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

Structure of relatively bi-exact group von Neumann algebras

Changying Ding Department of Mathematics, Vanderbilt University, 1326 Stevenson Center, Nashville, TN 37240, USA [email protected]  and  Srivatsav Kunnawalkam Elayavalli Institute of Pure and Applied Mathematics, 460 Portola Plaza, Los Angeles, CA 90095 [email protected]
Abstract.

Using computations in the bidual of 𝔹(L2M)\mathbb{B}(L^{2}M) we develop a new technique at the von Neumann algebra level to upgrade relative proper proximality to full proper proximality. This is used to structurally classify subalgebras of LΓL\Gamma where Γ\Gamma is an infinite group that is biexact relative to a finite family of subgroups {Λi}iI\{\Lambda_{i}\}_{i\in I} such that each Λi\Lambda_{i} is almost malnormal in Γ\Gamma. This generalizes the result of [DKEP22] which classifies subalgebras of von Neumann algebras of biexact groups. By developing a combination with techniques from Popa’s deformation-rigidity theory we obtain a new structural absorption theorem for free products and a generalized Kurosh type theorem in the setting of properly proximal von Neumann algebras.

1. Introduction

Recently the authors and J. Peterson in [DKEP22] developed the theory of small at infinity compactifications a la Ozawa ([BO08]), in the setting of tracial von Neumann algebras. At the foundation of this work lies the theory of operator MM-bimodules and the several natural topologies that arise in this setting (see [ER88], [Mag97, Mag98, Mag05]). The small at infinity compactification is a canonical strong operator bimodule (in the sense of Magajna [Mag97]) containing the compact operators. By using the noncommutative Grothendieck inequality (similar to Ozawa in [Oza10a]) it was seen that this strong operator bimodule coincides with 𝕂,1(M){\mathbb{K}}^{\infty,1}(M), the closure 𝕂,1(M)\mathbb{K}^{\infty,1}(M) of 𝕂(L2M)\mathbb{K}(L^{2}M) with respect to the ,1\|\cdot\|_{\infty,1}-norm on 𝔹(L2M)\mathbb{B}(L^{2}M) given by T,1=supxM,x1Tx^1\|T\|_{\infty,1}=\sup_{x\in M,\|x\|\leq 1}\|T\hat{x}\|_{1}. The small at infinity compactification of a tracial von Neumann algebra MM is then given by

𝕊(M)={T𝔹(L2M)[T,JxJ]𝕂,1(M),forallxM}.\mathbb{S}(M)=\{T\in\mathbb{B}(L^{2}M)\mid[T,JxJ]\in\mathbb{K}^{\infty,1}(M),\ {\rm for\ all\ }x\in M\}.

It is easy to see that this operator MM-system 𝕊(M)\mathbb{S}(M) contains MM and 𝕂(L2M)\mathbb{K}(L^{2}M), and is an MM-bimodule. The advantage of the strong operator bimodule perspective is that it to identify an operator T𝕊(M)T\in\mathbb{S}(M) suffices to check that [T,JxJ]𝕂,1(M)[T,JxJ]\in\mathbb{K}^{\infty,1}(M) for all xx in some weakly closed subset of MM. This is what allows for the passage between the group and the von Neumann algebra settings. Using this technology [DKEP22] defined the notion of proper proximality for finite von Neumann algebras, extending the dynamical notion for groups [BIP18]: A finite von Neumann algebra (M,τ)(M,\tau) is properly proximal if there does not exist an MM-central state φ\varphi on 𝕊(M)\mathbb{S}(M) such that φ|M=τ\varphi_{|M}=\tau. By identifying and studying this property in various examples, the authors of [DKEP22] obtained applications to the structure theory of II1-factors. The goal of the present paper is to add to the list of applications.

The machinery underlying the results in this paper is built on is the notion of an MM-boundary piece developed in [DKEP22], as an analogue of the group theoretic notion introduced in [BIP18]. The motivation for considering this notion is that it allows for one to exploit the dynamics that is available only on certain locations of the Stone-Cech boundary of the group. For a group Γ\Gamma, a boundary piece is a closed left and right invariant subset of β(Γ)Γ\beta(\Gamma)\setminus\Gamma, whereas in the von Neumann algebra setting, it is denoted by 𝕏{\mathbb{X}} typically and is a certain hereditary CC^{*}-subalgebra of 𝔹(L2M)\mathbb{B}(L^{2}M) containing the compact operators (see Section 3.1). One then considers the small at infinity compactification relative to a boundary piece 𝕊𝕏(M)\mathbb{S}_{{\mathbb{X}}}(M) where 𝕂,1(M)\mathbb{K}^{\infty,1}(M) is replaced with 𝕂𝕏,1(M)\mathbb{K}_{{\mathbb{X}}}^{\infty,1}(M), a suitable analogue for the boundary piece. Then one can define the notion of proper proximality relative to 𝕏{\mathbb{X}}, demanding that there be no MM-central state restricting to the trace on 𝕊𝕏(M)\mathbb{S}_{{\mathbb{X}}}(M). The main example we will be working with is a boundary piece generated by a finite family of von Neumann subalgebras {Mi}i=1n\{M_{i}\}_{i=1}^{n} (see Example 3.1), which is adapted from the construction for a finite family of subgroups (see Example 3.3 in [BIP18]).

In [DKE22], the authors demonstrated an instance where relative proper proximality can be lifted to full proper proximality, i.e, when the boundary piece arises from subgroups that are almost malnormal 111A subgroup H<GH<G is almost malnormal if for all gGHg\in G\setminus H, gHg1HgHg^{-1}\cap H is finite. and not co-amenable (see Lemma 3.3 in [DKE22]). The authors used this idea to classify proper proximality for wreath product groups. In this paper, we develop an analogue of this idea in the setting of von Neumann algebras (Theorem 1.1). In both cases, one has to work in the bidual of the small at infinity compactification for technical reasons, and this brings about an extra layer of subtlety especially in the von Neumann setting. More specifically we show that one can map the basic construction into the bidual version of the relative small at infinity compactification, provided the boundary piece arises from a mixing subalgebra. Composing with an appropriate state on this space, we get the link with relative amenability in the von Neumann setting. This upgrading theorem is the main new technical tool we develop in the present work:

Theorem 1.1.

Let MM be a diffuse finite von Neumann algebra, MiMM_{i}\subset M, i=1,,ni=1,\dots,n diffuse von Neumann subalgebras such that the Jones projections eMie_{M_{i}} pairwise commute, MiMM_{i}\subset M admits a bounded Pimsner-Popa basis (see Definition 2.1), and ApMpA\subset pMp is a von Neumann subalgebra, for some p𝒫(M)p\in\mathcal{P}(M). Suppose that AA is properly proximal relative to 𝕏{\mathbb{X}} inside MM, where 𝕏{\mathbb{X}} is the boundary piece associated with {Mi}i=1n\{M_{i}\}_{i=1}^{n}, and MiMM_{i}\subset M is mixing for each i=1,,ni=1,\dots,n. Then there exist projections f0𝒵(A)f_{0}\in\mathcal{Z}(A) and fi𝒵(ApMp)f_{i}\in{\mathcal{Z}}(A^{\prime}\cap pMp), 1in1\leq i\leq n, such that Af0Af_{0} is properly proximal and AfiAf_{i} is amenable relative to MiM_{i} inside MM for each 1in1\leq i\leq n, and i=0nfi=p\sum_{i=0}^{n}f_{i}=p.

Using these ideas we are interested in classifying subalgebras of group von Neumann algebras arising from groups that are biexact relative to a family of subgroups (see e.g. [BO08, Chapter 15]). The first result of this kind was obtained in Theorem 7.2 of [DKEP22] where it was shown that every subalgebra of the von Neumann algebra of a biexact group either has an amenable direct summand or is properly proximal. As essentially observed there, what relative biexactness buys us is the relative proper proximality for any subalgebra, relative to the boundary piece arising from the subgroups. Combining this with our upgrading result above, we obtain our main result below which is a structure theorem for von Neumann subalgebras of group von Neumann algebras that are biexact relative to a family of subgroups where each subgroup is almost malnormal.

Theorem 1.2.

Let Γ\Gamma be a countable group with a family of almost malnormal subgroups {Λi}i=1n\{\Lambda_{i}\}_{i=1}^{n}. If Γ\Gamma is biexact relative to {Λi}i=1n\{\Lambda_{i}\}_{i=1}^{n}, then for any von Neumann subalgebra ALΓA\subset L\Gamma, there exists p𝒵(A)p\in{\mathcal{Z}}{(A)} and projections pj𝒵((Ap)pL(Γ)p)p_{j}\in\mathcal{Z}((Ap)^{\prime}\cap pL(\Gamma)p) such that j=1npj=p\bigvee_{j=1}^{n}p_{j}=p and ApiAp_{i} is amenable relative to LΛiL\Lambda_{i} inside LΓL\Gamma, for each i=1,,ni=1,\dots,n and ApAp^{\perp} is properly proximal.

There are two natural instances where such a phenomenon (a countable group Γ\Gamma with a family of almost malnormal subgroups {Λi}i=1n\{\Lambda_{i}\}_{i=1}^{n} where Γ\Gamma is biexact relative to {Λi}i=1n\{\Lambda_{i}\}_{i=1}^{n}) is observed: First is in the setting of free products, which we deal with in the present paper. Second is in the setting of wreath products, which is investigated in a follow-up work by the first author [Din22]. There is conjecturally a third setting of relative hyperbolicity, which we comment on in the end of the introduction.

Thanks to Bass-Serre theory [Ser80] we have a complete understanding of subgroups of a free product of groups. As a result, one can derive results of the following nature: If H<G1G2H<G_{1}*G_{2} such that |HG1|3|H\cap G_{1}|\geq 3, then HH is amenable only if H<G1H<G_{1}. This phenomenon is referred to as amenable absorption. Interestingly, the situation for von Neumann algebras is much more complicated. There is comparatively a very limited understanding of von Neumann subalgebras of free products. Whether every self adjoint operator in any finite von Neumann algebra is contained in a copy of the hyperfinite II1-factor was itself an open problem for many years222This is a question of Kadison, Problem 7 from ’Problems on von Neumann algebras, Baton Rouge Conference’ . Popa settled it in the negative in [Pop83] by discovering a surprising amenable absorption theorem for free product von Neumann algebras, thereby showing that a generator masa in L𝔽2L\mathbb{F}_{2} is maximally amenable.

Popa’s ideas been used to show maximal amenability in other situations (See for instance [CFRW10], [Wen16], [PSW18], [BW16]). In the past decade there have been other new ideas that have been used to prove absorption theorems: Boutonnet-Carderi’s approach [BC15] relies on elementary computations in a crossed-product CC^{*}-algebra; Boutonnet-Houdayer [BH18] use the study of non normal conditional expectations; [HJNS21] used a free probabilistic approach to study absorption. Ozawa in [Oza10b] then gave a short proof of amenable absorption in tracial free products. There have also been a variety of important free product absorption results which are of a different flavor, and are structural in nature. See for example [IPP08] and [CH10].

By applying our Theorem 1.2 in the setting of free products and using machinery from Popa’s deformation-rigidity theory (specifically work of Ioana [Ioa18]), we obtain a generalized structural absorption theorem below:

Corollary 1.3.

Let (M1,τ1)(M_{1},\tau_{1}) and (M2,τ2)(M_{2},\tau_{2}) be such that MiLΓiM_{i}\cong L\Gamma_{i} where Γi\Gamma_{i} are countable exact groups and M=M1M2M=M_{1}\ast M_{2} be the tracial free product. Let AMA\subset M be a von Neumann subalgebra with AM1A\cap M_{1} diffuse. If AMA\subset M has no properly proximal direct summand, then AM1A\subset M_{1}.

Remark 1.4.

Using results of the upcoming work [DP22], one can relax the assumption on MiM_{i}, from being infinite group von Neumann algebras of exact groups, to just that they are diffuse weakly exact von Neumann algebras. We do not comment more on this at the moment because for the sake of examples, the above setting already provides many.

The authors of [IPR19] showed that there are examples of groups that are neither inner amenable nor properly proximal. All of these group von Neumann algebras fit into the setting of the above corollary. Note that Vaes constructed in [Vae12] plenty of groups that are inner amenable, yet their group von Neumann algebras lack Property (Gamma). Hence our results give a strict generalization of (Gamma) absorption (see Houdayer’s Theorem 4.1 in [Hou14] and see also Theorem A in [HJNS21]) in these examples.

Remark 1.5.

The above result is false if one considers amalgamated free products. For instance, take M1=M2=L𝔽2¯RM_{1}=M_{2}=L\mathbb{F}_{2}\overline{\otimes}R and A=(LL)¯RM1RM2A=(L\mathbb{Z}\ast L\mathbb{Z})\overline{\otimes}R\subset M_{1}\ast_{R}M_{2}, where two copies of LL\mathbb{Z} in AA are from M1M_{1} and M2M_{2}, respectively.

Remark 1.6.

Shortly before the posting of this paper, Drimbe announced a paper (see [Dri22]) where he shows using Popa’s deformation-rigidity theory that for any nonamenable inner amenable group Γ\Gamma, if L(Γ)M1M2L(\Gamma)\subset M_{1}*M_{2}, then L(Γ)L(\Gamma) intertwines into MiM_{i} for some i=1,2i=1,2. This in particular generalizes Corollary 1.3 in the case that ALΓA\cong L\Gamma for some inner amenable group Γ\Gamma, because he doesn’t require any assumptions for MiM_{i}.

Our techniques also reveal the following new Kurosh type structure theorem for free products in the setting of proper proximality, (partially generalizing Corollary 8.1 in [Dri22]). See also [Oza06, IPP08, Pet09, HU16] for other important Kurosh type theorems.

Corollary 1.7.

Let M=LΓ1LΓm=LΛ1LΛnM=L\Gamma_{1}*\cdots*L\Gamma_{m}=L\Lambda_{1}*\cdots*L\Lambda_{n}, where all groups Γi\Gamma_{i} and Λj\Lambda_{j} are countable exact nonamenable non-properly proximal i.c.c. groups. Then m=nm=n and after a permutation of indices LΓiL\Gamma_{i} is unitarily conjugate to LΛiL\Lambda_{i}.

We conclude by state the following folklore conjecture (also stated in [Oya22]), which would provide another family of examples for applying Theorem 1.2. Indeed the peripheral subgroups below are almost malnormal (see Theorem 1.4 in [Osi06]).

Conjecture 1 ([Oya22]).

If GG is exact and hyperbolic relative to a family of peripheral subgroups {Hi}i=1n\{H_{i}\}_{i=1}^{n}, then GG is biexact relative to {Hi}i=1n\{H_{i}\}_{i=1}^{n}.

Acknowledgements

The authors thank J. Peterson for stimulating conversations and very helpful suggestions. The authors thank Ben Hayes for reading our very early draft and offering many comments.

2. Preliminaries

2.1. The basic construction and Pimsner-Popa orthogonal bases

Let MM be a finite von Neumann algebra and QMQ\subset M be a von Neumann subalgebra. The basic construction M,eQ\langle M,e_{Q}\rangle is defined as the von Neumann subalgebra of 𝔹(L2M)\mathbb{B}(L^{2}M) generated by MM and the orthogonal projection eQe_{Q} from L2(M)L^{2}(M) onto L2(Q)L^{2}(Q). There is a semifinite faithful normal trace on M,eQ\langle M,e_{Q}\rangle satisfying Tr(xeQy)=τ(xy)\text{Tr}(xe_{Q}y)=\tau(xy), for every x,yMx,y\in M.

Let NMN\subset M be a von Neumann subalgebra. Then a Pimsner-Popa basis (see [PP86]) of MM over NN is a family of elements denoted M/N={mj}jJMM/N=\{m_{j}\}_{j\in J}\subset M such that

  1. (1)

    EN(mjmk)=δj,kpj{E}_{N}(m_{j}^{*}m_{k})=\delta_{j,k}p_{j}, where pj𝒫(N)p_{j}\in\mathcal{P}(N) is a projection.

  2. (2)

    L2(M)=jJmjL2(N)L^{2}(M)=\bigoplus_{j\in J}m_{j}L^{2}(N) and every xMx\in M has a unique decomposition x=jmjEN(mjx)x=\sum_{j}m_{j}{E}_{N}(m_{j}^{*}x).

In the case that N=L(Λ)N=L(\Lambda) and M=L(Γ)M=L(\Gamma) where Λ<Γ\Lambda<\Gamma, we can identify a Pimnser-Popa basis in MM from a choice of coset-representatives i.e, Γ=k0tkΛ\Gamma=\bigsqcup_{k\geq 0}t_{k}\Lambda, and mk:=λtk𝒰(LΓ)m_{k}:=\lambda_{t_{k}}\in{\mathcal{U}}(L\Gamma): M/N={uj}jJM/N=\{u_{j}\}_{j\in J}.

For technical reasons, we will need the existence of the following type of Pimsner-Popa basis for our results:

Definition 2.1.

Say that an inclusion of separable finite von Neumann algebras NMN\subset M admits a bounded Pimnser-Popa basis in MM if there exists a Pimsner-Popa basis {mk}k\{m_{k}\}_{k\in\mathbb{N}} for the inclusion NMN\subset M such that supkmk<\sup_{k\in\mathbb{N}}\|m_{k}\|<\infty.

Note that if M/NM/N consists of unitaries in MM, then it clearly also satisfies that it is a bounded Pimnser-Popa basis. This is a technical property considered by Ceccherini-Silberstein [CS04], called the U-property. It is a well known open problem if such bases always exist.

Such a Pimnser-Popa basis satisfying the above U-property always exists for the inclusion LΛLΓL\Lambda\subset L\Gamma where Λ<Γ\Lambda<\Gamma is a subgroup of a countable group Γ\Gamma.

2.2. Popa’s intertwining-by-bimodules

Theorem 2.2 ([Pop06]).

Let (M,τ)(M,\tau) be a tracial von Neumann algebra and PpMp,QMP\subset pMp,Q\subset M be von Neumann subalgebras. Then the following are equivalent:

  1. (1)

    There exist projections p0P,q0Qp_{0}\in P,q_{0}\in Q, a *-homomorphism θ:p0Pp0q0Qq0\theta:p_{0}Pp_{0}\rightarrow q_{0}Qq_{0} and a non-zero partial isometry vq0Mp0v\in q_{0}Mp_{0} such that θ(x)v=vx\theta(x)v=vx, for all xp0Pp0x\in p_{0}Pp_{0}.

  2. (2)

    There is no sequence un𝒰(P)u_{n}\in\mathcal{U}(P) satisfying EQ(xuny)20\|E_{Q}(x^{*}u_{n}y)\|_{2}\rightarrow 0, for all x,ypMx,y\in pM.

If one of these equivalent conditions holds, we write PMQP\prec_{M}Q, and say that a corner of PP embeds into QQ inside MM.

2.3. Relative amenability

Let PMP\subset M and QMQ\subset M be a von Neumann subalgebras. We say that PP is amenable relative to QQ inside MM if there exists a sequence ξnL2(M,eQ)\xi_{n}\in L^{2}(\langle M,e_{Q}\rangle) such that xξn,ξnτ(x)\langle x\xi_{n},\xi_{n}\rangle\rightarrow\tau(x), for every xMx\in M, and yξnξny20\|y\xi_{n}-\xi_{n}y\|_{2}\rightarrow 0, for every yPy\in P. By [OP10], Theorem 2.1 PP is amenable relative to QQ inside MM if and only if there exists a PP-central state in the basic construction M,eQ\langle M,e_{Q}\rangle that is normal when restricted to MM, and faithful on 𝒵(PM)\mathcal{Z}(P^{\prime}\cap M).

2.4. Mixing subalgebras and free products of finite von Neumann algebras.

Let MM be a finite von Neumann algebra and NMN\subset M a von Neumann subalgebra. Recall the inclusion NMN\subset M is mixing if L2(MN)L^{2}(M\ominus N) is mixing as an NN-NN bimodule, i.e., for any sequence un𝒰(N)u_{n}\in\mathcal{U}(N) converging to 0 weakly, one has EN(xuny)20\|E_{N}(xu_{n}y)\|_{2}\to 0 for any x,yMNx,y\in M\ominus N. When MM and NN are both diffuse, we may replace sequence of unitaries with any sequence in NN converging to 0 weakly [DKEP22, Theorem 5.9].

Remark 2.3.

Let MM be a diffuse finite von Neumann algebra and NMN\subset M a diffuse von Neumann subalgebra. If NMN\subset M is mixing, then it is easy to check that eNxJyJeN𝔹(L2M)e_{N}xJyJe_{N}\in{\mathbb{B}}(L^{2}M) is a compact operator from MM to L2ML^{2}M assuming xx or yMNy\in M\ominus N.

Examples of mixing subalgebras include M1M_{1} and M2M1M2M_{2}\subset M_{1}\ast M_{2}, where M1M_{1} and M2M_{2} are diffuse [Jol12, Proposition 1.6] and LΛLΓL\Lambda\subset L\Gamma, where Λ<Γ\Lambda<\Gamma is almost malnormal (see Proposition 2.4 in [BC17]).

The following [Ioa15, Corollary 2.12] is crucial to the proof of Theorem 5.1.

Lemma 2.4 (Ioana).

Let M1M_{1}, M2M_{2} be two diffuse tracial von Neumann algebras and M=M1M2M=M_{1}*M_{2} be the tracial free product. Let AMA\subset M be a subalgebra such that AA is amenable relative to M1M_{1} in MM. Then either AMM1A\prec_{M}M_{1} or AA is amenable.

We also need the following case of the main result of [BH18]:

Theorem 2.5 (Boutonnet-Houdayer).

Let M=M1M2M=M_{1}*M_{2}, where MiM_{i} are diffuse tracial von Neumann algebras. If AMA\subset M is a von Neumann subalgebra that satisfies AM1A\cap M_{1} is diffuse and AA is amenable relative to M1M_{1} inside MM, then AM1A\subset M_{1}.

3. Proper proximality for von Neumann algebras and boundary pieces

3.1. Boundary pieces from von Neumann subalgebras

Let MM be a finite von Neumann algebra. An MM-boundary piece is a hereditary C{\rm C}^{*}-subalgebra 𝕏𝔹(L2M){\mathbb{X}}\subset{\mathbb{B}}(L^{2}M) such that 𝕄(𝕏)M{\mathbb{M}}({\mathbb{X}})\cap M and 𝕄(𝕏)JMJ{\mathbb{M}}({\mathbb{X}})\cap JMJ are weakly dense in MM and JMJJMJ, respectively, where 𝕄(𝕏){\mathbb{M}}(\mathbb{X}) is the multiplier algebra of 𝕏\mathbb{X}. To avoid pathological examples, we will always assume that 𝕏{0}{\mathbb{X}}\not=\{0\}, and it follows that 𝕂(L2M)𝕏{\mathbb{K}}(L^{2}M)\subset{\mathbb{X}}, by the assumption on 𝕄(𝕏){\mathbb{M}}({\mathbb{X}}).

The main example of an MM-boundary piece we use in this paper is one generated by von Neumann subalgebras. We recall some facts about hereditary C{\rm C}^{*}-algebras for what follows (see e.g. [Bla06, II.5]).

Let AA be a C{\rm C}^{*}-algebra. There is a one-to-one correspondence between the set of hereditary C{\rm C}^{*}-subalgebras of A and the set of closed left ideals in AA: given a hereditary C{\rm C}^{*}-subalgebra HAH\subset A , LH:=AH={ah|aA,hH}L_{H}:=AH=\{ah\ |\ a\in A,h\in H\} is a closed left ideal; and for a closed left ideal LAL\subset A, HL=LLH_{L}=L\cap L^{*} is a hereditary C{\rm C}^{*}-subalgebra of AA. Given a subset of operators {bi}iIA\{b_{i}\}_{i\in I}\subset A, the hereditary C{\rm C}^{*}-subalgebra generated by {bi}iI\{b_{i}\}_{i\in I} is BAB={bab|bB+,aA}BAB=\{bab\ |\ b\in B_{+},a\in A\}, where BB is the C{\rm C}^{*}-subalgebra generated by {bi}iI\{b_{i}\}_{i\in I}.

Example 3.1.

[Boundary piece generated by subalgebras] Let MM be a finite von Neumann algebra. Suppose MiMM_{i}\subset M, i=1,,ni=1,\ldots,n are von Neumann subalgebras and denote by eMi𝔹(L2M)e_{M_{i}}\in{\mathbb{B}}(L^{2}M) the orthogonal projection from L2ML^{2}M onto the space L2MiL^{2}M_{i}. The MM-boundary piece associated with the family of subalgebras {Mi}i=1n\{M_{i}\}_{i=1}^{n} is the hereditary C{\rm C}^{*}-subalgebra of 𝔹(L2M){\mathbb{B}}(L^{2}M) generated by operators of the form xJyJeMixJyJe_{M_{i}} with x,yMx,y\in M, i=1,,ni=1,\dots,n, and it is clear that MM and JMJJMJ are contained in its multiplier algebra.

Lemma 3.2.

Let MM be a finite von Neumann algebra and MiMM_{i}\subset M, i=1,,ni=1,\dots,n von Neumann subalgebras such that the projections {eMi}i=1n\{e_{M_{i}}\}_{i=1}^{n} are pairwise commuting. Let 𝕏{\mathbb{X}} be the hereditary C{\rm C}^{*}-subalgebra in 𝔹(L2M){\mathbb{B}}(L^{2}M) generated by {xJyJ(i=1neMi)|x,yM}\{xJyJ(\vee_{i=1}^{n}e_{M_{i}})\ |\ x,y\in M\} and 𝕐{\mathbb{Y}} the hereditary C{\rm C}^{*}-subalgebra in 𝔹(L2M){\mathbb{B}}(L^{2}M) generated by {xJyJeMi|i=1,n,x,yM}\{xJyJe_{M_{i}}\ |\ i=1,\cdots n,\ x,y\in M\}. Then 𝕏=𝕐{\mathbb{X}}={\mathbb{Y}}.

Proof.

First note that eMi𝕏e_{M_{i}}\in{\mathbb{X}} for each ii since 0eMii=1neMi0\leq e_{M_{i}}\leq\vee_{i=1}^{n}e_{M_{i}}. We also have i=1neMi𝕐\vee_{i=1}^{n}e_{M_{i}}\in{\mathbb{Y}}. In fact, for each pair i,ji,j, eMieMj𝕐e_{M_{i}}\wedge e_{M_{j}}\in{\mathbb{Y}} as 0eMieMjeMi0\leq e_{M_{i}}\wedge e_{M_{j}}\leq e_{M_{i}}, and eMieMj=eMi+eMjeMieMj𝕐e_{M_{i}}\vee e_{M_{j}=}e_{M_{i}}+e_{M_{j}}-e_{M_{i}}\wedge e_{M_{j}}\in{\mathbb{Y}} as [eMi,eMj]=0[e_{M_{i}},e_{M_{j}}]=0. To see that 𝕏𝕐{\mathbb{X}}\subset{\mathbb{Y}}, note that L=𝔹(L2M)𝕏L={\mathbb{B}}(L^{2}M){\mathbb{X}} is contained in K=𝔹(L2M)𝕐K={\mathbb{B}}(L^{2}M){\mathbb{Y}}. Indeed, for any x,yMx,y\in M and T𝔹(L2M)T\in{\mathbb{B}}(L^{2}M), we have T(i=1neMi)xJyJ𝔹(L2M)𝕐xJyJ=𝔹(L2M)𝕐T(\vee_{i=1}^{n}e_{M_{i}})xJyJ\in{\mathbb{B}}(L^{2}M){\mathbb{Y}}xJyJ={\mathbb{B}}(L^{2}M){\mathbb{Y}} as MM and JMJJMJ are in the multiplier algebra of 𝕐{\mathbb{Y}}. By a similar argument we see that 𝕐𝕏{\mathbb{Y}}\subset{\mathbb{X}}. ∎

Lemma 3.3.

Under the above assumption, {u,vFuJvJ(i=1neMi)JvJu}F\{\vee_{u,v\in F}uJvJ(\vee_{i=1}^{n}e_{M_{i}})Jv^{*}Ju^{*}\}_{F\in\mathcal{F}} is an approximate unit for 𝕏{\mathbb{X}}, where \mathcal{F} is the collection of finite subsets of 𝒰(M)\mathcal{U}(M) ordered by inclusion.

Proof.

Set e0=i=1neMie_{0}=\vee_{i=1}^{n}e_{M_{i}} and eF=u,vFuJvJe0JvJue_{F}=\vee_{u,v\in F}uJvJe_{0}Jv^{*}Ju^{*} for any FF\in\mathcal{F}. First we observe that eF𝕏e_{F}\in{\mathbb{X}} as 0eFu,vFuJvJe0JvJu0\leq e_{F}\leq\sum_{u,v\in F}uJvJe_{0}Jv^{*}Ju^{*}. Note that whenever 1F1\in F, e0eF=e0e_{0}e_{F}=e_{0} and hence (u0Jv0Je0)eF=u0Jv0Je0(u_{0}Jv_{0}Je_{0})e_{F}=u_{0}Jv_{0}Je_{0}, for any u0,v0𝒰(M)u_{0},v_{0}\in\mathcal{U}(M). On the other hand, if u0,v0Fu_{0},v_{0}\in F, we have eF(u0Jv0Je0)=eF(u0Jv0Je0Jv0Ju0)u0Jv0Je0=u0Jv0Je0e_{F}(u_{0}Jv_{0}Je_{0})=e_{F}(u_{0}Jv_{0}Je_{0}Jv_{0}^{*}Ju_{0}^{*})u_{0}Jv_{0}Je_{0}=u_{0}Jv_{0}Je_{0}. The result follows by writing arbitrary x,yMx,y\in M as sums of four unitaries. ∎

Fix an MM-boundary piece 𝕏{\mathbb{X}} and let 𝕂𝕏L(M)𝔹(L2M){\mathbb{K}}^{L}_{\mathbb{X}}(M)\subset{\mathbb{B}}(L^{2}M) denote the ,2\|\cdot\|_{\infty,2} closure of the closed left ideal 𝔹(L2M)𝕏{\mathbb{B}}(L^{2}M){\mathbb{X}}, i.e., 𝕂𝕏L(M)=𝔹(L2M)𝕏¯,2{\mathbb{K}}^{L}_{\mathbb{X}}(M)=\overline{{\mathbb{B}}(L^{2}M){\mathbb{X}}}^{\|\cdot\|_{\infty,2}}, where ,2\|\cdot\|_{\infty,2} on 𝔹(L2M){\mathbb{B}}(L^{2}M) is given by T,2=supx(M)1Tx^2\|T\|_{\infty,2}=\sup_{x\in(M)_{1}}\|T\hat{x}\|_{2} for T𝔹(L2M)T\in{\mathbb{B}}(L^{2}M).

We let 𝕂𝕏(M)=(𝕂𝕏L(M))(𝕂𝕏L(M)){\mathbb{K}}_{\mathbb{X}}(M)=({\mathbb{K}}_{\mathbb{X}}^{L}(M))^{*}\cap({\mathbb{K}}_{\mathbb{X}}^{L}(M)), which is a hereditary C-subalgebra of 𝔹(L2M){\mathbb{B}}(L^{2}M) with MM and JMJJMJ contained in 𝕄(𝕂𝕏L(M)){\mathbb{M}}({\mathbb{K}}_{\mathbb{X}}^{L}(M)) [DKEP22, Section 3]. Denote by 𝕂𝕏,1(M){\mathbb{K}}_{\mathbb{X}}^{\infty,1}(M) the ,1\|\cdot\|_{\infty,1} closure of 𝕂𝕏(M){\mathbb{K}}_{\mathbb{X}}(M) in 𝔹(L2M){\mathbb{B}}(L^{2}M), T,1=supx,y(M)1Tx^,y^\|T\|_{\infty,1}=\sup_{x,y\in(M)_{1}}\langle T\hat{x},\hat{y}\rangle for T𝔹(L2M)T\in{\mathbb{B}}(L^{2}M) and it coincides with 𝕏¯,1\overline{{\mathbb{X}}}^{\|\cdot\|_{\infty,1}}.

Now put 𝕊𝕏(M)𝔹(L2M){\mathbb{S}}_{{\mathbb{X}}}(M)\subset{\mathbb{B}}(L^{2}M) to be

𝕊𝕏(M)={T𝔹(L2M)[T,JxJ]𝕂𝕏,1(M)forallxM},{\mathbb{S}}_{{\mathbb{X}}}(M)=\{T\in{\mathbb{B}}(L^{2}M)\mid[T,JxJ]\in{\mathbb{K}}_{\mathbb{X}}^{\infty,1}(M){\rm\ for\ all\ }x\in M\},

which is an operator system that contains MM. In the case when 𝕏=𝕂(L2M){\mathbb{X}}={\mathbb{K}}(L^{2}M), we write 𝕊(M){\mathbb{S}}(M) instead of 𝕊𝕂(L2M)(M){\mathbb{S}}_{{\mathbb{K}}(L^{2}M)}(M).

Recall from [DKEP22, Theorem 6.2] that for a finite von Neumann subalgebra NMN\subset M and an MM-boundary piece 𝕏\mathbb{X}, we say NN is properly proximal relative to 𝕏{\mathbb{X}} in MM if there is no NN-central state φ\varphi on 𝕊𝕏(M){\mathbb{S}}_{\mathbb{X}}(M) that is normal on MM. And we say MM is properly proximal if MM is properly proximal relative to 𝕂(L2M)\mathbb{K}(L^{2}M) in MM.

Remark 3.4.

Let MM and QQ be finite von Neumann algebras, 𝕏{\mathbb{X}} an MM-boundary piece, and NpMpN\subset pMp be a von Neumann subalgebra, where 0p𝒫(M)0\neq p\in\mathcal{P}(M).

  1. (1)

    Consider the u.c.p. map N:=Ad(eN)Ad(pJpJ):𝔹(L2M)𝔹(L2N)\mathcal{E}_{N}:=\operatorname{Ad}(e_{N})\circ\operatorname{Ad}(pJpJ):{\mathbb{B}}(L^{2}M)\to{\mathbb{B}}(L^{2}N). Then by [DKEP22, Remark 6.3] that N(𝕂𝕏(M))𝔹(L2N)\mathcal{E}_{N}({\mathbb{K}}_{\mathbb{X}}(M))\subset{\mathbb{B}}(L^{2}N) forms an NN-boundary piece. And we say N(𝕂𝕏(M))\mathcal{E}_{N}({\mathbb{K}}_{\mathbb{X}}(M)) is the induced NN-boundary piece, which will be denoted by 𝕏N{\mathbb{X}}^{N}.

  2. (2)

    If NN is properly proximal relative to 𝕏{\mathbb{X}} inside MM, then zNzN is also properly proximal relative to 𝕏{\mathbb{X}} inside MM for any 0z𝒵(𝒫(N))0\neq z\in\mathcal{Z}(\mathcal{P}(N)), since Ad(z)𝕊𝕏(M)𝕊𝕏(M)\operatorname{Ad}(z)\circ{\mathbb{S}}_{\mathbb{X}}(M)\subset{\mathbb{S}}_{\mathbb{X}}(M).

  3. (3)

    If NN is properly proximal relative to 𝕏{\mathbb{X}} inside MM, then NN has no amenable direct summand. To see this, suppose qNqN is amenable for some 0q𝒵(𝒫(N)0\neq q\in{\mathcal{Z}}(\mathcal{P}(N) and let φ\varphi be a qNqN-central state on 𝔹(L2(qN)){\mathbb{B}}(L^{2}(qN)). Consider μ:=φAd(q)Ad(eN):𝔹(L2M)\mu:=\varphi\circ\operatorname{Ad}(q)\circ\operatorname{Ad}(e_{N}):{\mathbb{B}}(L^{2}M)\to{\mathbb{C}}, and one checks that μ\mu is a NN-central state with μ|M\mu_{|M} being normal.

  4. (4)

    Notice that from the definition it follows that proper proximality is stable under taking direct sum. Thus we may take f𝒵(𝒫(Q))f\in{\mathcal{Z}}(\mathcal{P}(Q)) so that QfQf is the maximal properly proximal direct summand of QQ.

3.2. Bidual formulation of proper proximality

Given a finite von Neumann algebra MM and a C-subalgebra A𝔹(L2M)A\subset{\mathbb{B}}(L^{2}M) such that MM and JMJJMJ are contained in 𝕄(A){\mathbb{M}}(A), we recall that AMMA^{M\sharp M} (resp. AJMJJMJ)A^{JMJ\sharp JMJ}) denotes the space of φA\varphi\in A^{*} such that for each TAT\in A the map M×M(a,b)φ(aTb)M\times M\ni(a,b)\mapsto\varphi(aTb) (resp. JMJ×JMJ(a,b)φ(aTb)JMJ\times JMJ\ni(a,b)\mapsto\varphi(aTb)) is separately normal in each variable and set AJ=AMMAJMJJMJA^{\sharp}_{J}=A^{M\sharp M}\cap A^{JMJ\sharp JMJ}. Moreover, we may view (AJ)(A^{\sharp}_{J})^{*} as a von Neumann algebra in the following way, as shown in [DKEP22, Section 2]. Denote by pnor𝔹(L2M)p_{\rm nor}\in{\mathbb{B}}(L^{2}M)^{**} the supremum of support projections of states in 𝔹(L2M){\mathbb{B}}(L^{2}M)^{*} that restrict to normal states on MM and JMJJMJ, so that MM and JMJJMJ may be viewed as von Neumann subalgebras of pnor𝕄(A)pnorp_{\rm nor}{\mathbb{M}}(A)^{**}p_{\rm nor}. Note that pnorp_{\rm nor} lies in 𝕄(A){\mathbb{M}}(A)^{**} and pnor𝕄(A)pnorp_{\rm nor}{\mathbb{M}}(A)^{**}p_{\rm nor} is canonically identified with (𝕄(A)J)({\mathbb{M}}(A)^{\sharp}_{J})^{*}. Let qA𝒫(𝕄(A))q_{A}\in\mathcal{P}({\mathbb{M}}(A)^{**}) be the central projection such that qA(𝕄(A))=Aq_{A}({\mathbb{M}}(A)^{**})=A^{**} and we may then identify (AJ)(A^{\sharp}_{J})^{*} with qApnor𝕄(A)pnor=pnorApnorq_{A}p_{\rm nor}{\mathbb{M}}(A)^{**}p_{\rm nor}=p_{\rm nor}A^{**}p_{\rm nor}, which is also a von Neumann algebra. Furthermore, if BAB\subset A is another C-subalgebra with MM, JMJ𝕄(B)JMJ\subset{\mathbb{M}}(B), we may identify (BJ)(B^{\sharp}_{J})^{*} with qBpnorApnorqBq_{B}p_{\rm nor}A^{**}p_{\rm nor}q_{B}, which is a non-unital subalgebra of (AJ)(A^{\sharp}_{J})^{*}.

We will need the following bidual characterization of properly proximal.

Lemma 3.5.

[DKEP22, Lemma 8.5] Let MM be a separable tracial von Neumann algebra with an MM-boundary piece 𝕏\mathbb{X}. Then MM is properly proximal relative to 𝕏{\mathbb{X}} if and only if there is no MM-central state φ\varphi on

𝕊~𝕏(M):={T(𝔹(L2M)J)[T,a](𝕂𝕏(M)J)forallaJMJ}\widetilde{{\mathbb{S}}}_{{\mathbb{X}}}(M):=\left\{T\in\left({\mathbb{B}}(L^{2}M)_{J}^{{\sharp}}\right)^{*}\mid[T,a]\in\left({\mathbb{K}}_{\mathbb{X}}(M)_{J}^{\sharp}\right)^{*}\ {\rm for\ all\ }a\in JMJ\right\}

such that φ|M\varphi_{|M} is normal.

Using the above notations, we observe that we may identify 𝕊~𝕏(M)\tilde{{\mathbb{S}}}_{{\mathbb{X}}}(M) in the following way:

𝕊~𝕏(M)\displaystyle\widetilde{{\mathbb{S}}}_{\mathbb{X}}(M) ={T(𝔹(L2M)J)|[T,a](𝕂𝕏(M)J),foranyaJMJ}\displaystyle=\{T\in\big{(}{\mathbb{B}}(L^{2}M)_{J}^{\sharp}\big{)}^{*}|\ [T,a]\in\big{(}{\mathbb{K}}_{\mathbb{X}}(M)_{J}^{\sharp}\big{)}^{*},{\rm\ for\ any}\ a\in JMJ\}
={Tpnor𝔹(L2M)pnor|[T,a]q𝕏pnor(M(𝕂𝕏(M)))pnorq𝕏,foranyaJMJ},\displaystyle=\{T\in p_{\rm nor}{\mathbb{B}}(L^{2}M)^{**}p_{\rm nor}\ |\ [T,a]\in q_{{\mathbb{X}}}p_{\rm nor}\big{(}M({\mathbb{K}}_{\mathbb{X}}(M))\big{)}^{**}p_{\rm nor}q_{\mathbb{X}},{\rm\ for\ any}\ a\in JMJ\},

where q𝕏q_{{\mathbb{X}}} is the identity of 𝕂𝕏(M)(M(𝕂𝕏(M))){\mathbb{K}}_{\mathbb{X}}(M)^{**}\subset\big{(}M({\mathbb{K}}_{\mathbb{X}}(M))\big{)}^{**}. If we set q𝕂=q𝕂(L2M)q_{\mathbb{K}}=q_{{\mathbb{K}}(L^{2}M)} to be the identity of 𝕂(L2M)𝔹(L2M){\mathbb{K}}(L^{2}M)^{**}\subset{\mathbb{B}}(L^{2}M)^{**}, then using the above description of 𝕊~𝕏(M)\tilde{\mathbb{S}}_{\mathbb{X}}(M), we have q𝕏𝕊~𝕏(M)q𝕏q𝕂𝕊~(M)q_{{\mathbb{X}}}^{\perp}\widetilde{{\mathbb{S}}}_{{\mathbb{X}}}(M)q_{{\mathbb{X}}}^{\perp}\subset q_{{\mathbb{K}}}^{\perp}\widetilde{{\mathbb{S}}}(M), as q𝕏q_{\mathbb{X}} commutes with JMJJMJ.

Remark 3.6.

Recall that we may embed 𝔹(L2M){\mathbb{B}}(L^{2}M) into (𝔹(L2M)J)({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*} through the u.c.p. map ιnor\iota_{\rm nor}, which is given by ιnor=Ad(pnor)ι\iota_{\rm nor}=\operatorname{Ad}(p_{\rm nor})\circ\iota, where ι:𝔹(L2M)𝔹(L2M)\iota:{\mathbb{B}}(L^{2}M)\to{\mathbb{B}}(L^{2}M)^{**} is the canonical *-homomorphism into the universal envelope, and pnorp_{\rm nor} is the projection in 𝔹(L2M){\mathbb{B}}(L^{2}M)^{**} such that pnor𝔹(L2M)pnor=(𝔹(L2M)J)p_{\rm nor}{\mathbb{B}}(L^{2}M)^{**}p_{\rm nor}=({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}. Restricting ιnor\iota_{\rm nor} to C{\rm C}^{*}-subalgebra A𝔹(L2M)A\subset{\mathbb{B}}(L^{2}M) satisfying M,JMJ𝕄(A)M,JMJ\subset{\mathbb{M}}(A) give rise to the embedding of AA into (AJ)(A^{\sharp}_{J})^{*}, and (ιnor)M(\iota_{\rm nor})_{\mid M}, (ιnor)JMJ(\iota_{\rm nor})_{\mid JMJ} are faithful normal representations of MM and JMJJMJ, respectively. Furthermore, although ιnor\iota_{\rm nor} is not a *-homomorphism, spMeBM{\rm sp}Me_{B}M (that is, the span of elements xeByxe_{B}y where x,yMx,y\in M) is in the multiplicative domain of ϕ0\phi_{0}.

Lemma 3.7.

Let MM be a finite von Neumann algebra and 𝕏{\mathbb{X}} an MM-boundary piece. Let 𝕏0𝕂𝕏(M){\mathbb{X}}_{0}\subset{\mathbb{K}}_{\mathbb{X}}(M) be a C{\rm C}^{*}-subalgebra and {en}nI\{e_{n}\}_{n\in I} an approximate unit of 𝕏0{\mathbb{X}}_{0}. If 𝕏0𝕂𝕏,1(M){\mathbb{X}}_{0}\subset{\mathbb{K}}_{\mathbb{X}}^{\infty,1}(M) is dense in ,1\|\cdot\|_{\infty,1} and ι(en)\iota(e_{n}) commutes with pnorp_{\rm nor} for each nIn\in I, then limnιnor(en)(𝕂𝕏(M)J)\lim_{n}\iota_{\rm nor}(e_{n})\in({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*} is the identity, where the limit is in the weak topology.

Proof.

Since ιnor(𝕂𝕏(M))(𝕂𝕏(M)J)\iota_{\rm nor}({\mathbb{K}}_{\mathbb{X}}(M))\subset({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*} is weak dense and functionals in 𝕂𝕏(M)J{\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J} are continuous in ,1\|\cdot\|_{\infty,1} topology by [DKEP22, Proposition 3.1], we have ιnor(𝕏0)(𝕂𝕏(M)J)\iota_{\rm nor}({\mathbb{X}}_{0})\subset({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*} is also weak dense. Let e=limnιnor(en)(𝕂𝕏(M)J)e=\lim_{n}\iota_{\rm nor}(e_{n})\in({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*} be a weak limit point and for any T𝕏0T\in{\mathbb{X}}_{0}, we have

eιnor(T)=limnpnorι(en)ι(T)pnor=limnpnorι(enT)pnor=ιnor(T),e\iota_{\rm nor}(T)=\lim_{n}p_{\rm nor}\iota(e_{n})\iota(T)p_{\rm nor}=\lim_{n}p_{\rm nor}\iota(e_{n}T)p_{\rm nor}=\iota_{\rm nor}(T),

and similarly ιnor(T)e=ιnor(T)\iota_{\rm nor}(T)e=\iota_{\rm nor}(T). By density of ιnor(𝕏0)(𝕂𝕏(M)J)\iota_{\rm nor}({\mathbb{X}}_{0})\subset({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*}, we conclude that ee is the identity in (𝕂𝕏(M)J)({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*}. ∎

Lemma 3.8.

Let MM be a finite von Neumann algebra and NMN\subset M a von Neumann subalgebra. Let eN𝔹(L2M)e_{N}\in{\mathbb{B}}(L^{2}M) be the orthogonal projection onto L2NL^{2}N. Then ι(eN)𝔹(L2M)\iota(e_{N})\in{\mathbb{B}}(L^{2}M)^{**} commutes with pnorp_{\rm nor}.

Proof.

Suppose 𝔹(L2M)𝔹(){\mathbb{B}}(L^{2}M)^{**}\subset{\mathbb{B}}({\mathcal{H}}) and notice that ξ\xi{\mathcal{H}} is in the range of pnorp_{\rm nor} if and only if Mxι(x)ξ,ξM\ni x\to\langle\iota(x)\xi,\xi\rangle and JMJxι(x)ξ,ξJMJ\ni x\to\langle\iota(x)\xi,\xi\rangle are normal. For ξpnor\xi\in p_{\rm nor}{\mathcal{H}}, we have φ(x):=ι(x)ι(eB)ξ,ι(eN)ξ=ι(EN(x))ξ,ξ\varphi(x):=\langle\iota(x)\iota(e_{B})\xi,\iota(e_{N})\xi\rangle=\langle\iota(E_{N}(x))\xi,\xi\rangle is also normal for xMx\in M and JMJJMJ, which implies that ι(eN)pnor=pnorι(eN)pnor\iota(e_{N})p_{\rm nor}=p_{\rm nor}\iota(e_{N})p_{\rm nor}. It follows that ι(eN)\iota(e_{N}) and pnorp_{\rm nor} commutes. ∎

Lemma 3.9.

Let NMN\subset M be a mixing von Neumann subalgebra admitting a Pimsner-Popa basis {mk}\{m_{k}\} where mkMm_{k}\in M. Let 𝕏N{\mathbb{X}}_{N} be the associated boundary piece (see Example 3.1), and q𝕂(𝕂(L2M)J)q_{{\mathbb{K}}}\in({\mathbb{K}}(L^{2}M)_{J}^{\sharp})^{*}, q𝕏N(𝕂𝕏N(M)J)q_{{\mathbb{X}}_{N}}\in({\mathbb{K}}_{{\mathbb{X}}_{N}}(M)_{J}^{\sharp})^{*} be the respective identity elements. Then

k,lq𝕂ιnor(mkJmlJeNJmlJmk)=q𝕂q𝕏N.\sum_{k,l}q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}^{*}Je_{N}Jm_{l}Jm_{k}^{*})=q_{{\mathbb{K}}}^{\perp}q_{{\mathbb{X}}_{N}}.
Proof.

For notational simplicity, denote by pk,l=q𝕂ιnor(mkJmlJeNJmlJmk)p_{k,l}=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}^{*}Je_{N}Jm_{l}Jm_{k}^{*}). By mixing property of the inclusion NMN\subset M, we see that pk,lp_{k,l} are pairwise orthogonal projections. Indeed, if NMN\subset M is mixing, we have eNxJyJeNeNEN(x)JEN(y)JeN𝕂(M)e_{N}xJyJe_{N}-e_{N}E_{N}(x)JE_{N}(y)Je_{N}\in{\mathbb{K}}(M), i.e, is a compact operator when viewed as a bounded operator from the normed space MM to L2(M)L^{2}(M). Now we compute

pk,lpk,l\displaystyle p_{k,l}p_{k^{\prime},l^{\prime}} =q𝕂ιnor(mkJmlJeNJmlmlJmkmkeNJmlJmk)\displaystyle=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}Je_{N}Jm_{l}^{*}m_{l^{\prime}}Jm_{k}^{*}m_{k^{\prime}}e_{N}Jm_{l^{\prime}}^{*}Jm_{k^{\prime}}^{*})
=q𝕂ιnor(mkJmlJeNJqlJqkeNJmlJmk)δk,kδl,l\displaystyle=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}Je_{N}Jq_{l}Jq_{k}e_{N}Jm_{l^{\prime}}^{*}Jm_{k^{\prime}}^{*})\delta_{k,k^{\prime}}\delta_{l,l^{\prime}}
=q𝕂ιnor(mkJmlJeNJmlJmk)δk,kδl,l\displaystyle=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}^{*}Je_{N}Jm_{l^{\prime}}^{*}Jm_{k^{\prime}})\delta_{k,k^{\prime}}\delta_{l,l^{\prime}}

where ql𝒫(N)q_{l}\in\mathcal{P}(N) such that ql=EN(mlml)q_{l}=E_{N}(m_{l}^{*}m_{l}) and automatically satisfies mlql=mlm_{l}q_{l}=m_{l} (see Section 2.1).

Denote by 𝕏0𝔹(L2M){\mathbb{X}}_{0}\subset{\mathbb{B}}(L^{2}M) the hereditary C{\rm C}^{*}-subalgebra generated by xJyJeNxJyJe_{N} for x,yx,y in the C{\rm C}^{*}-algebra AA generated by {mka}aN,k\{m_{k}a\}_{a\in N,k\in\mathbb{N}}. It is clear that 𝕏0{\mathbb{X}}_{0} is an MM-boundary piece and note that AA is weakly dense (see Section 2.1, (2)) in MM.

Observe that 𝕂𝕏0,1(M)=𝕂𝕏N,1(M){\mathbb{K}}_{{\mathbb{X}}_{0}}^{\infty,1}(M)={\mathbb{K}}_{{\mathbb{X}}_{N}}^{\infty,1}(M) , where 𝕂𝕏0,1(M){\mathbb{K}}_{{\mathbb{X}}_{0}}^{\infty,1}(M) is obtained from 𝕏0{\mathbb{X}}_{0}. Notice that 𝔹(L2M)𝕏0𝕂𝕏L(M){\mathbb{B}}(L^{2}M){\mathbb{X}}_{0}\subset{\mathbb{K}}_{\mathbb{X}}^{L}(M) is dense in ,2\|\cdot\|_{\infty,2}. Indeed, for any contractions T𝔹(L2M)T\in{\mathbb{B}}(L^{2}M) and x,yMx,y\in M, we may find a net of contractions Ti𝔹(L2M)𝕏0T_{i}\in{\mathbb{B}}(L^{2}M){\mathbb{X}}_{0} such that TiTeNxJyJT_{i}\to Te_{N}xJyJ in ,2\|\cdot\|_{\infty,2}, as it follows directly from [DKEP22, Proposition 3.1]. It then follows that 𝕂𝕏0(M)𝕂𝕏,1(M){\mathbb{K}}_{{\mathbb{X}}_{0}}(M)\subset{\mathbb{K}}_{\mathbb{X}}^{\infty,1}(M) is dense in ,1\|\cdot\|_{\infty,1} and hence 𝕏0¯,1=𝕂𝕏0,1(M)=𝕂𝕏,1(M)\overline{{\mathbb{X}}_{0}}^{\infty,1}={\mathbb{K}}_{{\mathbb{X}}_{0}}^{\infty,1}(M)={\mathbb{K}}_{\mathbb{X}}^{\infty,1}(M) by [DKEP22, Proposition 3.6]. Note that pk,l(𝕂𝕏0(M)J)=(𝕂𝕏N(M)J)p_{k,l}\in({\mathbb{K}}_{{\mathbb{X}}_{0}}(M)_{J}^{\sharp})^{*}=({\mathbb{K}}_{{\mathbb{X}}_{N}}(M)_{J}^{\sharp})^{*} and pk,lq𝕂q𝕏Np_{k,l}\leq q_{{\mathbb{K}}}^{\perp}q_{{\mathbb{X}}_{N}}.

By the above paragraph it suffices to check the following: (k,lpk,l)ιnor(mkJmJaJbJeN)=q𝕂ιnor(mkJmJaJbJeN)(\sum_{k^{\prime},l^{\prime}}p_{k^{\prime},l^{\prime}})\iota_{\rm nor}(m_{k}Jm_{\ell}JaJbJe_{N})=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{\ell}JaJbJe_{N}) and ιnor(eNJmJmkaJbJ)(k,lpk,l)=q𝕂ιnor(eNJmJmkaJbJ)\iota_{\rm nor}(e_{N}Jm_{\ell}Jm_{k}aJbJ)(\sum_{k^{\prime},l^{\prime}}p_{k^{\prime},l^{\prime}})=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(e_{N}Jm_{\ell}Jm_{k}aJbJ) for all a,bNa,b\in N and k,lk,l\in\mathbb{N}. Indeed, every element in 𝕏0{\mathbb{X}}_{0} can be written as a norm limit of linear spans consisting of elements of the from x1Jy1JTJy2Jx2x_{1}Jy_{1}JTJy_{2}Jx_{2}, where xi,yiAx_{i},y_{i}\in A and T𝔹(L2N)T\in{\mathbb{B}}(L^{2}N) Further we can assume xi=mkax_{i}=m_{k}a with aNa\in N from density. Then we will get that for all z𝕏0z\in{\mathbb{X}}_{0}, k,lq𝕂ιnor(mkJmlJeNJmlJmk)ιnor(z)=q𝕂ιnor(z)\sum_{k,l}q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}^{*}Je_{N}Jm_{l}Jm_{k}^{*})\iota_{\rm nor}(z)=q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(z) and since ιnor(𝕏0)\iota_{\rm nor}({\mathbb{X}}_{0}) is weak dense in (𝕂𝕏0(M)J)=(𝕂𝕏N(M)J)({\mathbb{K}}_{{\mathbb{X}}_{0}}(M)_{J}^{\sharp})^{*}=({\mathbb{K}}_{{\mathbb{X}}_{N}}(M)_{J}^{\sharp})^{*} by the previous paragraph, so we get that k,lq𝕂ιnor(mkJmlJeNJmlJmk)=q𝕂q𝕏N\sum_{k,l}q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{l}^{*}Je_{N}Jm_{l}Jm_{k}^{*})=q_{{\mathbb{K}}}^{\perp}q_{{\mathbb{X}}_{N}}.

The above equality holds by a simple computation

pk,lιnor(mkJmJaJbJeN)=δk,kδl,lq𝕂ιnor(mkJmJaJbJeN)p_{k^{\prime},l^{\prime}}\iota_{\rm nor}(m_{k}Jm_{\ell}JaJbJe_{N})=\delta_{k,k^{\prime}}\delta_{l,l^{\prime}}q_{{\mathbb{K}}}^{\perp}\iota_{\rm nor}(m_{k}Jm_{\ell}JaJbJe_{N})

as in the beginning of this proof wherein we verified that pk,lp_{k,l} are projections.∎

Lemma 3.10.

Let MM be a finite von Neumann algebra and MiMM_{i}\subset M, i=1,ni=1,\dots n be von Neumann subalgebras such that eMie_{M_{i}} are pairwise commuting. Let 𝕏{\mathbb{X}} denote the boundary piece associated to {Mi}i=1n\{M_{i}\}_{i=1}^{n} as in Example 3.1. Let 𝕏i{\mathbb{X}}_{i} denote the boundary pieces associated to MiM_{i}. Let qiq_{i} denote the identities of the von Neumann algebras (𝕂𝕏i(M)J)({\mathbb{K}}_{{\mathbb{X}}_{i}}(M)_{J}^{\sharp})^{*} and q𝕏q_{\mathbb{X}} denote the identity of (𝕂𝕏(M)J)({\mathbb{K}}_{\mathbb{X}}(M)_{J}^{\sharp})^{*}. Then we have that q𝕏=i=1nqiq_{\mathbb{X}}=\vee_{i=1}^{n}q_{i}.

Proof.

Recall from the beginning of this section that (𝕂𝕏(M)J)({\mathbb{K}}_{\mathbb{X}}(M)_{J}^{\sharp})^{*} is a von Neumann algebra, as M,JMJM,JMJ are in the multiplier algebra of 𝕄(𝕂𝕏(M)){\mathbb{M}}({\mathbb{K}}_{\mathbb{X}}(M)). It is easy to see that q𝕏qiq_{\mathbb{X}}\geq q_{i} for each ii. Now we show that q𝕏i=1nqiq_{\mathbb{X}}\leq\vee_{i=1}^{n}q_{i}. Fix an increasing family of finite subsets of unitaries in MM, FnF_{n} such that {nFn}′′=M\{\bigcup_{n}F_{n}\}^{\prime\prime}=M. Let en=u,vFnuJvJ(ieMi)JvJue_{n}=\vee_{u,v\in F_{n}}uJvJ(\vee_{i}e_{M_{i}})Jv^{*}Ju^{*}. Clearly we have that ιnor(en)i=1nqi\iota_{\rm nor}(e_{n})\leq\vee_{i=1}^{n}q_{i}. Indeed, see that

ιnor(u,vFnuJvJeMiJvJu)<qi\iota_{\rm nor}(\vee_{u,v\in F_{n}}uJvJe_{M_{i}}Jv^{*}Ju^{*})<q_{i}

and then ιnor(en)=ιnor(iu,vFnuJvJeMiJvJu)i=1nqi\iota_{\rm nor}(e_{n})=\iota_{\rm nor}(\vee_{i}\vee_{u,v\in F_{n}}uJvJe_{M_{i}}Jv^{*}Ju^{*})\leq\vee_{i=1}^{n}q_{i}. From Lemmas 3.3 and 3.7 we see that q𝕏=limnιnor(en)i=1nqiq_{{\mathbb{X}}}=\lim_{n}\iota_{\rm nor}(e_{n})\leq\vee_{i=1}^{n}q_{i} as required. ∎

3.3. Induced boundary pieces in the bidual

Lemma 3.11.

Let MM be a finite von Neumann algebra, 𝕏{\mathbb{X}} an MM-boundary piece, and NpMpN\subset pMp a von Neumann subalgebra for some 0p𝒫(M)0\neq p\in{\mathcal{P}}(M). Set E:=Ad(eN)Ad(pJpJ):𝔹(L2M)𝔹(L2N)E:=\operatorname{Ad}(e_{N})\circ\operatorname{Ad}(pJpJ):{\mathbb{B}}(L^{2}M)\to{\mathbb{B}}(L^{2}N). Then its restriction E|𝕊(M)E_{|{\mathbb{S}}(M)} maps 𝕊𝕏(M){\mathbb{S}}_{\mathbb{X}}(M) to 𝕊(N){\mathbb{S}}(N). Moreover, there exists a u.c.p. map E~:𝕊~(M)𝕊~(N)\tilde{E}:\tilde{\mathbb{S}}(M)\to\tilde{\mathbb{S}}(N) such that E~M\tilde{E}_{\mid M} agrees with the conditional expectation from MM to NN.

Proof.

To see E|𝕊𝕏(M):𝕊(M)𝕊(N)E_{|{\mathbb{S}}_{\mathbb{X}}(M)}:{\mathbb{S}}(M)\to{\mathbb{S}}(N), note that pJpJeNJNaJN=JaJpJpJeNpJpJe_{N}J_{N}aJ_{N}=JaJpJpJe_{N} for any aNa\in N. and E:𝔹(L2M)𝔹(L2N)E:{\mathbb{B}}(L^{2}M)\to{\mathbb{B}}(L^{2}N) is ,1\|\cdot\|_{\infty,1}-continuous. Thus for any T𝕊𝕏(M)T\in{\mathbb{S}}_{\mathbb{X}}(M) and any aNa\in N, we have

[E(T),JNaJN]=E([T,JaJ])E(𝕂(M)¯,1)=𝕂(N)¯,1=𝕂,1(N),[E(T),J_{N}aJ_{N}]=E([T,JaJ])\in E(\overline{{\mathbb{K}}(M)}^{\|\cdot\|_{\infty,1}})=\overline{{\mathbb{K}}(N)}^{\|\cdot\|_{\infty,1}}={\mathbb{K}}^{\infty,1}(N),

i.e., E(T)𝕊(N)E(T)\in{\mathbb{S}}_{(}N).

Note that E:𝔹(L2N)𝔹(L2M)E^{*}:{\mathbb{B}}(L^{2}N)^{*}\to{\mathbb{B}}(L^{2}M)^{*} maps 𝔹(L2N)J{\mathbb{B}}(L^{2}N)^{\sharp}_{J} to 𝔹(L2M)J{\mathbb{B}}(L^{2}M)^{\sharp}_{J} by [DKEP22, Lemma 5.3], and similarly E:(𝕂(L2N))J(𝕂(L2M))JE^{*}:({\mathbb{K}}(L^{2}N))^{\sharp}_{J}\to({\mathbb{K}}(L^{2}M))^{\sharp}_{J}. Therefore E~:=(E|𝔹(L2N)J):(𝔹(L2M)J)(𝔹(L2N)J)\tilde{E}:=({E^{*}}_{|{\mathbb{B}}(L^{2}N)^{\sharp}_{J}})^{*}:({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}\to({\mathbb{B}}(L^{2}N)^{\sharp}_{J})^{*} and E~(𝕂(L2M))J):(𝕂(L2M)J)(𝕂(L2N)J)\tilde{E}_{\mid({\mathbb{K}}(L^{2}M))^{\sharp}_{J})^{*}}:({\mathbb{K}}(L^{2}M)^{\sharp}_{J})^{*}\to({\mathbb{K}}(L^{2}N)^{\sharp}_{J})^{*}. Hence we conclude that E~:𝕊~(M)𝕊~(N)\tilde{E}:\tilde{\mathbb{S}}(M)\to\tilde{\mathbb{S}}(N) with E~M\tilde{E}_{\mid M} agrees with the conditional expectation from MM to NN. ∎

3.4. Relative biexactness and relative proper proximality

Given a countable discrete group Γ\Gamma, a boundary piece II is a Γ×Γ\Gamma\times\Gamma invariant closed ideal such that c0ΓIΓc_{0}\Gamma\subset I\subset\ell^{\infty}\Gamma [BIP18]. The small at infinity compactification of Γ\Gamma relative to II is the spectrum of the C{\rm C}^{*}-algebra 𝕊I(Γ)={fΓfRtfI,foranytΓ}{\mathbb{S}}_{I}(\Gamma)=\{f\in\ell^{\infty}\Gamma\mid f-R_{t}f\in I,{\rm\ for\ any\ }t\in\Gamma\}. Recall that Γ\Gamma is said to be biexact relative to XX if Γ𝕊I(Γ)/I\Gamma{\,\curvearrowright\,}{\mathbb{S}}_{I}(\Gamma)/I is topologically amenable [Oza04], [BO08, Chapter 15], [BIP18]. We remark that this is equivalent to Γ𝕊I(Γ)\Gamma{\,\curvearrowright\,}{\mathbb{S}}_{I}(\Gamma) is amenable. Indeed, since we may embed ΓI\ell^{\infty}\Gamma\hookrightarrow I^{**} in a Γ\Gamma-equivariant way, we have ΓI(𝕊I(Γ)/I)=𝕊I(Γ)\Gamma{\,\curvearrowright\,}I^{**}\oplus({\mathbb{S}}_{I}(\Gamma)/I)^{**}={\mathbb{S}}_{I}(\Gamma)^{**} is amenable, and it follows that Γ𝕊I(Γ)\Gamma{\,\curvearrowright\,}{\mathbb{S}}_{I}(\Gamma) is an amenable action [BEW19, Proposition 2.7].

The following is a general version of [DKEP22, Theorem 7.1], whose proof follows similarly. For the convenience of the reader we include the proof sketch below. A more general version of this is obtained in the upcoming work [DP22].

Theorem 3.12.

Let M=LΓM=L\Gamma where Γ\Gamma is an nonamenable group that is biexact relative to a finite family of subgroups {Λi}iI\{\Lambda_{i}\}_{i\in I}. Denote by 𝕏{\mathbb{X}} the MM-boundary piece associated with {LΛi}iI\{L\Lambda_{i}\}_{i\in I}. If ApMpA\subset pMp for some 0p𝒫(M)0\neq p\in\mathcal{P}(M) such that AA has no amenable direct summands, then AA is properly proximal relative to 𝕏A{\mathbb{X}}^{A}, where 𝕏A{\mathbb{X}}^{A} is the induced AA-boundary piece as in Remark 3.4).

Proof.

Consider the Γ\Gamma-equivariant diagonal embedding (Γ)𝔹(2Γ)\ell^{\infty}(\Gamma)\subset{\mathbb{B}}(\ell^{2}\Gamma). Note that under this embedding c0(Γ,{Λi}iI)c_{0}(\Gamma,\{\Lambda_{i}\}_{i\in I}) is sent to 𝕏{\mathbb{X}}. Denote by 𝕊𝕏(Γ)={f(Γ)|ffgc0(Γ,{Λi}iI),gΓ}{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)=\{f\in\ell^{\infty}(\Gamma)|\ f-fg\in c_{0}(\Gamma,\{\Lambda_{i}\}_{i\in I}),\ \forall g\in\Gamma\}, the relative small at infinity compactification at the group level. Restricting this embedding to 𝕊𝕏(Γ){\mathbb{S}}_{{\mathbb{X}}}(\Gamma) then gives a Γ\Gamma-equivariant embedding into 𝕊𝕏(M){\mathbb{S}}_{\mathbb{X}}(M). Therefore we obtain a *-homomorphism from 𝕊𝕏(Γ)rΓ𝔹(2(Γ)){\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma\to{\mathbb{B}}(\ell^{2}(\Gamma)) whose image is contained in 𝕊𝕏(M){\mathbb{S}}_{{\mathbb{X}}}(M). Composing this with the map EE from Lemma 3.11, we obtain a u.c.p map ϕ:𝕊𝕏(Γ)rΓ𝕊𝕏A(A)\phi:{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma\to{\mathbb{S}}_{{\mathbb{X}}^{A}}(A). By hypothesis we have a projection p0𝒵(A)p_{0}\in\mathcal{Z}(A) and an Ap0Ap_{0} bimodular u.c.p map Φ:𝕊𝕏A(A)Ap0\Phi:{\mathbb{S}}_{{\mathbb{X}}^{A}}(A)\to Ap_{0}. Further composing with this map we obtain a u.c.p map from ϕ~:𝕊𝕏(Γ)rΓAp0\widetilde{\phi}:{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma\to Ap_{0}.

Now set φ:𝕊𝕏(Γ)rΓ\varphi:{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma\to{\mathbb{C}}, by φ(x):=xp0^,p0^τ(p)\varphi(x):=\frac{\langle x\widehat{p_{0}},\widehat{p_{0}}\rangle}{\tau(p)}. We then get a representation πφ:𝕊𝕏(Γ)rΓφ\pi_{\varphi}:{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma\to\mathcal{H}_{\varphi} and a state φ~𝔹(φ)\widetilde{\varphi}\in{\mathbb{B}}(\mathcal{H}_{\varphi})_{*} such that φ=φ~πφ\varphi=\widetilde{\varphi}\circ\pi_{\varphi}. Since Cr(Γ)C^{*}_{r}(\Gamma) is weakly dense in MM, we see by an argument of Boutonnet-Carderi (see Propositon 4.1 in [BC15]) that there is a projection q(πφ(𝕊𝕏(Γ)rΓ))′′q\in(\pi_{\varphi}({\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma))^{\prime\prime} such that φ~(q)=1\widetilde{\varphi}(q)=1 and there exists a normal unital *-homomorphism ι:L(Γ)qπφ(𝕊𝕏(Γ)rΓ))′′q\iota:L(\Gamma)\to q\pi_{\varphi}({\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma))^{\prime\prime}q.

Since Γ\Gamma is biexact relative to 𝕏{\mathbb{X}}, we have that 𝕊𝕏(Γ)rΓ{\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma is a nuclear CC^{*}-algebra. Therefore there is a u.c.p map ι~:𝔹(2(Γ))q(𝕊𝕏(Γ)rΓ)′′q\widetilde{\iota}:{\mathbb{B}}(\ell^{2}(\Gamma))\to q({\mathbb{S}}_{{\mathbb{X}}}(\Gamma)\rtimes_{r}\Gamma)^{\prime\prime}q extending ι\iota. Now we see that φ~ι~\widetilde{\varphi}\circ\widetilde{\iota} is an Ap0Ap_{0} central state on 𝔹(2(Γ)){\mathbb{B}}(\ell^{2}(\Gamma)) showing that AA has an amenable direct summand, which is a contradiction. ∎

In the case of general free products of finite von Neumann algebras M=M1M2M=M_{1}*M_{2} it ought to be the case that that if AMA\subset M such that AA has no amenable direct summand, then AMA\subset M is properly proximal relative to the boundary piece generated by M1M_{1} and M2M_{2}. However, currently we are only able to obtain this with an additional technical assumption that MiLΓiM_{i}\cong L\Gamma_{i} where Γi\Gamma_{i} are exact, so that Γ1Γ2\Gamma_{1}\ast\Gamma_{2} is biexact relative to {Γ1,Γ2}\{\Gamma_{1},\Gamma_{2}\} [BO08, Proposition 15.3.12]. We record below a general result about subalgebras in free products which follows essentially from Theorem 9.1 in [DKEP22], however we do not get the boundary piece associated to the subalgebras MiM_{i}. We instead get the boundary piece associated to the word length:

Let M1M_{1}, M2M_{2} be two finite von Neumann algebras and M=M1M2M=M_{1}*M_{2} be the tracial free product. Let AMA\subset M be a nonamenable subalgebra. Consider the free product deformation from [IPP08], i.e., M~=ML𝔽2\tilde{M}=M\ast L\mathbb{F}_{2}, θt=Ad(u1t)Ad(u2t)Aut(M~)\theta_{t}=\operatorname{Ad}(u_{1}^{t})\ast\operatorname{Ad}(u_{2}^{t})\in{\rm Aut}(\tilde{M}), with u1t=exp(itα1)u_{1}^{t}=\exp(it\alpha_{1}), u2t=exp(itα2)u_{2}^{t}=\exp(it\alpha_{2}), where α1\alpha_{1}, α2\alpha_{2} are selfadjoint element in L𝔽2L\mathbb{F}_{2} such that exp(iα1)=u1\exp(i\alpha_{1})=u_{1}, exp(iα2)=u2\exp(i\alpha_{2})=u_{2} and u1u_{1}, u2u_{2} are Haar unitaries in L𝔽2L\mathbb{F}_{2}. For t>0t>0, we have EMαt=P0+n=1(sin(πt)/πt)2nPnE_{M}\circ\alpha_{t}=P_{0}+\sum_{n=1}^{\infty}(\sin(\pi t)/\pi t)^{2n}P_{n} (see Section 2.5 in [Ioa15]), where PnP_{n} is the orthogonal projection to n=(i1,,in)SnL2(Mi1)L2(Min)\mathcal{H}_{n}=\oplus_{(i_{1},\cdots,i_{n})\in S_{n}}L^{2}(M_{i_{1}}\ominus\mathbb{C})\otimes\cdots\otimes L^{2}(M_{i_{n}}\ominus\mathbb{C}) and SnS_{n} is the set of alternating sequences of length nn. Consider the hereditary CC^{*}-algebra 𝕏F\mathbb{X}_{F} generated by {Pn}n0\{P_{n}\}_{n\geq 0}.

Proposition 3.13.

In the above setup, there exists a projection pAp\in A such that ApAp is amenable and ApAp^{\perp} is properly proximal relative to 𝕏F{\mathbb{X}}_{F}.

Proof.

It follows from the proof of [DKEP22, Proposition 9.1] that there exists an MM-bimodular u.c.p. map ϕ:(Mop)𝔹(L2M~L2M)𝕊~𝕏F(M)\phi:(M^{\rm op})^{\prime}\cap\mathbb{B}(L^{2}\tilde{M}\ominus L^{2}M)\to\tilde{\mathbb{S}}_{\mathbb{X}_{F}}(M). Moreover, since L2M~L2ML2M¯𝒦L^{2}\tilde{M}\ominus L^{2}M\cong L^{2}M\overline{\otimes}\mathcal{K} as an MM-MM bimodule for some right MM module 𝒦\mathcal{K} [Ioa15, Lemma 2.10], we may restrict ϕ\phi to 𝔹(L2M)id𝒦\mathbb{B}(L^{2}M)\otimes\operatorname{id}_{\mathcal{K}}. Take p𝒵(A)p\in\mathcal{Z}(A) to be the maximal projection such that ApAp is amenable and p1p\neq 1 as AA is nonamenable. If ApAp^{\perp} is not properly proximal relative to 𝕏\mathbb{X} inside MM, i.e., there exists an AA-central state φ\varphi on 𝕊~𝕏F(M)\tilde{\mathbb{S}}_{\mathbb{X}_{F}}(M) which is normal when restricted to pMpp^{\perp}Mp^{\perp}. Then pick q𝒵(Ap)q\in\mathcal{Z}(Ap^{\perp}) be the support projection of (φϕ)Ap(\varphi\circ\phi)_{\mid Ap^{\perp}} and we have AqAq is amenable, which contradicts the maximality of pp.

4. The Upgrading Theorem

Proof of Theorem 1.1.

First notice that since AA is properly proximal relative to 𝕏{\mathbb{X}} inside MM, it has no amenable direct summand by (3) of Remark 3.4. Let f𝒵(A)f\in\mathcal{Z}(A) be the projection such that AfAf^{\perp} is the maximal properly proximal direct summand of AA by (4) of Remark 3.4, and we may assume f0f\neq 0 since otherwise AA would be properly proximal. Therefore AfAf has no amenable direct summand, is properly proximal relative to 𝕏{\mathbb{X}} inside MM by (2) of Remark 3.4 and has no properly proximal direct summand. It follows from Lemma 3.5 that there exists an AfAf-central state μ\mu on 𝕊~(Af)\widetilde{{\mathbb{S}}}(Af) such that μAf\mu_{\mid Af} is normal. Moreover, by a maximal argument, we may assume μ𝒵(Af)\mu_{\mid\mathcal{Z}(Af)} is faithful, as AfAf has no properly proximal direct summand.

Let E~:𝕊~(M)𝕊~(Af)\tilde{E}:\tilde{\mathbb{S}}(M)\to\tilde{\mathbb{S}}(Af) be the u.c.p. map as in Lemma 3.11. Define a state φ=μE~:𝕊~(M)\varphi=\mu\circ\tilde{E}:\tilde{\mathbb{S}}(M)\to{\mathbb{C}}, and it follows that φ\varphi is AfAf-central and φfMf\varphi_{\mid fMf} is a faithful normal state. Let q𝕂q_{\mathbb{K}} be the identity of the von Neumann algebra (𝕂(L2M)J)(𝔹(L2M)J)({\mathbb{K}}(L^{2}M)^{\sharp}_{J})^{*}\subset({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}, q𝕏q_{\mathbb{X}} the identity of von Neumann algebra (𝕂𝕏(M)J)(𝔹(L2M)J)({\mathbb{K}}_{\mathbb{X}}(M)^{\sharp}_{J})^{*}\subset({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}. Note that q𝕂q𝕏q_{\mathbb{K}}\leq q_{\mathbb{X}} as 𝕂(L2M)𝕂𝕏(M){\mathbb{K}}(L^{2}M)\subset{\mathbb{K}}_{\mathbb{X}}(M).

First we analyze the support of φ\varphi. Observe that φ(q𝕂)=1\varphi(q_{{\mathbb{K}}}^{\perp})=1. Indeed, if φ(q𝕂)>0\varphi(q_{{\mathbb{K}}})>0, i.e., φ\varphi does not vanish on (𝕂(L2M)J)(\mathbb{K}(L^{2}M)^{\sharp}_{J})^{*}, then we may restrict φ\varphi to 𝔹(L2M)\mathbb{B}(L^{2}M), which embeds into (𝕂(L2M)J)(\mathbb{K}(L^{2}M)^{\sharp}_{J})^{*} as a normal operator MM-system [DKEP22, Section 8], and this shows that AfAf would have an amenable direct summand. Moreover, we have φ(q𝕏)=1\varphi(q_{{\mathbb{X}}})=1. Indeed, if φ(q𝕏)>0\varphi(q_{{\mathbb{X}}}^{\perp})>0, then

1μ(q𝕏)φAd(q𝕏):𝕊~𝕏(M)\frac{1}{\mu(q_{\mathbb{X}}^{\perp})}\varphi\circ\operatorname{Ad}(q_{\mathbb{X}}^{\perp}):\tilde{\mathbb{S}}_{\mathbb{X}}(M)\to\mathbb{C}

would be an AfAf-central that restricts to a normal state on fMffMf. Since 𝕊𝕏(M){\mathbb{S}}_{\mathbb{X}}(M) naturally embeds into 𝕊~𝕏(M)\widetilde{{\mathbb{S}}}_{\mathbb{X}}(M), this contradicts that AfAf is properly proximal relative to 𝕏{\mathbb{X}} inside MM. Therefore we conclude that φ(q𝕏q𝕂)=1.\varphi(q_{\mathbb{X}}q_{\mathbb{K}}^{\perp})=1.

For each 1in1\leq i\leq n, denote by 𝕏i:=𝕏Mi𝔹(L2M){\mathbb{X}}_{i}:={\mathbb{X}}_{M_{i}}\subset{\mathbb{B}}(L^{2}M) the MM-boundary piece associated with MiM_{i} and qi(𝕂𝕏i(M)J)q_{i}\in({\mathbb{K}}_{{\mathbb{X}}_{i}}(M)^{\sharp}_{J})^{*} the identity. Since i=1nqi=q𝕏\vee_{i=1}^{n}q_{i}=q_{\mathbb{X}} by Lemma 3.10, we have φ(qjq𝕂)>0\varphi(q_{j}q_{{\mathbb{K}}}^{\perp})>0 for some 1jn1\leq j\leq n

Claim: there exists a u.c.p. map ϕ:M,eMjq𝕂qj𝕊~(M)qj\phi:\langle M,e_{M_{j}}\rangle\to q_{\mathbb{K}}^{\perp}q_{j}\tilde{\mathbb{S}}(M)q_{j} such that ϕ(x)=q𝕂qjx\phi(x)=q_{\mathbb{K}}^{\perp}q_{j}x for any xMx\in M.

Proof of the claim..

Denote by {mk}k0M\{m_{k}\}_{k\geq 0}\subset M a bounded Pimsner-Popa basis of MM over MiM_{i}. For each n0n\geq 0, consider the u.c.p. map ψn:M,eMjM,eMj\psi_{n}:\langle M,e_{M_{j}}\rangle\to\langle M,e_{M_{j}}\rangle given by

ψn(x)=(knmkeMjmk)x(nmeMjm),\psi_{n}(x)=(\sum_{k\leq n}m_{k}e_{M_{j}}m_{k}^{*})x(\sum_{\ell\leq n}m_{\ell}e_{M_{j}}m_{\ell}^{*}),

and notice that ψn\psi_{n} maps M,eMj\langle M,e_{M_{j}}\rangle into the *-subalgebra A0:=sp{mkaeMjmaMj,k,0}A_{0}:={\rm sp}\{m_{k}ae_{M_{j}}m_{\ell}^{*}\mid a\in M_{j},k,\ell\geq 0\}.

Recall notations from Remark 3.6. By Lemma 3.8, we have

{ιnor(JmkJeMjJmkJ)}k0(𝔹(L2M)J)\{\iota_{\rm nor}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)\}_{k\geq 0}\subset({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}

is a family of pairwise orthogonal projections. Set

ej=k0ιnor(JmkJeMjJmkJ)(𝔹(L2M)J)e_{j}=\sum_{k\geq 0}\iota_{\rm nor}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)\in({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}

and define the map

ϕ0:A0\displaystyle\phi_{0}:A_{0} q𝕂(𝔹(L2M)J)\displaystyle\to q_{{\mathbb{K}}}^{\perp}({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*}
mraeMjm\displaystyle m_{r}ae_{M_{j}}m_{\ell}^{*} q𝕂ιnor(mra)ejιnor(m).\displaystyle\mapsto q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(m_{r}a)e_{j}\iota_{\rm nor}(m_{\ell}^{*}).

It is easy to check that ϕ0\phi_{0} is well-defined. We then check that ϕ0\phi_{0} is a *-homomorphism. It suffices to show that for any xMx\in M, we have

(1) q𝕂ejιnor(x)ej=q𝕂ιnor(EMj(x))ej.q_{\mathbb{K}}^{\perp}e_{j}\iota_{\rm nor}(x)e_{j}=q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(E_{M_{j}}(x))e_{j}.

Now we compute,

q𝕂ejιnor(x)ej\displaystyle q_{\mathbb{K}}^{\perp}e_{j}\iota_{\rm nor}(x)e_{j}
=\displaystyle= q𝕂k,0ιnor((JmkJeMjJmkJ)x(JmJeMjJmJ))\displaystyle q_{\mathbb{K}}^{\perp}\sum_{k,\ell\geq 0}\iota_{\rm nor}\big{(}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)x(Jm_{\ell}Je_{M_{j}}Jm_{\ell}^{*}J)\big{)}
=\displaystyle= q𝕂k0ιnor((JmkJeMjJmkJ)x(JmkJeMjJmkJ))+kιnor((JmkJeMjJmkJ)x(JmJeMjJmJ)).\displaystyle q_{\mathbb{K}}^{\perp}\sum_{k\geq 0}\iota_{\rm nor}\big{(}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)x(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)\big{)}+\sum_{k\neq\ell}\iota_{\rm nor}\big{(}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)x(Jm_{\ell}Je_{M_{j}}Jm_{\ell}^{*}J)\big{)}.

By Remark 2.3, we have (JmkJeMjJmkJ)(xEMj(x))(JmJeMjJmJ)𝔹(L2M)(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)(x-E_{M_{j}}(x))(Jm_{\ell}Je_{M_{j}}Jm_{\ell}^{*}J)\in{\mathbb{B}}(L^{2}M) is a compact operator from MM to L2ML^{2}M for kk\neq\ell. Since (JmkJeMjJmkJ)EMj(x)(JmJeMjJmJ)=0(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)E_{M_{j}}(x)(Jm_{\ell}Je_{M_{j}}Jm_{\ell}^{*}J)=0 if k\ell\neq k, we have kq𝕂ιnor(JmkJeMjJmkJxJmJeMjJmJ)=0\sum_{k\neq\ell}q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}JxJm_{\ell}Je_{M_{j}}Jm_{\ell}^{*}J)=0. Similarly, one checks that q𝕂ιnor((JmkJeMjJmkJ)x(JmkJeMjJmkJ))=q𝕂ιnor(EMj(x)(JmkJeMjJmkJ))q_{\mathbb{K}}^{\perp}\iota_{\rm nor}\big{(}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)x(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)\big{)}=q_{\mathbb{K}}^{\perp}\iota_{\rm nor}\big{(}E_{M_{j}}(x)(Jm_{k}Je_{M_{j}}Jm_{k}^{*}J)\big{)}.

It then follows from (1) that ϕ0\phi_{0} is a *-homomorphism. Now we verify that ϕ0\phi_{0} is norm continuous.

Given i=1dmkiaieMjmiA0\sum_{i=1}^{d}m_{k_{i}}a_{i}e_{M_{j}}m_{\ell_{i}}^{*}\in A_{0}, we may assume that kikjk_{i}\neq k_{j} and ij\ell_{i}\neq\ell_{j} if iji\neq j. Consider Pk=q𝕂i=1dιnor(JmkJmieMjmiJmkJ)P_{k}=q_{\mathbb{K}}^{\perp}\sum_{i=1}^{d}\iota_{\rm nor}(Jm_{k}Jm_{\ell_{i}}e_{M_{j}}m_{\ell_{i}}^{*}Jm_{k}^{*}J) and Qk=q𝕂i=1dιnor(JmkJmkieMjmkiJmkJ)Q_{k}=q_{\mathbb{K}}^{\perp}\sum_{i=1}^{d}\iota_{\rm nor}(Jm_{k}Jm_{k_{i}}e_{M_{j}}m_{k_{i}}^{*}Jm_{k}^{*}J). We have PkP_{k} and QkQ_{k} are a projections and PkPr=QkQr=0P_{k}P_{r}=Q_{k}Q_{r}=0 if krk\neq r by Remark 2.3. And for the same reason, we have ιnor(eMjmiJmkJ)Pk=q𝕂ιnor(eMjmiJmkJ)\iota_{\rm nor}(e_{M_{j}}m_{\ell_{i}}^{*}Jm_{k}^{*}J)P_{k}=q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(e_{M_{j}}m_{\ell_{i}}^{*}Jm_{k}^{*}J) as well as ιnor(eMjmkiJmkJ)Qk=q𝕂ιnor(eMjmkiJmkJ)\iota_{\rm nor}(e_{M_{j}}m_{k_{i}}^{*}Jm_{k}^{*}J)Q_{k}=q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(e_{M_{j}}m_{k_{i}}^{*}Jm_{k}^{*}J) for each 1id1\leq i\leq d. Let {\mathcal{H}} be the Hilbert space where (𝔹(L2M)J)({\mathbb{B}}(L^{2}M)^{\sharp}_{J})^{*} is represented on. For ξ,η()1\xi,\eta\in({\mathcal{H}})_{1}, we compute

|ϕ0(i=1dmkiaieMjmi)ξ,η|\displaystyle|\langle\phi_{0}(\sum_{i=1}^{d}m_{k_{i}}a_{i}e_{M_{j}}m_{\ell_{i}}^{*})\xi,\eta\rangle| k0|i=1dq𝕂ιnor(eMjmiJmkJ)ξ,ιnor(JmkJmkieMjai)η|\displaystyle\leq\sum_{k\geq 0}|\sum_{i=1}^{d}\langle q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(e_{M_{j}}m_{\ell_{i}}^{*}Jm_{k}^{*}J)\xi,\iota_{\rm nor}(Jm_{k}Jm_{k_{i}}e_{M_{j}}a_{i})^{*}\eta\rangle|
=k0|i=1dιnor(eMjmiJmkJ)Pkξ,ιnor(JmkJmkieMjai)Qkη|\displaystyle=\sum_{k\geq 0}|\sum_{i=1}^{d}\langle\iota_{\rm nor}(e_{M_{j}}m_{\ell_{i}}^{*}Jm_{k}^{*}J)P_{k}\xi,\iota_{\rm nor}(Jm_{k}Jm_{k_{i}}e_{M_{j}}a_{i})^{*}Q_{k}\eta\rangle|
k0ιnor(JmkJ(i=1dmkiaieMjmi)JmkJ)PkξQkη\displaystyle\leq\sum_{k\geq 0}\|\iota_{\rm nor}(Jm_{k}J(\sum_{i=1}^{d}m_{k_{i}}a_{i}e_{M_{j}}m_{\ell_{i}}^{*})Jm_{k}^{*}J)\|\|P_{k}\xi\|\|Q_{k}\eta\|
(supkmk2)i=1dmkiaieMjmi(k0Pkξ2)1/2(k0Qkξ2)1/2\displaystyle\leq(\sup_{k\in\mathbb{N}}\|m_{k}\|^{2})\|\sum_{i=1}^{d}m_{k_{i}}a_{i}e_{M_{j}}m_{\ell_{i}}^{*}\|(\sum_{k\geq 0}\|P_{k}\xi\|^{2})^{1/2}(\sum_{k\geq 0}\|Q_{k}\xi\|^{2})^{1/2}
(supkmk2)i=1dmkiaieMjmi.\displaystyle\leq(\sup_{k\in\mathbb{N}}\|m_{k}\|^{2})\|\sum_{i=1}^{d}m_{k_{i}}a_{i}e_{M_{j}}m_{\ell_{i}}^{*}\|.

This shows that ϕ0\phi_{0} is norm continuous as required.

Lastly we show that ϕ0\phi_{0} maps into q𝕂𝕊~(M)q_{\mathbb{K}}^{\perp}\tilde{\mathbb{S}}(M). It suffices to show that [ej,ιnor(JmuJ)]=0[e_{j},\iota_{\rm nor}(Jm_{\ell}uJ)]=0 for all \ell\in\mathbb{N} and u𝒰(Mj)u\in\mathcal{U}(M_{j}), since ϕ0(A0)\phi_{0}(A_{0}) commutes with ιnor(JMJ)\iota_{\rm nor}(JMJ).

Without loss of generality, we may assume that m0=1m_{0}=1. We compute

=q𝕂(k0ιnor(JmkJeMjJmkmuJ)ιnor(JmumkJeMjJmkJ))\displaystyle=q_{\mathbb{K}}^{\perp}\big{(}\sum_{k\geq 0}\iota_{\rm nor}(Jm_{k}Je_{M_{j}}Jm_{k}^{*}m_{\ell}uJ)-\iota_{\rm nor}(Jm_{\ell}um_{k}Je_{M_{j}}Jm_{k}^{*}J)\big{)}
=q𝕂(k0ιnor(JmkeMjmkmeMjueMjJ)ιnor(JmeMjueMjmkeMjmkJ))\displaystyle=q_{\mathbb{K}}^{\perp}\big{(}\sum_{k\geq 0}\iota_{\rm nor}(Jm_{k}e_{M_{j}}m_{k}^{*}m_{\ell}e_{M_{j}}ue_{M_{j}}J)-\iota_{\rm nor}(Jm_{\ell}e_{M_{j}}ue_{M_{j}}m_{k}e_{M_{j}}m_{k}^{*}J)\big{)}
=q𝕂(ιnor(JmuJeMj)ιnor(JmuJeMj))=0.\displaystyle=q_{\mathbb{K}}^{\perp}\big{(}\iota_{\rm nor}(Jm_{\ell}uJe_{M_{j}})-\iota_{\rm nor}(Jm_{\ell}uJe_{M_{j}})\big{)}=0.

Combining all the above arugments, we may extend ϕ0:Aq𝕂𝕊~(M)\phi_{0}:A\to q_{\mathbb{K}}^{\perp}\tilde{\mathbb{S}}(M) to a *-homomorphism on AA, where A=A0¯A=\overline{A_{0}}^{\|\cdot\|} is a C{\rm C^{*}}-algebra.

The next step is to define the map ϕ\phi. For each n0n\geq 0, set ϕn:=ϕ0ψn:M,eMjq𝕂𝕊~(M)\phi_{n}:=\phi_{0}\circ\psi_{n}:\langle M,e_{M_{j}}\rangle\to q_{\mathbb{K}}^{\perp}\tilde{\mathbb{S}}(M), which is c.p. and subunital by construction. We may then pick ϕCB(M,eMj,q𝕂𝕊~(M))\phi\in CB(\langle M,e_{M_{j}}\rangle,q_{\mathbb{K}}^{\perp}\tilde{\mathbb{S}}(M)) a weak limit point of {ϕn}n\{\phi_{n}\}_{n\in\mathbb{N}}, which exists as q𝕂𝕊~(M)q_{\mathbb{K}}^{\perp}\tilde{\mathbb{S}}(M) is a von Neumann algebra.

We claim that

Ad(qj)ϕ:M,eMjq𝕂qj𝕊~(M)qj\operatorname{Ad}(q_{j})\circ\phi:\langle M,e_{M_{j}}\rangle\to q_{{\mathbb{K}}}^{\perp}q_{j}\widetilde{{\mathbb{S}}}(M)q_{j}

is an MM-bimodular u.c.p. map, which amounts to showing ϕ(x)=q𝕂qjιnor(x)\phi(x)=q_{{\mathbb{K}}}^{\perp}q_{j}\iota_{\rm nor}(x) for any xMx\in M.

In fact, for any xMx\in M, we have

ϕ(x)=\displaystyle\phi(x)= limnϕ0(0k,n(mkEMj(mkxm)eMjm))\displaystyle\lim_{n\to\infty}\phi_{0}\Big{(}\sum_{0\leq k,\ell\leq n}(m_{k}E_{M_{j}}(m_{k}^{*}xm_{\ell})e_{M_{j}}m_{\ell}^{*})\Big{)}
=\displaystyle= q𝕂limn0k,nιnor(mkEMj(mkxm))ejιnor(m)\displaystyle q_{{\mathbb{K}}}^{\perp}\lim_{n\to\infty}\sum_{0\leq k,\ell\leq n}\iota_{\rm nor}(m_{k}E_{M_{j}}(m_{k}^{*}xm_{\ell}))e_{j}\iota_{\rm nor}(m_{\ell}^{*})
=\displaystyle= q𝕂limn0k,n(ιnor(mk)ejιnor(mk))ιnor(x)(ιnor(m)ejιnor(m)),\displaystyle q_{{\mathbb{K}}}^{\perp}\lim_{n\to\infty}\sum_{0\leq k,\ell\leq n}\big{(}\iota_{\rm nor}(m_{k})e_{j}\iota_{\rm nor}(m_{k}^{*})\big{)}\iota_{\rm nor}(x)\big{(}\iota_{\rm nor}(m_{\ell})e_{j}\iota_{\rm nor}(m_{\ell}^{*})\big{)},

where the last equation follows from (1). Finally, note that {pk}k0\{p_{k}\}_{k\geq 0} is a family of pairwise orthogonal projections by Remark 2.3 , where

pk:=q𝕂ιnor(mk)ejιnor(mk)=q𝕂r0ιnor(JmrJmkeMjmkJmrJ),p_{k}:=q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(m_{k})e_{j}\iota_{\rm nor}(m_{k}^{*})=q_{\mathbb{K}}^{\perp}\sum_{r\geq 0}\iota_{\rm nor}(Jm_{r}Jm_{k}e_{M_{j}}m_{k}^{*}Jm_{r}^{*}J),

and k0pk=k,r0q𝕂ιnor(JmrJmkeMjmkJmrJ)=q𝕂qj\sum_{k\geq 0}p_{k}=\sum_{k,r\geq 0}q_{\mathbb{K}}^{\perp}\iota_{\rm nor}(Jm_{r}Jm_{k}e_{M_{j}}m_{k}^{*}Jm_{r}^{*}J)=q_{\mathbb{K}}^{\perp}q_{j} by Lemma 3.9. Therefore, we conclude that ϕ(x)=q𝕂qjιnor(x)\phi(x)=q_{\mathbb{K}}^{\perp}q_{j}\iota_{\rm nor}(x), as desired. ∎

Now consider ν=φϕM,eMj\nu=\varphi\circ\phi\in\langle M,e_{M_{j}}\rangle^{*} and notice that 1φ(q𝕂qj)ν\frac{1}{\varphi(q_{\mathbb{K}}^{\perp}q_{j})}\nu is an AfAf-central state, which is a normal state when restricted to fMffMf. Let fj𝒵((Af)fMf)f_{j}\in\mathcal{Z}((Af)^{\prime}\cap fMf) be the support projection of ν|𝒵((Af)fMf)\nu_{|\mathcal{Z}((Af)^{\prime}\cap fMf)} and then we have AfjAf_{j} is amenable relative to MjM_{j} inside MM [OP10, Theorem 2.1]. Apply the same argument for each ii with φ(q𝕂qi)>0\varphi(q_{\mathbb{K}}^{\perp}q_{i})>0, we then obtain projections fifMff_{i}\in fMf (possibly 0) such that AfiAf_{i} is amenable relative to MiM_{i} inside MM.

Finally, to show i=1nfi=f\vee_{i=1}^{n}f_{i}=f, note that φ(qifi)=φ(qi)\varphi(q_{i}f_{i})=\varphi(q_{i}) as

φ(qifi)=φ(q𝕂qifi)=φ(ϕ(fi))=ν(fi)=0.\varphi(q_{i}f_{i}^{\perp})=\varphi(q_{\mathbb{K}}^{\perp}q_{i}f_{i}^{\perp})=\varphi(\phi(f_{i}^{\perp}))=\nu(f_{i}^{\perp})=0.

Consequently we have

φ(i=1nfi)φ(i=1nqifi)φ(i=1nqi)=1,\varphi(\vee_{i=1}^{n}f_{i})\geq\varphi(\vee_{i=1}^{n}q_{i}f_{i})\geq\varphi(\vee_{i=1}^{n}q_{i})=1,

and hence i=1nfi=f\vee_{i=1}^{n}f_{i}=f by the faithfulness of φfMf\varphi_{\mid fMf}. Since fi𝒵((Af)fMf)f_{i}\in\mathcal{Z}((Af)^{\prime}\cap fMf), we may rearrange these projections so that i=1nfi=f\sum_{i=1}^{n}f_{i}=f. ∎

5. Proofs of main theorems

Proof of Theorem 1.2.

This follows from noticing that the Jones projections eLΛie_{L\Lambda_{i}} pairwise commute, and then applying Theorem 3.12 and Thoerem 1.1. ∎

Theorem 5.1.

Let (M1,τ1)(M_{1},\tau_{1}) and (M2,τ2)(M_{2},\tau_{2}) be such that MiLΓiM_{i}\cong L\Gamma_{i} where Γi\Gamma_{i} are countable exact groups and M=M1M2M=M_{1}\ast M_{2} be the tracial free product. Let AMA\subset M be von Neumann subalgebra, then there exists projections {pi}i=13𝒵(AM)\{p_{i}\}_{i=1}^{3}\in\mathcal{Z}(A^{\prime}\cap M) such that ApiMMiAp_{i}\prec_{M}M_{i} for each i=1i=1 and 22, Ap3Ap_{3} is amenable and A(i=13pi)A(\vee_{i=1}^{3}p_{i})^{\perp} is properly proximal.

Proof of Theorem 5.1.

First note that the free products of the exact groups Γi\Gamma_{i} is biexact relative to {Γ1,Γ2}\{\Gamma_{1},\Gamma_{2}\} [BO08, Proposition 15.3.12] and [eM1,eM2]=0[e_{M_{1}},e_{M_{2}}]=0. Then by Theorem 3.12, we may take f1f_{1} and f2f_{2} from Theorem 1.1 and let pi𝒵(Afi)p_{i}^{\prime}\in\mathcal{Z}(Af_{i}) be the maximal projection such that ApiAp_{i}^{\prime} is amenable for each i=1,2i=1,2. Set pi=fipip_{i}=f_{i}-p_{i}^{\prime} for i=1i=1 and 22, and p3=p1+p2p_{3}=p_{1}^{\prime}+p_{2}^{\prime} and the rest follows from Lemma 2.4. ∎

Proof of Corollary 1.3.

Since AMA\subset M has no properly proximal direct summand, it follows from Theorem 1.1 and Theorem 3.12 that there exists central projections f1f_{1} and f2f_{2} in 𝒵(AM){\mathcal{Z}}(A^{\prime}\cap M) such that AfiAf_{i} is amenable relative to MiM_{i} inside MM for each ii, and f1+f2=1f_{1}+f_{2}=1.

If Af2Af_{2} is not amenable, then by Lemma 2.4 we have that Af2MM2Af_{2}\prec_{M}M_{2}. However, since AM1A\cap M_{1} is diffuse, we may pick a sequence of trace zero unitaries {un}\{u_{n}\} in AM1A\cap M_{1} converging to 0. One then checks that EM2(xunf2y)20\|E_{M_{2}}(xu_{n}f_{2}y)\|_{2}\to 0 for any xx, yMy\in M, which is a contradiction. Therefore AA is amenable relative to M1M_{1} inside MM. And then it follows from Theorem 2.5 that AM1A\subset M_{1}. ∎

Proof of Corollary 1.7.

Note that in the case of A=LΓ1A=L\Gamma_{1}, we have 𝒵(AM)={\mathcal{Z}}(A^{\prime}\cap M)={\mathbb{C}} and hence Theorem 5.1 implies that either LΓ1MLΛ1L\Gamma_{1}\prec_{M}L\Lambda_{1} or LΓ1MLΛ2LΛmL\Gamma_{1}\prec_{M}L\Lambda_{2}\ast\cdots\ast L\Lambda_{m}. The same argument as in [Dri22, Corollary 8.1] deduces the desired result. ∎

References

  • [BC15] Rémi Boutonnet and Alessandro Carderi, Maximal amenable von Neumann subalgebras arising from maximal amenable subgroups, Geom. Funct. Anal. 25 (2015), no. 6, 1688–1705. MR 3432155
  • [BC17] Rémi Boutonnet and Alessandro Carderi, Maximal amenable subalgebras of von neumann algebras associated with hyperbolic groups, Mathematische Annalen 367 (2017), no. 3, 1199–1216.
  • [BEW19] Alcides Buss, Siegfried Echterhoff, and Rufus Willett, Injectivity, crossed products, and amenable group actions, 2019, arXiv:1904.06771.
  • [BH18] Rémi Boutonnet and Cyril Houdayer, Amenable absorption in amalgamated free product von Neumann algebras, Kyoto J. Math. 58 (2018), no. 3, 583–593. MR 3843391
  • [BIP18] Rémi Boutonnet, Adrian Ioana, and Jesse Peterson, Properly proximal groups and their von Neumann algebras, 2018, arXiv:1809.01881.
  • [Bla06] B. Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122, Springer-Verlag, Berlin, 2006, Theory of CC^{*}-algebras and von Neumann algebras, Operator Algebras and Non-commutative Geometry, III.
  • [BO08] Nathanial P. Brown and Narutaka Ozawa, CC^{*}-algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. MR 2391387
  • [BW16] Arnaud Brothier and Chenxu Wen, The cup subalgebra has the absorbing amenability property, Internat. J. Math. 27 (2016), no. 2, 1650013, 6. MR 3464393
  • [CFRW10] Jan Cameron, Junsheng Fang, Mohan Ravichandran, and Stuart White, The radial masa in a free group factor is maximal injective, J. Lond. Math. Soc. (2) 82 (2010), no. 3, 787–809. MR 2739068
  • [CH10] Ionut Chifan and Cyril Houdayer, Bass-Serre rigidity results in von Neumann algebras, Duke Math. J. 153 (2010), no. 1, 23–54. MR 2641939
  • [CS04] Tullio Ceccherini-Silberstein, On subfactors with a unitary orthonormal basis, Sovrem. Mat. Prilozh. (2004), no. 22, Algebra i Geom., 102–125. MR 2462073
  • [Din22] Changying Ding, First 2\ell^{2}-betti number and proper proximality, 2022, In preparation.
  • [DKE22] Changying Ding and Srivatsav Kunnawalkam Elayavalli, Proper proximality for various families of groups, 2022, arXiv: 2107.02917.
  • [DKEP22] Changying Ding, Srivatsav Kunnawalkam Elayavalli, and Jesse Peterson, Properly proximal von Neumann algebras, 2022, arXiv: https://arxiv.org/abs/2204.00517.
  • [DP22] Changying Ding and Jesse Peterson, Biexact von Neumann algebras, 2022, In preparation.
  • [Dri22] Daniel Drimbe, Measure equivalence rigidity via s-malleable deformations, 2022.
  • [ER88] Edward G. Effros and Zhong-Jin Ruan, Representations of operator bimodules and their applications, J. Operator Theory 19 (1988), no. 1, 137–158.
  • [HJNS21] Ben Hayes, David Jekel, Brent Nelson, and Thomas Sinclair, A random matrix approach to absorption in free products, Int. Math. Res. Not. IMRN (2021), no. 3, 1919–1979. MR 4206601
  • [Hou14] Cyril Houdayer, Gamma stability in free product von neumann algebras, Communications in Mathematical Physics 336 (2014).
  • [HU16] Cyril Houdayer and Yoshimichi Ueda, Rigidity of free product von neumann algebras, Compositio Mathematica 152 (2016), no. 12, 2461–2492.
  • [Ioa15] Adrian Ioana, Cartan subalgebras of amalgamated free product II1{\rm II}_{1} factors, Ann. Sci. Éc. Norm. Supér. (4) 48 (2015), no. 1, 71–130, With an appendix by Ioana and Stefaan Vaes. MR 3335839
  • [Ioa18] by same author, Rigidity for von neumann algebras, Proceedings of the International Congress of Mathematicians. Volume II, 2018, pp. 1635–1668.
  • [IPP08] Adrian Ioana, Jesse Peterson, and Sorin Popa, Amalgamated free products of weakly rigid factors and calculation of their symmetry groups, Acta Math. 200 (2008), no. 1, 85–153. MR 2386109
  • [IPR19] Ishan Ishan, Jesse Peterson, and Lauren Ruth, Von Neumann equivalence and properly proximal groups, arXiv:1910.08682, 2019.
  • [Jol12] Paul Jolissaint, Examples of mixing subalgebras of von Neumann algebras and their normalizers, Bull. Belg. Math. Soc. Simon Stevin 19 (2012), no. 3, 399–413. MR 3027351
  • [Mag97] Bojan Magajna, Strong operator modules and the Haagerup tensor product, Proc. London Math. Soc. (3) 74 (1997), no. 1, 201–240.
  • [Mag98] by same author, A topology for operator modules over WW^{*}-algebras, J. Funct. Anal. 154 (1998), no. 1, 17–41.
  • [Mag05] by same author, Duality and normal parts of operator modules, J. Funct. Anal. 219 (2005), no. 2, 306–339.
  • [OP10] Narutaka Ozawa and Sorin Popa, On a class of II1{\rm II}_{1} factors with at most one Cartan subalgebra, Ann. of Math. (2) 172 (2010), no. 1, 713–749. MR 2680430
  • [Osi06] Denis V. Osin, Relatively hyperbolic groups: intrinsic geometry, algebraic properties, and algorithmic problems, Mem. Amer. Math. Soc. 179 (2006), no. 843, vi+100. MR 2182268
  • [Oya22] Koichi Oyakawa, Bi-exactness of relatively hyperbolic groups, 2022.
  • [Oza04] Narutaka Ozawa, Solid von Neumann algebras, Acta Math. 192 (2004), no. 1, 111–117. MR 2079600
  • [Oza06] by same author, A Kurosh-type theorem for type II1\rm II_{1} factors, Int. Math. Res. Not. (2006), Art. ID 97560, 21. MR 2211141
  • [Oza10a] by same author, A comment on free group factors, Noncommutative harmonic analysis with applications to probability II, Banach Center Publ., vol. 89, Polish Acad. Sci. Inst. Math., Warsaw, 2010, pp. 241–245. MR 2730894
  • [Oza10b] by same author, A comment on free group factors, Noncommutative harmonic analysis with applications to probability II, Banach Center Publ., vol. 89, Polish Acad. Sci. Inst. Math., Warsaw, 2010, pp. 241–245. MR 2730894
  • [Pet09] Jesse Peterson, L2L^{2}-rigidity in von Neumann algebras, Invent. Math. 175 (2009), no. 2, 417–433. MR 2470111
  • [Pop83] Sorin Popa, Maximal injective subalgebras in factors associated with free groups, Advances in Mathematics 50 (1983), 27–48.
  • [Pop06] Sorin Popa, Strong rigidity of II1\rm II_{1} factors arising from malleable actions of ww-rigid groups. I, Invent. Math. 165 (2006), no. 2, 369–408. MR 2231961
  • [PP86] Mihai Pimsner and Sorin Popa, Entropy and index for subfactors, Ann. Sci. École Norm. Sup. (4) 19 (1986), no. 1, 57–106. MR 860811
  • [PSW18] Sandeepan Parekh, Koichi Shimada, and Chenxu Wen, Maximal amenability of the generator subalgebra in qq-Gaussian von Neumann algebras, J. Operator Theory 80 (2018), no. 1, 125–152. MR 3835452
  • [Ser80] Jean-Pierre Serre, Trees, Springer-Verlag, Berlin-New York, 1980, Translated from the French by John Stillwell. MR 607504
  • [Vae12] Stefaan Vaes, An inner amenable group whose von Neumann algebra does not have property Gamma, Acta Math. 208 (2012), no. 2, 389–394. MR 2931384
  • [Wen16] Chenxu Wen, Maximal amenability and disjointness for the radial masa, J. Funct. Anal. 270 (2016), no. 2, 787–801. MR 3425903