representation π\pi of AA the operator π(a)\pi(a) has only countably many eigenvectors. By a counting argument, AA has at least 202^{\aleph_{0}} spatial equivalence classes of irreducible representations. ∎

8. Concluding Remarks

Our title was inspired by the title of the ground-breaking Shelah’s paper [27], but the answers to the questions posed in these titles are quite different. Solovay’s inaccessible may or may not be taken away depending on whether one requires the Baire-measurability alone, or the Lebesgue-measurability as well. In our case, the 1\diamondsuit_{\aleph_{1}} is not necessary for the construction. The question whether a counterexample to Naimark’s problem can be constructed in 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} alone, in 𝖹𝖥𝖢+𝖢𝖧\operatorname{\mathsf{ZFC}}+\operatorname{\mathsf{CH}}, or using κ\diamondsuit_{\kappa} for some κ2\kappa\geq\aleph_{2}, remains open.

Around 2010, the senior author conjectured that Naimark’s problem has positive answer in a model obtained by adding a sufficient number (supercompact cardinal, if need be) of Cohen reals. Theorem A and its proof give some (inconclusive) support for the negation of this conjecture. Additional support would be provided by a proof that a forcing notion with the properties of 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) and the countable chain condition, can be defined for tuples of inequivalent pure states for every simple and unital (not necessarily separable) C\textrm{C}^{*}-algebra. The experience suggests that the countable chain condition and non-commutativity do not mix well: Note the increase of complexity between [14, §3] and [14, §4], as well as the proof that a natural noncommutative generalization of the poset for adding many Cohen reals does not have the countable chain condition given in [8, Lemma 4.1]. Moreover, if AA has irreducible representations on both separable and non-separable Hilbert spaces (see e.g., [10, Theorem 10.4.3]), then adding an automorphism of AA that moves one of the associated pure states to another, necessarily collapses 1\aleph_{1}. Thus, the relevant question is whether such forcing can be constructed for C\textrm{C}^{*}-algebras that are inductive limits that appear in the course of the proof of Theorem A.

Another possible route towards constructing a counterexample to Naimark’s problem would be the following: Instead of forcing with 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) or a modification thereof, find a separable C\textrm{C}^{*}-subalgebra BB of AA such that the restrictions φ¯\bar{\varphi}^{\prime} and ψ¯\bar{\psi}^{\prime} to BB of all the tuples of pure states involved, uniquely determine their extensions to AA. Force with 𝔼B(φ¯,ψ¯)\mathbb{E}_{B^{\circ}}(\bar{\varphi}^{\prime},\bar{\psi}^{\prime}) to produce a generic automorphism Φ\Phi of BB such that φ¯\bar{\varphi}^{\prime} and ψ¯\bar{\psi}^{\prime} have unique, and equivalent, extensions to BΦB\rtimes_{\Phi}\mathbb{Z}. This plan hinges on the answer to the following purely C\textrm{C}^{*}-algebraic question.121212Note, however, that we may assume BB is an elementary submodel of AA; see [10, Appendix D]. For simplicity, it is stated for single pure states instead of mm-tuples.

Question 8.1.

Suppose that AA is a unital C\textrm{C}^{*}-algebra, BB is a unital C\textrm{C}^{*}-subalgebra of AA, both AA and BB are simple, and φ\varphi^{\prime} and ψ\psi^{\prime} are pure states of BB with the unique pure state extensions to AA. In addition, suppose that Φ\Phi is a sufficiently generic automorphism of BB such that φ\varphi^{\prime} and ψ\psi^{\prime} have unique and unitarily equivalent pure state extensions to BΦB\rtimes_{\Phi}\mathbb{Z}. Is there an amalgamation CC of BΦB\rtimes_{\Phi}\mathbb{Z} and AA such that φ\varphi and ψ\psi have unique pure state extensions to CC?

Such amalgamation would be a ‘partial crossed product’ of sorts of AA by an automorphism of BB. There is a rich literature on partial crossed products (see [6] and the references thereof), but our situation does not satisfy the requirements imposed on partial dynamical systems in [6, Definition 6.4].

Question 8.2.

Let AA be a separable, simple, non-type I C\textrm{C}^{*}-algebra. Does there exist an automorphism Θ\Theta of AA such that σΘσ\sigma\circ\Theta\nsim\sigma for every pure state σ\sigma of AA?

If a C\textrm{C}^{*}-algebra AA has an automorphism Θ\Theta as in Question 8.2, then for every m1m\geq 1 and tuples φ¯\bar{\varphi} in ψ¯\bar{\psi} in 𝖯m(A)\mathsf{P}_{m}(A), it has an automorphism Φ\Phi with the same property that in addition satisfies φ¯Φ=ψ¯\bar{\varphi}\circ\Phi=\bar{\psi}. Such an automorphism can be obtained by conjugating Θ\Theta by a Kishimoto–Ozawa–Sakai-type automorphism as in Theorem 6.1. Assuming that in addition one could assure that all pure states of AA have unique pure state extensions to a crossed product associated with Θ\Theta, one would secure the assumptions of the following.

Proposition 8.3.

Suppose that there exists a class 𝒜\mathcal{A} of separable, simple, unital C\textrm{C}^{*}-algebras such that:

  1. (1)

    𝒜\mathcal{A} is closed under inductive limits, and

  2. (2)

    For every A𝒜A\in\mathcal{A} and pure states φ\varphi and ψ\psi of AA, there exists an extension B𝒜B\in\mathcal{A} of AA such that (i) φ\varphi and ψ\psi have equivalent pure state extensions to BB and (ii) every pure state of AA has a unique pure state extension to BB.

Then 𝖢𝖧\operatorname{\mathsf{CH}} implies that there is a counterexample to Naimark’s problem.

Proof.

Suppose that 𝖢𝖧\operatorname{\mathsf{CH}} holds, and fix X1X\subseteq\aleph_{1} such that the inner model L[X]L[X] (see [23, Definition II.6.29]) contains all reals. Then 1\diamondsuit_{\aleph_{1}} holds in L[X]L[X] (see [23, Exercise III.7.21]). Working in L[X]L[X], modify the construction of a counterexample as in [1] (see also [10, Theorem 11.2.2]) as follows: One constructs an inductive system of C\textrm{C}^{*}-algebras AαA_{\alpha}, for α<1\alpha<\aleph_{1}, in 𝒜\mathcal{A} so that at every successor step of the construction the extension Aα+1A_{\alpha+1} of AαA_{\alpha} is chosen using 1\diamondsuit_{\aleph_{1}} and (2). At limit stages, take inductive limits. The inductive limit AA of this system is a counterexample to Naimark’s problem in L[X]L[X], by a proof analogous to those in [1] or [10, Theorem 11.2.2].

We claim that AA remains a counterexample to Naimark’s problem in the universe VV. Assume otherwise. Since it is a counterexample to Naimark’s problem in L[X]L[X], there exists a pure state η\eta of AA that belongs to VV but not to L[X]L[X]. The set

𝖢:={α<1: the restriction of η to Aα is pure}\mathsf{C}:=\{\alpha<\aleph_{1}:\text{ the restriction of }\eta\text{ to }A_{\alpha}\text{ is pure}\}

includes a club (see [10, Proposition 7.3.10]). Let α:=min(𝖢)\alpha:=\min(\mathsf{C}). By induction on countable ordinals βα\beta\geq\alpha, one proves that ηAβ\eta\upharpoonright A_{\beta} is the unique pure state extension of ηAα\eta\upharpoonright A_{\alpha} to AβA_{\beta} in L[X]L[X], for every β<1\beta<\aleph_{1}. At the successor stages this is a consequence of the choice of Aβ+1A_{\beta+1}, and at the limit stages it is automatic. This provides a definition of η\eta in L[X]L[X]; contradiction. ∎

The proof of Proposition 8.3 begs the question: Is it possible to add a new pure state to a counterexample to Naimark’s without adding new reals? The answer is, at least assuming 1\diamondsuit_{\aleph_{1}}, positive (see [10, Exercise 11.4.11]); one can see that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} suffices for this construction).

The main result of [12] is a construction (under the assumption of 1\diamondsuit_{\aleph_{1}}) of a nuclear, simple C\textrm{C}^{*}-algebra not isomorphic to its opposite algebra. We do not know whether the existence of an algebra with this property follows from 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}}. In the same theorem, a counterexample to Glimm’s dichotomy with exactly 0\aleph_{0} inequivalent pure states was constructed using 1\diamondsuit_{\aleph_{1}}. Such construction using 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} would require a generalization of the forcing 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) to countable sequences of inequivalent pure states.

As pointed out in [15, Section 7], there is a possibility that the methods developed there may give a 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} construction of a simple C\textrm{C}^{*}-algebra not isomorphic to its opposite. These methods are however unlikely to be relevant to Naimark’s problem (see the discussion of [29] below).

In [29], it was shown that a counterexample to Naimark’s problem cannot be a graph C\textrm{C}^{*}-algebra (not to be confused with the ‘graph CCR algebras’ of [10, §10]). We conjecture that a sweeping generalization of this result holds: If a C\textrm{C}^{*}-algebra A(Γ)A(\Gamma) is defined from a discrete object (graph, group, groupoid, semigroup, etc.) Γ\Gamma in a way that assures that (using the notation of Appendix A) A(Γ)A(\Gamma) as computed in MM is dense in A(Γ)A(\Gamma) as computed in M[𝔊]M[\mathfrak{G}] for all MM and all MM-generic filter 𝔊\mathfrak{G}, then A(Γ)A(\Gamma) is (provably in 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}}) not a counterexample to Naimark’s problem.

By [21, Corollary 2.3], an automorphism Φ\Phi of a separable and simple C\textrm{C}^{*}-algebra satisfies φΦφ\varphi\circ\Phi\sim\varphi for all pure states φ\varphi of AA if, and only if, it is inner. Theorem 4.4 and Theorem 6.1 imply that both the generic automorphism Φ𝔊\Phi_{\mathfrak{G}} and the conjugate of a ground-model outer automorphism by Φ𝔊\Phi_{\mathfrak{G}} send every ground-model pure state to an inequivalent pure state. We conjecture that this property is shared by every reduced word in outer automorphisms of AA and Φ𝔊\Phi_{\mathfrak{G}} in which the latter occurs. This resembles the properties of generic automorphisms (and anti-automorphisms) of II1 factors as exhibited in [18, Lemma A.2] and [31], and used there to construct interesting examples of II1 factors with a separable predual. These lemmas, combined with an iterated crossed product construction à la Akemann–Weaver propelled by 1\diamondsuit_{\aleph_{1}} was used in [13] to construct a hyperfinite II1 factor with non-separable predual and not isomorphic to its opposite. It is not difficult to see that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}} in place of 1\diamondsuit_{\aleph_{1}} suffices for this construction.

Appendix A C\textrm{C}^{*}-algebras in generic extensions

In this appendix, we state and prove a few straightforward results on C\textrm{C}^{*}-algebras in models of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}. In particular, we will prove results about the relation between AMA^{M} and AM[𝔊]A^{M[\mathfrak{G}]} that we could not find in the literature.

Recall that a property P of C\textrm{C}^{*}-algebras is said to be absolute if for every AA, and every MM and M[𝔊]M[\mathfrak{G}] as in §2.7 and §2.8, AMA^{M} has P if, and only if, AM[𝔊]A^{M[\mathfrak{G}]} has P.131313Purists may prefer the term ‘forcing absolute’ but the difference can be ignored in the context of this paper.

Lemma A.1.

Both being simple and being non-type I are absolute properties of C\textrm{C}^{*}-algebras.

Proof.

To prove that simplicity is absolute, we first consider the case when AA is separable and unital. Fix a countable norm-dense subset DD of AA. Then some xAx\in A generates a proper two-sided ideal if, and only if, for every mm\in\mathbb{N} and all mm-tuples (aj:j<m)(a_{j}:j<m) and (bj:j<m)(b_{j}:j<m) of elements of DD, we have that 1Ai<maixbi1\|1_{A}-\sum_{i<m}a_{i}xb_{i}\|\geq 1. This is because 1a<1\|1-a\|<1 implies that aa is invertible (see [10, Lemma 1.2.6]). Thus, the assertion ‘AA has a proper two-sided ideal’ is a 𝚺11\mathbf{\Sigma}^{1}_{1}-statement (with some code for AA as a parameter) and is therefore absolute between all transitive models of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}(see [19, Theorem 25.20] or [10, Theorem B.2.11]).

If AA is not necessarily unital, then a similar argument show that the assertion ‘some xAx\in A generates a proper ideal of AA’ is 𝚺21\mathbf{\Sigma}^{1}_{2}, and therefore absolute between all transitive models of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}} that contain all countable ordinals (see [19, Theorem 25.20] or [10, Theorem B.2.11]).

If AA is not necessarily separable, then standard reflection arguments show that it is simple if, and only if, it is an inductive limit of a directed family of its separable, simple C\textrm{C}^{*}-subalgebras (see [10, §7.3]), and the conclusion follows.

We can now prove the absoluteness of being non-type I. Since the completion in a forcing extension of the ground-model CAR algebra is isomorphic to the CAR algebra as calculated in the forcing extension, the upwards absoluteness of being non-type I follows by Glimm’s theorem (see [10, Theorem 3.7.2]). ∎

It is worth mentioning that every axiomatizable (in logic of metric structures, see [11]) property of C\textrm{C}^{*}-algebras is absolute. The reason for this is that the ball of radius nn in AMA^{M} is dense in the ball of radius nn in AM[𝔊]A^{M[\mathfrak{G}]} for all nn, and therefore the suprema and infima of continuous functions on these two sets agree. More generally, properties definable by uniform families of formulas (see [11, Definition 5.7.1.1]) are absolute. Together with [11, Theorem 2.5.0.1 and Theorem 5.7.1.3], this implies that many important properties of C\textrm{C}^{*}-algebras are absolute.

An example of a non-absolute property is separability. Also, being isomorphic to (H)\mathscr{B}(H) or to the Calkin algebra is not absolute, since (for example) adding new subsets of \mathbb{N} adds new projections to the atomic masa that are not in the norm-closure of the ground-model projections. This shows that (H)M\mathscr{B}(H)^{M} is not dense in (H)M[𝔊]\mathscr{B}(H)^{M[\mathfrak{G}]}; in order to assure that they are not even isomorphic, one can for example increase the cardinality of the continuum.

Corollary A.2.

If AA is a separable, simple, unital, and non-type I C\textrm{C}^{*}-algebra, then 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}) forces that AΦ𝔊A\rtimes_{\Phi_{\mathfrak{G}}}\mathbb{Z} has all of these properties.

Proof.

By Lemma A.1, these properties of AA are absolute. The crossed product is therefore unital and non-type I, and it remains to prove that it is simple. Being non-type I, AA has continuum many inequivalent pure states and Theorem 4.4 implies that Φ𝔊\Phi_{\mathfrak{G}} moves a pure state to an inequivalent pure state, and is therefore outer. By [21, Theorem 3.1], this implies that the crossed product is simple. ∎

Lemma A.3.

For a C\textrm{C}^{*}-algebra AA, and models MM and M[𝔊]M[\mathfrak{G}] as described in §2.8, we have the following:

  1. (1)

    𝖴(AM)\mathsf{U}(A^{M}) is norm-dense in 𝖴(AM[𝔊])\mathsf{U}(A^{M[\mathfrak{G}]}).

  2. (2)

    If φ\varphi and ψ\psi are pure states of AA in MM, then their (unique) pure state extensions to AM[𝔊]A^{M[\mathfrak{G}]} are equivalent in M[𝔊]M[\mathfrak{G}] if, and only if, φ\varphi and ψ\psi are equivalent in MM.

Proof.

Note that (1) is a consequence of the continuous functional calculus, as follows: Suppose that (an:n)(a_{n}:n\in\mathbb{N}) is a sequence of elements of AA in M[𝔊]M[\mathfrak{G}] that converges to a unitary uu. Then anan10\|a_{n}^{*}a_{n}-1\|\to 0 and anan10\|a_{n}a_{n}^{*}-1\|\to 0 as nn\to\infty. Therefore, anana_{n}^{*}a_{n} is invertible for a large enough nn, and hence |an|:=(anan)1/2|a_{n}|:=(a_{n}^{*}a_{n})^{1/2} is invertible for a large enough nn. The unitary from the polar decomposition of ana_{n}, un:=an|an|1u_{n}:=a_{n}|a_{n}|^{-1}, satisfies unu0\|u_{n}-u\|\to 0.

To see (2), as pointed out in §2.4, φ\varphi and ψ\psi are equivalent if, and only if, there is a unitary uu such that φAduψ<2\|\varphi\circ\operatorname{Ad}u-\psi\|<2, hence the conclusion follows from (1). ∎

Every definable (in the model-theoretic sense, see [11, §3]) subset of AA has the absoluteness property proved for 𝖴(A)\mathsf{U}(A) in Lemma A.3 by a proof analogous to that of Lemma A.3.

The following was essentially proved in [1, Proposition 6]. We include its proof for completeness.

Proposition A.4.

Let MM be a countable transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}}, let \mathbb{P} be a forcing notion in MM, and let AMA\in M be a unital, non-type I C\textrm{C}^{*}-algebra. If \mathbb{P} adds a new real to MM, then it adds a new pure state to AA which is inequivalent to any ground-model pure state of AA.

Proof.

The construction is very similar to the one in the proof of Theorem 5.5.4 in [10], where additional details can be found. For i<2i<2 define a linear functional δi\delta_{i} on M2()M_{2}(\mathbb{C}) by

δi((λ00λ01λ10λ11))=λii.\delta_{i}\left(\begin{pmatrix}\lambda_{00}&\lambda_{01}\\ \lambda_{10}&\lambda_{11}\end{pmatrix}\right)=\lambda_{ii}.

This is a pure state of M2()M_{2}(\mathbb{C}). For r2r\in 2^{\mathbb{N}} define a linear functional φr\varphi_{r} on the CAR algebra as follows: on the elementary tensors (note that in nan\bigotimes_{n\in\mathbb{N}}a_{n}, we have that an=1a_{n}=1 for all but finitely many nn) let

φr(nan)=nδr(n)(an).\varphi_{r}\left(\bigotimes_{n\in\mathbb{N}}a_{n}\right)=\prod_{n\in\mathbb{N}}\delta_{r(n)}(a_{n}).

The linear extension of φr\varphi_{r} (still denoted φr\varphi_{r}) is a pure state on M2M_{2^{\infty}}. By Glimm’s theorem (see [10, Theorem 3.7.2]), AA includes some separable C\textrm{C}^{*}-subalgebra BB which has the CAR algebra as a quotient. The composition of φr\varphi_{r} with the quotient map is a pure state of BB, and this pure state can be extended to a pure state ψr\psi_{r} of AA. Clearly, rr can be recovered from ψr\psi_{r} by evaluation.

Suppose that ψr\psi_{r} is equivalent to a ground-model pure state. Since 𝖴(A)M\mathsf{U}(A)^{M} is norm-dense in 𝖴(A)M[𝔊]\mathsf{U}(A)^{M[\mathfrak{G}]}, there exists a ground-model pure state σ\sigma of AA such that ψrσ<1\|\psi_{r}-\sigma\|<1. Then the restriction of σ\sigma to BB still factors through the quotient map to a state of the CAR algebra and rr can be recovered from this state. But this implies that rr belongs to the ground model; contradiction. ∎

Suppose MM is a transitive model of 𝖹𝖥𝖢𝖯\operatorname{\mathsf{ZFC-P}} and XX is a Polish space with a code in MM. By a result of Solovay, an element rr of XX is Cohen-generic over MM if, and only if, it belongs to every dense open subset of XX coded in MM (see [19, Lemma 26.24]). Thus, there exists a Cohen-generic element of XX over MM if, and only if, the closed nowhere dense subsets of XX with codes in MM do not cover XX.

The minimal cardinality of a family of nowhere dense sets that cover the real line is denoted by cov()\operatorname{cov}(\mathscr{M}) (see [10, §8.4]). The Baire category theorem implies that this cardinal is uncountable, and Martin’s Axiom for κ\kappa dense sets implies that cov()>κ\operatorname{cov}(\mathscr{M})>\kappa (see [23]).

Corollary A.5.

Let AA be a separable, simple, unital and non-type I C\textrm{C}^{*}-algebra, and let X𝖯(A)X\subseteq\mathsf{P}(A) with |X|<cov()|X|<\operatorname{cov}(\mathscr{M}). Then for every pair of inequivalent pure states φ\varphi and ψ\psi on AA there exists some ΦAut(A)\Phi\in\text{Aut}(A) such that φΦ=ψ\varphi\circ\Phi=\psi and every ρX\rho\in X has a unique pure state extension to AΦA\rtimes_{\Phi}\mathbb{Z}.

Proof.

Let MM be an elementary submodel of H((20)+)H((2^{\aleph_{0}})^{+}) such that AMA\in M, XMX\subseteq M, and |M|<cov()|M|<\operatorname{cov}(\mathscr{M}). Let M¯\bar{M} be the transitive collapse of MM. Then the nowhere dense subsets of \mathbb{R} coded in M¯\bar{M} are too few to cover \mathbb{R}. By a result of Solovay, a real that does not belong to any of these sets is Cohen-generic over M¯\bar{M} (see [19, Lemma 26.24]). By Theorem 3.8, there exists an M¯\bar{M}-generic filter 𝔊\mathfrak{G} on 𝔼A(φ¯,ψ¯)\mathbb{E}_{A^{\circ}}(\bar{\varphi},\bar{\psi}), and by Corollary 4.5, Φ:=Φ𝔊\Phi:=\Phi_{\mathfrak{G}} is as required. ∎

Appendix B The combinatorial principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}

This section contains only set-theoretic considerations: we prove that the combinatorial principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} does not imply Jensen’s 1\diamondsuit_{\aleph_{1}} and that it does not decide the cardinality of the continuum.

As the attentive reader may have noticed during the proof of Theorem 5.4 (or Theorem 7.1), the combinatorial principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} can be thought as an oracle in which the required tasks at successor steps can be done by the mean of a Cohen real. On the upside, and in opposition to the usual application of Jensen’s 1\diamondsuit_{\aleph_{1}}, such tasks can be delayed (this is the job of the book-keeping) and they do not have to be handled at the moment they are captured by the oracle. Our weakening of 1\diamondsuit_{\aleph_{1}} is, at the end of the day, a sort of guessing-plus-forcing axiom in which the generic objects exist (in a prescribed extension) only for countable forcing notions that are elements of models whose job is to capture subsets of 1\aleph_{1} correctly.

Lemma B.1.

It is relatively consistent with 𝖹𝖥𝖢\operatorname{\mathsf{ZFC}} that 𝖢𝗈𝗁𝖾𝗇+𝖢𝖧+¬1\diamondsuit^{\mathsf{Cohen}}+\operatorname{\mathsf{CH}}+\neg\diamondsuit_{\aleph_{1}} holds.

Proof.

Let M0M_{0} be a countable, transitive model of a large enough fragment of 𝖹𝖥𝖢+𝖢𝖧\operatorname{\mathsf{ZFC}}+\operatorname{\mathsf{CH}} in which 1\diamondsuit_{\aleph_{1}} fails. Such model was first constructed by Jensen (see [4], also [28, §V]). Let (α,˙α)α<1(\mathbb{P}_{\alpha},\dot{\mathbb{Q}}_{\alpha})_{\alpha<\aleph_{1}} be a finite support iteration of non-trivial ccc forcing notions, each of which has cardinality at most 1\aleph_{1}. Let 𝔊1\mathfrak{G}\subseteq\mathbb{P}_{\aleph_{1}} be an M0M_{0}-generic filter. By the countable chain condition, 1\diamondsuit_{\aleph_{1}} fails in M0[𝔊]M_{0}[\mathfrak{G}] (see [23, Exercise IV.7.57]) and the standard ‘counting of names’ argument shows that the Continuum Hypothesis holds in M0[𝔊]M_{0}[\mathfrak{G}].

For α<1\alpha<\aleph_{1}, Mα:=M0[𝔊ωα]M_{\alpha}:=M_{0}[\mathfrak{G}\cap\mathbb{P}_{\omega\cdot\alpha}] (here ωα\omega\cdot\alpha is the α\alphath limit ordinal) is the intermediate forcing extension. By the countable chain condition again, no reals are added at stages of uncountable cofinality (see [17, Lemma 18.9]), and therefore every real in M0[𝔊]M_{0}[\mathfrak{G}] belongs to some MαM_{\alpha} for α<1\alpha<\aleph_{1}. Since a finite support iteration of non-trivial ccc forcings adds a Cohen real at every limit stage of countable cofinality (see [23, Exercise V.4.25]), for every α<1\alpha<\aleph_{1} the model Mα+1M_{\alpha+1} contains a real that is Cohen-generic over MαM_{\alpha}.

Fix a name for a subset XX of 1\aleph_{1}. Yet again, by the countable chain condition and the standard closing off argument, there is a club C1C\subseteq\aleph_{1} such that for every αC\alpha\in C, the forcing notion α\mathbb{P}_{\alpha} adds XαX\cap\alpha. Therefore, XαMαX\cap\alpha\in M_{\alpha} for stationary many α\alpha, and 1\mathbb{P}_{\aleph_{1}} forces that 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} holds. ∎

The following corollary exhibits a substantial difference between the two principles 1\diamondsuit_{\aleph_{1}} and 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}.

Corollary B.2.

The principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} does not decide the value of 202^{\aleph_{0}}.

Proof.

If in the proof of Lemma B.1 we begin with a model of 20=κ2^{\aleph_{0}}=\kappa, then M0[𝔊]M_{0}[\mathfrak{G}] is a model of 𝖢𝗈𝗁𝖾𝗇+20=κ\diamondsuit^{\mathsf{Cohen}}+2^{\aleph_{0}}=\kappa. ∎

To see that 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} is not a consequence of 𝖢𝖧\operatorname{\mathsf{CH}}, we will show that, unlike the Continuum Hypothesis, 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} implies the existence of a Suslin tree.

In [25], Moore, Hrušák, and Džamonja introduced a variety of parametrized 1\diamondsuit_{\aleph_{1}} principles based on the weak diamond (see [5]) which have a similar relationship to 1\diamondsuit_{\aleph_{1}} as cardinal invariants of the continuum have to 𝖢𝖧\operatorname{\mathsf{CH}}.

Definition B.3 ([25]).

The principle (𝗇𝗈𝗇())\diamondsuit(\operatorname{\mathsf{non}}(\mathscr{M})) holds if for every function

F:2<1F\colon 2^{<\aleph_{1}}\to\mathscr{M}

such that F2αF\upharpoonright 2^{\alpha}, for α<1\alpha<\aleph_{1}, is Borel, there exists some g:1g\colon\aleph_{1}\to\mathbb{R} such that for all f:12f\colon\aleph_{1}\to 2, the set {α<1:g(α)F(fα)}\{\alpha<\aleph_{1}:g(\alpha)\notin F(f\upharpoonright\alpha)\} is stationary.

Proposition B.4.

The principle (𝗇𝗈𝗇())\diamondsuit(\operatorname{\mathsf{non}}(\mathscr{M})) is a consequence of 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}.

Proof.

Let (Mα:α<1)(M_{\alpha}:\alpha<\aleph_{1}) be a 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}-chain, and let F:2<1F\colon 2^{<\aleph_{1}}\to\mathscr{M} be such that for all α<1\alpha<\aleph_{1} the restriction F2αF\upharpoonright 2^{\alpha} is Borel. For each α<1\alpha<\aleph_{1}, let rαr_{\alpha}\in\mathbb{N}^{\mathbb{N}} be such that F2αF\upharpoonright 2^{\alpha} is definable from rαr_{\alpha} and let αϕ(α)<1\alpha\leq\phi(\alpha)<\aleph_{1} be such that rαMϕ(α)r_{\alpha}\in M_{\phi(\alpha)}. Define g:1g\colon\aleph_{1}\to\mathbb{R} by choosing g(α)g(\alpha) to be Cohen-generic over Mϕ(α)M_{\phi(\alpha)}. Let f:12f\colon\aleph_{1}\to 2 be arbitrary. Since {α<1:fαMα}\{\alpha<\aleph_{1}:f\upharpoonright\alpha\in M_{\alpha}\} is stationary, and g(α)g(\alpha) is Cohen-generic over a model containing both fαf\upharpoonright\alpha and rαr_{\alpha}, then {α<1:g(α)F(fα)}\{\alpha<\aleph_{1}:g(\alpha)\notin F(f\upharpoonright\alpha)\} is stationary as well. ∎

Corollary B.5.

Following the notation above.

  1. (1)

    If 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} holds then there is a Suslin tree.

  2. (2)

    The principle 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}} is not a consequence of 𝖢𝖧\operatorname{\mathsf{CH}}.

Proof.

By [25, Theorem 3.1], (𝗇𝗈𝗇())\diamondsuit(\operatorname{\mathsf{non}}(\mathscr{M})) implies that there is a Suslin tree and therefore (1) follows from Proposition B.4. (2) follows from (1) and the fact that 𝖢𝖧\operatorname{\mathsf{CH}} does not imply the existence of a Suslin tree ([4]). ∎

One could consider 𝖱𝖺𝗇𝖽𝗈𝗆\diamondsuit^{\operatorname{\mathsf{Random}}}, 𝖧𝖾𝖼𝗁𝗅𝖾𝗋\diamondsuit^{\operatorname{\mathsf{Hechler}}}, or diamonds associated to other Suslin ccc forcings. The countable chain condition of the forcing is used in order to assure the property 𝖢𝗈𝗁𝖾𝗇\diamondsuit^{\mathsf{Cohen}}(b) in Definition 5.2. We are not aware of any applications of these axioms.

References

  • [1] C. Akemann and N. Weaver, Consistency of a counterexample to Naimark’s problem, Proc. Natl. Acad. Sci. USA 101 (2004), no. 20, 7522–7525.
  • [2] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122, Springer-Verlag, Berlin, 2006, Theory of C\textrm{C}^{*}-algebras and von Neumann algebras, Operator Algebras and Non-commutative Geometry, III.
  • [3] N. Brown and N. Ozawa, C\textrm{C}^{*}-algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008.
  • [4] K.J. Devlin and H. Johnsbraten, The Souslin problem, Lecture Notes in Math., vol. 405, Springer, 1974.
  • [5] K.J. Devlin and S. Shelah, A weak version of \diamondsuit which follows from 20<212^{\aleph_{0}}<2^{\aleph_{1}}, Israel J. Math. 29 (1978), 239–247.
  • [6] R. Exel, Partial dynamical systems, Fell bundles and applications, Mathematical Surveys and Monographs, vol. 224, Amer. Math. Soc., 2017.
  • [7] I. Farah, Combinatorial set theory and C\textrm{C}^{*}-algebras errata, available at https://ifarah.mathstats.yorku.ca/combinatorial-set-theory-of-c-algebras-errata/.
  • [8] by same author, All automorphisms of all Calkin algebras, Math. Research Letters 18 (2011), 489–503.
  • [9] by same author, Logic and operator algebras, Proceedings of the Seoul ICM, volume II (Sun Young Jang et al., eds.), Kyung Moon SA, 2014, pp. 15–39.
  • [10] by same author, Combinatorial set theory and C\textrm{C}^{*}-algebras, Springer Monographs in Mathematics, Springer, 2019.
  • [11] I. Farah, B. Hart, M. Lupini, L. Robert, A. Tikuisis, A. Vignati, and W. Winter, Model theory of C\textrm{C}^{*}-algebras, Memoirs of the AMS 271 (2021), no. 1324.
  • [12] I. Farah and I. Hirshberg, Simple nuclear C\textrm{C}^{*}-algebras not isomorphic to their opposites, Proc. Natl. Acad. Sci. USA 114 (2017), no. 24, 6244–6249.
  • [13] by same author, A rigid hyperfinite type II1 factor, International Mathematics Research Notices 2022 (2020), no. 24, 846–861.
  • [14] I. Farah, G. Katsimpas, and A. Vaccaro, Embedding C\textrm{C}^{*}-algebras into the Calkin algebra, IMRN 2021 (2021), no. 11, 8188–8224.
  • [15] Ilijas Farah and Najla Manhal, Nonseparable ccr algebras, International Journal of Mathematics (2021), 2150094.
  • [16] H. Futamura, N. Kataoka, and A. Kishimoto, Homogeneity of the pure state space for separable CC^{\ast}-algebras, Internat. J. Math. 12 (2001), no. 7, 813–845.
  • [17] L.J. Halbeisen, Combinatorial Set Theory with a gentle Introduction to Forcing, Springer Monographs in Mathematics, Springer, 2017.
  • [18] A. Ioana, J. Peterson, and S. Popa, Amalgamated free products of weakly rigid factors and calculation of their symmetry groups, Acta Math. 200 (2008), no. 1, 85–153. MR 2386109
  • [19] T. Jech, Set theory, Springer, 2003.
  • [20] A. Kanamori, The higher infinite: large cardinals in set theory from their beginnings, Perspectives in Mathematical Logic, Springer, Berlin–Heidelberg–New York, 1995.
  • [21] A. Kishimoto, Outer automorphisms and reduced crossed products of simple C\textrm{C}^{*}-algebras, Communications in Mathematical Physics 81 (1981), no. 3, 429–435.
  • [22] A. Kishimoto, N. Ozawa, and S. Sakai, Homogeneity of the pure state space of a separable CC^{*}-algebra, Canad. Math. Bull. 46 (2003), no. 3, 365–372.
  • [23] K. Kunen, Set theory, Studies in Logic (London), vol. 34, College Publications, London, 2011.
  • [24] D.A. Martin, Borel determinacy, Annals of Math. 102 (1975), no. 2, 363–371.
  • [25] J.T. Moore, M. Hrušák, and M. Džamonja, Parametrized \diamondsuit principles, Trans. Amer. Math. Soc. 356 (2004), no. 6, 2281–2306.
  • [26] A. Rinot, Jensen’s diamond principle and its relatives, Set theory and its applications, Contemp. Math., vol. 533, Amer. Math. Soc., 2011, pp. 125–156.
  • [27] S Shelah, Can you take Solovay’s inaccessible away?, Isr. J. Math. 48 (1984), 1–47.
  • [28] S. Shelah, Proper and improper forcing, Perspectives in Mathematical Logic, Springer, 1998.
  • [29] N. Suri and M. Tomforde, Naimark’s problem for graph C\mathrm{C}^{\ast}-algebras, Illinois J. Math. 61 (2017), no. 3-4, 479–495.
  • [30] A. Vaccaro, Trace spaces of counterexamples to Naimark’s problem, J. Funct. Anal. 275 (2018), no. 10, 2794–2816.
  • [31] S. Vaes, Factors of type II1 without non-trivial finite index subfactors, Trans. Amer. Math. Soc. 361 (2009), no. 5, 2587–2606.
  • [32] E. Wofsey, P(ω)/finP(\omega)/{\rm fin} and projections in the Calkin algebra, Proc. Amer. Math. Soc. 136 (2008), no. 2, 719–726.