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

Asymptotic Cohomology and Uniform Stability
for Lattices in Semisimple Groups

Lev Glebsky, Alexander Lubotzky, Nicolas Monod, Bharatram Rangarajan Universidad Aut´onoma de San Luis Potosi, Mexico Email:glebsky@cactus.iico.uaslp.mxWeizmann Institute of Science, Rehovot, Israel Email: alex.lubotzky@mail.huji.ac.ilÉcole Polytechnique Fédérale de Lausanne (EPFL), Lausanne, Switzerland Email:nicolas.monod@epfl.chHebrew University of Jerusalem, Jerusalem, Israel Email: bharatrm.rangarajan@mail.huji.ac.il
Abstract

It is, by now, classical that lattices in higher rank semisimple groups have various rigidity properties. In this work, we add another such rigidity property to the list: uniform stability with respect to the family of unitary operators on finite-dimensional Hilbert spaces equipped with submultiplicative norms. Namely, we show that for (most) high-rank lattices, every finite-dimensional unitary ”almost-representation” of Γ\Gamma is a small deformation of a (true) unitary representation. This extends a result of Kazhdan [31] for amenable groups and of Burger-Ozawa-Thom [15] for SL(n,Z)SL(n,Z) (for n>2n>2). Towards this goal, we first build an elaborate cohomological theory capturing the obstruction to such stability, and show that the vanishing of second cohomology implies uniform stability in this setting. This cohomology can be roughly thought of as an asymptotic version of bounded cohomology, and sheds light on a question raised in [39] about a possible connection between vanishing of second bounded cohomology and Ulam stability.

Keywords— group stability, bounded cohomology, amenability

Mathematics Subject Classification— 20J05, 22D40

In memory of Robert J. Zimmer

Introduction

Consider a semisimple group G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}\mathbf{G}_{i}(K_{i}), where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is an almost KiK_{i}-simple group. If the rank i=1krkKi(𝐆i)2\sum^{k}_{i=1}rk_{K_{i}}(\mathbf{G}_{i})\geq 2, such a GG is referred to as a higher rank semisimple group. The class of irreducible lattices Γ\Gamma in such groups GG (referred to as higher rank lattices) form an interesting class of groups, which over the years, have been shown to satisfy many rigidity properties, such as local rigidity, Mostow strong-rigidity, Margulis super-rigidity (implying that they are arithmetic groups), Zimmer cocycle rigidity, quasi-isometric rigidity, first-order rigidity, etc. (see [20], [4], [34], [11] and the references therein). A common feature of the classical rigidity results is that such a higher rank lattice Γ\Gamma has some clear family of representations, and all other representations are just easy variants of them.
The goal of this paper is to demonstrate another type of rigidity phenomena of these lattices. Before stating the exact formulation, let us recall that Margulis super-rigidity, while usually not formulated this way, also gives a full classification of all the finite dimensional unitary representations of a higher rank lattice Γ\Gamma as above. Margulis super-rigidity implies that all such irreducible representations come from a combination of those that factor through finite quotients (and these are the only ones if Γ\Gamma is a non-uniform lattice) and from the representations of Γ\Gamma appearing naturally in its definition as an arithmetic group by Galois twisting (see §1.31.3 in [34]). The rigidity phenomenon we study here, which is called uniform stability, is the property that every unitary \sayalmost-representation of Γ\Gamma is a small deformation of a unitary representation.

Uniform Stability of Groups

Let Γ\Gamma be a discrete group and (G,dG)(G,d_{G}) be a metric group (where dGd_{G} is a bi-invariant metric on GG). For ϵ>0\epsilon>0, a map ϕ:ΓG\phi:\Gamma\to G is said to be an ϵ\epsilon-almost homomorphism (or ϵ\epsilon-homomorphism) if dG(ϕ(xy),ϕ(x)ϕ(y))ϵd_{G}(\phi(xy),\phi(x)\phi(y))\leq\epsilon for every x,yΓx,y\in\Gamma. The value supx,yΓdG(ϕ(xy),ϕ(x)ϕ(y))\sup_{x,y\in\Gamma}d_{G}(\phi(xy),\phi(x)\phi(y)) is called the defect of ϕ\phi.
Let 𝒢\mathcal{G} be a family of metric groups. We say that Γ\Gamma is uniformly stable with respect to 𝒢\mathcal{G} if for any ϵ>0\epsilon>0, there exists δ=δ(ϵ)\delta=\delta(\epsilon) with limϵ0δ(ϵ)=0\lim_{\epsilon\to 0}\delta(\epsilon)=0 such that for any ϵ\epsilon-homomorphism ϕ:ΓG\phi:\Gamma\to G (for G𝒢G\in\mathcal{G}), there exists a homomorphism ψHom(Γ,G)\psi\in Hom(\Gamma,G) with supxΓdG(ϕ(x),ψ(x))δ\sup_{x\in\Gamma}\>d_{G}(\phi(x),\psi(x))\leq\delta. In other words, Γ\Gamma is uniformly stable with respect to 𝒢\mathcal{G} if any almost homomorphism of Γ\Gamma to any group in the family 𝒢\mathcal{G} is close to a (true) homomorphism.

Questions of this nature were first raised and studied in [48], [50] and [49], and of particular interest is the case when 𝒢\mathcal{G} is the family of unitary operators on Hilbert spaces and the metric is given by a norm (on the space of bounded operators), as studied in [31] and [15]. Note that in this work, we will be interested solely in uniform stability, as opposed to pointwise stability, as studied in [19], [3] and the references therein.

The notion of uniform stability with respect to unitary operators on Hilbert spaces equipped with the operator norm is referred to in [15] as strong Ulam stability, while if we restrict the family to unitary operators on finite-dimensional Hilbert spaces, it is referred to as Ulam stability. In the pioneering work of Kazhdan [31] (and clarified further in [46] and [30]), it is shown that

Theorem 0.0.1 ([31]).

Every (discrete) amenable group Γ\Gamma is Ulam stable (in fact, even strongly Ulam stable).

It is worth noting that the only known examples of strongly Ulam stable groups are amenable, and it is natural to ask if strong Ulam stability characterizes amenability.
Let us mention at this point that one of the (innumerable) equivalent characterizations of amenability is given in terms of the vanishing of bounded cohomology with dual coefficients: Γ\Gamma is amenable iff Hbn(Γ,V)=0\operatorname{H}_{b}^{n}(\Gamma,V)=0 for every dual Banach Γ\Gamma-module VV and n>0n>0. Here Hbn(Γ,V)\operatorname{H}_{b}^{n}(\Gamma,V) denotes the nn-th bounded cohomology group of Γ\Gamma with coefficients in the Banach Γ\Gamma-module VV. Kazhdan’s proof does not use this result explicitly but does use a notion of ϵ\epsilon-cocycles and approximate cohomology in degree 22.

Ulam stability was further studied in [15] where they show more examples (and non-examples) of Ulam stable groups. It is shown there that if a group contains a non-abelian free subgroup, then it is not strongly Ulam stable. In particular, this means that higher rank lattices are not strongly Ulam stable. On the positive side, they show:

Theorem 0.0.2 ([15]).

Let 𝒪\mathcal{O} be the ring of integers of a number field, SS a finite set of primes, and 𝒪S\mathcal{O}_{S} the corresponding localization. Then for every n3n\geq 3, SL(n,𝒪S)SL(n,\mathcal{O}_{S}) is Ulam stable.

The proof of this result in [15] uses the fact that SL(n,𝒪S)SL(n,\mathcal{O}_{S}) (for n3n\geq 3) is boundedly generated by elementary matrices, and makes no reference to bounded cohomology (this result is further extended in [26] in the case of n=2n=2 when 𝒪S\mathcal{O}_{S} has infinitely many units). However, note that for Γ=SL(n,𝒪S)\Gamma=SL(n,\mathcal{O}_{S}), Hb2(Γ,V)=0\operatorname{H}_{b}^{2}\left(\Gamma,V\right)=0 for every dual separable Γ\Gamma-module VV. In fact, it is shown in [13] that for every higher rank lattice Γ\Gamma and any dual, separable Banach Γ\Gamma-module VV with VΓ={0}V^{\Gamma}=\{0\}, Hb2(Γ,V)=0\operatorname{H}_{b}^{2}(\Gamma,V)=0. All this hints at a possible connection between bounded cohomology and Ulam stability, as raised by Monod in his ICM talk [39, Problem F], and serves as one of the starting points for our current work.

Main Results and Methods

In this paper, we generalize LABEL:Kaz and Theorem 0.0.2 to a wider class of groups and metrics. We shall consider the question of uniform stability with respect to the family 𝔘\mathfrak{U} of groups of unitary operators on finite-dimensional Hilbert spaces, with the metrics induced from submultiplicative norms on matrices (which we shall denote uniform 𝔘\mathfrak{U}-stability). These include the pp-Schatten norms for 1p1\leq p\leq\infty (and in particular, uniform 𝔘\mathfrak{U}-stability subsumes Ulam stability). Furthermore, we shall show uniform 𝔘\mathfrak{U}-stability with a linear estimate, which means that the distance of an almost homomorphism from a homomorphism is linearly bounded by its defect, and all our results are proved in this stronger notion of stability.

To this end, we build a new type of bounded cohomological theory that can capture obstructions to uniform 𝔘\mathfrak{U}-stability, so that uniform 𝔘\mathfrak{U}-stability follows as a consequence of the vanishing of the second cohomology group in this theory. While we shall develop this in full detail in §4, our technique involves the following two main steps:

  • Defect Diminishing: Expressing the problem of uniform stability as a homomorphism lifting problem, we can treat it as a culmination of intermediate lifts so that at each step, the lifting kernel is abelian. This is a uniform variant of defect diminishing that was introduced in [19] in the non-uniform setting, and is applicable when the relevant norms in the target groups are submultiplicative.

  • Asymptotic Cohomology: Such a homomorphism lifting problem with abelian kernel naturally leads to a cohomological reformulation (as in [19]). However, unlike in the non-uniform setting where ordinary group cohomology comes up, in our uniform setting we need to carefully construct a new cohomology theory such that the vanishing of the second cohomology group in this model implies (uniform) defect diminishing, and hence uniform stability.

The cohomological theory we construct is an asymptotic variant of the bounded cohomology of the ultrapower Γ{}^{*}\Gamma with coefficients in a suitable ultraproduct Banach space 𝒲\mathcal{W}, but restricted to the \sayinternal objects in this universe, which we shall call the asymptotic cohomology of Γ\Gamma denoted Ha(Γ,𝒲)\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W}).

Theorem 0.0.3.

Suppose Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0, then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

This new cohomology theory bears some similarity to the theory of bounded cohomology, and sometimes we can easily adapt arguments there to our model (for instance, we can show that Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0 for an amenable group Γ\Gamma, immediately implying that amenable groups are Ulam-stable as in [31],[46],[30]), though other times serious difficulties arise (which are responsible for the length of this paper).

The groups Γ\Gamma that we will be particularly interested in are lattices in higher rank semisimple groups. Unlike some of the other rigidity results which sometimes hold for some lattices in rank one simple groups, we also first show the following result:

Proposition 0.0.4.

If Γ\Gamma is a lattice in a semisimple group of rank 11, then Γ\Gamma is not uniformly 𝔘\mathfrak{U}-stable.

It is shown in [25] that for such a Γ\Gamma, Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) is infinite dimensional. More precisely, Fujiwara constructs (many) non-trivial quasi-homomorphisms witnessing that the comparison map c:Hb2(Γ,𝐑)H2(Γ,𝐑)c:\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}) is not injective. By exponentiation, such quasimorphisms yield almost homomorphisms to U(1)U(1) that are not close to any homomorphism (we shall discuss this in more detail in §1). In particular, a lattice of rank one is not uniformly U-stable, and to hope for uniform 𝔘\mathfrak{U}-stability, the condition that rank of Γ\Gamma is at least 22 is necessary.

For the main result of the paper, we need some definitions capturing properties of the class of semisimple groups we will be interested in. For a locally compact group GG, we denote by Hb(G,𝐑)\operatorname{H}_{b}^{\bullet}(G,\mathbf{R}) the continuous bounded cohomology of GG with trivial coefficients.

  • A locally compact group GG is said to have the 2½property (of vanishing bounded cohomology) if Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff.

  • Let GG be a non-compact simple Lie group, and fix a minimal parabolic subgroup PGP\leq G. The group GG is said to have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) if there exist two proper parabolic subgroups Q1Q_{1} and Q2Q_{2} containing PP, both having the 2½-property, such that GG is boundedly generated by the union Q1Q2Q_{1}\cup Q_{2}. A semisimple group GG is said to have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) if all its simple factors have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

Note that if a semisimple group has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), then by definition, it must be of rank at least 22. But note that not all simple groups have the property (for instance, SL3(𝐑)SL_{3}(\mathbf{R})). However, in §6.3, we will show that many classes of groups do have this property, for example, all simple groups (of rank at least 22) over 𝐂\mathbf{C} or over a non-archimedean field, and SLn(𝐑)SL_{n}(\mathbf{R}) for n4n\geq 4.
We can now state our main result:

Theorem 0.0.5.

Let Γ\Gamma be a lattice in a semisimple group GG that has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}). Then Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0, so in particular, Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable.

The main result of our paper is thus concerned with showing that Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0 for Γ\Gamma being a lattice in a higher rank Lie group (satisfying certain conditions). For this, we take inspiration from the results of [13] about the vanishing of bounded cohomology for such lattices. More specifically, our approach is inspired by a proof of [43] specifically in degree two, and a version of that result and proof technique are outlined below.

Theorem 0.0.6 ([43]).

Let GG be a higher rank simple group, and PGP\leq G be a minimal parabolic subgroup. Suppose GG contains two proper parabolic subgroups Q1Q_{1} and Q2Q_{2} such that PQ1Q2P\subseteq Q_{1}\cap Q_{2}, GG is generated by Q1Q2Q_{1}\cup Q_{2}, and Hb2(Q1,𝐑)=Hb2(Q2,𝐑)=0\operatorname{H}_{b}^{2}(Q_{1},\mathbf{R})=\operatorname{H}_{b}^{2}(Q_{2},\mathbf{R})=0. Then for any lattice Γ\Gamma in GG and a dual separable Banach Γ\Gamma-module WW, Hb2(Γ,W)=0\operatorname{H}_{b}^{2}(\Gamma,W)=0.

The proof of Theorem 0.0.6 proceeds in several steps briefly sketched below, where we also mention the corresponding steps and difficulties in the proof of Theorem 0.0.5 even when GG is simple:

  • The first step is to use an Eckman-Shapiro induction to construct a dual, separable, continuous Banach GG-module VV so that Hb2(Γ,W)=Hb2(G,V)\operatorname{H}_{b}^{2}(\Gamma,W)=\operatorname{H}_{b}^{2}(G,V), thus reducing the problem to showing that Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0. A similar inductive procedure, described in §5, allows us to construct an ultraproduct Banach space 𝒱\mathcal{V} with an asymptotic action of GG so that Ha2(Γ,𝒲)=Ha2(G,𝒱)\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=\operatorname{H}_{a}^{2}(G,\mathcal{V}). Note that in the setting of asymptotic cohomology, we actually work with ultrapowers Γ{}^{*}\Gamma and G{}^{*}G, and so G/Γ{}^{*}G/{}^{*}\Gamma is not locally compact. However, the restriction to internal objects allows us to carefully work out an induction procedure as needed. The induced module 𝒱\mathcal{V} also has an internal continuity property (defined in §4.1) that we establish in §5.

  • Since PP is amenable, the bounded cohomology Hb(G,V)\operatorname{H}_{b}^{\bullet}(G,V) can be computed as the cohomology of the complex

    0{0}VG{V^{G}}L(G/P,V)G{L^{\infty}(G/P,V)^{G}}L((G/P)2,V)G{L^{\infty}((G/P)^{2},V)^{G}}L((G/P)3,V)G{L^{\infty}((G/P)^{3},V)^{G}}{\dots}ϵ\scriptstyle{\epsilon}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}

    Furthermore, for a parabolic subgroup PQGP\leq Q\leq G, the bounded cohomology Hb(Q,V)\operatorname{H}_{b}^{\bullet}(Q,V) can be computed as the cohomology of the complex

    0{0}VQ{V^{Q}}L(G/P,V)Q{L^{\infty}(G/P,V)^{Q}}L((G/P)2,V)Q{L^{\infty}((G/P)^{2},V)^{Q}}L((G/P)3,V)Q{L^{\infty}((G/P)^{3},V)^{Q}}{\dots}ϵ\scriptstyle{\epsilon}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}

    These steps too can be reworked in the asymptotic setting, analogous to the procedure in bounded cohomology theory, again thanks to the restriction on internality, and this is described in §4.

  • The motivation behind the preceding step is that we have at our disposal the following double ergodicity theorem: let VV be a continuous GG-module and αL((G/P)2,V)G\alpha\in L^{\infty}\left((G/P)^{2},V\right)^{G}, then α\alpha is essentially constant. This theorem follows from the Mautner’s lemma: let VV be a continuous GG-module and NGN\leq G be a non-compact subgroup, then VN=VGV^{N}=V^{G}. Both these results are particularly useful in the context of Hb2(G,V)\operatorname{H}_{b}^{2}(G,V).
    In our setting, there are particular difficulties in obtaining an analoguous Maunter lemma due to the asymptotic nature of our model. We overcome them for our specific Banach module 𝒱\mathcal{V} by applying a suitable correction to exact cocycles using structure results for the semisimple group GG; this is worked out in §6.

  • For the parabolic subgroups Q1Q_{1} and Q2Q_{2} as in the hypothesis, an inflation-restriction sequence argument implies that Hb2(Qi,V)=Hb2(Qi/Ni,VNi)\operatorname{H}_{b}^{2}(Q_{i},V)=\operatorname{H}_{b}^{2}(Q_{i}/N_{i},V^{N_{i}}) for NiN_{i} being the (amenable) radical of QiQ_{i} for i{1,2}i\in\{1,2\}. By Mautner’s lemma and the hypothesis that Hb2(Qi,𝐑)=0\operatorname{H}_{b}^{2}(Q_{i},\mathbf{R})=0, one concludes that Hb2(Qi,V)=0\operatorname{H}_{b}^{2}(Q_{i},V)=0.
    The analogous hypothesis in our asymptotic setting is Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), where the conditions that Hb2(Qi,𝐑)=0\operatorname{H}_{b}^{2}(Q_{i},\mathbf{R})=0 and Hb3(Qi,𝐑)\operatorname{H}_{b}^{3}(Q_{i},\mathbf{R}) is Hausdorff together are used to conclude that Ha2(Qi,𝒱)=0\operatorname{H}_{a}^{2}(Q_{i},\mathcal{V})=0 in §6.1.

  • Let ωL((G/P)3,V)G\omega\in L^{\infty}\left((G/P)^{3},V\right)^{G} be a bounded 22-cocycle for GG. Since Hb2(Q1,V)=Hb2(Q2,V)=0\operatorname{H}_{b}^{2}(Q_{1},V)=\operatorname{H}_{b}^{2}(Q_{2},V)=0, there exist α1L((G/P)2,V)Q1\alpha_{1}\in L^{\infty}\left((G/P)^{2},V\right)^{Q_{1}} and α2L((G/P)2,V))Q2\alpha_{2}\in L^{\infty}\left((G/P)^{2},V)\right)^{Q_{2}} such that ω=dα1=dα2\omega=d\alpha_{1}=d\alpha_{2}. In particular, α1α2\alpha_{1}-\alpha_{2} is a 11-cocycle for Q1Q2Q_{1}\cap Q_{2}. Since PQ1Q2P\leq Q_{1}\cap Q_{2}, α1α2\alpha_{1}-\alpha_{2} is a 11-cocycle for PP as well. But since Hb1(P,V)=0\operatorname{H}_{b}^{1}(P,V)=0, one can show using the double ergodicity theorem, that α1=α2\alpha_{1}=\alpha_{2} (=α=\alpha, say), implying that α\alpha is equivariant with respect to both Q1Q_{1} and Q2Q_{2}, and hence is GG-equivariant. Thus ω=dα\omega=d\alpha for αL((G/P)2,V)G\alpha\in L^{\infty}\left((G/P)^{2},V\right)^{G}, proving that Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0. This step too goes through in our setting once we have our asymptotic variant of the ergodicity theorem used in the classical case.

While in this paper, we focus on using the theory of asymptotic cohomology to prove uniform 𝔘\mathfrak{U}-stability for lattices in semisimple groups, the framework and tools developed here have also been used in subsequent work [23] to prove uniform 𝔘\mathfrak{U}-stability for other classes of groups such as lamplighter groups ΓΛ\Gamma\wr\Lambda where Λ\Lambda is infinite and amenable, as well as several groups of dynamical origin such as Thompson’s group FF. The techniques there too are analogous to corresponding vanishing results of bounded cohomology in [41], yet again highlighting the connections between the theories of bounded cohomology and asymptotic cohomology.

Outline of the Paper

We begin with the much simple setting of uniform stability with respect to the fixed group U(1)U(1) (equipped with the absolute value metric) in §1. In this case, we can reduce the question of uniform U(1)U(1)-stability of Γ\Gamma to the injectivity of the comparison map c:Hb2(Γ,𝐑)H2(Γ,𝐑)c:\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}). After recalling how classical results from [25] imply that lattices in Lie groups of rank 11 are not uniformly U(1)U(1)-stable, we then show that lattices in higher rank Lie groups are uniformly U(1)U(1)-stable. While the connection between uniform U(1)U(1)-stability of a group Γ\Gamma and the second bounded cohomology Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) is classical, it motivates the idea of using the logarithm map on an almost homomorphism to construct a bounded 22-cocycle of Γ\Gamma in a Banach space, which we develop in §3.2 for a more general setting.

In §2, we begin by defining the basic notions in full detail and rigor in §2.1. In particular we focus on interpreting stability in terms of sequences of maps and asymptotic homomorphisms, which we then refine further in §2.2 in the language of non-standard analysis. This formulation will allow us to reinterpret the question of uniform stability as a homomorphism lifting problem on the lines of the approach used in [19] [3]. While such a lifting problem motivates the attempt at constructing a cohomology, an obstacle here is that the kernel of the extension is not abelian. This issue is resolved in [19] by considering lifts in small increments so that the kernel at each step is abelian. This idea, known as defect diminishing, is explored in §2.3, and can be shown to imply uniform stability.

In §3 we begin by focusing on a particular family of metric groups for which defect diminishing corresponds to a homomorphism lifting problem with an abelian kernel. This is the family of unitary groups equipped with submultiplicative norms, discussed in §3.1. In [19], defect diminishing combined with ordinary group cohomology with coefficients in the abelian kernel (which turns out to be a Banach Γ\Gamma-module) is sufficient to study non-uniform stability. But in our setting, the uniformity condition involves subtleties that necessitate transfering to the internal Lie algebra with an internal asymptotic-action of Γ{}^{*}\Gamma (the ultrapower of Γ\Gamma), and the formulation of an \sayinternal and asymptotic bounded cohomology with cofficients in that internal space, denoted 𝒲\mathcal{W}. This is motivated in §3.2, and we conclude the section by demonstrating the machinery built so far in (re)proving the result of Kazhdan [31] that discrete amenable groups are Ulam stable.

§4 begins with the rigorous definitions of an internal Banach spaces and asymptotic G{}^{*}G-modules for a locally compact group GG (defining our notions in the category of topological groups is necessary in order to deal with lattices in Lie groups, which shall involve an Eckmann-Shapiro induction of cohomologies explored in §5) in §4.1. In §4.2, we formally define Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) using an internal LL^{\infty}-spaces, and study some functorial properties and different cochain complexes that can be used to compute Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) in §4.3. Many of the techniques used here have parallels in the theory of continuous bounded cohomology as in [37].

In §5, we restrict our attention to a lattice Γ\Gamma in a Lie group GG, and begin by studying an intermediate structure b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}(G,\mathcal{W})^{\sim{}^{*}\Gamma} that is not quite the induction module 𝒱=(D,𝒲)\mathcal{V}=\mathcal{L}^{\infty}(D,\mathcal{W}) in our machinery, but comes with an internal (true) G{}^{*}G-action (as opposed to an action upto infinitesimals). This structure leads to useful results proved in §5.1, and the actual induction module and an Eckmann-Shapiro induction procedure are discussed in §5.2. We conclude the section with a continuity property of our module 𝒱\mathcal{V}, which we use to define contracting elements to lay the groundwork for a Mautner’s lemma to be proved in the next section.

Finally, in §6, we come to the higher rank semisimple groups of interest, and begin by discussing an analogue of the Mautner’s lemma in our setting, along with double ergodicity lemmas in §6.1, and use these to prove that Ha2(Q,𝒱)=0\operatorname{H}_{a}^{2}(Q,\mathcal{V})=0 for QGQ\leq G being a proper parabolic subgroup. All these techniques come together in §6.2 where we prove that Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(G,\mathcal{V})=0 for semisimple groups GG that have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), thus implying uniform stability with a linear estimate for lattices in such groups. In §6.3, we list out classes of simple groups GG that have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), and conclude in §7 with some discussion on the limitations of our method and related open questions.

Acknowledgements

The second author acknowledges with gratitude the hospitality and support of the Fields Institute (Toronto) and the Institute for Advanced Study (Princeton) where part of this work was carried on, as well as a grant by the European Research Council (ERC) under the European Union’s Horizon 2020 (grant agreement No 882751882751). The results presented here are part of the fourth author’s PhD thesis, also supported by the same grant. The authors would like to thank Andrei Rapinchuk for his guidance in the proof of Proposition 6.1.1.

This paper is dedicated to the memory of Bob Zimmer in honor of his remarkable achievements and influence on the study of rigidity of lattices in semisimple Lie groups.

1 U(1)U(1)-Stability of Groups

We begin with a simpler setting, namely uniform stability of a discrete group Γ\Gamma with respect to the abelian group U(1)U(1). This section can be read independently of the rest.

Definition 1.0.1.

For ϵ>0\epsilon>0, a map ϕ:ΓU(1)\phi:\Gamma\to U(1) is said to be an ϵ\epsilon-homomorphism if

supx,yΓ|ϕ(xy)ϕ(x)ϕ(y)|ϵ\sup_{x,y\in\Gamma}\lvert\phi(xy)-\phi(x)\phi(y)\rvert\leq\epsilon

The value supx,yΓ|ϕ(xy)ϕ(x)ϕ(y)|\sup_{x,y\in\Gamma}\lvert\phi(xy)-\phi(x)\phi(y)\rvert is called the defect of ϕ\phi.

Definition 1.0.2.

A group Γ\Gamma is said to be uniformly U(1)U(1)-stable if for every δ>0\delta>0, there exists ϵ>0\epsilon>0 such that if ϕ:ΓU(1)\phi:\Gamma\to U(1) is an ϵ\epsilon-homomorphism, there exists a homomorphism ψ:ΓU(1)\psi:\Gamma\to U(1) such that supxΓ|ϕ(x)ψ(x)|<δ\sup_{x\in\Gamma}\lvert\phi(x)-\psi(x)\rvert<\delta.

A closely related notion is that of a quasimorphism. A quasimorphism is a map f:Γ𝐑f:\Gamma\to\mathbf{R} such that there exists D0D\geq 0 such that for every x,yΓx,y\in\Gamma,

|f(x)+f(y)f(xy)|D\lvert f(x)+f(y)-f(xy)\rvert\leq D

Let QM(Γ)QM(\Gamma) denote the space of all quasimorphisms of Γ\Gamma. A trivial example of a quasimorphism is obtained by perturbing a homomorphism by a bounded function, and such quasimorphisms form the subspace Hom(Γ,𝐑)Cb(Γ,𝐑)Hom(\Gamma,\mathbf{R})\oplus C_{b}(\Gamma,\mathbf{R}) of QM(Γ)QM(\Gamma) (where Cb(Γ,𝐑)C_{b}(\Gamma,\mathbf{R}) denotes the space of all bounded functions from Γ\Gamma to 𝐑\mathbf{R}) A quasimorphism that is not at a bounded distance from any homomorphism is called a non-trivial quasimorphism. In this setting, a question analogous to uniform stability is whether every quasimorphism is at a bounded distance from a homomorphism. That is, is QM(Γ)=Hom(Γ,𝐑)Cb(Γ,𝐑)QM(\Gamma)=Hom(\Gamma,\mathbf{R})\oplus C_{b}(\Gamma,\mathbf{R})?
It is known that every quasimorphism class contains a unique homogenous quasimorphism (a quasimorphism ff is said to be homogenous if for every gΓg\in\Gamma and n𝐍n\in\mathbf{N}, f(gn)=nf(g)f(g^{n})=nf(g)). Suppose ff is a homogenous quasimorphism of Γ\Gamma that is not a homomorphism. Then its exponent μe2πiϵf:ΓU(1)\mu\coloneq e^{2\pi i\epsilon f}:\Gamma\to U(1) is a function whose defect can be made arbitrarily small (by choosing ϵ0\epsilon\to 0), but whose distance from homomorphisms is bounded below by a positive constant.

Proposition 1.0.3 ([15] [44]).

If Γ\Gamma admits a non-trivial quasimorphism, then Γ\Gamma is not uniformly U(1)U(1)-stable.

Quasimorphisms are also closely related to group cohomology (this is well-known and classical, see [24],[39] for references). Observe that given a quasimorphism f:Γ𝐑f:\Gamma\to\mathbf{R}, the map

df:Γ×Γ𝐑df:\Gamma\times\Gamma\to\mathbf{R}
df(x,y)=f(x)+f(y)f(xy)df(x,y)=f(x)+f(y)-f(xy)

is a 22-coboundary for Γ\Gamma in 𝐑\mathbf{R}, while also being a bounded function that satisfies the 22-cocycle condition (and hence a bounded 22-cocycle). This leads to the following characterization of quasimorphisms classes ([24], [39]):

Proposition 1.0.4.

The kernel, denoted EHb2(Γ,𝐑)E\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}), of the comparison map c:Hb2(Γ,𝐑)H2(Γ,𝐑)c:\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}) is isomorphic to the space of quasimorphisms modulo the suabspace of trivial quasimorphisms.

EHb2(Γ,𝐑)QM(Γ)Cb(Γ,𝐑)Hom(Γ,𝐑)E\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\cong\frac{QM(\Gamma)}{C_{b}(\Gamma,\mathbf{R})\oplus Hom(\Gamma,\mathbf{R})}

Hence to show that Γ\Gamma is not U(1)U(1)-stable, it is sufficient to show that the comparison map c:Hb2(Γ,𝐑)H2(Γ,𝐑)c:\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}) is not injective. In [25], it is shown that for a lattice in a rank one semisimple Lie group, its comparison map has non-zero kernel, and hence

Theorem 1.0.5.

Let HH be a semisimple group (as in the introduction) of rank one, and let Γ\Gamma be a lattice in HH. Then Γ\Gamma is not uniformly U(1)U(1)-stable.

The rest of this section is devoted to showing that higher rank lattices are U(1)U(1)-stable. For this goal the bounded cohomology plays a central role. For simplicity, we endow U(1)U(1) with the distance coming from seeing it as 𝐑/𝐙\mathbf{R}/\mathbf{Z} (so we just have to apply a trigonometric formula if we prefer the norm distance).
Recall the long exact sequences associated to 𝐙𝐑𝐑/𝐙\mathbf{Z}\to\mathbf{R}\to\mathbf{R}/\mathbf{Z} for HnH^{n} and Hbn\operatorname{H}_{b}^{n}. Together with the comparison maps, we get a commutative diagram:

{\cdots}H1(Γ,𝐑/𝐙){H^{1}(\Gamma,\mathbf{R}/\mathbf{Z})}Hb2(Γ,𝐙){\operatorname{H}_{b}^{2}(\Gamma,\mathbf{Z})}Hb2(Γ,𝐑){\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})}H2(Γ,𝐑/𝐙){H^{2}(\Gamma,\mathbf{R}/\mathbf{Z})}{\cdots}{\cdots}H1(Γ,𝐑/𝐙){H^{1}(\Gamma,\mathbf{R}/\mathbf{Z})}H2(Γ,𝐙){H^{2}(\Gamma,\mathbf{Z})}H2(Γ,𝐑){H^{2}(\Gamma,\mathbf{R})}H2(Γ,𝐑/𝐙){H^{2}(\Gamma,\mathbf{R}/\mathbf{Z})}{\cdots}

Consider the subset KHb2(Γ,𝐑)K\subseteq\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) given by

K=Ker(Hb2(Γ,𝐑)H2(Γ,𝐑/𝐙))=Image(Hb2(Γ,𝐙)Hb2(Γ,𝐑))K=\mathrm{Ker}\Big{(}\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}/\mathbf{Z})\Big{)}=\mathrm{Image}\Big{(}\operatorname{H}_{b}^{2}(\Gamma,\mathbf{Z})\to\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\Big{)}

Recall that Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) is a Banach space; we can therefore consider the norm of elements in KK.

Proposition 1.0.6.

The following are equivalent for a group Γ\Gamma.

  1. 1.

    (Lipschitz U(1)U(1) stability): Every ϵ\epsilon-representation to U(1)U(1) with ϵ\epsilon small enough is at distance less than ϵ\epsilon from a representation.

  2. 2.

    (Linear U(1)U(1) stability): There is a constant cc such that every ϵ\epsilon-representation to U(1)U(1) is at distance less than cϵc\epsilon from a representation.

  3. 3.

    (U(1)U(1) stability): For each δ>0\delta>0 there is ϵ>0\epsilon>0 so that every ϵ\epsilon-representation to U(1)U(1) is at distance less than δ\delta from a representation.

  4. 4.

    (U(1)U(1) 1/31/3-stability): There is δ<1/3\delta<1/3 such that every ϵ\epsilon-representation to U(1)U(1) with ϵ\epsilon small enough is at distance less than δ\delta from a representation.

  5. 5.

    (Cohomology gap): Non-zero elements of KK have norm bounded below by a positive constant.

Remark 1.0.7.

The kernel of the comparison map Hb2(Γ,𝐑)H2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}) is contained in KK, see above diagram. Since this kernel is a vector space, point (5) fails as soon as this kernel is non-zero. This explains why quasimorphisms imply non-stability: it is a special case of the above as quasimorphisms \saylive in KK.

Proof.

Trivially (1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(4).

We prove (4)\Rightarrow(5) for the positive constant ϵ\epsilon as in (4) which, without loss of generality, can be made to satisfy ϵ<13δ\epsilon<1-3\delta. Consider a class [ω][\omega] in KK with [ω]<ϵ\|[\omega]\|<\epsilon. We can choose the representative ω(Γ2)\omega\in\ell^{\infty}(\Gamma^{2}) such that ω<ϵ\|\omega\|_{\infty}<\epsilon. Since [ω][\omega] is in the kernel to H2(Γ,𝐑/𝐙)H^{2}(\Gamma,\mathbf{R}/\mathbf{Z}), there is f:Γ𝐑f\colon\Gamma\to\mathbf{R} such that ωdf\omega\equiv df modulo 𝐙\mathbf{Z}. The map π=exp(2πif)\pi=\exp(2\pi if) is an ϵ\epsilon-representation. Thus π\pi is at distance less than δ\delta of an actual representation, which means that there is b:Γ[δ,δ]b\colon\Gamma\to[-\delta,\delta] with d(f+b)0d(f+b)\equiv 0 modulo 𝐙\mathbf{Z}. This means that ω+db\omega+db is integer-valued. On the other hand, db\|db\|_{\infty} is at most 3δ3\delta. Thus, since ϵ+3δ<1\epsilon+3\delta<1 we deduce that ω+db\omega+db actually vanishes and thus [ω][\omega] is zero in KK.

We now prove (5)\Rightarrow(1). We show it for 0<ϵ<1/40<\epsilon<1/4 smaller than the constant given by (5). For any π:ΓU(1)\pi\colon\Gamma\to U(1), choose f:Γ𝐑f\colon\Gamma\to\mathbf{R} such that π=exp(2πif)\pi=\exp(2\pi if). If π\pi is an ϵ\epsilon-representation, then there is ω:Γ2[ϵ,ϵ]\omega\colon\Gamma^{2}\to[-\epsilon,\epsilon] such that ωdf\omega\equiv df modulo 𝐙\mathbf{Z}. In particular, dω0d\omega\equiv 0 modulo 𝐙\mathbf{Z}, i.e. dωd\omega is integer-valued. Given that dω4ϵ\|d\omega\|_{\infty}\leq 4\epsilon, we have in fact dω=0d\omega=0 and hence we obtain a class [ω][\omega] in KK of norm less than ϵ\epsilon. Therefore, we can assume that [ω][\omega] is trivial, which means ω=db\omega=db for some b(Γ)b\in\ell^{\infty}(\Gamma).

In fact the operator dd on (Γ)\ell^{\infty}(\Gamma) has norm one: this was first observed by [36, p. 468] in a special case; the short proof is given in general in (the proof of) Corollary 2.7 in  [35]. We can therefore choose bb such that bϵ\|b\|_{\infty}\leq\epsilon. Now exp(2πi(fb))\exp(2\pi i(f-b)) is a representation at distance less than ϵ\epsilon of π\pi. ∎

Theorem 1.0.8.

Let Γ\Gamma be a group such that Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) is finite-dimensional and injects into H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}). Then Γ\Gamma is uniformly U(1)U(1)-stable.

Lemma 1.0.9.

Let AA be an abelian group and let BB be a subgroup of Hom𝐙(A,𝐙)Hom_{\mathbf{Z}}(A,\mathbf{Z}). If the image of BB in Hom𝐙(A,𝐑)Hom_{\mathbf{Z}}(A,\mathbf{R}) spans a subspace of finite 𝐑\mathbf{R}-dimension dd, then BB is free abelian of rank dd.

Proof of Lemma 1.0.9.

We can consider BB as a subgroup spanning a space of finite 𝐐\mathbf{Q}-dimension dd in Hom𝐙(A,𝐐)Hom_{\mathbf{Z}}(A,\mathbf{Q}). Viewing AA as a quotient of a free abelian group on some set XX, the group Hom𝐙(A,𝐙)Hom_{\mathbf{Z}}(A,\mathbf{Z}) is contained in 𝐙X\mathbf{Z}^{X}. While 𝐙X\mathbf{Z}^{X} is not free abelian in general, in Specker (Satz I p. 133 in [47]) it is proved that countable subgroups of 𝐙X\mathbf{Z}^{X} are free abelian . To be more precise, Specker proved it for XX countable but his proof works in general; alternatively, the statement immediately reduces to the case XX countable by taking a subset of XX separating the points of BB. Our BB, being a subgroup of a finite dimensional 𝐐\mathbf{Q}-space, is countable and hence free (from the above mentioned result from  [47]). Spanning a dd-dimensional 𝐐\mathbf{Q}-vector space, its rank must be also dd. ∎

Proof of Theorem 1.0.8.

It suffices to show that Γ\Gamma satisfies the cohomology gap — i.e., condition (5) of Proposition 1.0.6. Let B~\tilde{B} denote the image of Hb2(Γ,𝐙)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{Z}) in H2(Γ,𝐙)H^{2}(\Gamma,\mathbf{Z}). Thus the image of B~\tilde{B} in H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}) is precisely the image of KK.

Hb2(Γ,𝐙){\operatorname{H}_{b}^{2}(\Gamma,\mathbf{Z})}KHb2(Γ,𝐑){K\subseteq\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})}B~H2(Γ,𝐙){\tilde{B}\subseteq H^{2}(\Gamma,\mathbf{Z})}H2(Γ,𝐑){H^{2}(\Gamma,\mathbf{R})}

Let d<d<\infty be the dimension of the space spanned by KK in Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}). To prove the cohomology gap (5), we first claim that KK is free abelian of rank dd. This claim implies that KK is discrete in the finite dimensional space Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) since it spans a space of dimension dd (we use here the comparison with the standard lattice 𝐙d\mathbf{Z}^{d} in 𝐑d\mathbf{R}^{d} and the fact that linear maps are continuous in finite dimensions). Then, discreteness implies the desired gap. To prove the claim that KK is free abelian of rank dd, we can work with H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}) since Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) injects there. The universal coefficient theorem gives the following commutative diagram with exact rows.

0{0}Ext𝐙1(H1(Γ,𝐙),𝐙){Ext_{\mathbf{Z}}^{1}(H_{1}(\Gamma,\mathbf{Z}),\mathbf{Z})}H2(Γ,𝐙){H^{2}(\Gamma,\mathbf{Z})}Hom𝐙(H2(Γ,𝐙),𝐙){Hom_{\mathbf{Z}}(H_{2}(\Gamma,\mathbf{Z}),\mathbf{Z})}0{0}0{0}H2(Γ,𝐑){H^{2}(\Gamma,\mathbf{R})}Hom𝐙(H2(Γ,𝐙),𝐑){Hom_{\mathbf{Z}}(H_{2}(\Gamma,\mathbf{Z}),\mathbf{R})}0{0}

Therefore it suffices to show that the image BB of B~\tilde{B} in Hom𝐙(H2(Γ,𝐙),𝐙)Hom_{\mathbf{Z}}(H_{2}(\Gamma,\mathbf{Z}),\mathbf{Z}) is free abelian of rank dd, which follows from Lemma 1.0.9 applied to A=H2(Γ,𝐙)A=H_{2}(\Gamma,\mathbf{Z}). ∎

Since finitely presented groups have finite-dimensional H2H^{2}, we obtain the following clean necessary and sufficient condition:

Corollary 1.0.10.

Let Γ\Gamma be a finitely presented group. Then Γ\Gamma is uniformly U(1)U(1)-stable if and only if every quasi-morphism of Γ\Gamma is at bounded distance from a homomorphism.

Proof.

As explained above if QM(Γ)0QM(\Gamma)\neq 0 (i.e., not every quasi-morphism of Γ\Gamma is at bounded distance from a homomorphism) then Γ\Gamma is not U(1)U(1)-stable. On the other hand, if QM(Γ)=0QM(\Gamma)=0, then Ker(Hb2(Γ,𝐑)H2(Γ,𝐑))=0Ker(\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}))=0, i.e., Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) injects into H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}). If in addition Γ\Gamma is finitely presented H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}) is of finite dimension. Thus all the assumptions of Theorem 1.0.8 are satisfied and Γ\Gamma is uniformly U(1)U(1)-stable. ∎

Now, back to the case in which Γ\Gamma is a lattice in a higher rank group. For such Γ\Gamma, Burger and Monod [13] showed that Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) injects into H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}). Such Γ\Gamma is \sayusually finitely presented (and then we can apply the last corollary) except when Γ\Gamma is a lattice in (rank 2) semisimple Lie group of positive characteristic. But in this case Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) is anyway zero, by another result of Burger and Monod [13], so Theorem 1.0.8 applies and we can deduce:

Theorem 1.0.11.

If Γ\Gamma is a higher rank lattice then it is uniformly U(1)U(1)-stable.

The main result of this paper will be a far reaching extension of the above theorem (with some additional assumptions on the lattice Γ\Gamma) to a larger family of metric groups (namely any unitary group U(n)U(n) equipped with any submultiplicative matrix norm \|\cdot\|). In this general case, bounded cohomology theory would not suffice, so we will motivate and build a more appropriate cohomology theory in the upcoming sections.

We conclude this section with an observation in the case of groups of Hermitian type:

Proposition 1.0.12.

Let GG be a split Lie group of Hermitian type (for example, Sp(2m,𝐑)Sp(2m,\mathbf{R})), with universal central extension G~\tilde{G}. Let Γ\Gamma be a cocompact lattice in GG, and Γ~\tilde{\Gamma} be its preimage in G~\tilde{G}. Then Γ~\tilde{\Gamma} is not uniformly U(1)U(1)-stable (and hence, not uniformly 𝔘\mathfrak{U}-stable).

Proof.

Consider the non-trivial 22-cocycle αH2(Γ,𝐙)\alpha\in H^{2}(\Gamma,\mathbf{Z}) corresponding to the central extension Γ~\tilde{\Gamma} of Γ\Gamma. From [28], we know that α\alpha is actually an element of Hb2(Γ,𝐑)\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R}) that does not vanish in H2(Γ,𝐑)H^{2}(\Gamma,\mathbf{R}). In particular, α\alpha is not contained in the kernel of the comparison map c:Hb2(Γ,𝐑)H2(Γ,𝐑)c:\operatorname{H}_{b}^{2}(\Gamma,\mathbf{R})\to H^{2}(\Gamma,\mathbf{R}).
Since the extension Γ~\tilde{\Gamma} of Γ\Gamma is central (in particular, has amenable kernel), α\alpha has a pullback α~Hb2(Γ~,𝐑)\tilde{\alpha}\in\operatorname{H}_{b}^{2}(\tilde{\Gamma},\mathbf{R}) which is a non-trivial bounded 22-cocycle. However, α~\tilde{\alpha} is trivial in cohomology, hence α~\tilde{\alpha} is an element of the kernel of the comparison map c:Hb2(Γ~,𝐑)H2(Γ~,𝐑)c:\operatorname{H}_{b}^{2}(\tilde{\Gamma},\mathbf{R})\to H^{2}(\tilde{\Gamma},\mathbf{R}). Thus, from Propositions Proposition 1.0.3 and Proposition 1.0.4, we conclude that Γ~\tilde{\Gamma} is not uniformly U(1)U(1)-stable. ∎

Remark 1.0.13.

A non-trivial quasimorphism in the above proof of Proposition 1.0.12 can be explicitly described: let j:ΓΓ~j:\Gamma\to\tilde{\Gamma} be a section corresponding to the cocycle αH2(Γ,𝐙)\alpha\in H^{2}(\Gamma,\mathbf{Z}) so that as a set, Γ~=𝐙×j(Γ)\tilde{\Gamma}=\mathbf{Z}\times j(\Gamma). Consider the map ϕ:Γ~𝐑\phi:\tilde{\Gamma}\to\mathbf{R} defined to be ϕ(m,j(γ))m\phi(m,j(\gamma))\coloneqq m. One can check that this is a non-trivial quasimorphism of Γ~\tilde{\Gamma}, and note that this map is \saytrivial on Γ\Gamma and we do know, from Theorem 1.0.11, that Γ\Gamma is indeed uniformly U(1)U(1)-stable if it has rank at least 22, even if Γ~\tilde{\Gamma} is not.

2 Preliminaries and Basic Constructions

In this section, we establish the connection between uniform stability and a homomorphism lifting problem, which is then further explored in §3 for discrete groups, and in §4 for topological groups.
In the first §2.1, we define our central notion of uniform stability, and describe it using sequences of maps. This then allows for a reformulation of the notion of stability as a homomorphism lifting problem using the language of ultrafilters in §2.2. Finally, in §2.3, we introduce the idea of defect diminishing, which is a relaxation of the lifting problem with abelian kernels. This property can naturally be related to a cohomological problem that is then studied in the subsequent §3.

2.1 Uniform Stability and Asymptotic Homomorphisms

Let Γ\Gamma be a countable discrete group, and let (G,dG)(G,d_{G}) be a metric group (that is, a group GG equipped with a bi-invariant metric dGd_{G}). We use the metric to define the (uniform) distance between maps from Γ\Gamma to GG as follows: for f1,f2:ΓGf_{1},f_{2}:\Gamma\to G, the distance between f1f_{1} and f2f_{2}, denoted distΓ,G(f1,f2)dist_{\Gamma,G}(f_{1},f_{2}), as

distΓ,G(f1,f2)supxΓdG(f1(x),f2(x))dist_{\Gamma,G}(f_{1},f_{2})\coloneq\sup_{x\in\Gamma}d_{G}\left(f_{1}(x),f_{2}(x)\right)

This allows us to define the distance of a function f:ΓGf:\Gamma\to G from a homomorphism as follows:

Definition 2.1.1.

The homomorphism distance of a function f:ΓGf:\Gamma\to G, denoted DΓ,G(f)D_{\Gamma,G}(f), is defined as

DΓ,G(f)inf{distΓ,G(f,ψ):ψHom(Γ,G)}D_{\Gamma,G}(f)\coloneq\inf\{dist_{\Gamma,G}(f,\psi):\psi\in Hom(\Gamma,G)\}

The function ff is said to be δ\delta-close to a homomorphism if DΓ,G(f)δD_{\Gamma,G}(f)\leq\delta.

There is another invariant of a function ff that also quantifies its distance from being a homomorphism.

Definition 2.1.2.

For a function f:ΓGf:\Gamma\to G, we define its (uniform) defect defΓ,G(f)def_{\Gamma,G}(f) as

defΓ,G(f)supx,yΓdG(f(xy),f(x)f(y))def_{\Gamma,G}(f)\coloneq\sup_{x,y\in\Gamma}d_{G}\left(f(xy),f(x)f(y)\right)

The function ff is said to be an ϵ\epsilon-homomorphism if defΓ,G(f)ϵdef_{\Gamma,G}(f)\leq\epsilon.

Note that a priori both defΓ,G(f)def_{\Gamma,G}(f) and DΓ,G(f)D_{\Gamma,G}(f) could be \infty. It is easy to show (by the triangle inequality) that defΓ,G(f)3DΓ,G(f)def_{\Gamma,G}(f)\leq 3D_{\Gamma,G}(f) for any function ff, so if ff is close to a homomorphism, then it has small defect. Uniform stability is the question of whether the converse is true: is any function with small defect necessarily close to a homomorphism?
Uniform stability is usually studied with respect to not just one metric group (G,dG)(G,d_{G}) but a family 𝒢\mathcal{G} of metric groups. In this case, we can define

FΓ,𝒢:[0,][0,]F_{\Gamma,\mathcal{G}}:[0,\infty]\to[0,\infty]
FΓ,𝒢(ϵ)supG𝒢supf{DΓ,G(f):defΓ,G(f)ϵ}F_{\Gamma,\mathcal{G}}(\epsilon)\coloneq\sup_{G\in\mathcal{G}}\sup_{f}\{D_{\Gamma,G}(f):def_{\Gamma,G}(f)\leq\epsilon\}
Definition 2.1.3.

The group Γ\Gamma is said to be uniformly 𝒢\mathcal{G}-stable if

limϵ0+FΓ,𝒢(ϵ)=0\lim\limits_{\epsilon\to 0^{+}}F_{\Gamma,\mathcal{G}}(\epsilon)=0

A further refinement of the above definition involves a quanitification of the above convergence:

Definition 2.1.4.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate if ϵ0>0\exists\epsilon_{0}>0 and M0\exists M\geq 0 such that ϵ<ϵ0\forall\epsilon<\epsilon_{0} and G𝒢\forall G\in\mathcal{G}, every ϵ\epsilon-homomorphism ϕ:ΓG\phi:\Gamma\to G is MϵM\epsilon-close to a homomorphism.

It is helpful to rephrase these notions also in terms of sequences of maps, especially since we shall further refine this view in §2.2 and §2.3. Consider a sequence of functions {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} where Gn𝒢G_{n}\in\mathcal{G} for every n𝐍n\in\mathbf{N}.

  • A sequence {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} is said to be a (uniform) asymptotic homomorphism of Γ\Gamma to 𝒢\mathcal{G} if limndefΓ,Gn(ϕn)=0\lim\limits_{n\to\infty}def_{\Gamma,G_{n}}(\phi_{n})=0.

  • A sequence {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} is said to be (uniformly) asymptotically close to a homomorphism if limnDΓ,Gn(ϕn)=0\lim\limits_{n\to\infty}D_{\Gamma,G_{n}}(\phi_{n})=0.

It is easy to see that a group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable iff every asymptotic homomorphism of Γ\Gamma to 𝒢\mathcal{G} is asymptotically close to a homomorphism.
Uniform 𝒢\mathcal{G}-stability with a linear estimate too can be rephrased in terms of sequences of maps. Recall the Laundau big-O notation: for sequences {xn}n𝐍\{x_{n}\}_{n\in\mathbf{N}} and {yn}n𝐍\{y_{n}\}_{n\in\mathbf{N}} of positive real numbers, we denote xn=O(yn)x_{n}=O(y_{n}) if there exists a constant C0C\geq 0 and N𝐍N\in\mathbf{N} such that for all nNn\geq N, xnCynx_{n}\leq Cy_{n}. We denote by xn=o(yn)x_{n}=o(y_{n}) if there exists a sequence {ϵn}n𝐍\{\epsilon_{n}\}_{n\in\mathbf{N}} with limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0 and such that xn=ϵnynx_{n}=\epsilon_{n}y_{n}. Firstly, note that if Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate, then for any asymptotic homomorphism {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} of Γ\Gamma to 𝒢\mathcal{G}, DΓ,Gn(ϕn)=O(defΓ,Gn(ϕn))D_{\Gamma,G_{n}}(\phi_{n})=O\left(def_{\Gamma,G_{n}}(\phi_{n})\right). The following lemma shows that the converse is also true:

Lemma 2.1.5.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate iff for every asymptotic homomorphism {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} of Γ\Gamma to 𝒢\mathcal{G},

DΓ,Gn(ϕn)=O(defΓ,Gn(ϕn))D_{\Gamma,G_{n}}(\phi_{n})=O\left(def_{\Gamma,G_{n}}(\phi_{n})\right)
Proof.

Suppose Γ\Gamma is not uniformly 𝒢\mathcal{G}-stable with a linear estimate. Then for any M>0M>0 and any ϵ>0\epsilon>0, there exists a map ϕ:ΓG\phi:\Gamma\to G (for some G𝒢G\in\mathcal{G}) such that defΓ,G(ϕ)ϵdef_{\Gamma,G}(\phi)\leq\epsilon and DΓ,G(ϕ)>MϵD_{\Gamma,G}(\phi)>M\epsilon. Now consider a sequence {Mn}n𝐍\{M_{n}\}_{n\in\mathbf{N}} with limnMn=\lim_{n\to\infty}M_{n}=\infty and a sequence {ϵn}n𝐍\{\epsilon_{n}\}_{n\in\mathbf{N}} with limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0. For each n𝐍n\in\mathbf{N}, let ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} be the map with defΓ,Gn(ϕn)ϵndef_{\Gamma,G_{n}}(\phi_{n})\leq\epsilon_{n} but DΓ,Gn(ϕn)>MnϵnD_{\Gamma,G_{n}}(\phi_{n})>M_{n}\epsilon_{n}. Then {ϕn}n𝐍\{\phi_{n}\}_{n\in\mathbf{N}} is an asymptotic homomorphism of Γ\Gamma to 𝒢\mathcal{G} such that DΓ,Gn(ϕn)D_{\Gamma,G_{n}}(\phi_{n}) is clearly not O(defΓ,Gn(ϕn))O\left(def_{\Gamma,G_{n}}(\phi_{n})\right). ∎

We shall now conclude this section by informally introducing the following relaxation of the hypothesis of Lemma 2.1.5, which we shall call asymptotic defect diminishing. The idea is to look for improvements to the asymptotic homomorphism, rather than a true homomorphism.

Definition 2.1.6.

The group Γ\Gamma is said to have the asymptotic defect diminishing property with respect to the family 𝒢\mathcal{G} if for any asymptotic homomorphism {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}}, there exists an asymptotic homomorphism {ψn:ΓGn}n𝐍\{\psi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} such that

  • The defect defΓ,Gn(ψn)=o(defΓ,Gn(ϕn))def_{\Gamma,G_{n}}(\psi_{n})=o\left(def_{\Gamma,G_{n}}(\phi_{n})\right).

  • The distance distΓ,Gn(ϕn,ψn)=O(defΓ,Gn(ϕn))dist_{\Gamma,G_{n}}(\phi_{n},\psi_{n})=O\left(def_{\Gamma,G_{n}}(\phi_{n})\right).

The following results motivate this notion, which we shall formally study in more detail in §2.2 in an ultraproduct setting.

Lemma 2.1.7.

Suppose Γ\Gamma has the asymptotic defect diminishing property with respect to 𝒢\mathcal{G}. Then there exists ϵ0>0\epsilon_{0}>0 and M>0M>0 such that for ϵ<ϵ0\epsilon<\epsilon_{0} and any ϵ\epsilon-homomorphism ϕ:ΓG\phi:\Gamma\to G, there exists a map ψ:ΓG\psi:\Gamma\to G with defect defΓ,G(ψ)<12defΓ,G(ϕ)def_{\Gamma,G}(\psi)<\frac{1}{2}def_{\Gamma,G}(\phi) and distΓ,G(ψ)<MdefΓ,G(ϕ)dist_{\Gamma,G}(\psi)<Mdef_{\Gamma,G}(\phi).

Proof.

We shall prove this by contradiction, so suppose for every ϵ>0\epsilon>0 and every M>0M>0, there exists an ϵ\epsilon-homomorphism ϕ:ΓG\phi:\Gamma\to G such that for any ψ:ΓG\psi:\Gamma\to G, either defΓ,G(ψ)12defΓ,G(ϕ)def_{\Gamma,G}(\psi)\geq\frac{1}{2}def_{\Gamma,G}(\phi) or distΓ,G(ψ)MdefΓ,Gn(ϕn)dist_{\Gamma,G}(\psi)\geq Mdef_{\Gamma,G_{n}}(\phi_{n}).
Consider a sequence {Mn}n𝐍\{M_{n}\}_{n\in\mathbf{N}} with limnMn=\lim_{n\to\infty}M_{n}=\infty and a sequence {ϵn}n𝐍\{\epsilon_{n}\}_{n\in\mathbf{N}} with limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0. For each n𝐍n\in\mathbf{N}, let ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} (with Gn𝒢G_{n}\in\mathcal{G}) be a map with defΓ,Gn(ϕn)ϵndef_{\Gamma,G_{n}}(\phi_{n})\leq\epsilon_{n} such that for any map ψ:ΓGn\psi:\Gamma\to G_{n}, either defΓ,Gn(ψn)12defΓ,Gn(ϕn)def_{\Gamma,G_{n}}(\psi_{n})\geq\frac{1}{2}def_{\Gamma,G_{n}}(\phi_{n}) or distΓ,Gn(ψn)MndefΓ,Gn(ϕn)dist_{\Gamma,G_{n}}(\psi_{n})\geq M_{n}def_{\Gamma,G_{n}}(\phi_{n}). Then the asymptotic homomorphism {ϕn}n𝐍\{\phi_{n}\}_{n\in\mathbf{N}} as constructed proves that that Γ\Gamma does not have the asymptotic defect diminishing property with respect to 𝒢\mathcal{G}. ∎

Theorem 2.1.8.

Suppose 𝒢\mathcal{G} is such that every G𝒢G\in\mathcal{G} is a complete metric space (with respect to its metric dGd_{G}). Then group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate iff Γ\Gamma has the asymptotic defect diminishing property with respect to 𝒢\mathcal{G}

Proof.

Suppose Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate, then the implication is immediate. Conversely, suppose Γ\Gamma has the asymptotic defect diminishing property with respect to 𝒢\mathcal{G}. From the previous lemma, this means that there exists ϵ0>0\epsilon_{0}>0 and M>0M>0 such that for any ϵ\epsilon-homomorphism ϕ:ΓG\phi:\Gamma\to G (with ϵ<ϵ0\epsilon<\epsilon_{0}), there exists a map ψ:ΓG\psi:\Gamma\to G with defect defΓ,G(ψ)<12defΓ,G(ϕ)def_{\Gamma,G}(\psi)<\frac{1}{2}def_{\Gamma,G}(\phi) and distΓ,G(ψ)<MdefΓ,G(ϕ)dist_{\Gamma,G}(\psi)<Mdef_{\Gamma,G}(\phi). Set ϕ(0)ϕ\phi^{(0)}\coloneq\phi and ϕ(1)ψ\phi^{(1)}\coloneq\psi. Applying Lemma 2.1.7 inductively on ϕ(i)\phi^{(i)} to get ϕ(i+1)\phi^{(i+1)}, we obtain a sequence of maps {ϕ(j):ΓG}j𝐍\{\phi^{(j)}:\Gamma\to G\}_{j\in\mathbf{N}} where for each j𝐍j\in\mathbf{N}, ϕ(j)\phi^{(j)} has defect at most ϵ/2j\epsilon/2^{j} and is of distance at most Mϵ/2jM\epsilon/2^{j} from ϕ(j1)\phi^{(j-1)}. This gives us a Cauchy sequence of maps from ΓG\Gamma\to G. Since GG is a complete, the sequence {ϕ(j)\{\phi^{(j)} has a limit ϕ:ΓG\phi^{\infty}:\Gamma\to G with defect def(ϕ)=0def(\phi^{\infty})=0 (hence it is a homomorphism). As for its distance from ϕ\phi,

distΓ,G(ϕ,ϕ)j=0Mϵ2j=2Mϵdist_{\Gamma,G}(\phi,\phi^{\infty})\leq\sum\limits_{j=0}^{\infty}\frac{M\epsilon}{2^{j}}=2M\epsilon

Hence Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate. ∎

2.2 Ultraproducts and Internal Maps

We can quantify the asymptotic rates succintly using ultraproducts. We shall now briefly review some concepts from the theory of ultraproducts and non-standard analysis that would be of relevance to our constructions (for more details, refer [27] and [2]).
Let 𝒰\mathcal{U} be a non-principal ultrafilter on 𝐍\mathbf{N}, which is fixed throughout. A subset S𝐍S\subseteq\mathbf{N} is said to be large if S𝒰S\in\mathcal{U}.

Definition 2.2.1.

The (algebraic) ultraproduct 𝒰Xn\prod_{\mathcal{U}}X_{n} (or alternately, {Xn}𝒰\{X_{n}\}_{\mathcal{U}}) of an indexed collection {Xn}n𝐍\{X_{n}\}_{n\in\mathbf{N}} of sets is defined to be

𝒰Xnn𝐍Xn/\prod_{\mathcal{U}}X_{n}\coloneq\prod_{n\in\mathbf{N}}X_{n}/\sim

where for {xn}n𝐍,{yn}n𝐍n𝐍Xn\{x_{n}\}_{n\in\mathbf{N}},\{y_{n}\}_{n\in\mathbf{N}}\in\prod_{n\in\mathbf{N}}X_{n}, {xn}n𝐍{yn}n𝐍\{x_{n}\}_{n\in\mathbf{N}}\sim\{y_{n}\}_{n\in\mathbf{N}} if {n:xn=yn}𝒰\{n:x_{n}=y_{n}\}\in\mathcal{U}.

In other words, we identify two sequences {xn}n𝐍,{yn}n𝐍n𝐍Xn\{x_{n}\}_{n\in\mathbf{N}},\{y_{n}\}_{n\in\mathbf{N}}\in\prod_{n\in\mathbf{N}}X_{n} if they agree on a large set of indices. The image of a sequence {xn}n𝐍n𝐍Xn\{x_{n}\}_{n\in\mathbf{N}}\in\prod_{n\in\mathbf{N}}X_{n} under this equivalence relation shall be denoted {xn}𝒰\{x_{n}\}_{\mathcal{U}}. Conversely, given an element of 𝒰Xn\prod_{\mathcal{U}}X_{n}, we shall always regard it as {xn}𝒰\{x_{n}\}_{\mathcal{U}} for some sequence {xn}n𝐍n𝐍Xn\{x_{n}\}_{n\in\mathbf{N}}\in\prod_{n\in\mathbf{N}}X_{n}.
If Xn=XX_{n}=X for every n𝐍n\in\mathbf{N}, then 𝒰X\prod_{\mathcal{U}}X is called the (algebraic) ultrapower of XX, denoted X{}^{*}X. Note that XX can be embedded in X{}^{*}X via a diagonal embedding (for xXx\in X, x{x}𝒰Xx\mapsto\{x\}_{\mathcal{U}}\in{}^{*}X).
Ultraproducts can be made to inherit algebraic structures of their building blocks. More precisely, let {Xn}n𝐍\{X_{n}\}_{n\in\mathbf{N}}, {Yn}n𝐍\{Y_{n}\}_{n\in\mathbf{N}}, and {Zn}n𝐍\{Z_{n}\}_{n\in\mathbf{N}} be indexed families of sets with operations n:Xn×YnZn*_{n}:X_{n}\times Y_{n}\to Z_{n} for every n𝐍n\in\mathbf{N}. This naturally defines an operation :𝒰Xn×𝒰Yn𝒰Zn*:\prod_{\mathcal{U}}X_{n}\times\prod_{\mathcal{U}}Y_{n}\to\prod_{\mathcal{U}}Z_{n} by

{xn}𝒰{yn}𝒰={xnnyn}𝒰\{x_{n}\}_{\mathcal{U}}*\{y_{n}\}_{\mathcal{U}}=\{x_{n}*_{n}y_{n}\}_{\mathcal{U}}

We shall frequently encounter the following examples of ultraproducts:

  • The ultrapower G{}^{*}G of a group GG is itself a group. This can be seen by noting that G{}^{*}G is the quotient of the direct product group n𝐍G\prod_{n\in\mathbf{N}}G by the normal subgroup comprising elements {gn}n𝐍\{g_{n}\}_{n\in\mathbf{N}} with {n:gn=1}𝒰\{n:g_{n}=1\}\in\mathcal{U}.

  • The ultrapower 𝐑{}^{*}\mathbf{R} of 𝐑\mathbf{R} is a non-archimedean ordered field called the hyperreals.

  • Let {Wn}n𝐍\{W_{n}\}_{n\in\mathbf{N}} be a family of Banach spaces. Then 𝒰Wn\prod_{\mathcal{U}}W_{n} can be given the structure of a 𝐑{}^{*}\mathbf{R}-vector space. In fact, it also comes equipped with a 𝐑{}^{*}\mathbf{R}-valued norm.

One of the standard tools of non-standard analysis is the transfer principle which relates the truth of statements concerning objects and their counterparts in the ultraproduct universe. Intuitively, our standard universe comprises all objects under normal consideration like 𝐑\mathbf{R}, 𝐂\mathbf{C}, etc. but maybe formally modeled as follows: define

V0(𝐑)=𝐑V_{0}(\mathbf{R})=\mathbf{R}
Vn+1(𝐑)=Vn(𝐑)𝒫(Vn(𝐑))V_{n+1}(\mathbf{R})=V_{n}(\mathbf{R})\cup\mathcal{P}\left(V_{n}(\mathbf{R})\right)

where 𝒫(Vn(𝐑))\mathcal{P}\left(V_{n}(\mathbf{R})\right) denotes the power set of (Vn(𝐑))\left(V_{n}(\mathbf{R})\right). Then

V(𝐑)=n0Vn(𝐑)V(\mathbf{R})=\cup_{n\geq 0}V_{n}(\mathbf{R})

is called the superstructure over 𝐑\mathbf{R}, and can be interpreted as comprising all the natural structures we study in mathematics. This shall comprise our standard universe UnivUniv.
We can construct a mapping :V(𝐑)V(𝐑){}^{*}:V(\mathbf{R})\to V({}^{*}\mathbf{R}) that takes an object in the superstructure of 𝐑\mathbf{R} (our standard universe UnivUniv) to an object in the superstructure of 𝐑{}^{*}\mathbf{R} satisfying the following:

  • (Extension Principle) The mapping * maps 𝐑\mathbf{R} to 𝐑{}^{*}\mathbf{R}.

  • (Transfer Principle) For any first-order formula ϕ\phi involving kk variables, and A1,,AkV(𝐑)A_{1},\dots,A_{k}\in V(\mathbf{R}), the statement ϕ(A1,,Ak)\phi(A_{1},\dots,A_{k}) is true in V(𝐑)V(\mathbf{R}) iff ϕ(A1,,Ak)\phi({}^{*}A_{1},\dots,{}^{*}A_{k}) is true in V(𝐑)V({}^{*}\mathbf{R}).

  • (Countable Saturation) Suppose {Xn}n𝐍\{X_{n}\}_{n\in\mathbf{N}} is a collection of sets in (V(𝐑)){}^{*}\left(V(\mathbf{R})\right) such that the intersection of any finite subcollection is non-empty. Then Xn\cap X_{n} is non-empty.

It is a basic result of non-standard analysis that such a mapping * exists, and can be constructed using a non-principal ultrafilter on 𝐍\mathbf{N}. The image of this mapping shall be referred to as our non-standard universeUniv{}^{*}Univ, which is

Univ={{Xn}𝒰|XnUniv}{}^{*}Univ=\{\{X_{n}\}_{\mathcal{U}}\lvert X_{n}\in Univ\}

comprising ultraproducts of elements of UnivUniv. Note that 𝐑,𝐂,Γ{}^{*}\mathbf{R},{}^{*}\mathbf{C},{}^{*}\Gamma are all contained in Univ{}^{*}Univ.
Objects contained in UnivUniv are called standard, while objects contained in Univ{}^{*}Univ are called internal. In particular, a subset SS of an ultraproduct 𝒰Xn\prod_{\mathcal{U}}X_{n} is an internal subset if there exist subsets SnXnS_{n}\subseteq X_{n} for every n𝐍n\in\mathbf{N} such that S=𝒰SnS=\prod_{\mathcal{U}}S_{n}. A function f:𝒰Xn𝒰Ynf:\prod_{\mathcal{U}}X_{n}\to\prod_{\mathcal{U}}Y_{n} is said to be an internal function if there exists a sequence {fn:XnYn}n𝐍\{f_{n}:X_{n}\to Y_{n}\}_{n\in\mathbf{N}} such that f={fn}𝒰f=\{f_{n}\}_{\mathcal{U}}.
Objects that are not contained in Univ{}^{*}Univ are called external. Note that standard objects like 𝐑\mathbf{R} and 𝐂\mathbf{C} too are external. We can always consider objects that are neither in UnivUniv nor in Univ{}^{*}Univ. Two important examples of non-standard external subsets are

  • The set of bounded hyperreals, denoted 𝐑b{}^{*}\mathbf{R}_{b}, is the subset comprising elements {xn}𝒰𝐑\{x_{n}\}_{\mathcal{U}}\in{}^{*}\mathbf{R} for which there exists C𝐑0C\in\mathbf{R}_{\geq 0} such that |xn|C\lvert x_{n}\rvert\leq C for every n𝐍n\in\mathbf{N}.

  • The set of infinitesimal hyperreals, denoted 𝐑inf{}^{*}\mathbf{R}_{inf}, is the subset comprising elements {xn}𝒰𝐑\{x_{n}\}_{\mathcal{U}}\in{}^{*}\mathbf{R} such that for every real ϵ>0\epsilon>0, there exists a large set S𝒰S\in\mathcal{U} such that |xn|<ϵ\lvert x_{n}\rvert<\epsilon for every nSn\in S.

  • For x,y𝐑x,y\in{}^{*}\mathbf{R}, denote by x=O𝒰(y)x=O_{\mathcal{U}}(y) if x/y𝐑bx/y\in{}^{*}\mathbf{R}_{b}, and by x=o𝒰(y)x=o_{\mathcal{U}}(y) if x/y𝐑infx/y\in{}^{*}\mathbf{R}_{inf} (in particular,any bounded element x𝐑bx\in{}^{*}\mathbf{R}_{b} is x=O𝒰(1)x=O_{\mathcal{U}}(1) while any infinitesimal element ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} is ϵ=o𝒰(1)\epsilon=o_{\mathcal{U}}(1)).

Note that the preimage of 𝐑b{}^{*}\mathbf{R}_{b} under the map n𝐍𝐑𝐑\prod_{n\in\mathbf{N}}\mathbf{R}\to{}^{*}\mathbf{R} includes the subset of all bounded sequences, while the preimage of 𝐑inf{}^{*}\mathbf{R}_{inf} includes the subset of infinitesimal sequences (that is, sequences that converge to 0).
The subset 𝐑b{}^{*}\mathbf{R}_{b} forms a valuation ring with 𝐑inf{}^{*}\mathbf{R}_{inf} being the unique maximal ideal, with quotient 𝐑b/𝐑inf𝐑{}^{*}\mathbf{R}_{b}/{}^{*}\mathbf{R}_{inf}\cong\mathbf{R}. The quotient map st:𝐑b𝐑st:{}^{*}\mathbf{R}_{b}\to\mathbf{R} is known as the standard part map or limit along the ultrafilter 𝒰\mathcal{U}.
The previous construction can also be replicated for metric spaces with specified base points. Let {(Xn,dn,pn)}n𝐍\{(X_{n},d_{n},p_{n})\}_{n\in\mathbf{N}} be an indexed family of metric spaces (where the space XnX_{n} comes with the metric dnd_{n} and the base point pnp_{n}). The ultraproduct 𝒰Xn\prod_{\mathcal{U}}X_{n} comes equipped with an 𝐑{}^{*}\mathbf{R}-values metric 𝒰dn\prod_{\mathcal{U}}d_{n} and base point 𝒰pn\prod_{\mathcal{U}}p_{n}. Consider the subset, denoted (𝒰Xn)b\left(\prod_{\mathcal{U}}X_{n}\right)_{b}, of 𝒰Xn\prod_{\mathcal{U}}X_{n} comprising {xn}𝒰\{x_{n}\}_{\mathcal{U}} such that {dn(xn,pn)}𝒰𝐑b\{d_{n}(x_{n},p_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{b} (such elements are referred to as bounded or admissible), and a subset, denoted (𝒰Xn)inf\left(\prod_{\mathcal{U}}X_{n}\right)_{inf}, comprising {xn}𝒰\{x_{n}\}_{\mathcal{U}} such that {dn(xn,pn)}𝒰𝐑inf\{d_{n}(x_{n},p_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}. Define an equivalence relation \sim on (𝒰Xn)b\left(\prod_{\mathcal{U}}X_{n}\right)_{b} by {xn}𝒰{yn}𝒰\{x_{n}\}_{\mathcal{U}}\sim\{y_{n}\}_{\mathcal{U}} if {dn(xn,yn)}𝒰𝐑inf\{d_{n}(x_{n},y_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}. The set of equivalence classes (𝒰Xn)b/\left(\prod_{\mathcal{U}}X_{n}\right)_{b}/\sim is called the ultralimit of {(Xn,dn,pn)}n𝐍\{(X_{n},d_{n},p_{n})\}_{n\in\mathbf{N}}. We will return to this notion in the context of Banach spaces in §4.

Remark 2.2.2.

For convenience, we shall henceforth denote by \sayfor n𝒰n\in\mathcal{U} to mean \sayfor nSn\in S for some S𝒰S\in\mathcal{U}.

Returning to our setting, consider a sequence {ϕn:ΓGn}n𝐍\{\phi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} where Gn𝒢G_{n}\in\mathcal{G}. This can be used to construct an internal map {ϕn}𝒰:Γ𝒰Gn\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, and allows us to redefine asymptotic homomorphisms and closeness to homomorphisms in the ultraproduct.

  • Given two internal maps ϕ(1):Γ𝒰Gn\phi^{(1)}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} and ϕ(2):Γ𝒰Gn\phi^{(2)}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, we denote by dist(ϕ(1),ϕ(2)){distΓ,Gn(ϕn(1),ϕn(2))}𝒰dist(\phi^{(1)},\phi^{(2)})\coloneq\{dist_{\Gamma,G_{n}}(\phi^{(1)}_{n},\phi^{(2)}_{n})\}_{\mathcal{U}}. The maps ϕ(1)\phi^{(1)} and ϕ(2)\phi^{(2)} are said to be (internally) asymptotically close to each other if dist(ϕ(1),ϕ(2))𝐑infdist(\phi^{(1)},\phi^{(2)})\in{}^{*}\mathbf{R}_{inf}.

  • An internal map ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is said to be an internal homomorphism if there exists a sequence {ψn}n𝐍\{\psi_{n}\}_{n\in\mathbf{N}} of homomorphisms such that ψ={ψn}𝒰\psi=\{\psi_{n}\}_{\mathcal{U}}.

  • An internal map ϕ{ϕn}𝒰:Γ𝒰Gn\phi\coloneq\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is said to be (internally) asymptotically close to an internal homomorphism if there exists an internal homomorphism ψ={ψn:ΓGn}n𝐍\psi=\{\psi_{n}:\Gamma\to G_{n}\}_{n\in\mathbf{N}} such that dist(ϕ,ψ)𝐑infdist(\phi,\psi)\in{}^{*}\mathbf{R}_{inf} (in other words, {DΓ,Gn}𝒰𝐑inf\{D_{\Gamma,G_{n}}\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}).

  • An internal map ϕ{ϕn}𝒰:Γ𝒰Gn\phi\coloneq\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is called an internal asymptotic homomorphism if def(ϕ){def(ϕn)}𝒰𝐑infdef(\phi)\coloneq\{def(\phi_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}. For ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}, we shall call an internal asymptotic homomorphism ϕ\phi an internal ϵ\epsilon-homomorphism if def(ϕ)=ϵdef(\phi)=\epsilon (we similarly define internal O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-homomorphisms and internal o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-homomorphisms).

The following lemmas are variants of Lemma 2.1.5 and Theorem 2.1.8 in the ultraproduct setting:

Lemma 2.2.3.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable iff every internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} for Gn𝒢G_{n}\in\mathcal{G} is asymptotically close to an internal homomorphism.

Proof.

We shall prove both directions by contradiction.
Let ϕ={ϕn}𝒰:Γ𝒰Gn\phi=\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal asymptotic homomorphism that is not asymptotically close to any internal homomorphism. This means that {def(ϕn)}𝒰𝐑inf\{def(\phi_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}, but there exists some real c>0c>0 and S𝒰S\in\mathcal{U} such that DΓ,Gn(ϕn)cD_{\Gamma,G_{n}}(\phi_{n})\geq c for nSn\in S. In particular, for this cc and any ϵ>0\epsilon>0, there exists a map ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} with defect def(ϕn)ϵdef(\phi_{n})\leq\epsilon and DΓ,GncD_{\Gamma,G_{n}}\geq c, thus proving that Γ\Gamma is not uniformly 𝒢\mathcal{G}-stable.
Conversely, suppose Γ\Gamma is not uniformly 𝒢\mathcal{G}-stable. This means, by definition, that there exists δ>0\delta>0 such that for every ϵ>0\epsilon>0, there exists G𝒢G\in\mathcal{G} and ϕ:ΓG\phi:\Gamma\to G such that def(ϕ)ϵdef(\phi)\leq\epsilon and distΓ,G(ϕ)>δdist_{\Gamma,G}(\phi)>\delta. Let us fix this δ\delta, and consider a sequence {ϵn}n𝐍\{\epsilon_{n}\}_{n\in\mathbf{N}} with limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0. For each such ϵn\epsilon_{n}, there exists Gn𝒢G_{n}\in\mathcal{G} and ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} with def(ϕn)ϵndef(\phi_{n})\leq\epsilon_{n} and distΓ,Gn(ϕn)>δdist_{\Gamma,G_{n}}(\phi_{n})>\delta. Consider the internal asymptotic homomorphism {ϕn}𝒰\{\phi_{n}\}_{\mathcal{U}}. Clearly it is not asymptotically close to any internal homomorphism. ∎

Lemma 2.2.4.

The group Γ\Gamma is 𝒢\mathcal{G}-uniformly stable with a linear estimate iff for every internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, there exists an internal homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} with dist(ϕ,ψ)=O𝒰(def(ϕ))dist(\phi,\psi)=O_{\mathcal{U}}\left(def(\phi)\right).

Proof.

(The proof is similar to Lemma 2.1.5) Suppose Γ\Gamma is not uniformly 𝒢\mathcal{G}-stable with a linear estimate. Consider a sequence {Mn}n𝐍\{M_{n}\}_{n\in\mathbf{N}} with limnMn=\lim_{n\to\infty}M_{n}=\infty and a sequence {ϵn}n𝐍\{\epsilon_{n}\}_{n\in\mathbf{N}} with limnϵn=0\lim_{n\to\infty}\epsilon_{n}=0. For each n𝐍n\in\mathbf{N}, let ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} be the map with defΓ,Gn(ϕn)ϵndef_{\Gamma,G_{n}}(\phi_{n})\leq\epsilon_{n} but DΓ,Gn(ϕn)>MnϵnD_{\Gamma,G_{n}}(\phi_{n})>M_{n}\epsilon_{n}. Then the ultraproduct ϕ={ϕn}𝒰:Γ𝒰Gn\phi=\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is an internal asymptotic homomorphism with def(ϕ)=ϵ{ϵn}𝒰def(\phi)=\epsilon\coloneq\{\epsilon_{n}\}_{\mathcal{U}} such that for any internal homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, dist(ϕ,ψ)/def(ϕ)dist(\phi,\psi)/def(\phi) is not in 𝐑b{}^{*}\mathbf{R}_{b}.
Conversely, suppose Γ\Gamma is 𝒢\mathcal{G}-uniformly stable with a linear estimate, and let ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal asymptotic homomorphism. There exists ϵ0>0\epsilon_{0}>0, M>0M>0 and a large subset Sϵ0𝒰S_{\epsilon_{0}}\in\mathcal{U} such that for every nSϵ0n\in S_{\epsilon_{0}}, ϕn:ΓGn\phi_{n}:\Gamma\to G_{n} has defect def(ϕn)ϵ0def(\phi_{n})\leq\epsilon_{0}, and DΓ,Gn(ϕn)Mdef(ϕn)D_{\Gamma,G_{n}}(\phi_{n})\leq Mdef(\phi_{n}), allowing us to construct an internal homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} that is asymptotically close to ϕ\phi. ∎

We now reformulate the notion of defect diminishing in this ultraproduct setting:

Definition 2.2.5.

The group Γ\Gamma is said to have the defect diminishing property with respect to the family 𝒢\mathcal{G} if for any internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, there exists an asymptotic homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} such that

  • The defect def(ψ)=o𝒰(def(ϕ))def(\psi)=o_{\mathcal{U}}\left(def(\phi)\right).

  • The distance dist(ϕ,ψ)=O𝒰(def(ϕ))dist(\phi,\psi)=O_{\mathcal{U}}\left(def(\phi)\right).

The following lemma reformulates Theorem 2.1.8 in the ultraproduct setting, and the proof is on the same lines.

Theorem 2.2.6.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate iff Γ\Gamma has the defect diminishing property with respect to 𝒢\mathcal{G}.

2.3 Internal Liftings and Defect Diminishing

In this §, we shall reinterpret an internal asymptotic homomorphism as a (true) homomorpism to a quotient group. This will allow us to describe defect diminishing as a homomorphism lifting problem. In §3.1 we will restrict to a family of metric groups of interest such that this homomorphism lifting problem has an abelian kernel, allowing us to build a cohomological theory capturing the obstruction to uniform stability.
Let ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal asymptotic homomorphism. Consider the external subset (𝒰Gn)inf\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} defined as

(𝒰Gn)inf{{gn}𝒰:{dn(gn,1)}𝒰𝐑inf}\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}\coloneq\Big{\{}\{g_{n}\}_{\mathcal{U}}:\{d_{n}(g_{n},1)\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}\Big{\}}

In other words, (𝒰Gn)inf\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} comprises all elements that are infinitesimally close to the identity in 𝒰Gn\prod_{\mathcal{U}}G_{n}. It is easy to check that (𝒰Gn)inf\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} is not just a subset but a normal subgroup of 𝒰Gn\prod_{\mathcal{U}}G_{n}. Since ϕ\phi has defect def(ϕ)𝐑infdef(\phi)\in{}^{*}\mathbf{R}_{inf}, this means that for every x,yΓx,y\in{}^{*}\Gamma, ϕ(xy)1ϕ(x)ϕ(y)(𝒰Gn)inf\phi(xy)^{-1}\phi(x)\phi(y)\in\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}, making ϕ~:Γ𝒰Gn/(𝒰Gn)inf\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} a homomorphism. Thus,

Lemma 2.3.1.

Given an internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, its composition, denoted ϕ~\tilde{\phi}, with the canonical quotient homomorphism 𝒰Gn𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} is a homomorphism from Γ{}^{*}\Gamma to the group 𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}.

Note that a homomorphism from Γ{}^{*}\Gamma to 𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} does not necessarily arise as the composition of an internal map Γ𝒰Gn{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} and the quotient homomorphism 𝒰Gn𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}. However, we will be specifically interested in homomorphisms that arise this way.

Definition 2.3.2.

Let F:Γ𝒰Gn/(𝒰Gn)infF:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} be a homomorphism. We say that FF has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} if F^\hat{F} is internal, and its composition with the canonical quotient homomorphism 𝒰Gn𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} is FF.
We say that FF has an internal lift homomorphism if there exists an internal lift F^\hat{F} of FF that is also a homomorphism from Γ{}^{*}\Gamma to 𝒰Gn\prod_{\mathcal{U}}G_{n}.

Observe that for an internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, ϕ\phi itself is an internal lift of the homomorphism ϕ~\tilde{\phi}. Conversely,

Lemma 2.3.3.

Suppose a homomorphism F:Γ𝒰Gn/(𝒰Gn)infF:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}. Then F^\hat{F} is an internal asymptotic homomorphism. Furthermore, suppose F^(1)\hat{F}^{(1)} and F^(2)\hat{F}^{(2)} are two internal lifts of a homomorphism F:Γ𝒰Gn/(𝒰Gn)infF:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}, then F^(1)\hat{F}^{(1)} and F^(2)\hat{F}^{(2)} are asymptotically close to each other.

Proof.

The map F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} , being internal, is of the form F^={F^n}𝒰\hat{F}=\{\hat{F}_{n}\}_{\mathcal{U}} for F^n:ΓGn\hat{F}_{n}:\Gamma\to G_{n} for every n𝐍n\in\mathbf{N}. Since FF is a homomorphism, for every x={xn}𝒰,y={yn}𝒰Γx=\{x_{n}\}_{\mathcal{U}},y=\{y_{n}\}_{\mathcal{U}}\in{}^{*}\Gamma, {F^n(xnyn)1F^n(xn)F^n(yn)}𝒰(𝒰Gn)inf\{\hat{F}_{n}(x_{n}y_{n})^{-1}\hat{F}_{n}(x_{n})\hat{F}_{n}(y_{n})\}_{\mathcal{U}}\in\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} which means that {def(F^n)}𝒰𝐑inf\{def(\hat{F}_{n})\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}, making F^\hat{F} an internal asymptotic homomorphism.
Since both F^(1)\hat{F}^{(1)} and F^(2)\hat{F}^{(2)} are internal lifts of the homomorphism F:Γ𝒰Gn/(𝒰Gn)infF:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}, for every x={xn}𝒰Γx=\{x_{n}\}_{\mathcal{U}}\in{}^{*}\Gamma, {dn(F^n(1)(xn),F^n(2)(xn))}𝒰𝐑inf\{d_{n}(\hat{F}^{(1)}_{n}(x_{n}),\hat{F}^{(2)}_{n}(x_{n}))\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{inf}. ∎

This motivates the following equivalent condition for uniform 𝒢\mathcal{G}-stability in terms of internal lifts.

Lemma 2.3.4.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable iff every homomorphism ϕ~:Γ𝒰Gn/(𝒰Gn)inf\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} that has an internal lift also has an internal lift homomorphism.

Proof.

Let ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal asymptotic homomorphism, and ϕ~:Γ𝒰Gn/(𝒰Gn)inf\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} be the homomorphism obtained by composing ϕ\phi with the quotient map 𝒰Gn𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf}. Let ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal lift homomorphism of ϕ~\tilde{\phi}. Then by the previous Lemma 2.3.3, ϕ\phi is asymptotically close to the internal homomorphism ψ\psi. By Lemma 2.2.3, we conclude that Γ\Gamma is uniformly 𝒢\mathcal{G}-stable. The converse follows by definition of uniform 𝒢\mathcal{G}-stability. ∎

Remark 2.3.5.

The quotient group 𝒰Gn/(𝒰Gn)inf\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} is called the metric ultraproduct of the sequence {Gn}n𝐍\{G_{n}\}_{n\in\mathbf{N}} of groups. For (pointwise) stability of groups, it is shown in [3] that Γ\Gamma is (pointise) 𝒢\mathcal{G}-stable if any homomorphism ϕ~:Γ𝒰Gn/(𝒰Gn)inf\tilde{\phi}:\Gamma\to\prod_{\mathcal{U}}G_{n}/\left(\prod_{\mathcal{U}}G_{n}\right)_{inf} from Γ\Gamma to the metric ultraproduct, can be lifted to a homomorphism ψ:Γ𝒰Gn\psi:\Gamma\to\prod_{\mathcal{U}}G_{n}. In our setting, the uniformity requirement forces us to work with internal maps from the ultrapower Γ{}^{*}\Gamma as opposed to just maps from Γ\Gamma.

We shall further refine the internal lifting property parametrizing it by the precise defect. Let ϕ={ϕn}𝒰:Γ𝒰Gn\phi=\{\phi_{n}\}_{\mathcal{U}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} be an internal asymptotic homomorphism with defect def(ϕ)=ϵ𝐑infdef(\phi)=\epsilon\in{}^{*}\mathbf{R}_{inf}. Consider the subset B(ϵ)B(\epsilon) (elements bounded by ϵ\epsilon) of 𝒰Gn\prod_{\mathcal{U}}G_{n} defined as follows:

B(ϵ){{gn}𝒰𝒰Gn:{dn(gn,1n)}𝒰=O𝒰(ϵ)}B(\epsilon)\coloneqq\{\{g_{n}\}_{\mathcal{U}}\in\prod_{\mathcal{U}}G_{n}:\{d_{n}(g_{n},1_{n})\}_{\mathcal{U}}=O_{\mathcal{U}}(\epsilon)\}

Note that B(ϵ)B(\epsilon) is an externally defined subset. Since the metric on GnG_{n} is bi-invariant, the subset B(ϵ)B(\epsilon) is a normal subgroup of 𝒰Gn\prod_{\mathcal{U}}G_{n}. Let qB(ϵ):𝒰Gn𝒰Gn/B(ϵ)q_{B(\epsilon)}:\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) be the canonical quotient homomorphism, and denote by ϕ~\tilde{\phi} the composition map qB(ϵ)ϕ:Γ(𝒰Gn)/B(ϵ)q_{B(\epsilon)}\cdot\phi:{}^{*}\Gamma\to\left(\prod_{\mathcal{U}}G_{n}\right)/B(\epsilon). The following lemma is a parametrized reformulation of Lemma 2.3.3, and the proof is on the same lines, which we omit here:

Lemma 2.3.6.

The map ϕ~:Γ(𝒰Gn)/B(ϵ)\tilde{\phi}:{}^{*}\Gamma\to\left(\prod_{\mathcal{U}}G_{n}\right)/B(\epsilon) is a group homomorphism.

Conversely, let F:Γ𝒰Gn/B(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon). An internal map F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is said to be an internal lift of ϕ~\tilde{\phi} if qB(ϵ)F^=Fq_{B(\epsilon)}\cdot\hat{F}=F. We again have an analogue of Lemma 2.3.3 here too, whose proof is similar:

Lemma 2.3.7.

Suppose a homomorphism F:Γ𝒰Gn/B(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}. Then F^\hat{F} is an internal O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-homomorphism. Futhermore, if F^(1)\hat{F}^{(1)} and F^(2)\hat{F}^{(2)} are two such internal lifts of FF, then F^(1)\hat{F}^{(1)} and F^(2)\hat{F}^{(2)} are internally O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-close to each other.

Lemma 2.3.8.

The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate iff for every ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}, every homomorphism ϕ~:Γ𝒰Gn/B(ϵ)\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) that has an internal lift also has an internal lift homomorphism.

Proof.

Suppose ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is an internal asymptotic homomorphism with defect ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}. Then ϕ~::Γ𝒰Gn/B(ϵ)\tilde{\phi}::{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) is a homomorphism which has internal lift ϕ\phi. Thus, it also has an internal lift homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} which is internally O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-close to ϕ\phi. Hence, by Lemma 2.3.4, Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate. The converse is immediate. ∎

Thus, for an infinitesimal ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}, if a homomorphism ϕ~:Γ𝒰Gn/B(ϵ)\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) that has an internal lift, can be internally lifted to an internal homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, then Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate. We shall now try to obtain such a lift by a sequence of intermediate lifts.
For ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}, denote by I(ϵ)I(\epsilon) the subset of 𝒰Gn\prod_{\mathcal{U}}G_{n} (elements infinitesimal with respect to ϵ\epsilon) defined as

I(ϵ){{gn}𝒰𝒰Gn:{dn(gn,1n)}𝒰=o𝒰(ϵ)}I(\epsilon)\coloneq\{\{g_{n}\}_{\mathcal{U}}\in\prod_{\mathcal{U}}G_{n}:\{d_{n}(g_{n},1_{n})\}_{\mathcal{U}}=o_{\mathcal{U}}(\epsilon)\}

Note that by the bi-invariance of the metric, I(ϵ)B(ϵ)I(\epsilon)\subseteq B(\epsilon) is a normal subgroup of 𝒰Gn\prod_{\mathcal{U}}G_{n}. Let qI(ϵ):𝒰Gn𝒰Gn/I(ϵ)q_{I(\epsilon)}:\prod_{\mathcal{U}}G_{n}\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) be the canonical quotient homomorphism. The following lemma is similar to Lemma 2.3.3, and the proof too is on the same lines:

Lemma 2.3.9.

Suppose ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} is an internal o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-homomorphism, then qI(ϵ)ϕ:Γ𝒰Gn/I(ϵ)q_{I(\epsilon)}\cdot\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) is a homomorphism. Conversely, suppose a homomorphism F:Γ𝒰Gn/I(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}, then F^\hat{F} is an internal o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-homomorphism, and any two internal lifts of FF are internally o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-close to one another.

We can now reformulate the defect diminishing property in terms of internal lifts.

Lemma 2.3.10.

The group Γ\Gamma has the defect diminishing property with respect to 𝒢\mathcal{G} iff for every ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} and every homomorphism F:Γ𝒰Gn/B(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) that has an internal lift, FF has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} such that qI(ϵ)F^:Γ𝒰Gn/I(ϵ)q_{I(\epsilon)}\cdot\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) is a homomorphism.

Proof.

Suppose Γ\Gamma has the defect diminishing property, then it is immediate from Definition 2.2.5 that for every ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} and every homomorphism F:Γ𝒰Gn/B(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) that has an internal lift, FF has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} such that qI(ϵ)F:Γ𝒰Gn/I(ϵ)q_{I(\epsilon)}\cdot F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) is a homomorphism.
Conversely, consider an internal asymptotic homomorphism ϕ:Γ𝒰Gn\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} with defect def(ϕ)=ϵdef(\phi)=\epsilon, and the induced homomorphism ϕ~:Γ𝒰Gn/B(ϵ)\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon). By the hypothesis of the lemma (and Lemma 2.3.9), there exists an internal o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-homomorphism ψ:Γ𝒰Gn\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} which is O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-close to ϕ\phi. This shows that Γ\Gamma has the defect diminishing property with respect to 𝒢\mathcal{G}. ∎

We now recap the results obtained so far:

Theorem 2.3.11.

Let Γ\Gamma be a discrete group, and 𝒢\mathcal{G} be a family of metric groups such that for every group G𝒢G\in\mathcal{G}, its metric dGd_{G} is complete. Then the following are equivalent:

  1. 1.

    The group Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate.

  2. 2.

    The group Γ\Gamma has the defect diminishing property with respect to 𝒢\mathcal{G}.

  3. 3.

    For every ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} and every homomorphism F:Γ𝒰Gn/B(ϵ)F:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/B(\epsilon) that has an internal lift, FF has an internal lift F^:Γ𝒰Gn\hat{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} such that qI(ϵ)F~:Γ𝒰Gn/I(ϵ)q_{I(\epsilon)}\cdot\tilde{F}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}/I(\epsilon) is a homomorphism.

3 A Cohomological Interpretation of Stability

We concluded the previous section by noting that in order to prove that Γ\Gamma is uniformly 𝒢\mathcal{G}-stable with a linear estimate, it is sufficient to show that it has the defect diminishing property with respect to 𝒢\mathcal{G}. Furthermore, we interpreted this property in terms of internal lifts of homomorphisms to quotient groups.
Recall that a matrix norm \|\cdot\| on Mn(𝐂)M_{n}(\mathbf{C}) is said to be submultiplicative if for every A,BMn(𝐂)A,B\in M_{n}(\mathbf{C}), ABAB\|AB\|\leq\|A\|\cdot\|B\|. From now on, we shall work exclusively with the family of unitary groups, each equipped with a unitarily bi-invariant submultiplicative matrix norm, and shall denote this family by 𝔘\mathfrak{U}.

𝔘{(U(n),):n𝐍 and  is submultiplicative}\mathfrak{U}\coloneq\Big{\{}\left(U(n),\|\cdot\|\right):n\in\mathbf{N}\text{ and }\|\cdot\|\text{ is submultiplicative}\Big{\}}

In particular, these include the pp-Schatten norms given by

Ap={(Tr|A|p)1/p1p<supν=1Aνp=\|A\|_{p}=\begin{cases}(Tr|A|^{p})^{1/p}&1\leq p<\infty\\ \sup_{\|\nu\|=1}\|A\nu\|&p=\infty\end{cases}

Note that for p=p=\infty, this is the operator norm (as studied in [31] and [15]), while for p=2p=2, this is the Frobenius norm or the (unnormalized) Hilbert-Schmidt norm (as studied in [19] and [1]).

Before we proceed further, we state and sketch the proof of the following useful transference lemma for uniform 𝔘\mathfrak{U}-stability with a linear estimate, which we shall use in §6. The proof is on the lines of a relative version of [15, Theorem 3.2] (which reproves Kazhdan’s result on the Ulam stability of amenable groups), which we can adapt here in the simpler setting of finite index.

Lemma 3.0.1.

Let ΛΓ\Lambda\leq\Gamma be a subgroup of finite index. Then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate iff Λ\Lambda is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Sketch.

Suppose Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate. Then the proof of [15, Corollary 2.7] (further explained in [26, Lemma II.22]) implies that Λ\Lambda too is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.
Conversely, suppose Λ\Lambda is uniformly 𝔘\mathfrak{U}-stable with a linear estimate, and let ϕ:ΓU(n)\phi:\Gamma\to U(n) be an ϵ\epsilon-homomorphism. Since the finite index subgroup Λ\Lambda is uniformly 𝔘\mathfrak{U}-stable with a linear estimate, we can assume that the restriction of ϕ\phi to Λ\Lambda is a homomorphism, and furthermore, ϕ(gδ)=ϕ(g)ϕ(δ)\phi(g\delta)=\phi(g)\phi(\delta) for every gΓg\in\Gamma, δΛ\delta\in\Lambda. Now define ϕ:ΓM(n)\phi^{\prime}:\Gamma\to M(n) as

ϕ(g)1|Γ:Λ|xΓ/Λϕ(gx)ϕ(x)\phi^{\prime}(g)\coloneq\frac{1}{|\Gamma:\Lambda|}\sum_{x\in\Gamma/\Lambda}\phi(gx)\phi(x)^{*}

Note that ϕ(g)\phi^{\prime}(g) is just the average over coset representatives in Γ/Λ\Gamma/\Lambda, and M(n)M(n) is the space of n×nn\times n matrices. Just as in the proof of [15, Theorem 3.2], this ϕ\phi^{\prime} can be normalized to obtain ϕ1:ΓU(n)\phi_{1}:\Gamma\to U(n) such that ϕ1\phi_{1} has defect Cϵ2C\epsilon^{2} (for a universal constant CC not depending on nn), and we repeat the process to obtain a (true) homomorphism as the limit. ∎

Remark 3.0.2.

In fact, it is further shown in [23, Proposition 1.5] that for a subgroup ΛΓ\Lambda\leq\Gamma that is co-amenable in Γ\Gamma, if Λ\Lambda is uniformly 𝔘\mathfrak{U}-stable with a linear estimate, then Γ\Gamma too is uniformly 𝔘\mathfrak{U}-stable with a linear estimate. This generalizes one direction of Lemma 3.0.1, though the converse is not true in this level of generality.

Remark 3.0.3.

In particular, if Γ1\Gamma_{1} and Γ2\Gamma_{2} are commensurable, then Γ1\Gamma_{1} is uniformly 𝔘\mathfrak{U}-stable with a linear estimate iff Γ2\Gamma_{2} is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Given a sequence of unitary groups {U(kn)}n𝐍\{U(k_{n})\}_{n\in\mathbf{N}}, we denote its ultraproduct by 𝒰U(kn)\prod_{\mathcal{U}}U(k_{n}),and given an element u={un}𝒰𝒰U(kn)u=\{u_{n}\}_{\mathcal{U}}\in\prod_{\mathcal{U}}U(k_{n}) (where for each n𝐍n\in\mathbf{N}, unU(kn)u_{n}\in U(k_{n})), we denote its distance from the identity 1𝒰U(kn)1\in\prod_{\mathcal{U}}U(k_{n}) by u1{unI}𝒰𝐑b\|u-1\|\coloneq\{\|u_{n}-I\|\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{b}, for notational convenience.
Note that for ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} and an ultraproduct 𝒰U(kn)\prod_{\mathcal{U}}U(k_{n}), the subsets B(ϵ)𝒰GnB(\epsilon)\subseteq\prod_{\mathcal{U}}G_{n} and I(ϵ)B(ϵ)I(\epsilon)\subseteq B(\epsilon) can now be written as

B(ϵ)={u𝒰U(kn):uI=O𝒰(ϵ)}B(\epsilon)=\Big{\{}u\in\prod_{\mathcal{U}}U(k_{n}):\|u-I\|=O_{\mathcal{U}}(\epsilon)\Big{\}}
I(ϵ)={u𝒰U(kn):uI=o𝒰(ϵ)}I(\epsilon)=\Big{\{}u\in\prod_{\mathcal{U}}U(k_{n}):\|u-I\|=o_{\mathcal{U}}(\epsilon)\Big{\}}

Furthermore, the submultiplicativity of the norms implies that:

Lemma 3.0.4.

The group B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) is abelian.

Proof.

Let a,bB(ϵ)a,b\in B(\epsilon), and consider the commutator abababa^{*}b^{*}. We shall prove that ababI(ϵ)aba^{*}b^{*}\in I(\epsilon). Observe that ababI=abba\|aba^{*}b^{*}-I\|=\|ab-ba\| since the norm is unitarily invariant.

abba=(aI)(bI)(bI)(aI)2aIbI\|ab-ba\|=\|(a-I)(b-I)-(b-I)(a-I)\|\leq 2\|a-I\|\|b-I\|

This is because of the submultiplicativity of the norm. Since a,bB(ϵ)a,b\in B(\epsilon), we conclude that ababI=O𝒰(ϵ2)=o𝒰(ϵ)\|aba^{*}b^{*}-I\|=O_{\mathcal{U}}(\epsilon^{2})=o_{\mathcal{U}}(\epsilon). ∎

The fact that B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) is an abelian group will allow us to rephrase the lifting property discussed in §2.3 in terms of the vanishing of a cohomology which we shall develop in detail in §3.2 and §4.
In §3.1, we explicitly work out the \saycocycles corresponding to possible lifts of a homomorphism, which are then transferred to the linearized setting in §3.2 where a cohomological theory begins to reveal itself. Finally, in §3.3, we demonstrate the notions discussed in the case of discrete abelian groups, showing that the vanishing of our second cohomology implies uniform stability.

3.1 Lifting with an Abelian Kernel

Let ϕ:Γ𝒰U(kn)\phi:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) be an internal ϵ\epsilon-homomorphism that induces a homomorphism ϕ~:Γ𝒰U(kn)/B(ϵ)\tilde{\phi}:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n})/B(\epsilon). Let ψ:Γ𝒰U(kn)\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) be an internal lift of ϕ~\tilde{\phi}.
To an internal lift ψ:Γ𝒰U(kn)\psi:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) of ϕ~\tilde{\phi}, we associate an internal map

ρψ:Γ×𝒰U(kn)𝒰U(kn)\rho_{\psi}:{}^{*}\Gamma\times\prod_{\mathcal{U}}U(k_{n})\to\prod_{\mathcal{U}}U(k_{n})
ρψ(g)(u)ψ(g)uψ(g)1\rho_{\psi}(g)(u)\coloneq\psi(g)\cdot u\cdot\psi(g)^{-1} (3.1)

For every gΓg\in{}^{*}\Gamma and uB(ϵ)u\in B(\epsilon), ρψ(g)(u)B(ϵ)\rho_{\psi}(g)(u)\in B(\epsilon), while for uI(ϵ)u\in I(\epsilon), ρψ(g)(u)I(ϵ)\rho_{\psi}(g)(u)\in I(\epsilon).

Lemma 3.1.1.

The internal map ρψ:Γ×𝒰U(kn)𝒰U(kn)\rho_{\psi}:{}^{*}\Gamma\times\prod_{\mathcal{U}}U(k_{n})\to\prod_{\mathcal{U}}U(k_{n}) induces an action, denoted ρ~ψ\tilde{\rho}_{\psi}, of Γ{}^{*}\Gamma on the abelian group B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon).

Proof.

For g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma and uB(ϵ)u\in B(\epsilon), we want to show that ρψ(g1)(ρψ(g2)(u))ρψ(g1g2)(u)I(ϵ)\rho_{\psi}(g_{1})(\rho_{\psi}(g_{2})(u))-\rho_{\psi}(g_{1}g_{2})(u)\in I(\epsilon). Note that for every g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma, ψ(g1g2)1ψ(g1)ψ(g2)B(ϵ)\psi(g_{1}g_{2})^{-1}\psi(g_{1})\psi(g_{2})\in B(\epsilon), so

ψ(g1)ψ(g2)uψ(g2)1ψ(g1)1ψ(g1g2)uψ(g1g2)1=ψ(g1g2)1ψ(g1)ψ(g2)uuψ(g1g2)1ψ(g1)ψ(g2)\|\psi(g_{1})\psi(g_{2})u\psi(g_{2})^{-1}\psi(g_{1})^{-1}-\psi(g_{1}g_{2})u\psi(g_{1}g_{2})^{-1}\|=\|\psi(g_{1}g_{2})^{-1}\psi(g_{1})\psi(g_{2})u-u\psi(g_{1}g_{2})^{-1}\psi(g_{1})\psi(g_{2})\|

Since the elements ψ(g1g2)1ψ(g1)ψ(g2)\psi(g_{1}g_{2})^{-1}\psi(g_{1})\psi(g_{2}) and uu are in B(ϵ)B(\epsilon), their commutator is in I(ϵ)I(\epsilon) (as in proof of Lemma 3.0.4). ∎

We shall call ρ~ψ\tilde{\rho}_{\psi} the action induced from ψ\psi. Observe that we have defined ρψ\rho_{\psi} using a given internal lift ψ\psi. However, this induced action of Γ{}^{*}\Gamma on B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) is independent of the choice of internal lift of ϕ~\tilde{\phi}.

Lemma 3.1.2.

For two internal lifts ψ1\psi_{1} and ψ2\psi_{2} of ψ~\tilde{\psi}, ρ~ψ1=ρ~ψ2\tilde{\rho}_{\psi_{1}}=\tilde{\rho}_{\psi_{2}}.

Proof.

Note that for gΓg\in{}^{*}\Gamma, ψ2(g)1ψ1(g)B(ϵ)\psi_{2}(g)^{-1}\psi_{1}(g)\in B(\epsilon) since they are both internal lifts of ϕ~\tilde{\phi}. Hence for uB(ϵ)u\in B(\epsilon), again as in the proof of Lemma 3.0.4, ψ2(g)1ψ1(g)uuψ2(g)1ψ1(g)I(ϵ)\psi_{2}(g)^{-1}\psi_{1}(g)u-u\psi_{2}(g)^{-1}\psi_{1}(g)\in I(\epsilon), which implies that ψ1(g)uψ1(g)1ψ2(g)uψ2(g)1I(ϵ)\psi_{1}(g)u\psi_{1}(g)^{-1}-\psi_{2}(g)u\psi_{2}(g)^{-1}\in I(\epsilon). ∎

So while the internal map ρψ:Γ×𝒰U(kn)𝒰U(kn)\rho_{\psi}:{}^{*}\Gamma\times\prod_{\mathcal{U}}U(k_{n})\to\prod_{\mathcal{U}}U(k_{n}) depends on ψ\psi, the induced action ρ~ψ\tilde{\rho}_{\psi} is independent of the choice of internal lift of ϕ~\tilde{\phi}. Hence we can denote this action by ρ~ϕ\tilde{\rho}_{\phi}.
Define the internal map

αψ:Γ×Γ𝒰U(kn)\alpha_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n})
αψ(g1,g2)ψ(g1)ψ(g2)ψ(g1g2)1\alpha_{\psi}(g_{1},g_{2})\coloneq\psi(g_{1})\psi(g_{2})\psi(g_{1}g_{2})^{-1}

Note that by construction αψ\alpha_{\psi} takes values in B(ϵ)B(\epsilon) (and our goal is to find some lift ψ\psi for which αψ\alpha_{\psi} takes values only in I(ϵ)I(\epsilon)). Observe that

Lemma 3.1.3.

For g1,g2,g3Γg_{1},g_{2},g_{3}\in{}^{*}\Gamma,

αψ(g1,g2)αψ(g1g2,g3)αψ(g1,g2g3)1=ψ(g1)αψ(g2,g3)ψ(g1)1\alpha_{\psi}(g_{1},g_{2})\cdot\alpha_{\psi}(g_{1}g_{2},g_{3})\cdot\alpha_{\psi}(g_{1},g_{2}g_{3})^{-1}=\psi(g_{1})\cdot\alpha_{\psi}(g_{2},g_{3})\psi(g_{1})^{-1}

Let qB/I:B(ϵ)B(ϵ)/I(ϵ)q_{B/I}:B(\epsilon)\to B(\epsilon)/I(\epsilon) be the canonical quotient homomorphism. Since αψ\alpha_{\psi} takes values in B(ϵ)B(\epsilon), let α~ψqB/Iαψ:Γ×ΓB(ϵ)/I(ϵ)\tilde{\alpha}_{\psi}\coloneq q_{B/I}\cdot\alpha_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to B(\epsilon)/I(\epsilon). Since B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) is abelian, Lemma 3.1.3 implies the following corollary:

Corollary 3.1.4.

For g1,g2,g3Γg_{1},g_{2},g_{3}\in{}^{*}\Gamma,

ρ~ϕ(g1)α~ψ(g2,g3)α~ψ(g1g2,g3)+α~ψ(g1,g2g3)α~ψ(g1,g2)=0\tilde{\rho}_{\phi}(g_{1})\cdot\tilde{\alpha}_{\psi}(g_{2},g_{3})-\tilde{\alpha}_{\psi}(g_{1}g_{2},g_{3})+\tilde{\alpha}_{\psi}(g_{1},g_{2}g_{3})-\tilde{\alpha}_{\psi}(g_{1},g_{2})=0

While we noted that the action ρ~ψ\tilde{\rho}_{\psi} of Γ{}^{*}\Gamma on B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) does not depend on the choice of lift ψ\psi, the map α~ψ:Γ×ΓB(ϵ)/I(ϵ)\tilde{\alpha}_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to B(\epsilon)/I(\epsilon) does depend on the choice of lift ψ\psi. Our hope is to prove the existence of some choice of lift ψ\psi such that α~ψ\tilde{\alpha}_{\psi} is trivial.
Consider α~ψ1\tilde{\alpha}_{\psi_{1}} and α~ψ2\tilde{\alpha}_{\psi_{2}} for two different internal lifts ψ1\psi_{1} and ψ2\psi_{2} of ϕ~\tilde{\phi}. Define an internal map

βψ1,ψ2:Γ𝒰Gn\beta_{\psi_{1},\psi_{2}}:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n}
βψ1,ψ2(g)ψ2(g)ψ1(g)1\beta_{\psi_{1},\psi_{2}}(g)\coloneq\psi_{2}(g)\psi_{1}(g)^{-1}

Since βψ1,ψ2\beta_{\psi_{1},\psi_{2}} takes values in B(ϵ)B(\epsilon), denote by β~ψ1,ψ2:ΓB(ϵ)/I(ϵ)\tilde{\beta}_{\psi_{1},\psi_{2}}:{}^{*}\Gamma\to B(\epsilon)/I(\epsilon) its composition with the quotient map B(ϵ)B(ϵ)/I(ϵ)B(\epsilon)\to B(\epsilon)/I(\epsilon). Then for g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma, a careful computation shows that

α~ψ2α~ψ1=β~ψ1,ψ2(g1)+ρ~ϕ(g1)β~ψ1,ψ2(g2)β~ψ1,ψ2(g1g2)\tilde{\alpha}_{\psi_{2}}-\tilde{\alpha}_{\psi_{1}}=\tilde{\beta}_{\psi_{1},\psi_{2}}(g_{1})+\tilde{\rho}_{\phi}(g_{1})\cdot\tilde{\beta}_{\psi_{1},\psi_{2}}(g_{2})-\tilde{\beta}_{\psi_{1},\psi_{2}}(g_{1}g_{2}) (3.2)

Suppose there exists an internal map β:Γ𝒰Gn\beta:{}^{*}\Gamma\to\prod_{\mathcal{U}}G_{n} that takes values in B(ϵ)B(\epsilon) such that for the lift ψ\psi, the following equation holds for all g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma:

α~ψ(g1,g2)=β~(g1)+ρ~ϕ(g1)β~(g2)β~(g1g2)\tilde{\alpha}_{\psi}(g_{1},g_{2})=\tilde{\beta}(g_{1})+\tilde{\rho}_{\phi}(g_{1})\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2})

Then the internal map

ψ:Γ𝒰U(kn)\psi^{\sim}:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n})
ψ(g)ψ(g)β(g)1\psi^{\sim}(g)\coloneq\psi(g)\beta(g)^{-1}

is also an internal lift of ϕ~\tilde{\phi} such that for every g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma,

α~ψ(g1,g2)=0\tilde{\alpha}_{\psi^{\sim}}(g_{1},g_{2})=0

In particular, this means that ψ\psi^{\sim} is the internal lift that we want.
The above discussion hints at a cohomological theory that captures the obstruction to such lifts. The idea is as follows: any candidate internal lift ψ\psi of ϕ~\tilde{\phi}, with defect O𝒰(ϵ)O_{\mathcal{U}}(\epsilon), gives us a type of 22-cocycle of Γ{}^{*}\Gamma with coefficients in B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon), and if that cocycle happens to be a 11-couboundary, then the lift ψ\psi can be corrected to obtain another lift that has defect o𝒰(ϵ)o_{\mathcal{U}}(\epsilon) and is still O𝒰(ϵ)O_{\mathcal{U}}(\epsilon)-close to ψ\psi (and ϕ\phi), thus implying the defect diminishing property that we want.

Remark 3.1.5.

In [19], it is shown (using the idea mentioned in Remark 2.3.5 and defect diminishing) that Γ\Gamma is (pointwise) stable (with respect to unitary matrices equipped with the Frobenius norm) if H2(Γ,B(ϵ)/I(ϵ))H^{2}\left(\Gamma,B(\epsilon)/I(\epsilon)\right) vanishes. From Lemma 3.1.3, it might be tempting to simply consider α~ψ\tilde{\alpha}_{\psi} as a bounded 22-cocycle for the group Γ{}^{*}\Gamma with coefficients in the abelian group B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon), and interpret Eq. 3.2 (and the ensuing discussion) as insisting that α~ψ\tilde{\alpha}_{\psi} is the coboundary of a bounded 11-cochain of Γ{}^{*}\Gamma in B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon). But we cannot simply work with Hb2(Γ,B(ϵ)/I(ϵ))\operatorname{H}_{b}^{2}\left({}^{*}\Gamma,B(\epsilon)/I(\epsilon)\right), since we need to ensure that the bounded 22-cocycle α~ψ\tilde{\alpha}_{\psi} (which was induced from an internal map αψ\alpha_{\psi}) is the coboundary of a bounded 11-cochain β~:ΓB(ϵ)/I(ϵ)\tilde{\beta}:{}^{*}\Gamma\to B(\epsilon)/I(\epsilon) that is itself also induced from an internal map β:Γ𝒰U(kn)\beta:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}). This insistence on our maps being induced from internal maps is essential in our setting of uniform stability, and leads to the definition of an internal and asymptotic bounded cohomology machinery that we construct in §4.

3.2 Linearization and the Lie Algebra

In the previous section we observed that the defect diminishing property could be interpreted as a cohomological problem based on an action of Γ{}^{*}\Gamma on the abelian group B(ϵ)/IϵB(\epsilon)/I_{\epsilon}. However, as pointed out in Remark 3.1.5 the subtlety here involves the requirement that we deal only with maps that are induced from some internal mapping to 𝒰U(kn)\prod_{\mathcal{U}}U(k_{n}). At that level we do not have the abelianness that would allow us to properly formulate a cohomology theory. In this §, we transfer to the Lie algebra allowing us to work with spaces of maps from the group to Banach spaces.
For a matrix AMn(𝐂)A\in M_{n}(\mathbf{C}), consider the matrix logarithm given by

logAj=1(1)j1(AI)jj\log A\coloneq\sum\limits_{j=1}^{\infty}(-1)^{j-1}\frac{(A-I)^{j}}{j}

The series above converges if AI<1\|A-I\|<1 (for some submultiplicative matrix norm \|\cdot\|). By subadditivity and submultiplicativity of the norm \|\cdot\|, if ϵ1/2\epsilon\leq 1/2 and uIϵ\|u-I\|\leq\epsilon,

loguj=1uIj2ϵ\|\log{u}\|\leq\sum\limits_{j=1}^{\infty}\|u-I\|^{j}\leq 2\epsilon
Lemma 3.2.1.

For every ϵ<1/2\epsilon<1/2, n𝐍n\in\mathbf{N}, uU(n)u\in U(n) and every submultiplicative norm \|\cdot\| on Mn(𝐂)M_{n}(\mathbf{C}), if uIϵ\|u-I\|\leq\epsilon, then logu2ϵ\|\log{u}\|\leq 2\epsilon.

It is a classical result that for a unitary matrix uU(n)u\in U(n), its logarithm logu\log{u} (whenever it is defined) is an anti-Hermitian matrix. Denote by 𝔲(n)\mathfrak{u}(n) the (real) vector space of anti-Hermitian matrices in Mn(𝐂)M_{n}(\mathbf{C}).
In the other direction, we have the matrix exponential map defined as

exp(A)=j=0Ajj!\exp(A)=\sum\limits_{j=0}^{\infty}\frac{A^{j}}{j!}

which is well-defined for every AMn(𝐂)A\in M_{n}(\mathbf{C}). For an anti-hermitian matrix W𝔲(n)W\in\mathfrak{u}(n) of the form W=Udiag(iθj)j=1nUW=U\circ diag(i\theta_{j})_{j=1}^{n}\circ U^{*} for UU(n)U\in U(n), its exponential exp(W)=Udiag(eiθj)j=1nUexp(W)=U\circ diag(e^{i\theta_{j}})_{j=1}^{n}\circ U^{*}. The following trivial bound is sufficient for our purposes:

Lemma 3.2.2.

For every ϵ>0\epsilon>0, n𝐍n\in\mathbf{N} and a submultiplicative norm \|\cdot\| on Mn(𝐂)M_{n}(\mathbf{C}), for a matrix W𝔲(n)W\in\mathfrak{u}(n) with W<ϵ\|W\|<\epsilon, exp(W)Ieϵ1\|exp(W)-I\|\leq e^{\epsilon}-1.

Note that since 𝔲(n)\mathfrak{u}(n) is finite-dimensional, it is complete for any norm, making it a finite-dimensional real Banach space. It comes with an isometric (adjoint) action of U(n)U(n) given as follows: for v𝔲(n)v\in\mathfrak{u}(n) and UU(n)U\in U(n), Ad(U)(v)UvU𝔲(n)Ad(U)(v)\coloneq UvU^{*}\in\mathfrak{u}(n).
Consider the family of 𝐑\mathbf{R}-vector spaces of anti-hermitian matrices 𝔲(n)\mathfrak{u}(n) each equipped with a submultiplicative norm.

{(𝔲(n),):n𝐍 and  is submultiplicative}\Big{\{}\left(\mathfrak{u}(n),\|\cdot\|\right):n\in\mathbf{N}\text{ and }\|\cdot\|\text{ is submultiplicative}\Big{\}}

For an infinitesimal ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}, define the internal map

logϵ:𝒰U(kn)𝒰𝔲(kn){}_{\epsilon}\log:\prod_{\mathcal{U}}U(k_{n})\to\prod_{\mathcal{U}}\mathfrak{u}(k_{n})
logϵu1ϵ{logukn}𝒰{}_{\epsilon}\log{u}\coloneq\frac{1}{\epsilon}\{\log{u_{k_{n}}}\}_{\mathcal{U}}

and similarly, the internal map exp:𝒰𝔲(kn)𝒰U(kn)\exp:\prod_{\mathcal{U}}\mathfrak{u}(k_{n})\to\prod_{\mathcal{U}}U(k_{n}) given by

expϵu{expϵknukn}𝒰{}_{\epsilon}\exp{u}\coloneq\{\exp{\epsilon_{k_{n}}u_{k_{n}}}\}_{\mathcal{U}}

Let us denote the ultraproduct 𝒰𝔲(kn)\prod_{\mathcal{U}}\mathfrak{u}(k_{n}) by 𝒲\mathcal{W} from now on. Note that 𝒲\mathcal{W} comes with a 𝐑{}^{*}\mathbf{R}-valued norm, which we shall denote simply as \|\cdot\|, obtained as the ultraproduct of the respective norms of each 𝔲(kn)\mathfrak{u}(k_{n}). The bounded elements of 𝒲\mathcal{W} shall be denoted 𝒲b\mathcal{W}_{b} while the infinitesimal elements of 𝒲\mathcal{W} are denoted by 𝒲inf\mathcal{W}_{inf}. That is,

𝒲b{w𝒲:w𝐑b}\mathcal{W}_{b}\coloneq\Big{\{}w\in\mathcal{W}:\|w\|\in{}^{*}\mathbf{R}_{b}\Big{\}} (3.3)
𝒲inf{w𝒲:w𝐑inf}\mathcal{W}_{inf}\coloneq\Big{\{}w\in\mathcal{W}:\|w\|\in{}^{*}\mathbf{R}_{inf}\Big{\}} (3.4)

The motivation behind scaling the definitions of log\log and exp\exp by 1/ϵ1/\epsilon and ϵ\epsilon respectively is as follows:

Proposition 3.2.3.

The internal map logϵ:𝒰U(kn)𝒲{}_{\epsilon}\log:\prod_{\mathcal{U}}U(k_{n})\to\mathcal{W} when restricted to B(ϵ)B(\epsilon), takes values in 𝒲b\mathcal{W}_{b}, and elements in I(ϵ)I(\epsilon) are taken to 𝒲inf\mathcal{W}_{inf}. This induces an isomorphism of the abelian groups 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf} and B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon).

Proof.

It follows from Lemma 3.2.1 that for uB(ϵ)u\in B(\epsilon), logϵu𝒲b{}_{\epsilon}\log{u}\in\mathcal{W}_{b}, and for uI(ϵ)u\in I(\epsilon), logϵu𝒲inf{}_{\epsilon}\log{u}\in\mathcal{W}_{inf}. The map is surjective on 𝒲b\mathcal{W}_{b} since the map logϵ(ϵexpv)=v{}_{\epsilon}\log{(_{\epsilon}\exp{v})}=v for v𝒲bv\in\mathcal{W}_{b} (and similarly for 𝒲inf\mathcal{W}_{inf} as well).
From properties of the logarithm map, it follows that for u1,u2B(ϵ)u_{1},u_{2}\in B(\epsilon),logϵu1u2(ϵlogu1+ϵlogu2)𝒲inf{}_{\epsilon}\log{u_{1}u_{2}}-(_{\epsilon}\log{u_{1}}+_{\epsilon}\log{u_{2}})\in\mathcal{W}_{inf}. Hence logϵ{}_{\epsilon}\log induces a surjective group homomorphism from B(ϵ)B(\epsilon) to 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf} with kernel I(ϵ)I(\epsilon). ∎

The ultralimit 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf} shall be denoted 𝒲~\tilde{\mathcal{W}}. The above lemma tells us that 𝒲~B(ϵ)/I(ϵ)\tilde{\mathcal{W}}\cong B(\epsilon)/I(\epsilon). In fact, 𝒲~B(ϵ)/I(ϵ)\tilde{\mathcal{W}}\cong B(\epsilon)/I(\epsilon) is not just an abelian group but has the structure of a real Banach space. It is an example of a construction known as a Banach space ultralimit. We shall not prove this result here, but refer to [29] and [2] for more details:

Proposition 3.2.4 ([29]).

The space W~=B(ϵ)/I(ϵ)\tilde{W}=B(\epsilon)/I(\epsilon) is a real Banach space.

Recall that we had defined (Eq. 3.1) an internal map ρψ:Γ×𝒰U(kn)𝒰U(kn)\rho_{\psi}:{}^{*}\Gamma\times\prod_{\mathcal{U}}U(k_{n})\to\prod_{\mathcal{U}}U(k_{n}) defined as ρψ(g)v=ψ(g)vψ(g)1\rho_{\psi}(g)v=\psi(g)v\psi(g)^{-1} which had induced an action ρ~ϕ\tilde{\rho}_{\phi} of Γ{}^{*}\Gamma on B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon). We can similarly define an internal map

πψ:Γ×𝒲𝒲\pi_{\psi}:{}^{*}\Gamma\times\mathcal{W}\to\mathcal{W}

through the internal adjoint action of 𝒰U(kn)\prod_{\mathcal{U}}U(k_{n}) on 𝒲\mathcal{W} (that is, conjugation),

πψvψ(g)vψ(g)1\pi_{\psi}v\coloneqq\psi(g)v\psi(g)^{-1} (3.5)

Again, by the submultiplicativity of the norms, for v𝒲bv\in\mathcal{W}_{b} and g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma,

πψ(g1g2)vπψ(g1)πψ(g2)v𝒲inf\pi_{\psi}(g_{1}g_{2})v-\pi_{\psi}(g_{1})\pi_{\psi}(g_{2})v\in\mathcal{W}_{inf}

Thus, the internal map πψ\pi_{\psi} as defined above induces an action of Γ{}^{*}\Gamma on 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf}. Unless there is ambiguity, we shall denote the induced action of gΓg\in{}^{*}\Gamma on v~𝒲b/𝒲inf\tilde{v}\in\mathcal{W}_{b}/\mathcal{W}_{inf} through πψ\pi_{\psi} by gv~g\cdot\tilde{v}.

Lemma 3.2.5.

The internal map logϵ:𝒰U(kn)𝒲{}_{\epsilon}\log:\prod_{\mathcal{U}}U(k_{n})\to\mathcal{W} induces a Γ{}^{*}\Gamma-equivariant (additive) group isomorphism between B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) (with the action induced from ρψ\rho_{\psi}) and 𝒲~\tilde{\mathcal{W}} (with the action induced from πψ\pi_{\psi}).

Proof.

We already saw that logϵ{}_{\epsilon}\log induces a group isomorphism between B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) and 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf}. Let gΓg\in{}^{*}\Gamma and uB(ϵ)u\in B(\epsilon). Then ρψ(g)u=ψ(g)uψ(g)1\rho_{\psi}(g)u=\psi(g)u\psi(g)^{-1}, which means that logϵ(ρψ(g)u)=ψ(g)ϵloguψ(g)1{}_{\epsilon}\log(\rho_{\psi}(g)u)=\psi(g)_{\epsilon}\log{u}\psi(g)^{-1} since the matrix logarithm is invariant with respect to conjugation by a unitary matrix. ∎

In particular, 𝒲~\tilde{\mathcal{W}} is a real Banach space with an isometric action of Γ{}^{*}\Gamma (it is a real Banach Γ{}^{*}\Gamma-module).

Remark 3.2.6.

The fact that 𝒲=B(ϵ)/I(ϵ)\mathcal{W}=B(\epsilon)/I(\epsilon) is a real Banach Γ{}^{*}\Gamma-module is useful in reducing (pointwise) stability of Γ\Gamma with respect to the family 𝔘\mathfrak{U} to showing that H2(Γ,𝒲)=0H^{2}(\Gamma,\mathcal{W})=0, and this line of study is pursued in [19] and [33]. However, in our setting of uniform stability, the Banach structure of 𝒲\mathcal{W} is not as directly relevant.

Corresponding to the internal map αψ:Γ×Γ𝒰U(kn)\alpha_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}), define the internal map

α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W}
α(g1,g2)ϵlogαψ(g1,g2)\alpha(g_{1},g_{2})\coloneqq_{\epsilon}\log{\alpha_{\psi}(g_{1},g_{2})} (3.6)

Since αψ:Γ×Γ𝒰U(kn)\alpha_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) takes values only in B(ϵ)B(\epsilon), it is clear that α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} takes values only in 𝒲b\mathcal{W}_{b}. We shall denote by α~:Γ×Γ𝒲~\tilde{\alpha}:{}^{*}\Gamma\times{}^{*}\Gamma\to\tilde{\mathcal{W}} the map induced obtained by composing α\alpha with the canonical quotient map 𝒲b𝒲~\mathcal{W}_{b}\to\tilde{\mathcal{W}}.

Lemma 3.2.7.

The map α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} satisfies the following condition: for any g1,g2,g3Γg_{1},g_{2},g_{3}\in{}^{*}\Gamma,

πψ(g1)α(g2,g3)α(g1g2,g3)+α(g1,g2g3)α(g1,g2)𝒲inf\pi_{\psi}(g_{1})\alpha(g_{2},g_{3})-\alpha(g_{1}g_{2},g_{3})+\alpha(g_{1},g_{2}g_{3})-\alpha(g_{1},g_{2})\in\mathcal{W}_{inf}
Proof.

Recall that αψ\alpha_{\psi} satisfies the following property: for g1,g2,g3Γg_{1},g_{2},g_{3}\in{}^{*}\Gamma,

αψ(g1,g2)αψ(g1g2,g3)αψ(g1,g2g3)1=ρψ(g1)αψ(g2,g3)\alpha_{\psi}(g_{1},g_{2})\alpha_{\psi}(g_{1}g_{2},g_{3})\alpha_{\psi}(g_{1},g_{2}g_{3})^{-1}=\rho_{\psi}(g_{1})\alpha_{\psi}(g_{2},g_{3})

The conclusion then follows from the fact that the map logϵ:𝒰U(kn)𝒲{}_{\epsilon}\log:\prod_{\mathcal{U}}U(k_{n})\to\mathcal{W} induces a Γ{}^{*}\Gamma-equivariant group homomorphism between B(ϵ)/I(ϵ)B(\epsilon)/I(\epsilon) and 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf}. ∎

Thus, the induced map α~:Γ×Γ𝒲\tilde{\alpha}:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} satisifes the 22-cocycle condition given by: for g1,g2,g3Γg_{1},g_{2},g_{3}\in{}^{*}\Gamma,

g1α~(g2,g3)α~(g1g2,g3)+α~(g1,g2g3)α~(g1,g2)=0g_{1}\cdot\tilde{\alpha}(g_{2},g_{3})-\tilde{\alpha}(g_{1}g_{2},g_{3})+\tilde{\alpha}(g_{1},g_{2}g_{3})-\tilde{\alpha}(g_{1},g_{2})=0 (3.7)

So transfering to the internal Lie algebra through the logϵ{}_{\epsilon}\log map, we thus have a map α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} which takes values in 𝒲b\mathcal{W}_{b} and satisfies the 2-cocycle condition modulo 𝒲inf\mathcal{W}_{inf}. Recall that the defect diminishing condition was implied by the following statement: suppose αψ:Γ×Γ𝒰U(kn)\alpha_{\psi}:{}^{*}\Gamma\times{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) is an internal function taking values in B(ϵ)B(\epsilon) and such that α~ϕ\tilde{\alpha}_{\phi} satisfies the 22-cocycle condition. Then there exists an internal β:Γ𝒰U(kn)\beta:{}^{*}\Gamma\to\prod_{\mathcal{U}}U(k_{n}) taking values in B(ϵ)B(\epsilon) such that α~ϕ(g1,g2)=β~(g1)+g1β~(g2)β~(g1g2)\tilde{\alpha}_{\phi}(g_{1},g_{2})=\tilde{\beta}(g_{1})+g_{1}\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2}) Using the logϵ{}_{\epsilon}\log map to transfer to 𝒲\mathcal{W}, we conclude with the following proposition summarizing our work so far:

Proposition 3.2.8.

Suppose for every internal α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} with Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b} that satisfies the 22-cocycle condition (Eq. 3.7), there exists an internal β:Γ𝒲\beta:{}^{*}\Gamma\to\mathcal{W} taking values in 𝒲b\mathcal{W}_{b} such that

α~(g1,g2)=β~(g1)+g1β~(g2)β~(g1g2)\tilde{\alpha}(g_{1},g_{2})=\tilde{\beta}(g_{1})+g_{1}\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2})

then Γ\Gamma exhibits the defect diminishing property, and is therefore uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Since we shall work with internal maps from (Γ)2({}^{*}\Gamma)^{2} (or Γ{}^{*}\Gamma) to 𝒲\mathcal{W} that take values in 𝒲b\mathcal{W}_{b}, it is helpful to describe such a map as an ultraproduct of bounded maps. Let {αn(Γ2,𝔲(kn))}k=1\{\alpha_{n}\in\ell^{\infty}(\Gamma^{2},\mathfrak{u}(k_{n}))\}_{k=1}^{\infty} be a family of maps such that there exists a constant C>0C>0 such that αnC\|\alpha_{n}\|_{\infty}\leq C for every n𝐍n\in\mathbf{N}. Then it is clear that the ultraproduct α={αn}𝒰\alpha=\{\alpha_{n}\}_{\mathcal{U}} has image Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b}. Conversely,

Lemma 3.2.9.

Let α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} be an internal map with Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b}. Then there exists a family {αn(Γ2,𝔲(kn))}k=1\{\alpha_{n}\in\ell^{\infty}(\Gamma^{2},\mathfrak{u}(k_{n}))\}_{k=1}^{\infty} such that α={αn}𝒰\alpha=\{\alpha_{n}\}_{\mathcal{U}}. Conversely, if {αn(Γ2,𝔲(kn))}k=1\{\alpha_{n}\in\ell^{\infty}(\Gamma^{2},\mathfrak{u}(k_{n}))\}_{k=1}^{\infty} is a family of maps such that {αn}𝒰𝐑b\{\|\alpha_{n}\|_{\infty}\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{b}, then α={αn}𝒰Γ×Γ𝒲\alpha=\{\alpha_{n}\}_{\mathcal{U}}{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} is an internal map with Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b}.

Proof.

Since α\alpha is internal, it is of the form α={fn}𝒰\alpha=\{f_{n}\}_{\mathcal{U}} for a family of maps fn:Γ×Γ𝔲(kn)f_{n}:\Gamma\times\Gamma\to\mathfrak{u}(k_{n}). For a subset S𝒰S\in\mathcal{U}, suppose fnf_{n} is unbounded for every nSn\in S. Then for each nSn\in S, there exists xn,ynΓx_{n},y_{n}\in\Gamma such that fn(xn,yn)f_{n}(x_{n},y_{n}) has norm at least nn. In particular, for x={xn}𝒰x=\{x_{n}\}_{\mathcal{U}} and y={yn}𝒰y=\{y_{n}\}_{\mathcal{U}}, we have α(x,y)𝒲b\alpha(x,y)\notin\mathcal{W}_{b}. This is a contradiction to the hypothesis that Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b}. The converse is immediate from the definition of 𝒲b\mathcal{W}_{b}. ∎

Thus, an internal map α:Γ×Γ𝒲\alpha:{}^{*}\Gamma\times{}^{*}\Gamma\to\mathcal{W} with Im(α)𝒲bIm(\alpha)\subseteq\mathcal{W}_{b} can be described as {αn}𝒰\{\alpha_{n}\}_{\mathcal{U}} where for every k𝐍k\in\mathbf{N}, αn:Γ×Γ𝔲(kn)\alpha_{n}:\Gamma\times\Gamma\to\mathfrak{u}(k_{n}) is a bounded map, and such that {αn}𝒰𝐑b\{\|\alpha_{n}\|_{\infty}\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{b}. In other words, the internal map α\alpha is the ultraproduct of bounded maps, and is also bounded as an ultraproduct.
From now on, we shall regard α\alpha as an element of 𝒰((Γ)2,𝔲(kn)\prod_{\mathcal{U}}\ell^{\infty}(({}^{*}\Gamma)^{2},\mathfrak{u}(k_{n}), which we shall henceforth denote ((Γ)2,𝒲)\mathcal{L}^{\infty}(({}^{*}\Gamma)^{2},\mathcal{W}) (understood as the space of internal maps from (Γ)2({}^{*}\Gamma)^{2} to 𝒲\mathcal{W}. In fact, α\alpha is actually an element of ((Γ)2,𝒲)b\mathcal{L}^{\infty}(({}^{*}\Gamma)^{2},\mathcal{W})_{b} since not only is it internally bounded, but also {αn}𝒰𝐑b\{\|\alpha_{n}\|_{\infty}\}_{\mathcal{U}}\in{}^{*}\mathbf{R}_{b}. In general, we shall use the following notation:

  • For m𝐍m\in\mathbf{N}, the internal space ((Γ)m,𝒲)\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}) is defined as

    ((Γ)m,𝒲){(Γm,𝔲(kn))}𝒰\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W})\coloneq\Big{\{}\ell^{\infty}(\Gamma^{m},\mathfrak{u}(k_{n}))\Big{\}}_{\mathcal{U}}
  • The (external) subspace of ((Γ)m,𝒲)\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}) comprising internal functions with bounded (supremum) norm will be denoted b((Γ)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W}) while the (external) subspace of b((Γ)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W}) comprising internal functions with infinitesimal (supremum) norm will be denoted inf((Γ)m,𝒲)\mathcal{L}^{\infty}_{inf}(({}^{*}\Gamma)^{m},\mathcal{W}).

  • The quotient b((Γ)m,𝒲)/inf((Γ)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W})/\mathcal{L}^{\infty}_{inf}(({}^{*}\Gamma)^{m},\mathcal{W}) shall be denoted ~((Γ)m,𝒲)\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}). This space comprises bounded maps from (Γ)m({}^{*}\Gamma)^{m} to 𝒲~\tilde{\mathcal{W}} that are induced from internal maps in ((Γ)m,𝒲)\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}). As in Remark 3.2.6, ~((Γ)m,𝒲)\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}) is a real Banach space.

  • For a map fb((Γ)m,𝒲)f\in\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W}), we shall denote by f~~((Γ)m,𝒲)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}) the composition of ff with the canonical quotient map b((Γ)m,𝒲)~((Γ)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W})\to\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W}).

For convenience, we restate Proposition 3.2.8 in this notation:

Proposition 3.2.10.

Suppose for every αb((Γ)2,𝒲)\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{2},\mathcal{W}) that satisfies the 22-cocycle condition (Eq. 3.7), there exists a βb((Γ)1,𝒲)\beta\in\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{1},\mathcal{W}) such that

α~(g1,g2)=β~(g1)+g1β~(g2)β~(g1g2)\tilde{\alpha}(g_{1},g_{2})=\tilde{\beta}(g_{1})+g_{1}\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2})

then Γ\Gamma exhibits the defect diminishing property, and is therefore uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

3.3 Uniform Stability of Amenable groups

In this section, we shall demonstrate an application of Proposition 3.2.10 to amenable groups. The first step is to recognize the duality of 𝒲\mathcal{W} in an internal way.
Consider the space 𝒲=𝒰𝔲(kn)\mathcal{W}=\prod_{\mathcal{U}}\mathfrak{u}(k_{n}), and let 𝒲𝒰(𝔲(n))\mathcal{W}^{\sharp}\coloneqq\prod_{\mathcal{U}}(\mathfrak{u}(n))^{*}. For the Banach space 𝔲(n)\mathfrak{u}(n), consider its dual space (𝔲(n))(\mathfrak{u}(n))^{*} and let |\langle\cdot|\cdot\rangle the canonical duality (for instance, if 𝔲(n)\mathfrak{u}(n) is equipped with the Schatten pp-norm for p>1p>1, then (𝔲(n))(\mathfrak{u}(n))^{*} comes equipped with the Schatten qq-norm, where 1/p+1/q=11/p+1/q=1). Denote by 𝒲\mathcal{W}^{\sharp} the ultraproduct 𝒰(𝔲(kn))\prod_{\mathcal{U}}(\mathfrak{u}(k_{n}))^{*}, and its external subsets 𝒲b\mathcal{W}^{\sharp}_{b} and 𝒲inf\mathcal{W}^{\sharp}_{inf} as in (Eq. 3.3) and (Eq. 3.4). Note that the internal pairing

|𝒰:𝒲×𝒲𝐑\langle\cdot|\cdot\rangle_{\mathcal{U}}:\mathcal{W}^{\sharp}\times\mathcal{W}\to{}^{*}\mathbf{R}

induces a pairing

|:𝒲~×𝒲~𝐑\langle\cdot|\cdot\rangle:\tilde{\mathcal{W}}^{\sharp}\times\tilde{\mathcal{W}}\to\mathbf{R}

Equivalently, 𝒲b\mathcal{W}_{b}^{\sharp} comprises λ𝒲\lambda\in\mathcal{W}^{\sharp} such that λ(v)𝐑b\lambda(v)\in{}^{*}\mathbf{R}_{b} for every v𝒲bv\in\mathcal{W}_{b}, while 𝒲inf\mathcal{W}_{inf}^{\sharp} comprises λ𝒲\lambda\in\mathcal{W}^{\sharp} such that λ(v)𝐑inf\lambda(v)\in{}^{*}\mathbf{R}_{inf} for every v𝒲bv\in\mathcal{W}_{b}.
We can use the internal map πψ:Γ×𝒲𝒲\pi_{\psi}:{}^{*}\Gamma\times\mathcal{W}\to\mathcal{W} to define the internal map

πψ:Γ×𝒲𝒲\pi_{\psi}^{\sharp}:{}^{*}\Gamma\times\mathcal{W}^{\sharp}\to\mathcal{W}^{\sharp} (3.8)

which on λ𝒲\lambda\in\mathcal{W}^{\sharp} and v𝒲v\in\mathcal{W} is defined to be

πψ(g)(λ)(v)λ(πψ(g1)v)\pi_{\psi}^{\sharp}(g)(\lambda)(v)\coloneqq\lambda(\pi_{\psi}(g^{-1})v)
Lemma 3.3.1.

The internal map πψ\pi_{\psi}^{\sharp} restricts to a map on 𝒲b\mathcal{W}^{\sharp}_{b} that induces an action of Γ{}^{*}\Gamma on 𝒲b/𝒲inf\mathcal{W}^{\sharp}_{b}/\mathcal{W}^{\sharp}_{inf}.

Proof.

Let λ𝒲b\lambda\in\mathcal{W}_{b}^{\sharp}. For v𝒲v\in\mathcal{W}, since πψ(g)(λ)(v)=λ(πψ(g1)v)\pi_{\psi}^{\sharp}(g)(\lambda)(v)=\lambda(\pi_{\psi}(g^{-1})v), note that πψ(g1)v𝒲b\pi_{\psi}(g^{-1})v\in\mathcal{W}_{b} for v𝒲bv\in\mathcal{W}_{b}. Hence πψ(g)(λ)𝒲b\pi_{\psi}^{\sharp}(g)(\lambda)\in\mathcal{W}_{b}^{\sharp} for every gΓg\in{}^{*}\Gamma. Similarly, for λ𝒲inf\lambda\in\mathcal{W}_{inf}^{\sharp}, πψ(λ)𝒲inf\pi_{\psi}^{\sharp}(\lambda)\in\mathcal{W}_{inf}^{\sharp}. That this induces an action of Γ{}^{*}\Gamma on 𝒲b/𝒲inf\mathcal{W}_{b}^{\sharp}/\mathcal{W}_{inf}^{\sharp} follows easily from the fact that πψ\pi_{\psi} induces an action of Γ{}^{*}\Gamma on 𝒲b/𝒲inf\mathcal{W}_{b}/\mathcal{W}_{inf}. ∎

Going one step further, we can obtain a canonical identification of 𝒲\mathcal{W} with (𝒲)(\mathcal{W}^{\sharp})^{\sharp} (which has the same norm as 𝒲\mathcal{W}) so that we can regard 𝒲\mathcal{W} as (𝒲)(\mathcal{W}^{\sharp})^{\sharp} with (πψ)=πψ(\pi_{\psi}^{\sharp})^{\sharp}=\pi_{\psi}. This is true for 𝒲\mathcal{W} since 𝒲=𝒰𝔲(n)\mathcal{W}=\prod_{\mathcal{U}}\mathfrak{u}(n), and 𝔲(n)\mathfrak{u}(n) is finite-dimensional for each n𝐍n\in\mathbf{N}.

Remark 3.3.2.

We shall often use this reflexivity property of 𝒲\mathcal{W} to regard its dual 𝒲\mathcal{W}^{\sharp} as its predual, so that v𝒲v\in\mathcal{W} acts on λ𝒲\lambda\in\mathcal{W}^{\sharp} by vλ=λ(v)v\cdot\lambda=\lambda(v). However, note that for the following discussion of amenability, what we actually need is not reflexivity but merely the property that 𝒲\mathcal{W} is dual.

Let us now recall the definition of amenability for discrete groups. While there are innumerable equivalent definitions of amenability, here we shall see the definition that is most relevant to us (later on in §4.3, we shall study amenability and amenable actions in the locally compact case). Consider the Banach space (Γ)\ell^{\infty}(\Gamma) with the following action of Γ\Gamma: for g,xΓg,x\in\Gamma and f(Γ)f\in\ell^{\infty}(\Gamma), (gf)(x)=f(g1x)(g\cdot f)(x)=f(g^{-1}x). A mean on (Γ)\ell^{\infty}(\Gamma) is a bounded linear functional m:(Γ)𝐑m:\ell^{\infty}(\Gamma)\to\mathbf{R} such that m1\|m\|\leq 1, m(1)=1m(1)=1 and m(f)0m(f)\geq 0 whenever f0f\geq 0. The mean mm is said to be Γ\Gamma-invariant if for every gΓg\in\Gamma and f(Γ)f\in\ell^{\infty}(\Gamma), m(gf)=m(f)m(g\cdot f)=m(f).

Definition 3.3.3.

The discrete group Γ\Gamma is said to be amenable if there exists a Γ\Gamma-invariant mean on (Γ)\ell^{\infty}(\Gamma).

While the definition of amenability asks for a Γ\Gamma-invariant mean on (Γ)\ell^{\infty}(\Gamma), this can be easily extended to obtain a Γ\Gamma-equivariant mean on (Γ,W)\ell^{\infty}(\Gamma,W) for a dual normed WW-module (where the action of Γ\Gamma on (Γ,W)\ell^{\infty}(\Gamma,W) is given by (gf)(x)=gf(g1x)(g\cdot f)(x)=g\cdot f(g^{-1}x). The following lemma builds on this idea to construct an internal mean on (Γ,𝒲)\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W}).

Lemma 3.3.4.

Suppose Γ\Gamma is amenable. Then there exists an internal map min:(Γ,𝒲)𝒲m_{in}:\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W})\to\mathcal{W} such that minm_{in} induces a linear map m~:~(Γ,𝒲)𝒲~\tilde{m}:\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W})\to\tilde{\mathcal{W}} satisfying the following two conditions:

  • Suppose f~~(Γ,𝒲)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W}) is the constant function f~(g)=v~\tilde{f}(g)=\tilde{v} for every gΓg\in{}^{*}\Gamma, then m~(f~)=v~\tilde{m}(\tilde{f})=\tilde{v}.

  • For f~~(Γ,𝒲)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W}), m~(f~)f~\|\tilde{m}(\tilde{f})\|\leq\|\tilde{f}\|.

  • For gΓg\in{}^{*}\Gamma and f~~(Γ,𝒲)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W}), m~(gf~)=gm~(f~)\tilde{m}\left(g\cdot\tilde{f}\right)=g\cdot\tilde{m}(\tilde{f}).

Proof.

Consider f={fn}𝒰(Γ,𝒲)f=\{f_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W}). Since 𝒲=(𝒲)\mathcal{W}=(\mathcal{W}^{\sharp})^{\sharp}, for each λ𝒲\lambda\in\mathcal{W}^{\sharp}, we get an internal map

fλ:Γ𝐑f_{\lambda}:{}^{*}\Gamma\to{}^{*}\mathbf{R}
fλ(x)f(x)(λ)f_{\lambda}(x)\coloneqq f(x)(\lambda)

Note that fλf_{\lambda} being internal, is of the form {(fλ)n}𝒰\{(f_{\lambda})_{n}\}_{\mathcal{U}} where (fλ)n(Γ)(f_{\lambda})_{n}\in\ell^{\infty}(\Gamma). This allows us to construct the internal map minλ:(Γ,𝒲)𝐑m_{in}^{\lambda}:\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W})\to{}^{*}\mathbf{R} as

minλ(f)={m((fλ)n)}𝒰m_{in}^{\lambda}(f)=\{m\left((f_{\lambda})_{n}\right)\}_{\mathcal{U}}

and finally min:(Γ,𝒲)(𝒲)m_{in}:\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W})\to(\mathcal{W}^{\sharp})^{\sharp} as

min(f)(λ)minλ(f)m_{in}(f)(\lambda)\coloneqq m_{in}^{\lambda}(f)

It is straightforward to check that minm_{in} as defined induces a linear map m~:~(Γ,𝒲)𝒲b/𝒲inf\tilde{m}:\tilde{\mathcal{L}}({}^{*}\Gamma,\mathcal{W})\to\mathcal{W}_{b}/\mathcal{W}_{inf}. As for Γ{}^{*}\Gamma-equivariance, this follows from the observation that (gf)λ(x)=π(g)f(g1x)(λ)(g\cdot f)_{\lambda}(x)=\pi(g)f(g^{-1}x)(\lambda) while (gfλ)(x)=f(g1x)(λ)(g\cdot f_{\lambda})(x)=f(g^{-1}x)(\lambda). The conditions on m~\tilde{m} follow from the definition and properties of the Γ\Gamma-invariant mean mm on (Γ)\ell^{\infty}(\Gamma). ∎

We shall denote the internal mean above by minxm_{in}^{x} (or m~x\tilde{m}^{x}) when the mean is understood to be taken over xΓx\in\Gamma. This would be particularly useful when working with multivariate maps where we fix certain coordinates to obtain a univariate map which we can take a mean over (as in the following Proposition 3.3.5). Note that in this notation, it is easy to see that the Γ{}^{*}\Gamma-equivariance of the mean constructed above translates to the simpler (invariant) form: for f~~(Γ,𝒲)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W}) and gΓg\in{}^{*}\Gamma,

m~x(f~(gx))=m~x(f~(x))\tilde{m}^{x}\left(\tilde{f}(gx)\right)=\tilde{m}^{x}\left(\tilde{f}(x)\right)

We shall now use this internal map minm_{in} and the Γ{}^{*}\Gamma-equivariant map m~\tilde{m} to show the following:.

Proposition 3.3.5.

Suppose Γ\Gamma is amenable. Then for every αb((Γ)2,𝒲)\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{2},\mathcal{W}) that satisfies the 22-cocycle condition (Eq. 3.7), there exists a βb(Γ,𝒲)\beta\in\mathcal{L}^{\infty}_{b}({}^{*}\Gamma,\mathcal{W}) such that

α~(g1,g2)=β~(g1)+g1β~(g2)β~(g1g2)\tilde{\alpha}(g_{1},g_{2})=\tilde{\beta}(g_{1})+g_{1}\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2})
Proof.

Suppose αL((Γ)2,𝒲)\alpha\in\mathcal{L}^{\infty}L(({}^{*}\Gamma)^{2},\mathcal{W}) satisfies the 22-cocyle condition: for every g1,g2,xΓg_{1},g_{2},x\in{}^{*}\Gamma,

α~(g1,g2)=g1α~(g2,x)α~(g1g2,x)+α~(g1,g2x)\tilde{\alpha}(g_{1},g_{2})=g_{1}\cdot\tilde{\alpha}(g_{2},x)-\tilde{\alpha}(g_{1}g_{2},x)+\tilde{\alpha}(g_{1},g_{2}x)

For a fixed gg, the map αg:Γ𝒲\alpha_{g}:{}^{*}\Gamma\to\mathcal{W} defined as αg(x)α(g,x)\alpha_{g}(x)\coloneq\alpha(g,x) is clearly contained in (Γ,𝒲)\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W}). Define β(Γ,𝒲)\beta\in\mathcal{L}^{\infty}({}^{*}\Gamma,\mathcal{W}) as

β(g)min(αg)\beta(g)\coloneq m_{in}\left(\alpha_{g}\right)

In other words, β(g)=minx(α(g,x))\beta(g)=m_{in}^{x}\left(\alpha(g,x)\right). Then the 22-cocycle condition satisfied by α\alpha immediately implies that

α~(g1,g2)=g1β~(g2)β~(g1g2)+β~(g1)\tilde{\alpha}(g_{1},g_{2})=g_{1}\cdot\tilde{\beta}(g_{2})-\tilde{\beta}(g_{1}g_{2})+\tilde{\beta}(g_{1})

Observe that the proof of Proposition 3.3.5 is almost exactly on the lines of the proof that Hb2(Γ,V)=0\operatorname{H}_{b}^{2}(\Gamma,V)=0 for amenable Γ\Gamma and dual normed Γ\Gamma-module WW (refer to Theorem 3.63.6 in [24] for more details).
In light of Proposition 3.3.5 and Proposition 3.2.10, we conclude that:

Corollary 3.3.6.

If Γ\Gamma is a discrete amenable group, then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Note that while on the one hand Corollary 3.3.6 generalizes Kazhdan’s result [31] to a larger family (where we allow any submultiplicative matrix norm as opposed to just the operator norm as in [31] and [15]), we do not prove here the analogous result of strong Ulam stability, where the family comprises groups of unitary operators on (possibly infinite-dimensional) Hilbert spaces.

4 Asymptotic Cohomology of Groups

In this section, we shall formally define the asymptotic cohomology theory of (topological) groups, and study some basic properties along the lines of the theory of bounded cohomology. Recall that our goal is to prove uniform 𝔘\mathfrak{U}-stability for lattices in higher rank Lie groups, so this forces us to develop the cohomology theory for locally compact groups (as opposed to just discrete groups as we briefly saw in §3.2 and §3.3).
The basic objects we shall deal with are defined in §4.1, where we describe the category of asymptotic cohomology abstractly using tools from cohomological algebra. In §4.3, we define the asymptotic cohomology of groups and relate it to the way it was motivated in §3.3. Finally, in §4.3, we use Zimmer amenability and the functorial relations of §4.1 to obtain other complexes that compute the same cohomology, which we shall use in §5 and §6.

4.1 Basic Definitions and Some Cohomological Algebra

Recall, from §2.2, that we fix a non-principal ultrafilter 𝒰\mathcal{U} on 𝐍\mathbf{N} to define ultraproducts and internal objects. For convenience, we set some notation and conventions now. Let CC be a category, with CC-objects and CC-morphisms. We shall define a category C{}^{*}C in Univ{}^{*}Univ as follows:

  • The objects of C{}^{*}C, referred to as internal CC-objects, shall be ultraproducts 𝒰Xn\prod_{\mathcal{U}}X_{n} where {Xn}n𝐍\{X_{n}\}_{n\in\mathbf{N}} is an indexed collection of CC-objects.

  • The morphisms of C{}^{*}C, referred to as internal morphism, are of the form ϕ={ϕn}𝒰\phi=\{\phi_{n}\}_{\mathcal{U}} (also denoted 𝒰ϕn\prod_{\mathcal{U}}\phi_{n}), where {ϕn}n𝐍\{\phi_{n}\}_{n\in\mathbf{N}} is an indexed collection of CC-morphisms.

Given a category CC, the internal CC-objects and internal CC-morphisms form a category C{}^{*}C in Univ{}^{*}Univ, and these two categories have the same first-order theories, allowing us to use the transfer principle, as remarked in §2.2.

Definition 4.1.1.

Let AA be a property of CC-objects (resp, CC-morphisms). Then 𝒰Xn\prod_{\mathcal{U}}X_{n} (resp, ϕ:𝒰Xn𝒰Yn\phi:\prod_{\mathcal{U}}X_{n}\to\prod_{\mathcal{U}}Y_{n}) has internal AA if XnX_{n} (resp, ϕn\phi_{n}) has AA for every nSn\in S with S𝒰S\in\mathcal{U}.

For example, let GG be a locally compact, second countable topological group, and consider the ultrapower group G{}^{*}G and let ={En}𝒰\mathcal{E}=\{E_{n}\}_{\mathcal{U}} be an internal Banach space. An internal map π:G×\pi:{}^{*}G\times\mathcal{E}\to\mathcal{E}, where π={πn:G×EnEn}\pi=\{\pi_{n}:G\times E_{n}\to E_{n}\}, is an internal action of G{}^{*}G on \mathcal{E} (or \mathcal{E} is an internal G{}^{*}G-representation) if the map πn:G×EnEn\pi_{n}:G\times E_{n}\to E_{n} is an isometric GG-representation for every n𝐍n\in\mathbf{N}, and the internal map π\pi is internally continuous if πn:G×EnEn\pi_{n}:G\times E_{n}\to E_{n} is continuous for every n𝐍n\in\mathbf{N} (here G×EnG\times E_{n} is endowed with the product topology). All of these notions simply involve passing from standard categories to their internal counterparts in Univ{}^{*}Univ.
We shall now work with the category BanBan whose objects are (real) Banach spaces and whose morphisms are bounded linear maps, and study Ban{}^{*}Ban. Consider the ultraproduct =𝒰En\mathcal{E}=\prod_{\mathcal{U}}E_{n} of the real Banach spaces {En}n𝐍\{E_{n}\}_{n\in\mathbf{N}}, which is an internal Banach space. For an element vv\in\mathcal{E} where v={vn}𝒰v=\{v_{n}\}_{\mathcal{U}}, we denote by v{vn}𝒰𝐑\|v\|\coloneq\{\|v_{n}\|\}_{\mathcal{U}}\in{}^{*}\mathbf{R}. Given two internal Banach spaces ={En}𝒰\mathcal{E}=\{E_{n}\}_{\mathcal{U}} and ={Fn}𝒰\mathcal{F}=\{F_{n}\}_{\mathcal{U}}, we shall denote by 𝓂(,)\mathcal{Hom}(\mathcal{E},\mathcal{E}) the set of internal morphisms between \mathcal{E} and \mathcal{F}. These are exactly of the form ϕ{ϕn:EnFn}𝒰\phi\coloneq\{\phi_{n}:E_{n}\to F_{n}\}_{\mathcal{U}} where ϕn:EnFn\phi_{n}:E_{n}\to F_{n} is a bounded linear map for every n𝐍n\in\mathbf{N} (such maps are internal morphisms). Note that 𝓂(,)\mathcal{Hom}(\mathcal{E},\mathcal{E}) itself is an internal Banach space when endowed with the internal operator norm (that is, ϕ{ϕnop}𝒰\|\phi\|\coloneq\{\|\phi_{n}\|_{op}\}_{\mathcal{U}}).
Consider the set 𝓂(,𝐑)\mathcal{Hom}(\mathcal{E},{}^{*}\mathbf{R}), which is the internal dual of \mathcal{E}, denoted \mathcal{E}^{\sharp}. Explicitly, for each Banach space EnE_{n} as above, let EnE_{n}^{\sharp} denote its (constinuous) dual Banach space, and let |\langle\cdot|\cdot\rangle the canonical duality, and \mathcal{E}^{\sharp} denote the ultraproduct 𝒰En\prod_{\mathcal{U}}E_{n}^{\sharp}. Note that we have an internal pairing

|𝒰:×𝐑\langle\cdot|\cdot\rangle_{\mathcal{U}}:\mathcal{E}^{\sharp}\times\mathcal{E}\to{}^{*}\mathbf{R}

We shall call \mathcal{E} an internal dual Banach space if \mathcal{E} is the internal dual of some internal Banach space (which we shall denote =𝒰En\mathcal{E}^{\flat}=\prod_{\mathcal{U}}E_{n}^{\flat}). For ϕ𝓂(,)\phi\in\mathcal{Hom}(\mathcal{E},\mathcal{E}), we shall denote by ϕ𝓂(,)\phi^{\sharp}\in\mathcal{Hom}(\mathcal{E}^{\sharp},\mathcal{E}^{\sharp}) the internal adjoint map with respect to the internal pairing above: that is ϕ\phi^{\sharp} is such that for every vv\in\mathcal{E} and ww\in\mathcal{E}^{\sharp}, ϕw,v𝒰=w,ϕv𝒰\langle\phi^{\sharp}w,v\rangle_{\mathcal{U}}=\langle w,\phi v\rangle_{\mathcal{U}}.
We shall now see some general functorial aspects of (standard) real Banach spaces and extend them naturally to internal Banach spaces. Let XX, YY, EE and FF be (standard) real Banach spaces. Let B(X×E)B(X\times E) denote the Banach space of bounded bilinear forms X×E𝐑X\times E\to\mathbf{R}, and L(E,F)L(E,F) denote the Banach space of bounded linear functions from EE to FF. Through the canonical pairing, note that B(X×E)B(X\times E) is naturally isometrically isomorphic to the Banach space of bounded linear maps EXE\to X^{\sharp} (and also to the Banach space of bounded linear maps XEX\to E^{\sharp}). That is,

B(X×E)L(E,X)L(X,E)B(X\times E)\cong L(E,X^{\sharp})\cong L(X,E^{\sharp}) (4.1)

Using these identifications, we now consider two functors:

  • Given a bounded linear operator k:XYk:X^{\sharp}\to Y^{\sharp}, define a bounded linear operator k:B(X×E)B(Y×E)k_{*}:B(X\times E)\to B(Y\times E) by (kβ)(,e)=kβ(,e))(k_{*}\beta)(\cdot,e)=k\beta(\cdot,e)). Note that the correspondence kkk\to k_{*} is covariant (although kk_{*} depends on EE, for ease of notation, we assume the relevant Banach space from context).

  • Given a bounded linear operator σ:EF\sigma:E\to F, define a bounded linear operator σ:B(X×F):B(X×E)\sigma^{*}:B(X\times F)\to:B(X\times E) by (σβ)(x,e)=β(x,σ(e))(\sigma^{*}\beta)(x,e)=\beta(x,\sigma(e)). In this case, the correspondence σσ\sigma\to\sigma^{*} is contravariant (although σ\sigma^{*} depends on XX, for ease of notation, we assume the relevant Banach space from context).

Proposition 4.1.2.

Let k:XYk:X^{\sharp}\to Y^{\sharp} and σ:EF\sigma:E\to F, and consider the covariant k:B(X×E)B(Y×E)k_{*}:B(X\times E)\to B(Y\times E) and contravariant σ:B(X×F):B(X×E)\sigma^{*}:B(X\times F)\to:B(X\times E) as defined above. Then

  • kσ=σkk_{*}\sigma^{*}=\sigma^{*}k_{*}

  • kσkσ\|k_{*}\sigma^{*}\|\leq\|k_{*}\|\cdot\|\sigma^{*}\|

These notions extend easily to internal Banach spaces as well. Let 𝒳={Xn}𝒰\mathcal{X}=\{X_{n}\}_{\mathcal{U}}, 𝒴={Yn}𝒰\mathcal{Y}=\{Y_{n}\}_{\mathcal{U}} and ={En}𝒰\mathcal{E}=\{E_{n}\}_{\mathcal{U}}, ={Fn}𝒰\mathcal{F}=\{F_{n}\}_{\mathcal{U}} be internal Banach spaces. Denote by (𝒳×)\mathcal{B}(\mathcal{X}\times\mathcal{E}) the internal Banach space (𝒳×)𝒰B(Xn×En)\mathcal{B}(\mathcal{X}\times\mathcal{E})\coloneq\prod_{\mathcal{U}}B(X_{n}\times E_{n}) and by (𝒳,)𝒰L(Xn,En)\mathcal{L}(\mathcal{X},\mathcal{E})\coloneq\prod_{\mathcal{U}}L(X_{n},E_{n}). Then

(𝒳×)(,𝒳)(𝒳,)\mathcal{B}(\mathcal{X}\times\mathcal{E})\cong\mathcal{L}(\mathcal{E},\mathcal{X}^{\sharp})\cong\mathcal{L}(\mathcal{X},\mathcal{E}^{\sharp})

where the isomorphisms are internally isometric. Let k:𝒳𝒴k:\mathcal{X}^{\sharp}\to\mathcal{Y}^{\sharp} (where k={kn}𝒰k=\{k_{n}\}_{\mathcal{U}}), and σ:\sigma:\mathcal{E}\to\mathcal{F} (where σ={σn}𝒰\sigma=\{\sigma_{n}\}_{\mathcal{U}}) be internal morphisms. Then we have the covariant internal morphism k:(𝒳×)(𝒴×)k_{*}:\mathcal{B}(\mathcal{X}\times\mathcal{E})\to\mathcal{B}(\mathcal{Y}\times\mathcal{E}) defined as k={(kn)}𝒰k_{*}=\{(k_{n})_{*}\}_{\mathcal{U}}), and the contravariant internal morphism σ:(𝒳×)(𝒳×)\sigma^{*}:\mathcal{B}(\mathcal{X}\times\mathcal{F})\to\mathcal{B}(\mathcal{X}\times\mathcal{E}) defined by σ={σn}𝒰\sigma^{*}=\{\sigma_{n}^{*}\}_{\mathcal{U}} such that kσ=σkk_{*}\sigma^{*}=\sigma^{*}k_{*} and kσkσ\|k_{*}\sigma^{*}\|\leq\|k_{*}\|\cdot\|\sigma^{*}\|.
Let \mathcal{E} be an internal Banach space. Just as in as in (Eq. 3.3) and (Eq. 3.4), the internal norm :𝐑\|\cdot\|:\mathcal{E}\to{}^{*}\mathbf{R} allows us to define special external subsets as follows:

b{v:v𝐑b}\mathcal{E}_{b}\coloneq\{v\in\mathcal{E}:\|v\|\in{}^{*}\mathbf{R}_{b}\}
inf{v:v𝐑inf}\mathcal{E}_{inf}\coloneq\{v\in\mathcal{E}:\|v\|\in{}^{*}\mathbf{R}_{inf}\}

and denote the ultralimit b/inf\mathcal{E}_{b}/\mathcal{E}_{inf} by ~\tilde{\mathcal{E}}, which is a real Banach space [29]). The correspondence ~\mathcal{E}\mapsto\tilde{\mathcal{E}} is not a functor from Ban{}^{*}Ban to BanBan as such, because an internal morphism ϕ𝓂(,)\phi\in\mathcal{Hom}(\mathcal{E},\mathcal{F}) need not induce a morphism ϕ~:~~\tilde{\phi}:\tilde{\mathcal{E}}\to\tilde{\mathcal{F}} in general. However, if we restrict ourselves to bounded objects and morphisms in Ban{}^{*}Ban, then we do get a functorial correspondence. That is, consider the external subsets 𝓂b(,)\mathcal{Hom}_{b}(\mathcal{E},\mathcal{F}) and 𝓂inf(,)\mathcal{Hom}_{inf}(\mathcal{E},\mathcal{F}). Then

Proposition 4.1.3.

Any ϕ𝓂b(,)\phi\in\mathcal{Hom}_{b}(\mathcal{E},\mathcal{F}), induces a map ϕ~Hom(~,~)\tilde{\phi}\in Hom(\tilde{\mathcal{E}},\tilde{\mathcal{F}}).

Proof.

Note that ϕ{ϕn:EnFn}𝒰\phi\coloneq\{\phi_{n}:E_{n}\to F_{n}\}_{\mathcal{U}}, where each ϕn:EnFn\phi_{n}:E_{n}\to F_{n} is a bounded linear map for every n𝐍n\in\mathbf{N}. Since ϕ𝓂b(,)\phi\in\mathcal{Hom}_{b}(\mathcal{E},\mathcal{F}), this means that ϕ(b)b\phi(\mathcal{E}_{b})\subseteq\mathcal{F}_{b}, and ϕ(inf)inf\phi(\mathcal{E}_{inf})\subseteq\mathcal{F}_{inf}, thus inducing a bounded linear map ϕ~:~~\tilde{\phi}:\tilde{\mathcal{E}}\to\tilde{\mathcal{F}} between the real Banach spaces ~\tilde{\mathcal{E}} and ~\tilde{\mathcal{F}}. ∎

For example, if ϕ:\phi:\mathcal{E}\to\mathcal{F} is an internal isometry, then ϕ𝓂b(,)\phi\in\mathcal{Hom}_{b}(\mathcal{E},\mathcal{F}) induces ϕ~:~~\tilde{\phi}:\tilde{\mathcal{E}}\to\tilde{\mathcal{F}}.
Observe that 𝓂~(,)\tilde{\mathcal{Hom}}(\mathcal{E},\mathcal{F}) is a subspace of Hom(~,~)Hom(\tilde{\mathcal{E}},\tilde{\mathcal{F}}). The latter comprises the Banach space of all bounded linear maps between real Banach spaces ~\tilde{\mathcal{E}} and ~\tilde{\mathcal{F}}, while the former is a Banach subspace comprising those bounded linear maps that were induced from internal morphisms from \mathcal{E} to \mathcal{F}.

Proposition 4.1.4.

Let \mathcal{E} be an internal Banach space with dual \mathcal{E}^{\sharp}. Then the internal pairing 𝒰b(,)\langle\cdot\rangle_{\mathcal{U}}\in\mathcal{B}_{b}(\mathcal{E}^{\sharp},\mathcal{E}). Furthermore, for ϕ𝓂b(,)\phi\in\mathcal{Hom}_{b}(\mathcal{E},\mathcal{E}), its internal adjoint ϕ𝓂b(,)\phi^{\sharp}\in\mathcal{Hom}_{b}(\mathcal{E}^{\sharp},\mathcal{E}^{\sharp}).

Thus, many of our functorial results about internal Banach spaces in Ban{}^{*}Ban pass through when we restrict to bounded elements, and induce corresponding results in BanBan.
Our main structure of interest is an asymptotic variant of internal G{}^{*}G-representations using the external subsets b\mathcal{E}_{b} and inf\mathcal{E}_{inf}:

Definition 4.1.5.

Let GG be a locally compact, second countable topological group, and \mathcal{E} be an internal Banach space. An internal isometry π:G×\pi:{}^{*}G\times\mathcal{E}\to\mathcal{E} be an internal isometry such that it induces an action π~:G×~~\tilde{\pi}:{}^{*}G\times\tilde{\mathcal{E}}\to\tilde{\mathcal{E}} of G{}^{*}G on the real Banach space ~\tilde{\mathcal{E}}. The internal Banach space \mathcal{E}, equipped with such an internal map π\pi, is called an asymptotic Banach G{}^{*}G-module with asymptotic G{}^{*}G-representation π\pi.

We shall denote an asymptotic Banach G{}^{*}G-module either by (π,)(\pi,\mathcal{E}), or just \mathcal{E} if the map π\pi can implicitly be assumed in context.
Observe that the real Banach space ~\tilde{\mathcal{E}} is a (true) representation of G{}^{*}G through π~\tilde{\pi}. An element vbv\in\mathcal{E}_{b} is said to be asymptotically fixed by gGg\in{}^{*}G if its image v~~\tilde{v}\in\tilde{\mathcal{E}} is (truly) fixed by gg. The set of asymptotically G{}^{*}G-fixed elements of \mathcal{E} shall be denoted bG\mathcal{E}_{b}^{\sim{}^{*}G}. More generally, for an internal subgroup 𝒩G\mathcal{N}\leq{}^{*}G, the set of asymptotically 𝒩\mathcal{N}-fixed elements of \mathcal{E} shall be denoted b𝒩\mathcal{E}_{b}^{\sim\mathcal{N}}
For an asymptotic Banach G{}^{*}G-module \mathcal{E}, consider its dual internal Banach space =(,𝐑)\mathcal{E}^{\sharp}=\mathcal{L}(\mathcal{E},{}^{*}\mathbf{R}). In this case, the internal map π:G×\pi:{}^{*}G\times\mathcal{E}\to\mathcal{E} defines an internal map

π:G×\pi^{\sharp}:{}^{*}G\times\mathcal{E}^{\sharp}\to\mathcal{E}^{\sharp}
π(g)(λ)(v)=λ(π(g)1v)\pi^{\sharp}(g)(\lambda)(v)=\lambda\left(\pi(g)^{-1}v\right)

in the usual way, and it is easy to check that \mathcal{E}^{\sharp} an asymptotic Banach G{}^{*}G-module (π,)(\pi^{\sharp},\mathcal{E}^{\sharp}). Essentially, we are simply defining the internal adjoint of π(g)\pi(g) with respect to the pairing of \mathcal{E} and \mathcal{E}^{\sharp}, and the functorial results of Proposition 4.1.4 and Proposition 4.1.3 ensure that the map π\pi^{\sharp} defined this way is an asymptotic G{}^{*}G-representation of G{}^{*}G on \mathcal{E}^{\sharp}. An asymptotic Banach G{}^{*}G-module (π,)(\pi,\mathcal{E}) is called a dual asymptotic Banach G{}^{*}G-module if (π,)(\pi,\mathcal{E}) is the dual of an asymptotic Banach G{}^{*}G-module (π,)(\pi^{\flat},\mathcal{E}^{\flat}).

More generally, let (π,)(\pi,\mathcal{E}) and (ρ,)(\rho,\mathcal{F}) be asymptotic Banach G{}^{*}G-modules. Then (,)\mathcal{L}(\mathcal{E},\mathcal{F}) and (,)\mathcal{B}(\mathcal{E},\mathcal{F}) can also be regarded as asymptotic Banach G{}^{*}G-modules in a natural way.

Definition 4.1.6.

Let (π,)(\pi,\mathcal{E}) and (ρ,)(\rho,\mathcal{F}) be asymptotic Banach G{}^{*}G-modules. An asymptotic G{}^{*}G-morphism from \mathcal{E} to \mathcal{F} is a map ϕb(,)\phi\in\mathcal{L}_{b}(\mathcal{E},\mathcal{F}) such that the induced ϕ~:~~\tilde{\phi}:\tilde{\mathcal{E}}\to\tilde{\mathcal{F}} is G{}^{*}G-equivariant (that is, ϕ~\tilde{\phi} is a morphism of G{}^{*}G-representations ~\tilde{\mathcal{E}} and ~\tilde{\mathcal{F}}).

In other words, an asymptotic G{}^{*}G-morphism is an element of b(,)\mathcal{L}_{b}(\mathcal{E},\mathcal{F}) whose image in ~(,)\tilde{\mathcal{L}}(\mathcal{E},\mathcal{F}) is a G{}^{*}G-fixed point, and the set of such elements is denoted 𝓂G(,)\mathcal{Hom}_{G}(\mathcal{E},\mathcal{F}). The following proposition tells us that the functors kkk\mapsto k_{*} and σσ\sigma\mapsto\sigma^{*}, defined for internal Banach spaces, respect the asymptotic G{}^{*}G-representations.

Proposition 4.1.7.

Let ,,𝒳\mathcal{E},\mathcal{F},\mathcal{X} and 𝒴\mathcal{Y} be asymptotic Banach G{}^{*}G-modules, and k:𝒳𝒴k:\mathcal{X}^{\sharp}\to\mathcal{Y}^{\sharp} and σ:\sigma:\mathcal{E}\to\mathcal{F} be asymptotic G{}^{*}G-morphisms. Then k:(𝒳,)(𝒴,)k_{*}:\mathcal{B}(\mathcal{X},\mathcal{E})\to\mathcal{B}(\mathcal{Y},\mathcal{E}) and σ:(𝒳,)(𝒳,)\sigma^{*}:\mathcal{B}(\mathcal{X},\mathcal{F})\to\mathcal{B}(\mathcal{X},\mathcal{E}) are also asymptotic G{}^{*}G-morphisms.

Definition 4.1.8.

An asymptotic G{}^{*}G-cochain complex (,d)(\mathcal{E}^{\bullet},d^{\bullet}) is a 0\mathbb{Z}_{\small{\geq 0}}-indexed sequence

0{0}0{\mathcal{E}^{0}}1{\mathcal{E}^{1}}2{\mathcal{E}^{2}}3{\mathcal{E}^{3}}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

where for every n0n\geq 0, n\mathcal{E}^{n} is an asymptotic Banach G{}^{*}G-module, dnd^{n} is an asymptotic G{}^{*}G-morphism, and dn+1dn(bn)infn+2d^{n+1}d^{n}(\mathcal{E}^{n}_{b})\subseteq\mathcal{E}^{n+2}_{inf}.

We shall also denote (,d)(\mathcal{E}^{\bullet},d^{\bullet}) by just \mathcal{E}^{\bullet} if the maps dnd^{n} are assumed from context, and also assume that 1=0\mathcal{E}^{-1}=0 to make statements with indices easier. Observe that the condition dn+1dn(bn1)infn+1d^{n+1}d^{n}(\mathcal{E}^{n-1}_{b})\subseteq\mathcal{E}^{n+1}_{inf} is a relaxation to infinitesimals of the usual condition on differential maps. In fact, since the maps dnd^{n} are asymptotic G{}^{*}G-morphisms, the asymptotic G{}^{*}G-cochain complex induces a (true) cochain complex

0{0}(~0)G{(\tilde{\mathcal{E}}^{0})^{{}^{*}G}}(~1)G{(\tilde{\mathcal{E}}^{1})^{{}^{*}G}}(~2)G{(\tilde{\mathcal{E}}^{2})^{{}^{*}G}}(~3)G{(\tilde{\mathcal{E}}^{3})^{{}^{*}G}}{\dots}d~0\scriptstyle{\tilde{d}^{0}}d~1\scriptstyle{\tilde{d}^{1}}d~2\scriptstyle{\tilde{d}^{2}}d~3\scriptstyle{\tilde{d}^{3}}d~4\scriptstyle{\tilde{d}^{4}}

allowing us to define the following:

Definition 4.1.9.

The asymptotic cohomology of the asymptotic G{}^{*}G-cochain complex (,d)(\mathcal{E}^{\bullet},d^{\bullet}), denoted Ha(,d)\operatorname{H}_{a}^{\bullet}(\mathcal{E}^{\bullet},d^{\bullet}) or Ha()\operatorname{H}_{a}^{\bullet}(\mathcal{E}^{\bullet}), is defined to be

Ham()ker(d~m+1)/Im(d~m)\operatorname{H}_{a}^{m}(\mathcal{E}^{\bullet})\coloneq ker(\tilde{d}^{m+1})/Im(\tilde{d}^{m})

An element αb\alpha\in\mathcal{E}_{b} such that α~ker(d~m+1)\tilde{\alpha}\in ker(\tilde{d}^{m+1}) is called an asymptotic mm-cocycle, and will be called an asymptotic mm-coboundary if α~Im(d~m)\tilde{\alpha}\in Im(\tilde{d}^{m}).

To understand the correspondence Ha()\mathcal{E}^{\bullet}\mapsto\operatorname{H}_{a}^{\bullet}(\mathcal{E}^{\bullet}), we now define morphisms and homotopies of asymptotic G{}^{*}G-cochain complexes.

Definition 4.1.10.

Let (,d)(\mathcal{E}^{\bullet},d_{\mathcal{E}}^{\bullet}) and (,d)(\mathcal{F}^{\bullet},d_{\mathcal{F}}^{\bullet}) be asymptotic G{}^{*}G-cochain complexes. An asymptotic G{}^{*}G-morphism α:\alpha^{\bullet}:\mathcal{E}^{\bullet}\to\mathcal{F}^{\bullet} between (,d)(\mathcal{E}^{\bullet},d_{\mathcal{E}}^{\bullet}) and (,d)(\mathcal{F}^{\bullet},d_{\mathcal{F}}^{\bullet}) is a family of asymptotic G{}^{*}G-morphisms αn:nn\alpha^{n}:\mathcal{E}^{n}\to\mathcal{F}^{n} for every n0n\in\mathbb{Z}_{\small{\geq 0}} such that for every n0n\in\mathbb{Z}_{\small{\geq 0}},

(dn+1αnαn+1dn)(bn)infn+1(d^{n+1}_{\mathcal{F}}\alpha^{n}-\alpha^{n+1}d^{n}_{\mathcal{E}})(\mathcal{E}^{n}_{b})\subseteq\mathcal{F}^{n+1}_{inf}

Again, this simply means that we have a G{}^{*}G-morphism of the cochain complexes

0{0}(~0)G{(\tilde{\mathcal{E}}^{0})^{{}^{*}G}}(~1)G{(\tilde{\mathcal{E}}^{1})^{{}^{*}G}}(~2)G{(\tilde{\mathcal{E}}^{2})^{{}^{*}G}}(~3)G{(\tilde{\mathcal{E}}^{3})^{{}^{*}G}}{\dots}0{0}(~0)G{(\tilde{\mathcal{F}}^{0})^{{}^{*}G}}(~1)G{(\tilde{\mathcal{F}}^{1})^{{}^{*}G}}(~2)G{(\tilde{\mathcal{F}}^{2})^{{}^{*}G}}(~3)G{(\tilde{\mathcal{F}}^{3})^{{}^{*}G}}{\dots}d~0\scriptstyle{\tilde{d}^{0}}α~0\scriptstyle{\tilde{\alpha}^{0}}d~1\scriptstyle{\tilde{d}^{1}}α~1\scriptstyle{\tilde{\alpha}^{1}}d~2\scriptstyle{\tilde{d}^{2}}α~2\scriptstyle{\tilde{\alpha}^{2}}d~3\scriptstyle{\tilde{d}^{3}}α~3\scriptstyle{\tilde{\alpha}^{3}}d~4\scriptstyle{\tilde{d}^{4}}d~0\scriptstyle{\tilde{d}^{0}}d~1\scriptstyle{\tilde{d}^{1}}d~2\scriptstyle{\tilde{d}^{2}}d~3\scriptstyle{\tilde{d}^{3}}d~4\scriptstyle{\tilde{d}^{4}}

Let α\alpha^{\bullet} be an asymptotic G{}^{*}G-morphism between the asymptotic G{}^{*}G-cochain complexes 𝒳\mathcal{X}^{\bullet} and 𝒴\mathcal{Y}^{\bullet}. While this clearly gives us an induced map Ha(𝒳)Ha(𝒴)\operatorname{H}_{a}^{\bullet}(\mathcal{X}^{\bullet})\to\operatorname{H}_{a}^{\bullet}(\mathcal{Y}^{\bullet}), we now describe when two such asymptotic G{}^{*}G-morphisms α\alpha^{\bullet} and β\beta^{\bullet} correspond to the same induced maps of cohomologies.

Definition 4.1.11.

Let α,β:𝒳𝒴\alpha^{\bullet},\beta^{\bullet}:\mathcal{X}^{\bullet}\to\mathcal{Y}^{\bullet} be two asymptotic G{}^{*}G-morphisms between the asymptotic G{}^{*}G-cochain complex 𝒳\mathcal{X}^{\bullet} and 𝒴\mathcal{Y}^{\bullet}. Then α\alpha^{\bullet} is said to be asymptotically G{}^{*}G-homotopic to β\beta^{\bullet} if there exists a family of asymptotic G{}^{*}G-morphisms σn:nn1\sigma^{n}:\mathcal{E}^{n}\to\mathcal{F}^{n-1} such that

d~nσ~n+σ~n+1d~n+1=α~nβ~n\tilde{d}^{n}\tilde{\sigma}^{n}+\tilde{\sigma}^{n+1}\tilde{d}^{n+1}=\tilde{\alpha}^{n}-\tilde{\beta}^{n}

for every n0n\in\mathbb{Z}_{\small{\geq 0}}.

The following lemma lists results that follow from standard cohomological techniques:

Proposition 4.1.12.

Let 𝒳\mathcal{X}^{\bullet} and 𝒴\mathcal{Y}^{\bullet} be asymptotic G{}^{*}G-cochain complexes.

  • Suppose α,β:𝒳𝒴\alpha^{\bullet},\beta^{\bullet}:\mathcal{X}^{\bullet}\to\mathcal{Y}^{\bullet} are asymptotic G{}^{*}G-morphisms such that α\alpha^{\bullet} is asymptotically G{}^{*}G-homotopic to β\beta^{\bullet}. Then they induce the same map Ha(𝒳)Ha(𝒴)\operatorname{H}_{a}^{\bullet}(\mathcal{X}^{\bullet})\to\operatorname{H}_{a}^{\bullet}(\mathcal{Y}^{\bullet}) at the level of cohomology.

  • Suppose α:𝒳𝒴\alpha^{\bullet}:\mathcal{X}^{\bullet}\to\mathcal{Y}^{\bullet} and β:𝒴𝒳\beta^{\bullet}:\mathcal{Y}^{\bullet}\to\mathcal{X}^{\bullet} are asymptotic G{}^{*}G-morphisms such that αβ:𝒴𝒴\alpha^{\bullet}\circ\beta^{\bullet}:\mathcal{Y}^{\bullet}\to\mathcal{Y}^{\bullet} is asymptotically G{}^{*}G-homotopic to the identity on 𝒴\mathcal{Y}^{\bullet}, and βα:𝒳𝒳\beta^{\bullet}\circ\alpha^{\bullet}:\mathcal{X}^{\bullet}\to\mathcal{X}^{\bullet} is asymptotically G{}^{*}G-homotopic to the identity on 𝒳\mathcal{X}^{\bullet}. Then Ha(𝒳)\operatorname{H}_{a}^{\bullet}(\mathcal{X}^{\bullet}) is isomorphic to Ha(𝒴)\operatorname{H}_{a}^{\bullet}(\mathcal{Y}^{\bullet}). In this case, the maps α\alpha^{\bullet} and β\beta^{\bullet} are called asymptotic G{}^{*}G-homotopy equivalences between 𝒳\mathcal{X}^{\bullet} and 𝒴\mathcal{Y}^{\bullet}.

Given an asymptotic G{}^{*}G-cochain complex (𝒳,d)(\mathcal{X}^{\bullet},d^{\bullet}) and another asymptotic Banach G{}^{*}G-module \mathcal{E}, we can use the functors dndnd^{n}\to d^{n}_{*} to construct an asymptotic G{}^{*}G-cochain complex

0{0}(𝒳0×){\mathcal{B}(\mathcal{X}^{0}\times\mathcal{E})}(𝒳1×){\mathcal{B}(\mathcal{X}^{1}\times\mathcal{E})}(𝒳2×){\mathcal{B}(\mathcal{X}^{2}\times\mathcal{E})}(𝒳3×){\mathcal{B}(\mathcal{X}^{3}\times\mathcal{E})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

which we shall denote (𝒳,)\mathcal{B}(\mathcal{X}^{\bullet},\mathcal{E}). Combining the functorial properties of kkk\to k_{*} and σσ\sigma\to\sigma^{*} and the fact that they respect the structure of the asymptotic G{}^{*}G-representations, we get the following:

Proposition 4.1.13.

Let kk^{\bullet} be an asymptotic G{}^{*}G-morphism between the asymptotic G{}^{*}G-cochain complexes (𝒳)(\mathcal{X}^{\bullet})^{\sharp} and (𝒴)(\mathcal{Y}^{\bullet})^{\sharp}, and \mathcal{E} be an asymptotic Banach G{}^{*}G-module. Then kk^{\bullet}_{*} is an asymptotic G{}^{*}G-morphism between the asymptotic G{}^{*}G-cochain complexes (𝒳×)\mathcal{B}(\mathcal{X}^{\bullet}\times\mathcal{E}) and (𝒴×)\mathcal{B}(\mathcal{Y}^{\bullet}\times\mathcal{E}).

Note that while we cannot conclude anything about Ha((𝒳,))\operatorname{H}_{a}^{\bullet}(\mathcal{B}(\mathcal{X}^{\bullet},\mathcal{E})) from Ha(𝒳)\operatorname{H}_{a}^{\bullet}(\mathcal{X}^{\bullet}), we can still use Proposition 4.1.12 to show that:

Lemma 4.1.14.

Let k:(𝒳)(𝒴)k^{\bullet}:(\mathcal{X}^{\bullet})^{\sharp}\to(\mathcal{Y}^{\bullet})^{\sharp} and j:(𝒴)(𝒳)j^{\bullet}:(\mathcal{Y}^{\bullet})^{\sharp}\to(\mathcal{X}^{\bullet})^{\sharp} be asymptotic G{}^{*}G-homotopy equivalences (as in Proposition 4.1.12) between the asymptotic G{}^{*}G-cochain complexes (𝒳)(\mathcal{X}^{\bullet})^{\sharp} and (𝒴)(\mathcal{Y}^{\bullet})^{\sharp}. Let \mathcal{E} be an asymptotic Banach G{}^{*}G-module. Then k:(𝒳×)(𝒴×)k_{*}^{\bullet}:\mathcal{B}(\mathcal{X}\times\mathcal{E})\to\mathcal{B}(\mathcal{Y}\times\mathcal{E}) and j:(𝒴×)(𝒳×)j_{*}^{\bullet}:\mathcal{B}(\mathcal{Y}\times\mathcal{E})\to\mathcal{B}(\mathcal{X}\times\mathcal{E}) be asymptotic G{}^{*}G-homotopy equivalences between the asymptotic G{}^{*}G-cochain complexes (𝒳,)\mathcal{B}(\mathcal{X}^{\bullet},\mathcal{E}) and (𝒴,)\mathcal{B}(\mathcal{Y}^{\bullet},\mathcal{E}).

4.2 The LL^{\infty}-cohomology and Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V})

We now come to our definition of the asymptotic cohomology of the group GG with coefficients in a dual asymptotic Banach G{}^{*}G-module (π,𝒱)(\pi,\mathcal{V}), where 𝒱={Vn}𝒰\mathcal{V}=\{V_{n}\}_{\mathcal{U}}. We shall first define it explicitly, and then relate it to the notions discussed in §4.1 to derive further results.
Our objects of interest shall be ultraproducts of LL^{\infty}-spaces. Since these are not spaces of functions but spaces of equivalence classes of functions upto null sets, we now quickly review some notions regarding its essential image. For m0m\geq 0 and dual Banach space VV (with predual VV^{\flat}, and equipped with the weak-* topology), Lw(Gm,V)L_{w*}^{\infty}(G^{m},V) denotes the Banach space of (equivalence classes of) essentially bounded weak-* measurable maps from GmG^{m} to VV with (essential) supremum norm \|\cdot\|.
For a weak-* measurable function f:GmVf:G^{m}\to V, the set f1(Ou)f^{-1}(O_{u}) is measurable in GmG^{m}. Define the essential image Im(f)Im(f) to be

Im(f)={uV|μ(f1(Ou))>0 for every weak- neighborhood Ou of v}Im(f)=\{u\in V\;|\;\mu(f^{-1}(O_{u}))>0\text{ for every weak-$*$ neighborhood }O_{u}\mbox{ of }v\}

where μ\mu is the Haar measure on GmG^{m}. We extend this notion to function classes in Lw(Gm,V)L_{w*}^{\infty}(G^{m},V) as well. Let fLw(Gm,V)f\in L_{w*}^{\infty}(G^{m},V).

  • For any functions f1,f2f_{1},f_{2} in the class of ff, Im(f1)=Im(f2)Im(f_{1})=Im(f_{2}). That is, essential image is independent of the representative of ff, allowing us to define Im(f)Im(f) as the essential image of any representative in its class.

  • Let f0f_{0} be a representative of ff, then Im(f)range(f0)¯Im(f)\subseteq\overline{range(f_{0})} (for uIm(f)u\in Im(f), any OuO_{u} intersects with range(f0)range(f_{0})).

  • If uIm(f)u\not\in Im(f) then there is a representative f0f_{0} such that urange(f0)¯u\not\in\overline{range(f_{0})} (in this case, there is a neighborhood OuO_{u} such that f1(Ou)f^{-1}(O_{u}) has measure 0, and so, one may redefine ff on f1(Ou)f^{-1}(O_{u}) to avoid OuO_{u}).

  • Im(f)={range(g)¯|g is a representative off}Im(f)=\bigcap\{\overline{range(g)}\;|\;g\mbox{ is a representative of}f\}.

  • There is a representative f0f_{0} of ff such that range(f0)Im(f)range(f_{0})\subseteq Im(f) (this is because Im(f)Im(f), being a norm bounded closed subset of VV, is second countable, and so f1(Im(f))f^{-1}(Im(f)) is measurable and of the full measure).

Formally, we shall call fLw(Gm,V)f\in L_{w*}^{\infty}(G^{m},V) essentially constant if there exists vVv\in V with Im(f)={v}Im(f)=\{v\}.
Consider the internal Banach space

((G)m,𝒱)𝒰Lw(Gm,Vn)\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{V}\right)\coloneq\prod_{\mathcal{U}}L_{w*}^{\infty}(G^{m},V_{n})

For f={fn}𝒰((G)m,𝒱)f=\{f_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{V}\right), we denote by f\|f\| the hyperreal {fn}𝒰𝐑\{\|f_{n}\|\}_{\mathcal{U}}\in{}^{*}\mathbf{R} and by Im(f)Im(f) the subset {Im(fn)}𝒰𝒱\{Im(f_{n})\}_{\mathcal{U}}\subseteq\mathcal{V}. Note that the external subset b((G)m,𝒱)\mathcal{L}^{\infty}_{b}(({}^{*}G)^{m},\mathcal{V}) is the subset of function classes f={fn}𝒰f=\{f_{n}\}_{\mathcal{U}} such that Im(f)𝒱bIm(f)\subseteq\mathcal{V}_{b}, while while inf((G)m,𝒱)\mathcal{L}^{\infty}_{inf}(({}^{*}G)^{m},\mathcal{V}) is the subset of function classes f={fn}𝒰f=\{f_{n}\}_{\mathcal{U}} such that Im(f)𝒱infIm(f)\subseteq\mathcal{V}_{inf}. Observe that Im(f)Im(f) is an internal subset of 𝒱\mathcal{V}, while 𝒱b\mathcal{V}_{b} and 𝒱inf\mathcal{V}_{inf} are external.

Claim 4.2.1.

For f={fn}𝒰b((G)m,𝒱)f=\{f_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}_{b}(({}^{*}G)^{m},\mathcal{V}), Im(f)𝒱bGIm(f)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G} iff there exists an internal function f={fn}𝒰f^{\prime}=\{f^{\prime}_{n}\}_{\mathcal{U}} with fnf^{\prime}_{n} in the class of fnf_{n} for n𝒰n\in\mathcal{U} such that range(f)={range(fn)}𝒰𝒱bGrange(f^{\prime})=\{range(f_{n}^{\prime})\}_{\mathcal{U}}\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}.

Proof.

Suppose Im(f)𝒱bGIm(f)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. Then since there exists an internal representative ff^{\prime} with range(f)Im(f)range(f^{\prime})\subseteq Im(f), we have range(f)𝒱bGrange(f^{\prime})\subseteq\mathcal{V}_{b}^{\sim{}^{*}G} as well for this ff^{\prime}.
Conversely, suppose there exists an internal representative function ff^{\prime} with range(f)𝒱bGrange(f^{\prime})\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. Let ϵ={ϵn}𝒰\epsilon=\{\epsilon_{n}\}_{\mathcal{U}} where ϵn=ess.supπn(g)vv|vrange(fn),gG)\epsilon_{n}=ess.sup\left\|\pi_{n}(g)v-v\|\;|\;v\in range(f^{\prime}_{n}),g\in G\right). Note that ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf} since range(f)𝒱bGrange(f^{\prime})\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. This allows us to express the internal subset range(f)range(f^{\prime}) as a subset of the internal set E={En}𝒰E=\{E_{n}\}_{\mathcal{U}} where

En={vVn|vfn and gG,πn(g)vvϵn}E_{n}=\{v\in V_{n}\;|\;\|v\|\leq\|f^{\prime}_{n}\|\text{ and }\forall g\in G,\|\pi_{n}(g)v-v\|\leq\epsilon_{n}\}

That is, range(f)E𝒱bGrange(f^{\prime})\subseteq E\subseteq\mathcal{V}_{b}^{\sim{}^{*}G} with EE being internal. Also note that for n𝒰n\in\mathcal{U}, EnE_{n} is closed in the weak-* topology. Hence Im(f)E𝒱bGIm(f)\subseteq E\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. ∎

The internal map π:G×𝒱𝒱\pi:{}^{*}G\times\mathcal{V}\to\mathcal{V} can be used to define the internal map τm:G×((G)m,𝒱)((G)m,𝒱)\tau^{m}:{}^{*}G\times\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V}) for each m0m\geq 0 defined as

(τm(g)(f))(g1,g2,,gm)π(g)f(g1g1,,g1gm)(\tau^{m}(g)(f))(g_{1},g_{2},\dots,g_{m})\coloneqq\pi(g)f(g^{-1}g_{1},\dots,g^{-1}g_{m}) (4.2)

for every gGg\in{}^{*}G and every g1,,gmGg_{1},\dots,g_{m}\in{}^{*}G. This internal map τm:G×((G)m,𝒱)((G)m,𝒱)\tau^{m}:{}^{*}G\times\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V}) induces an action τ~m\tilde{\tau}^{m} of G{}^{*}G on the ultralimit space ~((G)m,𝒱)\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{V}). In particular, this means that (τm,((G)m,𝒱))(\tau^{m},\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})) is an asymptotic Banach G{}^{*}G-module.
For simplicity, we shall denote the induced action of gGg\in{}^{*}G on f~~((G)m,𝒱)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{V}) simply by gf~g\cdot\tilde{f}. We shall denote by ((G)m,𝒱)G\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})^{\sim{}^{*}G} the set of elements f((G)m,𝒱)f\in\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V}) such that f~~((G)m,𝒱)G\tilde{f}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{V})^{{}^{*}G}. Such an element ff shall be referred to as asymptotically G{}^{*}G-equivariant.

Remark 4.2.2.

Later we shall also deal with (truly) G{}^{*}G-invariantmaps f(Gm,𝒱)f\in\mathcal{L}^{\infty}({}^{*}G^{m},\mathcal{V}) with an internal right action of GG on the domain. In this case, we mean that f(x)=f(xg)f(x)=f(xg) for every gGg\in{}^{*}G, and for x={xn}𝒰{Xn}𝒰Gx=\{x_{n}\}_{\mathcal{U}}\subseteq\{X_{n}\}_{\mathcal{U}}\subseteq{}^{*}G where XnGX_{n}\subseteq G is co-null in GG. We shall refer to this as f(x)=f(xg)f(x)=f(xg) for every gGg\in{}^{*}G, and almost every xGx\in{}^{*}G, for convenience.

For each m0m\geq 0, define the internal map

dm:((G)m,𝒱)((G)m+1,𝒱)d^{m}:\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}G)^{m+1},\mathcal{V})
dmf(g0,,gm)j=0m(1)jf(g0,,gj^,,gm)d^{m}f(g_{0},\dots,g_{m})\coloneq\sum\limits_{j=0}^{m}(-1)^{j}f(g_{0},\dots,\hat{g_{j}},\dots,g_{m})

It is clear that dmdm1=0d^{m}\circ d^{m-1}=0 for every m1m\geq 1. Note that dmd^{m} induces a map

d~m:~((G)m,𝒱)~((G)m+1,𝒱)\tilde{d}^{m}:\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{V})\to\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{V})

Furthermore, when we restrict to ((G)m,𝒱)G\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})^{\sim{}^{*}G},

Lemma 4.2.3.

For f((G)m,𝒱)Gf\in\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})^{\sim{}^{*}G}, dmf((G)m+1,𝒱)Gd^{m}f\in\mathcal{L}^{\infty}(({}^{*}G)^{m+1},\mathcal{V})^{\sim{}^{*}G}. That is, dd^{\bullet} is an asymptotic G{}^{*}G-morphism of asymptotic Banach G{}^{*}G-modules.

Proof.

Consider

π(g)(dmf(g0,,gm))=j=0m(1)jπ(g)f(g0,,gj^,,gm)\pi(g)(d^{m}f(g_{0},\dots,g_{m}))=\sum\limits_{j=0}^{m}(-1)^{j}\pi(g)f(g_{0},\dots,\hat{g_{j}},\dots,g_{m})

Note that since f((G)m,𝒱)Gf\in\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{V})^{\sim{}^{*}G}, π(g)f(g0,,gm^,,gm)f(gg0,,ggm)𝒱inf\pi(g)f(g_{0},\dots,\hat{g_{m}},\dots,g_{m})-f(gg_{0},\dots,gg_{m})\in\mathcal{V}_{inf}, thus implying the conclusion. ∎

Since the maps d~m:~((G)m,𝒱)~((G)m+1,𝒱)\tilde{d}^{m}:\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{V})\to\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{V}) are G{}^{*}G-equivariant for every m1m\geq 1, with d~md~m1=0\tilde{d}^{m}\circ\tilde{d}^{m-1}=0, we have an asymptotic G{}^{*}G-cochain complex

0{0}(G,𝒱){\mathcal{L}^{\infty}({}^{*}G,\mathcal{V})}((G)2,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{V})}((G)3,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{V})}((G)4,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

and the induced G{}^{*}G-cochain complex

0{0}~(G,𝒱)G{\tilde{\mathcal{L}}^{\infty}({}^{*}G,\mathcal{V})^{{}^{*}G}}~((G)2,𝒱)G{\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{2},\mathcal{V})^{{}^{*}G}}~((G)3,𝒱)G{\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{3},\mathcal{V})^{{}^{*}G}}{\dots}d~0\scriptstyle{\tilde{d}^{0}}d~1\scriptstyle{\tilde{d}^{1}}d~2\scriptstyle{\tilde{d}^{2}}d~3\scriptstyle{\tilde{d}^{3}}
Definition 4.2.4.

The asymptotic cohomology group of GG with coefficients in a dual asymptotic Banach G{}^{*}G-module 𝒱\mathcal{V}, denoted Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}), is defined to be the asymptotic cohomology of the asymptotic G{}^{*}G-cochain complex

0{0}(G,𝒱){\mathcal{L}^{\infty}({}^{*}G,\mathcal{V})}((G)2,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{V})}((G)3,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{V})}((G)4,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

Let us illustrate the above definitions in the case of a discrete group Γ\Gamma and the coefficients being the asymptotic Banach Γ{}^{*}\Gamma-module (πΓ,𝒲)(\pi_{\Gamma},\mathcal{W}) where 𝒲=𝒰𝔲kn\mathcal{W}=\prod_{\mathcal{U}}\mathfrak{u}_{k_{n}} and πΓ=πψ\pi_{\Gamma}=\pi_{\psi} is as defined in (Eq. 3.5), and relate the construction to Proposition 3.2.10. In this case, the asymptotic bounded cohomology Ha(Γ,𝒲)\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W}) is the cohomology of the complex given by

0{0}~(Γ,𝒲)Γ{\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W})^{{}^{*}\Gamma}}~((Γ)2,𝒲)Γ{\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{2},\mathcal{W})^{{}^{*}\Gamma}}~((Γ)3,𝒲)Γ{\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{3},\mathcal{W})^{{}^{*}\Gamma}}{\dots}d~0\scriptstyle{\tilde{d}^{0}}d~1\scriptstyle{\tilde{d}^{1}}d~2\scriptstyle{\tilde{d}^{2}}d~3\scriptstyle{\tilde{d}^{3}}

We shall now construct the same cohomology in another equivalent way which relates to Proposition 3.2.10. For m0m\geq 0, define an internal map

δm:((Γ)m,𝒲)((Γ)m+1,𝒲)\delta^{m}:\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W})\to\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m+1},\mathcal{W})
δm(f)(g1,,gm+1)πΓ(g1)f(g2,,gm+1)+j=1m(1)jf(g1,,gjgj+1,,gm+1)+(1)m+1f(g1,,gm)\delta^{m}(f)(g_{1},\dots,g_{m+1})\coloneqq\pi_{\Gamma}(g_{1})f(g_{2},\dots,g_{m+1})+\sum_{j=1}^{m}(-1)^{j}f(g_{1},\dots,g_{j}g_{j+1},\dots,g_{m+1})+(-1)^{m+1}f(g_{1},\dots,g_{m})

Again, δm\delta^{m} restricts to a map from b((Γ)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W}) to b((Γ)m+1,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m+1},\mathcal{W}) (which we shall continue to call δm\delta^{m}), which maps inf((Γ)m,𝒲)\mathcal{L}^{\infty}_{inf}(({}^{*}\Gamma)^{m},\mathcal{W}) to inf((Γ)m+1,𝒲)\mathcal{L}^{\infty}_{inf}(({}^{*}\Gamma)^{m+1},\mathcal{W}). Since πΓ\pi_{\Gamma} induces an asymptotic action of Γ{}^{*}\Gamma on 𝒲\mathcal{W}, the induced map

δ~m:~((Γ)m,𝒲)~((Γ)m+1,𝒲)\tilde{\delta}^{m}:\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W})\to\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m+1},\mathcal{W})
δ~m(f~)(g1,,gm+1)=g1f~(g2,,gm+1)+j=1m(1)jf~(g1,,gjgj+1,,gm+1)+(1)m+1f~(g1,,gm)\tilde{\delta}^{m}(\tilde{f})(g_{1},\dots,g_{m+1})=g_{1}\tilde{f}(g_{2},\dots,g_{m+1})+\sum_{j=1}^{m}(-1)^{j}\tilde{f}(g_{1},\dots,g_{j}g_{j+1},\dots,g_{m+1})+(-1)^{m+1}\tilde{f}(g_{1},\dots,g_{m})

is exactly the coboundary map on ~((Γ)m,𝒲)\tilde{\ell}(({}^{*}\Gamma)^{m},\mathcal{W}). Essentially, we now work with the Γ{}^{*}\Gamma-complex 𝒞\mathcal{C}^{\bullet}

0{0}𝒲~{\tilde{\mathcal{W}}}~(Γ,𝒲){\tilde{\mathcal{L}}^{\infty}({}^{*}\Gamma,\mathcal{W})}~((Γ)2,𝒲){\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{2},\mathcal{W})}~((Γ)3,𝒲){\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{3},\mathcal{W})}{\dots}δ~0\scriptstyle{\tilde{\delta}^{0}}δ~1\scriptstyle{\tilde{\delta}^{1}}δ~2\scriptstyle{\tilde{\delta}^{2}}δ~3\scriptstyle{\tilde{\delta}^{3}}

Since δ~mδ~m1=0\tilde{\delta}^{m}\cdot\tilde{\delta}^{m-1}=0, for m1m\geq 1, denote the mm-th cohomology group of this complex by Hm(𝒞)H^{m}(\mathcal{C}^{\bullet}). In this notation, it is immediate that Proposition 3.2.10 can be restated as

Theorem 4.2.5.

Suppose H2(𝒞)=0H^{2}(\mathcal{C}^{\bullet})=0, then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

We now conclude this subsection by showing that H(𝒞)Ha(Γ,𝒲)H^{\bullet}(\mathcal{C}^{\bullet})\cong\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W}).

Theorem 4.2.6.

For every m1m\geq 1, Hm(𝒞)Ham(Γ,𝒲)H^{m}(\mathcal{C}^{\bullet})\cong\operatorname{H}_{a}^{m}(\Gamma,\mathcal{W}). In particular, suppose Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0, then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Proof.

Consider the internal map

hm:((Γ)m+1,𝒲)((Γ)m,𝒲)h^{m}:\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m+1},\mathcal{W})\to\mathcal{L}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W})
hmf(g1,,gm)f(1,g1,g1g2,,g1g2gm)h^{m}f(g_{1},\dots,g_{m})\coloneqq f(1,g_{1},g_{1}g_{2},\dots,g_{1}g_{2}\cdots g_{m})

This is simply a reparametrization, and its restriction to b((Γ)m,𝒲)Γ\mathcal{L}^{\infty}_{b}(({}^{*}\Gamma)^{m},\mathcal{W})^{\sim{}^{*}\Gamma} induces an isomorphism

h~m:~((Γ)m+1,𝒲)Γ~((Γ)m,𝒲)\tilde{h}^{m}:\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m+1},\mathcal{W})^{{}^{*}\Gamma}\to\tilde{\mathcal{L}}^{\infty}(({}^{*}\Gamma)^{m},\mathcal{W})

such that the homogenous coboundary map d~m\tilde{d}^{m} translates to the coboundary map δ~m\tilde{\delta}^{m} thus making Ham(Γ,𝒲)H_{a}^{m}(\Gamma,\mathcal{W}) canonically isomorphic to Hm(𝒞)H^{m}(\mathcal{C}^{\bullet}). ∎

Remark 4.2.7.

This complex 𝒞\mathcal{C}^{\bullet} can be thought of as the bar resolution in our context, where we work with an inhomogenous differential map, and the proof of Theorem 4.2.6 goes through to show that in general, Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) can be computed using an inhomogenous cochain complex just as with Γ\Gamma. We shall work with this bar resolution for the rest of this subsection.

We conclude this subsection with some results on Ha(G,𝒱)\operatorname{H}_{a}(G,\mathcal{V}) when the map π:G×𝒱𝒱\pi:{}^{*}G\times\mathcal{V}\to\mathcal{V} is an internal trivial action of G{}^{*}G on 𝒱\mathcal{V} (that is, for every gGg\in{}^{*}G and every v𝒱v\in\mathcal{V}, π(g)v=v\pi(g)v=v). Such an asymptotic Banach G{}^{*}G-module shall be referred to as a trivial Banach G{}^{*}G-module.
We first recall the following facts about the vanishing moduli of GG which are implicit in [35] and clarified further in [42, Definition 2.5] and [22, Lemma 4.12]:

Theorem 4.2.8 ([35][42][22]).

Let VV be trivial dual Banach GG-module.

  • Suppose Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0. Then there exists a constant C1>0C_{1}>0 such that for every (inhomogenous) 22-cocycle αLw(G2,V)\alpha^{\prime}\in L^{\infty}_{w*}(G^{2},V), there exists an (inhomogenous) 11-cochain βLw(G,V)\beta\in L^{\infty}_{w*}(G,V) such that α=δ1β\alpha^{\prime}=\delta^{1}\beta and βC1α\|\beta\|\leq C_{1}\|\alpha\|.

  • Suppose Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff. Then there exists a constant C2>0C_{2}>0 such that for every (inhomogenous) 22-cochain αLw(G2,V)\alpha\in L^{\infty}_{w*}(G^{2},V), there exists an (inhomogenous) 22-cocycle αLw(G2,V)\alpha^{\prime}\in L^{\infty}_{w*}(G^{2},V) such that ααC2δ2α\|\alpha-\alpha^{\prime}\|\leq C_{2}\|\delta^{2}\alpha\|.

We can now use Theorem 4.2.8 for V=𝐑V=\mathbf{R} to show that:

Proposition 4.2.9.

Suppose Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff. Then Ha2(G,𝐑)=0\operatorname{H}_{a}^{2}(G,{}^{*}\mathbf{R})=0

Proof.

Let α={αn}𝒰b((G)2,𝐑)\alpha=\{\alpha_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}_{b}(({}^{*}G)^{2},{}^{*}\mathbf{R}) with δ2αinf((G)2,𝐑)\delta^{2}\alpha\in\mathcal{L}^{\infty}_{inf}(({}^{*}G)^{2},{}^{*}\mathbf{R}). From Theorem 4.2.8, there exist constants C1,C2>0C_{1},C_{2}>0 such that for n𝒰n\in\mathcal{U}, there exists βnL(G,𝐑)\beta_{n}\in L^{\infty}(G,\mathbf{R}) such that αnδ1βnC2δ2αn\|\alpha_{n}-\delta^{1}\beta_{n}\|\leq C_{2}\|\delta^{2}\alpha_{n}\| and βnC1δ1βn\|\beta_{n}\|\leq C_{1}\|\delta^{1}\beta_{n}\|. Setting β={βn}𝒰b(G,𝐑)\beta=\{\beta_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}_{b}({}^{*}G,{}^{*}\mathbf{R}), we see that αδ1βinf((G)2,𝐑)\alpha-\delta^{1}\beta\in\mathcal{L}^{\infty}_{inf}(({}^{*}G)^{2},{}^{*}\mathbf{R}), to conclude that Ha2(G,𝐑)=0\operatorname{H}_{a}^{2}(G,{}^{*}\mathbf{R})=0

To extend this lemma to show vanishing of Ha2(G,𝒱)\operatorname{H}_{a}^{2}(G,\mathcal{V}) for more general trivial Banach G{}^{*}G-modules 𝒱\mathcal{V}, we would need the constants C1C_{1} and C2C_{2} in Theorem 4.2.8 to work uniformly across all trivial Banach GG-modules.

Lemma 4.2.10.

Suppose Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0 and Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff for every trivial dual Banach GG-module VV.

  • There exists a constant C1>0C_{1}>0 such that for every trivial dual Banach GG-module WW and a (inhomogenous) 22-cocycle αLw(G2,W)\alpha^{\prime}\in L^{\infty}_{w*}(G^{2},W), there exists an (inhomogenous) 11-cochain βLw(G,W)\beta\in L^{\infty}_{w*}(G,W) such that α=δ1β\alpha^{\prime}=\delta^{1}\beta and βC1α\|\beta\|\leq C_{1}\|\alpha\|.

  • There exists a constant C2>0C_{2}>0 such that for any trivial for every trivial dual Banach GG-module WW and (inhomogenous) 22-cochain αLw(G2,W)\alpha\in L^{\infty}_{w*}(G^{2},W), there exists an (inhomogenous) 22-cocycle αLw(G2,W)\alpha^{\prime}\in L^{\infty}_{w*}(G^{2},W) such that ααC2δ2α\|\alpha-\alpha^{\prime}\|\leq C_{2}\|\delta^{2}\alpha\|.

Proof.

Suppose, for the sake of contradiction, that there exists a sequence {Wn}n𝐍\{W_{n}\}_{n\in\mathbf{N}} of dual Banach spaces (all with trivial actions of GG), and 22-cocyles {αnLw(G2,Wn)}n𝐍\{\alpha_{n}\in L^{\infty}_{w*}(G^{2},W_{n})\}_{n\in\mathbf{N}} with αn=1\|\alpha_{n}\|=1 for every n1n\geq 1, such that for every sequence {βnLw(G,Wn)}n𝐍\{\beta_{n}\in L^{\infty}_{w*}(G,W_{n})\}_{n\in\mathbf{N}} with δ1βn=αn\delta^{1}\beta_{n}=\alpha_{n} for every n1n\geq 1, we have βnn\|\beta_{n}\|\geq n. Consider the \ell^{\infty} direct sum WnWnW\coloneq\oplus_{n}W_{n} (which is a dual Banach space) and the 22-cocycle αnαn\alpha\coloneq\oplus_{n}\alpha_{n} with α=1\|\alpha\|=1. Then since Hb2(G,W)=0\operatorname{H}_{b}^{2}(G,W)=0, that there exists βLw(G,W)\beta\in L^{\infty}_{w*}(G,W) with δ1β=α\delta^{1}\beta=\alpha, and let β=C1\|\beta\|=C_{1}. Note that the projections of β\beta on WnW_{n} for n>C1n>C_{1} would give us βnLw(G,Wn)\beta_{n}\in L^{\infty}_{w*}(G,W_{n}) with δ1βn=αn\delta^{1}\beta_{n}=\alpha_{n} and βnC1\|\beta_{n}\|\leq C_{1}, contradicting our assumption. The proof of the second item is similar. ∎

We now use the universality of the constants C1C_{1} and C2C_{2} in Lemma 4.2.10 to show the vanishing of Ha2(G,𝒱)\operatorname{H}_{a}^{2}(G,\mathcal{V}) for a trivial Banach G{}^{*}G-module, just like in Proposition 4.2.9.

Proposition 4.2.11.

Suppose Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0 and Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff for every trivial dual Banach GG-module VV. Then Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(G,\mathcal{V})=0 for every trivial asymptotic Banach G{}^{*}G-module 𝒱\mathcal{V}.

While the assumption that Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0 and Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff for every trivial dual Banach GG-module VV, a priori, seems stronger than the assumption that Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff, we now show that the former is actually implied by the latter. Recall that a Banach space VV is said to be injective if for any Banach space embedding VWV\subset W, VV has a complement in WW.

Proposition 4.2.12.

Suppose Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff. Then the image δ2L(G2)\delta^{2}L^{\infty}(G^{2}) in L(G3)L^{\infty}(G^{3}) is injective.

Proof.

Note that the image δ1L(G)\delta^{1}L^{\infty}(G) in L(G2)L^{\infty}(G^{2}) is closed, since Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0. In particular, this image is isomorphic to L(G)/𝐑L^{\infty}(G)/\mathbf{R}, which is injective since L(G)L^{\infty}(G) and 𝐑\mathbf{R} are injective. Next, since Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff, the image δ2L(G2)\delta^{2}L^{\infty}(G^{2}) is closed, and isomorphic to L2(G)/δ1L(G)L^{2}(G)/\delta^{1}L^{\infty}(G), which is injective since L(G2)L^{\infty}(G^{2}) and δ1L(G)\delta^{1}L^{\infty}(G) are injective. ∎

Lemma 4.2.13.

Suppose Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff. Then for any trivial dual Banach GG-module VV, Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0 and Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff.

Proof.

Note (Eq. 4.3 and Eq. 4.1) that Lw(Gj,V)L(V,L(Gj))L^{\infty}_{w*}(G^{j},V)\cong L\left(V^{\flat},L^{\infty}(G^{j})\right). In this identification, the differential is just given by applying the scalar differential δ\delta to the image L(Gj)L^{\infty}(G^{j}). It follows that the complementing map given by Proposition 4.2.12 provides a complementing map for the image δ2Lw(G2,V)\delta^{2}L^{\infty}_{w*}(G^{2},V) of Lw(G2,V)L^{\infty}_{w*}(G^{2},V) in Lw(G3,V)L^{\infty}_{w*}(G^{3},V). In particular this image is closed and hence Hb3(G,V)\operatorname{H}_{b}^{3}(G,V) is Hausdorff. Similarly, we can show that Hb2(G,V)=0\operatorname{H}_{b}^{2}(G,V)=0 as well. ∎

Combining Lemma 4.2.13 and Proposition 4.2.11, we conclude that

Corollary 4.2.14.

Suppose Hb2(G,𝐑)=0\operatorname{H}_{b}^{2}(G,\mathbf{R})=0 and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff. Then Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(G,\mathcal{V})=0 for any trivial dual Banach G{}^{*}G-module 𝒱\mathcal{V}.

This property of vanishing Hb2(G,𝐑)\operatorname{H}_{b}^{2}(G,\mathbf{R}) and Hausdorffness of Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) will be very useful to us in Section 6.3, and in Section 6.3 we shall study this property (called the \say2½-property) in more detail.

4.3 Amenable Actions and Cohomology of Subgroups

Recall that we had presented the asymptotic cohomology of GG with coefficients in the dual asymptotic Banach G{}^{*}G-module 𝒱\mathcal{V} as asymptotic cohomology of the complex

0{0}(G,𝒱){\mathcal{L}^{\infty}({}^{*}G,\mathcal{V})}((G)2,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{V})}((G)3,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{V})}((G)4,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

In this subsection, we shall relate the functorial descriptions in §4.1 with Definition 4.2.4 to obtain other complexes that can also be used to compute the same cohomology.
The first step is to interpret Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) in a way that fits with Proposition 4.1.13. For this, we recall the following classical fact that follows from the Dunford-Pettis theorem: for a (standard) dual Banach space EE,

B(L1(Gm)×E)L(L1(Gm),E)Lw(Gm,E)B(L^{1}(G^{m})\times E)\cong L\left(L^{1}(G^{m}),E^{\sharp}\right)\cong L^{\infty}_{w*}(G^{m},E^{\sharp}) (4.3)

In fact, this goes through even when we consider the analogous asymptotic Banach G{}^{*}G-modules. Denoting the ultrapower L1(Gm){}^{*}L^{1}(G^{m}) by 1(Gm)\mathcal{L}^{1}(G^{m}), we note that ((G)m)\mathcal{L}^{\infty}(({}^{*}G)^{m}) is a dual asymptotic G{}^{*}G-module with predual 1((G)m)\mathcal{L}^{1}(({}^{*}G)^{m}). In fact, it is more than just an asymptotic G{}^{*}G-representation: G{}^{*}G actually has a (true) internal action on 1((G)m)\mathcal{L}^{1}(({}^{*}G)^{m}) and its dual ((G)m)\mathcal{L}^{\infty}(({}^{*}G)^{m}).
This allows us to construct Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) starting from the GG-cochain complex

0{0}L(G){L^{\infty}(G)}L(G2){L^{\infty}(G^{2})}L(G3){L^{\infty}(G^{3})}L(G4){L^{\infty}(G^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

extending it internally to the asymptotic G{}^{*}G-cochain complex (which, in this case, turns out to be a (true) internal cochain complex)

0{0}(G){\mathcal{L}^{\infty}({}^{*}G)}((G)2){\mathcal{L}^{\infty}(({}^{*}G)^{2})}((G)3){\mathcal{L}^{\infty}(({}^{*}G)^{3})}((G)4){\mathcal{L}^{\infty}(({}^{*}G)^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

and finally using the covariant functor ddd\mapsto d_{*} for the coefficient module 𝒱\mathcal{V}^{\flat} (the predual of 𝒱\mathcal{V}) to get

0{0}(1(G)×𝒱){\mathcal{B}(\mathcal{L}^{1}({}^{*}G)\times\mathcal{V}^{\flat})}(1((G)2)×𝒱){\mathcal{B}(\mathcal{L}^{1}(({}^{*}G)^{2})\times\mathcal{V}^{\flat})}(1((G)3)×𝒱){\mathcal{B}(\mathcal{L}^{1}(({}^{*}G)^{3})\times\mathcal{V}^{\flat})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}

which, in turn, is seen to be the same as

0{0}(G,𝒱){\mathcal{L}^{\infty}({}^{*}G,\mathcal{V})}((G)2,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{V})}((G)3,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{V})}((G)4,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

The advantage of this reformulation of Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) is that we can use Lemma 4.1.14 to construct other asymptotic G{}^{*}G-cochain complexes that compute the same cohomology Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}). In view of this approach, we review some definitions and facts from [37]:

Definition 4.3.1.

Let SS be a regular GG-space. A conditional expectation 𝔪:L(G×S)L(S)\mathfrak{m}:L^{\infty}(G\times S)\to L^{\infty}(S) is a measurable norm one linear map such that

  • 𝔪(1G×S)=1S\mathfrak{m}(1_{G\times S})=1_{S}

  • For every fL(G×S)f\in L^{\infty}(G\times S) and every measurable set ASA\subset S, 𝔪(f1G×A)=𝔪(f)1A\mathfrak{m}(f\cdot 1_{G\times A})=\mathfrak{m}(f)\cdot 1_{A}.

The GG-action on SS is said to be Zimmer amenable if there exists a GG-equivariant conditional expectation 𝔪:L(G×S)L(S)\mathfrak{m}:L^{\infty}(G\times S)\to L^{\infty}(S).

What we shall use is the following consequence of Zimmer amenability:

Proposition 4.3.2.

Let SS be a regular GG-space with a Zimmer-amenable action of GG. Then there exists a GG-homotopy equivalence between the GG-cochain complexes

0{0}L(G){L^{\infty}(G)}L(G2){L^{\infty}(G^{2})}L(G3){L^{\infty}(G^{3})}L(G4){L^{\infty}(G^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

and

0{0}L(S){L^{\infty}(S)}L(S2){L^{\infty}(S^{2})}L(S3){L^{\infty}(S^{3})}L(S4){L^{\infty}(S^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

Extending this internally, if SS a regular GG-space with a Zimmer-amenable action of GG, then this gives us an asymptotic G{}^{*}G-homotopy equivalences between the asymptotic G{}^{*}G-cochain complexes

0{0}(G){\mathcal{L}^{\infty}({}^{*}G)}((G)2){\mathcal{L}^{\infty}(({}^{*}G)^{2})}((G)3){\mathcal{L}^{\infty}(({}^{*}G)^{3})}((G)4){\mathcal{L}^{\infty}(({}^{*}G)^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

and

0{0}(S){\mathcal{L}^{\infty}({}^{*}S)}((S)2){\mathcal{L}^{\infty}(({}^{*}S)^{2})}((S)3){\mathcal{L}^{\infty}(({}^{*}S)^{3})}((S)4){\mathcal{L}^{\infty}(({}^{*}S)^{4})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

Now, since (L1(Sm))=L(Sm)(L^{1}(S^{m}))^{\sharp}=L^{\infty}(S^{m}) and B(L1(Sm)×E)L(L1(Sm),E)Lw(Sm,E)B(L^{1}(S^{m})\times E)\cong L\left(L^{1}(S^{m}),E^{\sharp}\right)\cong L^{\infty}_{w*}(S^{m},E^{\sharp}), applying Lemma 4.1.14 with 𝒳=1((G))\mathcal{X}^{\bullet}=\mathcal{L}^{1}(({}^{*}G)^{\bullet}), 𝒴=1((S))\mathcal{Y}^{\bullet}=\mathcal{L}^{1}(({}^{*}S)^{\bullet}) and =𝒱\mathcal{E}=\mathcal{V}^{\flat}, we get

Theorem 4.3.3.

Let SS be a regular GG-space with a Zimmer-amenable action of GG. Then Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) can be computed as the asymptotic cohomology of the asymptotic G{}^{*}G-cochain complex

0{0}(S,𝒱){\mathcal{L}^{\infty}({}^{*}S,\mathcal{V})}((S)2,𝒱){\mathcal{L}^{\infty}(({}^{*}S)^{2},\mathcal{V})}((S)3,𝒱){\mathcal{L}^{\infty}(({}^{*}S)^{3},\mathcal{V})}((S)4,𝒱){\mathcal{L}^{\infty}(({}^{*}S)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

We now state some observations that follow immediately from Theorem 4.3.3, which we shall use later in Proposition 6.1.12 in §6. Let SS and TT be two Zimmer-amenable regular GG-spaces, so that by Theorem 4.3.3, we have an asymptotic G{}^{*}G-homotopy equivalence between the asymptotic G{}^{*}G-cochain complexes ((S),𝒱)\mathcal{L}^{\infty}(({}^{*}S)^{\bullet},\mathcal{V}) and ((T),𝒱)\mathcal{L}^{\infty}(({}^{*}T)^{\bullet},\mathcal{V}) given by k:((S),𝒱)((T),𝒱)k^{\bullet}:\mathcal{L}^{\infty}(({}^{*}S)^{\bullet},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}T)^{\bullet},\mathcal{V}) and j:((T),𝒱)((S),𝒱)j^{\bullet}:\mathcal{L}^{\infty}(({}^{*}T)^{\bullet},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}S)^{\bullet},\mathcal{V}).

  • Let ωb((T)m+1,𝒱)G\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}T)^{m+1},\mathcal{V})^{\sim{}^{*}G} be an asymptotic mm-cocycle such that Im(ω)𝒱bGIm(\omega)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. Then Im(jmω)𝒱bGIm(j^{m}\omega)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. This follows from the construction of jmj^{m} and Proposition 4.1.2.

  • Let ωb((S)m+1,𝒱)G\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m+1},\mathcal{V})^{\sim{}^{*}G} be an asymptotic mm-cocycle such that Im(kmω)𝒱bGIm(k^{m}\omega)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. Then, setting ω1=jmkmωb((S)m+1,𝒱)G\omega_{1}=j^{m}k^{m}\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m+1},\mathcal{V})^{\sim{}^{*}G}, note that Im(ω1)𝒱bGIm(\omega_{1})\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. Furthermore, since kk^{\bullet} and jj^{\bullet} are asymptotic G{}^{*}G-homotopy equivalences, ω\omega and ω1\omega_{1} are asymptotically cohomologous, that is, there exists αb((S)m,𝒱)G\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{V})^{\sim{}^{*}G} such that

    ω~ω~1=d~mα~\tilde{\omega}-\tilde{\omega}_{1}=\tilde{d}^{m}\tilde{\alpha} (4.4)

Theorem 4.3.3 has the following two immediate corollaries which we shall use in §5 and §6. The first of these is an asymptotic analogue of the classical result that amenable groups have vanishing bounded cohomology, and follows from the fact that if GG is amenable, then the trivial space is Zimmer-amenable for GG.

Corollary 4.3.4.

Let GG be an amenable group and 𝒱\mathcal{V} be a dual asymptotic Banach G{}^{*}G-module. Then for every n1n\geq 1, Han(G,𝒱)=0\operatorname{H}_{a}^{n}(G,\mathcal{V})=0.

The second corollary uses the fact that for a lattice Γ\Gamma in a locally compact group GG, GG is Zimmer-amenable as a regular Γ\Gamma-space. This serves as the starting point for an induction procedure to go from Γ\Gamma to GG which we describe in §5.

Corollary 4.3.5.

Let ΓG\Gamma\leq G be a lattice in a locally compact group GG, and let 𝒲\mathcal{W} be a dual asymptotic Banach Γ{}^{*}\Gamma-module. Then Ha(Γ,𝒲)\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W}) can be computed as the asymptotic cohomology of the asymptotic Γ{}^{*}\Gamma-cochain complex

0{0}(G,𝒲){\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})}((G)2,𝒲){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{W})}((G)3,𝒲){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{W})}((G)4,𝒲){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{W})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

The next corollary uses the fact that for a closed subgroup QGQ\leq G (and in particular, when Q=GQ=G), and a closed amenable subgroup PGP\leq G, the space G/PG/P is a regular QQ-space that is Zimmer-amenable for the QQ-action.

Corollary 4.3.6.

Let PGP\leq G be a closed amenable subgroup of GG, QQ be a closed subgroup of GG, and 𝒱\mathcal{V} be a dual asymptotic Banach G{}^{*}G-module. Then Ha(Q,𝒱)\operatorname{H}_{a}^{\bullet}(Q,\mathcal{V}) can be computed as the asymptotic cohomology of the asymptotic Q{}^{*}Q-cochain complex

0{0}(((G/P)),𝒱){\mathcal{L}^{\infty}(({}^{*}(G/P)),\mathcal{V})}(((G/P))2,𝒱){\mathcal{L}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})}(((G/P))3,𝒱){\mathcal{L}^{\infty}(({}^{*}(G/P))^{3},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}

The second corollary uses the fact that for a closed subgroup QGQ\leq G, the space G/PG/P is a regular QQ-space that is Zimmer-amenable for the QQ-action.

5 The Induction Module

In the previous section, we noted (in Corollary 4.3.5) that for a lattice Γ\Gamma in a Lie group GG, the cohomology Ha(Γ,𝒲)\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W}) could be computed as the asymptotic cohomology of the asymptotic Γ{}^{*}\Gamma-cochain complex

0{0}(G,𝒱){\mathcal{L}^{\infty}({}^{*}G,\mathcal{V})}((G)2,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{2},\mathcal{V})}((G)3,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{3},\mathcal{V})}((G)4,𝒱){\mathcal{L}^{\infty}(({}^{*}G)^{4},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}d4\scriptstyle{d^{4}}

We shall now apply an induction procedure to go from Γ{}^{*}\Gamma-equivariance to G{}^{*}G-equivariance, and shall construct an asymptotic Banach G{}^{*}G-module 𝒱\mathcal{V} such that Ha(Γ,𝒲)Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(\Gamma,\mathcal{W})\cong\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}).
In §5.1, we begin by studying useful properties of an intermediary structure b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} that shall arise in the induction procedure worked out in §5.2. This structure is, upto infinitesimals, equal to the induced module (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) that we shall use, but has the additional useful feature of being equipped with a (true) internal action of G{}^{*}G. A 11-cohomology argument is used to pass between asymptotically equivariant maps in (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) and (truly) equivariant maps in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, which is a result we shall often use, especially in §6.1. Finally, the induction procedure is described in §5.2.

5.1 The G{}^{*}G-action on b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}

Recall that the internal Banach space 𝒲=𝒰𝔲(kn)\mathcal{W}=\prod_{\mathcal{U}}\mathfrak{u}(k_{n}) came with an internal map πΓ:Γ×𝒲𝒲\pi_{\Gamma}:{}^{*}\Gamma\times\mathcal{W}\to\mathcal{W} such that this map induced an action of Γ{}^{*}\Gamma on 𝒲~=𝒲b/𝒲inf\tilde{\mathcal{W}}=\mathcal{W}_{b}/\mathcal{W}_{inf}. Let SS be an Zimmer-amenable regular GG-space. For m0m\geq 0, consider the internal Banach space

((S)m,𝒲)\mathcal{L}^{\infty}(({}^{*}S)^{m},\mathcal{W})

equipped with the following internal G{}^{*}G-action: for gGg\in{}^{*}G, f((S)m,𝒲)f\in\mathcal{L}^{\infty}(({}^{*}S)^{m},\mathcal{W}),

(gf)(x1,,xm)=f(g1x1,,g1xm)(g\cdot f)(x_{1},\dots,x_{m})=f(g^{-1}x_{1},\dots,g^{-1}x_{m})

for x1,,xmSx_{1},\dots,x_{m}\in{}^{*}S. Clearly b((S)m,𝒲)\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W}) is invariant with respect to this G{}^{*}G-action, and induces a G{}^{*}G-action on L~((S)m,𝒲)\tilde{L}^{\infty}(({}^{*}S)^{m},\mathcal{W}).
Consider the subsets b((S)m,𝒲)G\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{{}^{*}G} of bounded G{}^{*}G-fixed points of this action, and the subset b((S)m,𝒲)G\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G} of bounded asymptotically G{}^{*}G-fixed points. Clearly,

b((S)m,𝒲)G+inf((S)m,𝒲)b((S)m,𝒲)G\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{{}^{*}G}+\mathcal{L}^{\infty}_{inf}(({}^{*}S)^{m},\mathcal{W})\subseteq\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G}

We shall now show that the containment goes through in the other direction too. That is,

Lemma 5.1.1.

For every fb((S)m,𝒲)Gf\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G}, there exists βinf((S)m,𝒲)\beta\in\mathcal{L}^{\infty}_{inf}(({}^{*}S)^{m},\mathcal{W}) such that fβb((S)m,𝒲)Gf-\beta\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{{}^{*}G}. Moreover, the map fβf\mapsto\beta is induced from an internal map from ((S)m,𝒲)\mathcal{L}^{\infty}(({}^{*}S)^{m},\mathcal{W}) to itself.

Proof.

Consider the internal map α:G((S)m,𝒲)\alpha:{}^{*}G\to\mathcal{L}^{\infty}(({}^{*}S)^{m},\mathcal{W}) defined as α(g)gff\alpha(g)\coloneq g\cdot f-f. Note that since fb((S)m,𝒲)Gf\in\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G}, Im(α)inf((S)m,𝒲)Im(\alpha)\subseteq\mathcal{L}^{\infty}_{inf}(({}^{*}S)^{m},\mathcal{W}) (hence α𝐑inf\|\alpha\|\in{}^{*}\mathbf{R}_{inf}). Let α={αn}𝒰\alpha=\{\alpha_{n}\}_{\mathcal{U}} where for every n𝐍n\in\mathbf{N}, αn:GLw(Sm,𝔲(kn))\alpha_{n}:G\to L^{\infty}_{w*}(S^{m},\mathfrak{u}(k_{n})). Observe that each such αn\alpha_{n} is a bounded function that satisfies the (inhomogenous) 11-cocycle condition. That is, αnHb1(G,Lw(Sm,𝔲(kn)))\alpha_{n}\in\operatorname{H}_{b}^{1}\left(G,L^{\infty}_{w*}(S^{m},\mathfrak{u}(k_{n}))\right). From [37], we know that Lw(Sm,𝔲(kn))L^{\infty}_{w*}(S^{m},\mathfrak{u}(k_{n})) is a relatively injective Banach GG-module, and hence Hb1(G,Lw(Sm,𝔲(kn)))=0\operatorname{H}_{b}^{1}\left(G,L^{\infty}_{w*}(S^{m},\mathfrak{u}(k_{n}))\right)=0. In fact, there exists a constant CC (independent of nn) such that for every n𝐍n\in\mathbf{N}, there exists βnLw(Sm,𝔲(kn))\beta_{n}\in L^{\infty}_{w*}(S^{m},\mathfrak{u}(k_{n})) with αn(g)=gβnβn\alpha_{n}(g)=g\cdot\beta_{n}-\beta_{n}, with βnCαn\|\beta_{n}\|\leq C\|\alpha_{n}\|. Set β={βn}𝒰\beta=\{\beta_{n}\}_{\mathcal{U}} so that for gGg\in{}^{*}G, α(g)=gββ\alpha(g)=g\cdot\beta-\beta implying that g(fβ)=fβg\cdot(f-\beta)=f-\beta. Since βCα\|\beta\|\leq C\|\alpha\|, βinf((S)m,𝒲)\beta\in\mathcal{L}^{\infty}_{inf}(({}^{*}S)^{m},\mathcal{W}). Note that the correspondence fβf\mapsto\beta is internal by construction. ∎

A special case of particular interest to us is the internal Banach space (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}). This space comes equipped with an internal G{}^{*}G-action and an asymptotic Γ{}^{*}\Gamma-action:

  • For gGg\in{}^{*}G and f(G,𝒲)f\in\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}), define (gf)(x)=f(xg)(g\cdot f)(x)=f(xg) for xGx\in{}^{*}G. This makes (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) an internal G{}^{*}G-representation.

  • Consider the internal map πΓ:Γ×(G,𝒲)(G,𝒲)\pi^{\prime}_{\Gamma}:{}^{*}\Gamma\times\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) defined as follows: for γΓ\gamma\in{}^{*}\Gamma and f(G,𝒲)f\in\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}), define

    (πΓ(γ)f)(x)=πΓ(γ)f(γ1x)(\pi^{\prime}_{\Gamma}(\gamma)f)(x)=\pi_{\Gamma}(\gamma)f(\gamma^{-1}x)

    where xGx\in{}^{*}G. This makes (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) into an asymptotic Banach Γ{}^{*}\Gamma-module.

Remark 5.1.2.

Note that G{}^{*}G acts internally on (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}), while Γ{}^{*}\Gamma does not act, through πΓ\pi^{\prime}_{\Gamma}, on (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}), but only induces an action on the quotient ~(G,𝒲)\tilde{\mathcal{L}}^{\infty}({}^{*}G,\mathcal{W}).

Let us restrict to the subspace b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} comprising functions that are asymptotically Γ{}^{*}\Gamma-equivariant. Note that b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} is not an internal space.

Lemma 5.1.3.

The subset b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} is invariant under the internal action of G{}^{*}G.

Consider the subspaces (b(G,𝒲)Γ)G\left(\mathcal{L}^{\infty}_{b}(G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} and (b(G,𝒲)Γ)G\left(\mathcal{L}^{\infty}_{b}(G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}. Observe that the proof of Lemma 5.1.1 goes through when restricted to asymptotically Γ{}^{*}\Gamma-equivariant elements, giving us:

Corollary 5.1.4.

For v(b(G,𝒲)Γ)Gv\in\left(\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}, there exists w(b(G,𝒲)Γ)Gw\in\left(\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} such that vwinf(G,𝒲)v-w\in\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W}). Moreover, the map vwv\mapsto w is induced from an internal map from (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) to itself.

An element f(b(G,𝒲)Γ)Gf\in\left(\mathcal{L}^{\infty}_{b}(G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} satisfies the following two conditions:

  • For every gGg\in{}^{*}G and almost every xGx\in{}^{*}G, f(xg)=f(x)f(xg)=f(x). In other words, ff is an essentially constant function, where the constant value is an element of 𝒲b\mathcal{W}_{b}.

  • For γΓ\gamma\in{}^{*}\Gamma and almost every xGx\in{}^{*}G, f(γx)γf(x)𝒲inff(\gamma x)-\gamma f(x)\in\mathcal{W}_{inf} (that is, the constant value of ff is an element of 𝒲b)Γ\mathcal{W}_{b})^{\sim{}^{*}\Gamma}).

Thus, f(b(G,𝒲)Γ)Gf\in\left(\mathcal{L}^{\infty}_{b}(G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} can be represented as an element of (𝒲b)Γ(\mathcal{W}_{b})^{\sim{}^{*}\Gamma}.

Lemma 5.1.5.

The internal map e:𝒲(G,𝒲)e:\mathcal{W}\to\mathcal{L}^{\infty}(G,\mathcal{W}) defined as e(w)(g)we(w)(g)\coloneq w for w𝒲bw\in\mathcal{W}_{b} and gGg\in{}^{*}G restricts to a bijection between 𝒲bΓ\mathcal{W}_{b}^{\sim{}^{*}\Gamma} and (b(G,𝒲)Γ)G\left(\mathcal{L}^{\infty}_{b}(G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}.

Corollary 5.1.4 and Lemma 5.1.5 together imply that, upto infinitesimals, (b(G,𝒲)Γ)G\left(\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G} can be identified with 𝒲bΓ\mathcal{W}_{b}^{\sim{}^{*}\Gamma}, and this bijection is (trivially) G{}^{*}G-equivariant and induced from an internal map between (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) and 𝒲\mathcal{W}.
We combine the results above to obtain a corollary that will be useful later in §6.

Proposition 5.1.6.

Let QQ be a closed subgroup of GG. Let fb((Q)m,(G,𝒲))Qf\in\mathcal{L}^{\infty}_{b}\left(({}^{*}Q)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)^{\sim{}^{*}Q} be such that Im(f)(b(G,𝒲)Γ)GIm(f)\subseteq(\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma})^{\sim{}^{*}G}. Then there exists βinf((Q)m,(G,𝒲))\beta\in\mathcal{L}_{inf}^{\infty}\left(({}^{*}Q)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) such that fβf-\beta is Q{}^{*}Q-fixed, and Im(fβ)(b(G,𝒲)Γ)G=𝒲bΓIm(f-\beta)\subseteq\left(\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}=\mathcal{W}_{b}^{{}^{*}\Gamma}. The map fβf\mapsto\beta is induced from an internal map from ((Q)m,(G,𝒲))\mathcal{L}^{\infty}\left(({}^{*}Q)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) to itself.

Proof.

Since Im(f)(b(G,𝒲)Γ)GIm(f)\subseteq(\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma})^{\sim{}^{*}G}, Corollary 5.1.4 gives us fb((Q)m,(G,𝒲))Qf^{\prime}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}Q)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)^{\sim{}^{*}Q} with ffinf((Q)m,(G,𝒲))f-f^{\prime}\in\mathcal{L}^{\infty}_{inf}\left(({}^{*}Q)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) and Im(f)(b(G,𝒲)Γ)G=𝒲bΓIm(f^{\prime})\subseteq(\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma})^{{}^{*}G}=\mathcal{W}_{b}^{\sim{}^{*}\Gamma}. The conclusion then follows from applying Lemma 5.1.1 to ff^{\prime}. ∎

Recall Lemma 5.1.1 where an element in b((S)m,𝒲)G\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G} was shown to be corrected to get a truly G{}^{*}G-fixed element in b((S)m,𝒲)G\mathcal{L}^{\infty}_{b}(({}^{*}S)^{m},\mathcal{W})^{\sim{}^{*}G}. We shall now see a similar result with coefficients being (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) instead, which we shall use in §5.2. For m0m\geq 0, consider the internal space

((G)m,(G,𝒲))\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)

The internal action of G{}^{*}G on (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) can be extended to the following internal action of G{}^{*}G on this space in the natural way: for g,hGg,h\in{}^{*}G, F((G)m,(G,𝒲))F\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right),

(gF)(g1,g2,,gm)(x)=F(g1g1,,g1gm)(xg)(g\cdot F)(g_{1},g_{2},\dots,g_{m})(x)=F(g^{-1}g_{1},\dots,g^{-1}g_{m})(xg)

For convenience, let us denote by

((G)m,b(G,𝒲)Γ)\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)

the space of internal maps in ((G)m,in(G,𝒲))\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{in}^{\infty}({}^{*}G,\mathcal{W})\right) whose image is contained in b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. From Lemma 5.1.3, this space is invariant under the internal action of G{}^{*}G, so denote by

((G)m,b(G,𝒲)Γ)G\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}

the subspace of G{}^{*}G-equivariant maps, and by

((G)m,b(G,𝒲)Γ)G\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}

the subspace of asymptotically G{}^{*}G-equivariant maps. That is, f((G)m,b(G,𝒲)Γ)Gf\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G} if for every gGg\in{}^{*}G, gffinf((G)m,(G,𝒲))g\cdot f-f\in\mathcal{L}_{inf}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right).
Note that

((G)m,b(G,𝒲)Γ)G+inf((G)m,(G,𝒲))((G)m,b(G,𝒲)Γ)G\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}+\mathcal{L}_{inf}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)\subseteq\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}

In other words, a perturbation of a G{}^{*}G-equivariant function by an infinitesimal function is clearly an asymptotically G{}^{*}G-equivariant function. We now show that the converse is true. That is, any asymptotically G{}^{*}G-equivariant map is infinitesimally close to a G{}^{*}G-equivariant map.

Proposition 5.1.7.

For f((G)m,b(G,𝒲)Γ)Gf\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}, there exists βinf((G)m,(G,𝒲))\beta\in\mathcal{L}_{inf}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) such that fβ((G)m,b(G,𝒲)Γ)Gf-\beta\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} (that is, fβf-\beta is (truly) G{}^{*}G-equivariant). The map fβf\mapsto\beta is induced from an internal map from ((G)m,(G,𝒲))\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) to itself.

Proof.

The argument is exactly as in the proof of Lemma 5.1.1, except that now we use the fact that Lw(Gm,Lw(G,𝔲(kn)))L^{\infty}_{w*}(G^{m},L^{\infty}_{w*}(G,\mathfrak{u}(k_{n}))) is relatively injective as a GG-module. ∎

In conclusion,

((G)m,b(G,𝒲)Γ)G+inf((G)m,(G,𝒲))=((G)m,b(G,𝒲)Γ)G\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}+\mathcal{L}_{inf}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)=\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G}

5.2 (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) and the Eckmann-Shapiro Induction

Consider the space b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. We shall now show that, upto infinitesimals, this space can be identified with the internal Banach space 𝒱b(D,𝒲)\mathcal{V}\coloneq\mathcal{L}_{b}^{\infty}({}^{*}D,\mathcal{W}) where DD is a Borelian left fundamental domain of Γ\Gamma in GG. This will then be used to serve as the coefficients to define the asymptotic cohomology Ha(G,𝒱)\operatorname{H}_{a}^{\bullet}(G,\mathcal{V}) of GG.
Consider the internal map

θ:(G,𝒲)(D,𝒲)\theta:\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})

given by restriction of a function to D{}^{*}D. That is, for f={fn}𝒰(G,𝒲)f=\{f_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}),

θf={fn|D}𝒰\theta f=\{f_{n}|_{D}\}_{\mathcal{U}} (5.1)

In the other direction, consider the internal map

ζ:(D,𝒲)(G,𝒲)\zeta:\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})
ζf(g)πΓ(γ)f(z)\zeta f(g)\coloneqq\pi_{\Gamma}(\gamma)f(z) (5.2)

where g=γzg=\gamma z for γΓ\gamma\in{}^{*}\Gamma and zDz\in D. Observe that θζ\theta\cdot\zeta is the identity on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}). As for ζθ\zeta\cdot\theta,

Lemma 5.2.1.

For any fb(G,𝒲)Γf\in\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, (ζθ)(f)finf(G,𝒲)(\zeta\cdot\theta)(f)-f\in\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W}).

Proof.

Since fb(G,𝒲)Γf\in\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, we note that for γΓ\gamma\in{}^{*}\Gamma and zDz\in{}^{*}D, f(γz)πΓ(γ)f(z)𝒲inff(\gamma z)-\pi_{\Gamma}(\gamma)f(z)\in\mathcal{W}_{inf}. ∎

Furthermore, since θ\theta maps inf(G,𝒲)\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W}) to inf(D,𝒲)\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}), and ζ\zeta maps inf(D,𝒲)\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}) to inf(G,𝒲)\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W}),

Lemma 5.2.2.

The internal maps θ\theta and ζ\zeta induce bijections

θ~:b(G,𝒲)Γ/inf(G,𝒲)~(D,𝒲)\tilde{\theta}:\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}/\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W})\to\tilde{\mathcal{L}}^{\infty}({}^{*}D,\mathcal{W})
ζ~:~(D,𝒲)b(G,𝒲)Γ/inf(G,𝒲)\tilde{\zeta}:\tilde{\mathcal{L}}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}/\mathcal{L}_{inf}^{\infty}({}^{*}G,\mathcal{W})

with ζ~=θ~1\tilde{\zeta}=\tilde{\theta}^{-1}.

Henceforth we shall restrict θ\theta to b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. We shall now define an internal map

πG:G×(D,𝒲)(D,𝒲)\pi_{G}:{}^{*}G\times\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})
πG(g)(f)(z)πΓ(γ)f(x)\pi_{G}(g)(f)(z)\coloneqq\pi_{\Gamma}(\gamma)f(x) (5.3)

where zg=γxzg=\gamma x for γΓ\gamma\in{}^{*}\Gamma and xDx\in{}^{*}D. Note that G×DD{}^{*}G\times{}^{*}D\to{}^{*}D given by (g,z)x(g,z)\mapsto x, where zg=γxzg=\gamma x, defines an internal right action of G{}^{*}G on D{}^{*}D. So the above map πG\pi_{G} in Eq. 5.3 can be denoted as

πG(g)(f)(z)πΓ(γ)f(zg)\pi_{G}(g)(f)(z)\coloneqq\pi_{\Gamma}(\gamma)f(zg)

This map πG\pi_{G} induces an action of G{}^{*}G on ~(D,𝒲)\tilde{\mathcal{L}}^{\infty}({}^{*}D,\mathcal{W}), which we shall denote πG~\tilde{\pi_{G}}. In particular, note that this gives the internal Banach space (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) the structure of an asymptotic Banach G{}^{*}G-module, with the asymptotic G{}^{*}G-representation being πG\pi_{G} as defined above.

Lemma 5.2.3.

The maps θ~\tilde{\theta} and ζ~\tilde{\zeta} are G{}^{*}G-equivariant.

Proof.

Note that while we have an internal action of G{}^{*}G on (G,𝒲)\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) that is invariant on b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, the internal map πG\pi_{G} is not an action of G{}^{*}G on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}). Nevertheless, let fb(G,𝒲)Γf\in\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} and gGg\in{}^{*}G. Then πG(g)(θf)(z)=πΓ(γ)(θf)(x)=πΓ(γ)f(x)\pi_{G}(g)(\theta f)(z)=\pi_{\Gamma}(\gamma)(\theta f)(x)=\pi_{\Gamma}(\gamma)f(x) where zg=γxzg=\gamma x. Also, (θgf)(z)=f(zg)=f(γx)(\theta gf)(z)=f(zg)=f(\gamma x). Since fb(G,𝒲)Γf\in\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, f(γx)πΓ(γ)f(x)𝒲inff(\gamma x)-\pi_{\Gamma}(\gamma)f(x)\in\mathcal{W}_{inf}. Thus, θgfπG(g)θfinf(D,𝒲)\theta gf-\pi_{G}(g)\theta f\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}). A similar argument hold for ζ\zeta as well. ∎

Remark 5.2.4.

Consider the space (D,𝒲)G\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})^{\sim{}^{*}G} of asymptotic G{}^{*}G-fixed elements. The restrictions of the maps ζ\zeta and θ\theta to (D,𝒲)G\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})^{\sim{}^{*}G} and ((G,𝒲)Γ)G\left(\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{\sim{}^{*}G} allows us to identify these two spaces upto infinitesimals. Furthermore, from Corollary 5.1.4 and Lemma 5.1.5, we see that, upto infinitesimals, (D,𝒲)G\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})^{\sim{}^{*}G} can be identified with 𝒲bΓ\mathcal{W}_{b}^{\sim{}^{*}\Gamma}.

One of the advantages of defining and working with (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) (instead of (G,𝒲)Γ\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}) is that, not only is it an asymptotic Banach G{}^{*}G-module, but is also easily seen to be dual, which we explicitly describe below. Consider the internal Banach space 1(D,𝒲)\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}) constructed as

1(D,𝒲)𝒰L1(D,(𝔲(kn)))\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat})\coloneqq\prod_{\mathcal{U}}L^{1}\left(D,(\mathfrak{u}(k_{n}))^{\flat}\right)

where L1(D,(𝔲(kn)))L^{1}\left(D,(\mathfrak{u}(k_{n}))^{\flat}\right) is the Bochner-Lebesgue space of Bochner-integrable functions from DD to (𝔲(kn))(\mathfrak{u}(k_{n}))^{\flat}. Note that L(D,𝔲(kn))L^{\infty}(D,\mathfrak{u}(k_{n})) is the dual space of L1(D,(𝔲(kn)))L^{1}(D,(\mathfrak{u}(k_{n}))^{\flat}), so in particular, an element of (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) is an internal linear map from 1(D,𝒲)\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}) to 𝐑{}^{*}\mathbf{R}. We now have an explicit dual pairing can be used to construct the predual asymptotic G{}^{*}G-action given πG\pi_{G}. For f={fn}𝒰(D,𝒲)f=\{f_{n}\}_{\mathcal{U}}\in\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) and η={ηn}𝒰1(D,𝒲)\eta=\{\eta_{n}\}_{\mathcal{U}}\in\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}), define f,η\langle f,\eta\rangle as

f,η={Dfn(x),ηn(x)𝑑x}𝒰\langle f,\eta\rangle=\left\{\int_{D}\langle f_{n}(x),\eta_{n}(x)\rangle dx\right\}_{\mathcal{U}}

This defines an internal pairing

,𝒰:(D,𝒲)×1(D,𝒲)𝐑\langle,\rangle_{\mathcal{U}}:\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\times\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat})\to{}^{*}\mathbf{R}

that induces a pairing between the 𝐑\mathbf{R}-spaces ~(D,𝒲)\tilde{\mathcal{L}}^{\infty}({}^{*}D,\mathcal{W}) and ~1(D,𝒲)\tilde{\mathcal{L}}^{1}({}^{*}D,\mathcal{W}^{\flat}). The space 1(D,𝒲)\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}) comes with an internal map

πG:G×1(D,𝒲)1(D,𝒲)\pi_{G}^{\flat}:{}^{*}G\times\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat})\to\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat})

such that:

  • The map π~G\tilde{\pi}_{G} is contragredient to π~G\tilde{\pi}_{G}^{\flat}, that is, for f(D,𝒲)f\in\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}), η1(D,𝒲)\eta\in\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}) and gGg\in{}^{*}G,

    πG(g)f,ηf,πG(g)η𝐑inf\langle\pi_{G}(g)f,\eta\rangle-\langle f,\pi_{G}^{\flat}(g)\eta\rangle\in{}^{*}\mathbf{R}_{inf}
  • The internal map πG\pi_{G}^{\flat} induces an action of G{}^{*}G, denoted π~G\tilde{\pi}_{G}^{\flat}, on the quotient ~1(D,𝒲)\tilde{\mathcal{L}}^{1}({}^{*}D,\mathcal{W}^{\flat}).

Thus,

Proposition 5.2.5.

The internal Banach space (πG,(D,𝒲))(\pi_{G},\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})) is a dual asymptotic Banach G{}^{*}G-module with predual 1(D,𝒲)\mathcal{L}^{1}({}^{*}D,\mathcal{W}^{\flat}).

We now have the dual asymptotic Banach G{}^{*}G-module 𝒱=(D,𝒲)\mathcal{V}=\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) and an internal map

πG:G×𝒱𝒱\pi_{G}:{}^{*}G\times\mathcal{V}\to\mathcal{V}

that induces an action of G{}^{*}G on 𝒱~\tilde{\mathcal{V}}. This allows us to define the asymptotic cohomology of GG with coefficients in 𝒱\mathcal{V} as in §4.1, as the cohomology of the complex

0{0}𝒱~G{\tilde{\mathcal{V}}^{{}^{*}G}}~(G,𝒱)G{\tilde{\mathcal{L}}^{\infty}({}^{*}G,\mathcal{V})^{{}^{*}G}}~((G)2,𝒱)G{\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{2},\mathcal{V})^{{}^{*}G}}{\dots}d~1\scriptstyle{\tilde{d}^{-1}}d~0\scriptstyle{\tilde{d}^{0}}d~1\scriptstyle{\tilde{d}^{1}}
Theorem 5.2.6.

For every m0m\geq 0, Ham(G,𝒱)Ham(Γ,𝒲)\operatorname{H}_{a}^{m}(G,\mathcal{V})\cong\operatorname{H}_{a}^{m}(\Gamma,\mathcal{W}).

The first step towards proving Theorem 5.2.6 involves a bijection between Γ{}^{*}\Gamma-invariants in ~((G)m,𝒲)\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W}), and G{}^{*}G-invariants in ~((G)m,𝒱)\tilde{\mathcal{L}}^{\infty}\left(({}^{*}G)^{m},\mathcal{V}\right). After that, we shall use the internal maps θ:(G,𝒲)(D,𝒲)\theta:\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) and ζ:(D,𝒲)(G,𝒲)\zeta:\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}) (defined in (Eq. 5.1) and (Eq. 5.2)) to pass between b(G,𝒲)Γ\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} and 𝒱\mathcal{V}.
Let α~~((G)m,𝒲)Γ\tilde{\alpha}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W})^{{}^{*}\Gamma}, and let α((G)m,𝒲)\alpha\in\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{W}) be an internal map that induces α~\tilde{\alpha} (note that αb((G)m,𝒲)Γ\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}G)^{m},\mathcal{W})^{\sim{}^{*}\Gamma}). Define the internal map

Aα:(G)m(G,𝒲)A_{\alpha}:({}^{*}G)^{m}\to\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})
Aα(g1,,gm)(x)α(xg1,,xgm)A_{\alpha}(g_{1},\dots,g_{m})(x)\coloneq\alpha(xg_{1},\dots,xg_{m})

Firstly it is clear that Aα((G)m,(G,𝒲))A_{\alpha}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right). Furthermore,

Proposition 5.2.7.

For every α~~((G)m,𝒲)Γ\tilde{\alpha}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W})^{{}^{*}\Gamma}, the map Aα((G)m,(G,𝒲))A_{\alpha}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma} and is G{}^{*}G-equivariant.

Proof.

For g,xGg,x\in{}^{*}G,

Aα(gg1,,ggm)(x)=α(xgg1,,xggm)=Aα(g1,,gm)(xg)A_{\alpha}(gg_{1},\dots,gg_{m})(x)=\alpha(xgg_{1},\dots,xgg_{m})=A_{\alpha}(g_{1},\dots,g_{m})(xg)

This proves that AαA_{\alpha} is G{}^{*}G-equivariant. Next, for γΓ\gamma\in{}^{*}\Gamma,

Aα(g1,,gm)(γx)=α(γxg1,,γxgm)A_{\alpha}(g_{1},\dots,g_{m})(\gamma x)=\alpha(\gamma xg_{1},\dots,\gamma xg_{m})

Since α~~((G)m,𝒲)Γ\tilde{\alpha}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W})^{{}^{*}\Gamma}, this means that

α(γxg1,,γxgm)πΓ(γ)α(xg1,,xgm)𝒲inf\alpha(\gamma xg_{1},\dots,\gamma xg_{m})-\pi_{\Gamma}(\gamma)\alpha(xg_{1},\dots,xg_{m})\in\mathcal{W}_{inf}

Hence Aα(g1,,gm)(γx)πΓ(γ)(Aα(g1,,gm)(γx))𝒲infA_{\alpha}(g_{1},\dots,g_{m})(\gamma x)-\pi_{\Gamma}(\gamma)\left(A_{\alpha}(g_{1},\dots,g_{m})(\gamma x)\right)\in\mathcal{W}_{inf}. This shows that Aα(g1,,gm)b(G,𝒲)ΓA_{\alpha}(g_{1},\dots,g_{m})\in\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. ∎

Thus, given a Γ{}^{*}\Gamma-equivariant map α~~((G)m,𝒲)\tilde{\alpha}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W}), we obtain an internal G{}^{*}G-equivariant map Aα((G)m,(G,𝒲))A_{\alpha}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) that takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}.
Conversely, suppose we have an internal G{}^{*}G-equivariant map A((G)m,(G,𝒲))A\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) that takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. Define the internal map

αA((G)m,𝒲)\alpha_{A}\in\mathcal{L}^{\infty}(({}^{*}G)^{m},\mathcal{W})
αA(g1,,gm)A(x1g1,,x1gm)(x)\alpha_{A}(g_{1},\dots,g_{m})\coloneqq A(x^{-1}g_{1},\dots,x^{-1}g_{m})(x)

for xGx\in{}^{*}G. Note that since AA is G{}^{*}G-equivariant, the map xA(x1g1,,x1gm)(x)x\mapsto A(x^{-1}g_{1},\dots,x^{-1}g_{m})(x) is essentially constant in 𝒲\mathcal{W}, making the above well-defined.

Lemma 5.2.8.

Given an internal G{}^{*}G-equivariant map A((G)m,(G,𝒲))A\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) that takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, the internal map αA\alpha_{A} as defined above induces the map α~A\tilde{\alpha}_{A} that is Γ{}^{*}\Gamma-equivariant. That is, α~A~((G)m,𝒲)Γ\tilde{\alpha}_{A}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W})^{{}^{*}\Gamma}.

Proof.

Let γΓ\gamma\in{}^{*}\Gamma, then since AA is G{}^{*}G-equivariant,

αA(γg1,,γgm)=A(x1γg1,,x1γgm)(x)=A(x1g1,,x1gm)(γx)\alpha_{A}(\gamma g_{1},\dots,\gamma g_{m})=A(x^{-1}\gamma g_{1},\dots,x^{-1}\gamma g_{m})(x)=A(x^{-1}g_{1},\dots,x^{-1}g_{m})(\gamma x)

Since AA takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}, we know that

A(x1g1,,x1gm)(γx)πΓ(γ)(A(x1g1,,x1gm)(x))𝒲infA(x^{-1}g_{1},\dots,x^{-1}g_{m})(\gamma x)-\pi_{\Gamma}(\gamma)\left(A(x^{-1}g_{1},\dots,x^{-1}g_{m})(x)\right)\in\mathcal{W}_{inf}

Thus, αA(γg1,,γgm)πΓ(γ)αA(g1,,gm)𝒲inf\alpha_{A}(\gamma g_{1},\dots,\gamma g_{m})-\pi_{\Gamma}(\gamma)\alpha_{A}(g_{1},\dots,g_{m})\in\mathcal{W}_{inf}, proving that α~A~((G)m,𝒲)Γ\tilde{\alpha}_{A}\in\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m},\mathcal{W})^{{}^{*}\Gamma}. ∎

Furthermore, the correspondences

α\displaystyle\alpha Aα\displaystyle\mapsto A_{\alpha} (5.4)
A\displaystyle A αA\displaystyle\mapsto\alpha_{A} (5.5)

are clearly inverses of each other, thus giving a bijection between internal maps αb((G)m,𝒲)\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}G)^{m},\mathcal{W}) that are asymptotically Γ{}^{*}\Gamma-equivariant, and internal G{}^{*}G-equivariant maps A((G)m,(G,𝒲))A\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right) that take values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}.

Proof of Theorem 5.2.6.

For m0m\in\mathbb{Z}_{\small{\geq 0}}, it is sufficient to show a bijection between the quotients ~((G)m+1,𝒲)Γ\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{W})^{{}^{*}\Gamma} and ~((G)m+1,𝒱)G\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{V})^{{}^{*}G} that commutes with the differentials d~\tilde{d}.
Let αb((G)m+1,𝒲)Γ\alpha\in\mathcal{L}^{\infty}_{b}(({}^{*}G)^{m+1},\mathcal{W})^{\sim{}^{*}\Gamma} and let

Aα((G)m+1,(G,𝒲))A_{\alpha}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m+1},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\right)

be its lift (as in (Eq. 5.4)) that is asymptotically G{}^{*}G-equivariant and takes values in b(G,𝒲)Γ\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}. Composing AαA_{\alpha} with θ\theta, we get

θAαb((G)m+1,𝒱)G\theta\cdot A_{\alpha}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}G)^{m+1},\mathcal{V}\right)^{\sim{}^{*}G}

In the other direction, for an internal map Ab((G)m+1,𝒱)GA^{\prime}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}G)^{m+1},\mathcal{V}\right)^{\sim{}^{*}G}, we first compose ζ\zeta (defined in (Eq. 5.2)) with AA^{\prime} to get ζA((G)m+1,b(G,𝒲)Γ)\zeta\cdot A^{\prime}\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m+1},\mathcal{L}^{\infty}_{b}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right), where ζA\zeta\cdot A^{\prime} is asymptotically G{}^{*}G-equivariant. Let A((G)m+1,b(G,𝒲)Γ)GA\in\mathcal{L}^{\infty}\left(({}^{*}G)^{m+1},\mathcal{L}_{b}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G} be the G{}^{*}G-equivariant map infinitesimally close to ζA\zeta\cdot A^{\prime} (as guaranteed by Proposition 5.1.7).
Now we can descend to Γ\Gamma by defining the internal map

αA((G)m+1,𝒲)\alpha_{A}\in\mathcal{L}^{\infty}(({}^{*}G)^{m+1},\mathcal{W})
αA(g0,,gm)A(x1g0,,x1gm)(x)\alpha_{A}(g_{0},\dots,g_{m})\coloneqq A(x^{-1}g_{0},\dots,x^{-1}g_{m})(x)

Since AA is G{}^{*}G-equivariant, from Lemma 5.2.8 we conclude that αA\alpha_{A} is asymptotically Γ{}^{*}\Gamma-equivariant. Thus, we have a bijection between the quotients ~((G)m+1,𝒲)Γ\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{W})^{{}^{*}\Gamma} and ~((G)m+1,𝒱)G\tilde{\mathcal{L}}^{\infty}(({}^{*}G)^{m+1},\mathcal{V})^{{}^{*}G} that commutes with the differentials d~.\tilde{d}.

5.3 Internal Contraction and Fixed Points

Recall that we have the dual asymptotic Banach G{}^{*}G-module (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) with the asymptotic G{}^{*}G-action of G{}^{*}G given by πG\pi_{G} as in (Eq. 5.3). The map πG\pi_{G} can be studied in terms of two internal maps: an internal right G{}^{*}G-action G×DD{}^{*}G\times{}^{*}D\to{}^{*}D of G{}^{*}G on D{}^{*}D given by (g,z)zg(g,z)\mapsto zg, and another internal twisting map G×DΓ{}^{*}G\times{}^{*}D\to{}^{*}\Gamma given by (g,z)γ(g,z)\mapsto\gamma (here γΓ\gamma\in{}^{*}\Gamma and xDx\in{}^{*}D are such that zg=γxzg=\gamma x). The latter map is then composed with πΓ\pi_{\Gamma} to give πG(g)(f)(z)πΓ(γ)f(x)\pi_{G}(g)(f)(z)\coloneqq\pi_{\Gamma}(\gamma)f(x) as in (Eq. 5.3).
We now see a continuity property of the asymptotic G{}^{*}G-action πG\pi_{G} on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}). Observe that the internal map (g,z)πΓ(γ)(g,z)\mapsto\pi_{\Gamma}(\gamma) is internally measurable (and internally locally constant), and as for the internal (true) action of G{}^{*}G on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) given by (gf)(z)f(x)(g\cdot f)(z)\coloneq f(x), this is internally continuous, but with respect to the internal L2L^{2}-norm defined on f={fn}𝒰f=\{f_{n}\}_{\mathcal{U}} as follows:

f22{Dfn(x)2𝑑x}𝒰\|f\|_{2}^{2}\coloneq\left\{\int_{D}\|f_{n}(x)\|^{2}dx\right\}_{\mathcal{U}}
Lemma 5.3.1.

The dual asymptotic Banach G{}^{*}G-module (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) is internally continuous with respect to the internal L2L^{2}-norm on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}).

We shall refer to internal continuity with respect to the L2L^{2}-norm on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) as internal L2L^{2}-continuity. The main consequence of this property that we shall use is the following: let g(1),g(2),g^{(1)},g^{(2)},\dots be a sequence of elements of G{}^{*}G (with g(m)={gn(m)}𝒰g^{(m)}=\{g_{n}{(m)}\}_{\mathcal{U}}) such that the internal limit limmg(m)={limmgn(m)}𝒰=gG\lim\limits_{m\to\infty}g^{(m)}=\{\lim\limits_{m\to\infty}g_{n}^{(m)}\}_{\mathcal{U}}=g\in{}^{*}G. Then for f(D,𝒲)f\in\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}), limmf(g(m))=f(g)\lim\limits_{m\to\infty}f(g^{(m)})=f(g) (where the limit is now with respect to the internal L2L^{2}-norm). A particular instance of an internal limit we shall use is given by contraction of elements.

Definition 5.3.2.

Let g,hGg,h\in G. We say that hh is contracted by gg (or gg contracts hh) if limmgmhgm=1\lim\limits_{m\to\infty}g^{-m}hg^{m}=1. Let g={gn}𝒰g=\{g_{n}\}_{\mathcal{U}} and h={hn}𝒰h=\{h_{n}\}_{\mathcal{U}} be elements of G{}^{*}G. We say hh is internally contracted by gg (or gg internally contracts hh) if for every n𝒰n\in\mathcal{U},

limmgnmhngnm=1\lim\limits_{m\to\infty}g_{n}^{-m}h_{n}g_{n}^{m}=1

In other words, gg internally contracts hh if the sequence gmhgmg^{-m}hg^{m} has internal limit being the identity in G{}^{*}G.

Definition 5.3.3.

An internal subgroup of the ultrapower G{}^{*}G is called an internally amenable subgroup if it is of the form 𝒰Hn\prod_{\mathcal{U}}H_{n} where HnGH_{n}\leq G is a closed amenable subgroup for every n𝐍n\in\mathbf{N}.

Remark 5.3.4.

Suppose g,hGg,h\in G such that gg contracts hh. Then the closed subgroup <g,h><g,h> generated by gg and hh is an amenable subgroup of GG (refer Propositions 6.56.5 and 6.176.17 in [16]). In particular, this implies that if g,hGg,h\in{}^{*}G is such that gg internally contracts hh, then the internal subgroup <g,h><g,h> generated by gg and hh in G{}^{*}G is an internally amenable subgroup of G{}^{*}G.

To use the full power of internal contraction and the internal L2L^{2}-continuity of the asymptotic G{}^{*}G-action πG\pi_{G}, our first step is to \saycorrect the asymptotic G{}^{*}G-action πG\pi_{G} to get a true action when restricted to an internally amenable subgroup. Note that the imperfection is in the map πG\pi_{G}^{\prime}, so consider the map α:G×D𝒰U(kn)\alpha:{}^{*}G\times{}^{*}D\to\prod_{\mathcal{U}}U(k_{n}) with α(g,z)πΓ(γ)\alpha(g,z)\coloneq\pi_{\Gamma}(\gamma). Note that α\alpha is a measurable map, and for every g1,g2Γg_{1},g_{2}\in{}^{*}\Gamma and zDz\in{}^{*}D,

α(g1,z)α(g2,zg1)α(g1g2,z)𝐑inf\|\alpha(g_{1},z)\alpha(g_{2},zg_{1})-\alpha(g_{1}g_{2},z)\|\in{}^{*}\mathbf{R}_{inf} (5.6)

Observe that this looks similar to the very classical question of Ulam stability, but with a twist provided by the action of GG on DD. We consider the following definitions which are analogues of uniform stability in the context of such twists:

Definition 5.3.5.

Let GG be a locally compact, second countable group, and XX be a non-singular GG-space. A measurable map ψ:G×XU(n)\psi:G\times X\to U(n) is called a twisted homomorphism of GG (with respect to DD) if for every g1,g2Gg_{1},g_{2}\in G and almost every zXz\in X,

ψ(g1g2,z)=ψ(g1,z)ψ(g2,zg1)\psi(g_{1}g_{2},z)=\psi(g_{1},z)\psi(g_{2},zg_{1})

For ϵ>0\epsilon>0 (the defect), a measurable map ϕ:G×XU(n)\phi:{}G\times X\to U(n) is called a twisted ϵ\epsilon-homomorphism of GG (with respect to DD) if for almost every g1,g2Gg_{1},g_{2}\in G and almost every zXz\in X

ϕ(g1,z)ϕ(g2,zg1)ϕ(g1g2,z)ϵ\|\phi(g_{1},z)\phi(g_{2},zg_{1})-\phi(g_{1}g_{2},z)\|\leq\epsilon

The following claim essentially retraces the arguments used in §3 (where we use the logarithm map to obtain an asymptotic cocycle, and use the vanishing of cohomology to diminish defect).

Claim 5.3.6.

There exists ϵ>0\epsilon>0 small enough, and a constant CC such that for an amenable subgroup HH of GG and any twisted ϵ\epsilon-homomorphism α:G×DU(n)\alpha:G\times D\to U(n), there exists a measurable map αH:H×DU(n)\alpha_{H}:H\times D\to U(n) of HH such that for every hHh\in H and almost every zDz\in D, α(h,z)αH(h,z)Cϵ\|\alpha(h,z)-\alpha_{H}(h,z)\|\leq C\epsilon, and for every h1,h2Hh_{1},h_{2}\in H and zDz\in D,

αH(h1,z)αH(h2,zh1)=αH(h1h2,z)\alpha_{H}(h_{1},z)\alpha_{H}(h_{2},zh_{1})=\alpha_{H}(h_{1}h_{2},z)
Proof.

As in §3, consider a sequence {αn:H×DU(kn)}n𝐍\{\alpha_{n}:H\times D\to U(k_{n})\}_{n\in\mathbf{N}} and its ultraproduct α\alpha, which is an asymptotic twisted homomorphism with defect ϵ𝐑inf\epsilon\in{}^{*}\mathbf{R}_{inf}. Define the internal map

ω:H×H×D𝒰𝔲(kn)\omega:{}^{*}H\times{}^{*}H\times{}^{*}D\to\prod_{\mathcal{U}}\mathfrak{u}(k_{n})
ω(h1,h2,z)=logϵ(α(h1,z)α(h2,zh1)α(h1h2,z)1)\omega(h_{1},h_{2},z)={}_{\epsilon}\log\left(\alpha(h_{1},z)\alpha(h_{2},zh_{1})\alpha(h_{1}h_{2},z)^{-1}\right)

As 𝔲(kn)=𝒲\mathfrak{u}(k_{n})=\mathcal{W}, note that we can regard ω\omega as an element of b(H×H×D,𝒲)\mathcal{L}^{\infty}_{b}({}^{*}H\times{}^{*}H\times{}^{*}D,\mathcal{W}), or equivalently, b((H)2,(D,𝒲))\mathcal{L}^{\infty}_{b}(({}^{*}H)^{2},\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})). Equipping (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) with an asymptotic action ρα:H×(D,𝒲)(D,𝒲)\rho_{\alpha}:{}^{*}H\times\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) of H{}^{*}H given by ρα(h)(f)(z)α(h,z)f(zh)α(h,z)1\rho_{\alpha}(h)(f)(z)\coloneq\alpha(h,z)f(zh)\alpha(h,z)^{-1}, making it a dual asymptotic H{}^{*}H-module. One can check that the map ω\omega satisfies the condition to be an (inhomogenous) asymptotic 22-cocycle in Ha2(H,(D,𝒲))\operatorname{H}_{a}^{2}(H,\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})). Now since HH is amenable and (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) is a dual asymptotic Banach H{}^{*}H-module, we know (from Corollary 4.3.4) that Ha2(H,(D,𝒲))=0\operatorname{H}_{a}^{2}(H,\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}))=0, implying that there exists βb(H,(D,𝒲))\beta\in\mathcal{L}^{\infty}_{b}({}^{*}H,\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})) such that αϵexpβ\alpha\cdot_{\epsilon}\exp{\beta} is an o𝒰(ϵ)o_{\mathcal{U}}(\epsilon)-twisted homomorphism (note that αϵexpβ\alpha\cdot_{\epsilon}\exp{\beta} is measurable). Repeating this process (as in defect diminishing), we obtain a measurable map αH:H×DU(n)\alpha_{H}^{\prime}:H\times D\to U(n) such that for almost every h1,h2Hh_{1},h_{2}\in H and almost every zDz\in D, we haveαH(h1,z)αH(h2,zh1)=αH(h1h2,z)\alpha_{H}^{\prime}(h_{1},z)\alpha_{H}^{\prime}(h_{2},zh_{1})=\alpha_{H}^{\prime}(h_{1}h_{2},z). From Theorem B99 (p.200) in [51], we conclude that there exists αH\alpha_{H} as desired. ∎

Remark 5.3.7.

The same argument also goes through for internally amenable subgroups \mathcal{H} by applying 5.3.6 internally.

Lemma 5.3.8.

Let G\mathcal{H}\leq{}^{*}G be an internally amenable subgroup of G{}^{*}G. Then there exists an internal map π:×(D,𝒲)(D,𝒲)\pi_{\mathcal{H}}:\mathcal{H}\times\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) such that

  • For every h1,h2Hh_{1},h_{2}\in H, π(h1h2)=πH(h1)πH(h2)\pi_{\mathcal{H}}(h_{1}h_{2})=\pi_{H}(h_{1})\pi_{H}(h_{2}). In other words, π\pi_{\mathcal{H}} is a (true) internal action of \mathcal{H} on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}).

  • For every fb(D,𝒲)f\in\mathcal{L}_{b}^{\infty}({}^{*}D,\mathcal{W}) and hh\in\mathcal{H}, π(h)fπG(h)finf(D,𝒲)\pi_{\mathcal{H}}(h)f-\pi_{G}(h)f\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}).

  • The internal action π\pi_{\mathcal{H}} of \mathcal{H} is internally 22-continuous.

Proof.

Consider the internal map α:G×D𝒰U(kn)\alpha:{}^{*}G\times{}^{*}D\to\prod_{\mathcal{U}}U(k_{n})

α(g,z)πΓ(γ)\alpha(g,z)\coloneq\pi_{\Gamma}(\gamma)

where γx=zg\gamma x=zg for γΓ\gamma\in{}^{*}\Gamma and xDx\in{}^{*}D. Since α\alpha is an asymptotic twisted homomorphism of GG, from Remark 5.3.7, there exists an internally measurable map α:×D𝒰U(kn)\alpha_{\mathcal{H}}:\mathcal{H}\times{}^{*}D\to\prod_{\mathcal{U}}U(k_{n}) as in 5.3.6.
Define π:×(D,𝒲)(D,𝒲)\pi_{\mathcal{H}}:\mathcal{H}\times\mathcal{L}^{\infty}({}^{*}D,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) by

π(g)(f)(z)α(g,z)f(x)\pi_{\mathcal{H}}(g)(f)(z)\coloneq\alpha_{\mathcal{H}}(g,z)f(x)

whereas always, xDx\in{}^{*}D such that zg=γxzg=\gamma x for γΓ\gamma\in{}^{*}\Gamma. Then π\pi_{\mathcal{H}} is a (true) internal action of \mathcal{H} on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}), and for every fb(D,𝒲)f\in\mathcal{L}_{b}^{\infty}({}^{*}D,\mathcal{W}) and hh\in\mathcal{H}, π(h)fπG(h)finf(D,𝒲)\pi_{\mathcal{H}}(h)f-\pi_{G}(h)f\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}). Since π\pi_{\mathcal{H}} is an internal action of \mathcal{H} that is internally measurable, it is also internally 22-continuous. ∎

Lemma 5.3.9.

Let x,yGx,y\in{}^{*}G such that xx internally contracts yy. Suppose fb(D,𝒲)f\in\mathcal{L}^{\infty}_{b}({}^{*}D,\mathcal{W}) is such that πG(x)ffinf(D,𝒲)\pi_{G}(x)f-f\in\mathcal{L}^{\infty}_{inf}({}^{*}D,\mathcal{W}) (in other words, xx fixes f~\tilde{f}). Then πG(y)ffinf(D,𝒲)\pi_{G}(y)f-f\in\mathcal{L}^{\infty}_{inf}({}^{*}D,\mathcal{W}), that is, yy too fixes f~\tilde{f}.

Proof.

Let \mathcal{H} be the closure of the subgroup generated by xx and yy. By Remark 5.3.4, \mathcal{H} is an internally amenable subgroup of G{}^{*}G. By Lemma 5.3.8, consider the correction π\pi_{\mathcal{H}} whose restriction to \mathcal{H} is a (true) internally 22-continuous action of \mathcal{H} on (D,𝒲)\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}). Now π(x)ffinf(D,𝒲)\pi_{\mathcal{H}}(x)f-f\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}), and let fb(D,𝒲)f_{\mathcal{H}}\in\mathcal{L}_{b}^{\infty}({}^{*}D,\mathcal{W}) such that ffinf(D,𝒲)f-f_{\mathcal{H}}\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}) and π(x)f=f\pi_{\mathcal{H}}(x)f_{\mathcal{H}}=f_{\mathcal{H}}.
Now we have

π(y)fHfH2=π(y)π(xm)fπ(xm)f2\|\pi_{\mathcal{H}}(y)f_{H}-f_{H}\|_{2}=\|\pi_{\mathcal{H}}(y)\pi_{\mathcal{H}}(x^{-m})f_{\mathcal{H}}-\pi_{\mathcal{H}}(x^{-m})f_{\mathcal{H}}\|_{2}

Since π\pi_{\mathcal{H}} too acts unitarily,

π(y)π(xm)fπ(xm)f2=π(xm)π(y)πH(xm)ff2\|\pi_{\mathcal{H}}(y)\pi_{\mathcal{H}}(x^{-m})f_{\mathcal{H}}-\pi_{\mathcal{H}}(x^{-m})f_{\mathcal{H}}\|_{2}=\|\pi_{\mathcal{H}}(x^{m})\pi_{\mathcal{H}}(y)\pi_{H}(x^{-m})f_{\mathcal{H}}-f_{\mathcal{H}}\|_{2}

Since π\pi_{\mathcal{H}} is a (true) internal action of {\mathcal{H}}, π(xm)π(y)π(xm)=π(xmyxm)\pi_{\mathcal{H}}(x^{m})\pi_{\mathcal{H}}(y)\pi_{\mathcal{H}}(x^{-m})=\pi_{\mathcal{H}}(x^{m}yx^{-m}), and so

π(xm)π(y)π(xm)ff2=π(xmyxm)ff2\|\pi_{\mathcal{H}}(x^{m})\pi_{\mathcal{H}}(y)\pi_{\mathcal{H}}(x^{-m})f_{\mathcal{H}}-f_{\mathcal{H}}\|_{2}=\|\pi_{\mathcal{H}}(x^{m}yx^{-m})f_{\mathcal{H}}-f_{\mathcal{H}}\|_{2}

Since π\pi_{\mathcal{H}} is internally 22-continuous and yy is internally contracted by xx, we conclude that π(y)f=f\pi_{\mathcal{H}}(y)f_{\mathcal{H}}=f_{\mathcal{H}}, implying that πG(y)ffinf(D,𝒲)\pi_{G}(y)f-f\in\mathcal{L}_{inf}^{\infty}({}^{*}D,\mathcal{W}). ∎

We can apply the above results in a slightly more general setting, which we shall develop into an internal Mautner’s Lemma in Subection §6.1.

Definition 5.3.10.

Let TGT\leq G be a closed subgroup and M>0M>0. The group GG is said to be MM-boundedly generated by TT-contracted elements if any element gGg\in G, there exist s1,,smGs_{1},\dots,s_{m}\in G with mMm\leq M, such that g=s1s2smg=s_{1}s_{2}\dots s_{m}, and each siGs_{i}\in G is contracted by some element of TT. More generally, for a family 𝔗\mathfrak{T} of closed subgroups of GG, the group GG is said to be boundedly generated by 𝔗\mathfrak{T}-contracted elements if there exists M>0M>0 such that GG is MM-boundedly generated by TT-contracted elements for every T𝔗T\in\mathfrak{T}.

Remark 5.3.11.

Suppose GG is boundedly generated by 𝔗\mathfrak{T}-contracted elements for a family 𝔗\mathfrak{T} of subgroups. Let 𝒯=𝒰Tn\mathcal{T}=\prod_{\mathcal{U}}T_{n} be an internal subgroup of G{}^{*}G, with Tn𝔗T_{n}\in\mathfrak{T} for every n𝐍n\in\mathbf{N}. Observe that there exists M>0M>0 such that any gGg\in{}^{*}G can be expressed as g=s1s2smg=s_{1}s_{2}\dots s_{m} such that each siGs_{i}\in{}^{*}G is internally contracted by some element of 𝒯\mathcal{T}.

With this definition, Lemma 5.3.9 implies the following corollary:

Corollary 5.3.12.

Suppose GG is boundedly generated by 𝔗\mathfrak{T}-contracted elements, and let 𝒯=𝒰Tn\mathcal{T}=\prod_{\mathcal{U}}T_{n} be an internal subgroup of G{}^{*}G, with Tn𝔗T_{n}\in\mathfrak{T} for every n𝐍n\in\mathbf{N}. Then if fb(D,𝒲)𝒯f\in\mathcal{L}^{\infty}_{b}({}^{*}D,\mathcal{W})^{\sim\mathcal{T}}, then fb(D,𝒲)Gf\in\mathcal{L}^{\infty}_{b}({}^{*}D,\mathcal{W})^{\sim{}^{*}G} (that is, if ff is asymptotically fixed by all elements of 𝒯\mathcal{T}, then it is asymptotically G{}^{*}G-fixed).

Proof.

Let fb(D,𝒲)𝒯f\in\mathcal{L}^{\infty}_{b}({}^{*}D,\mathcal{W})^{\sim\mathcal{T}} and gGg\in{}^{*}G. Let g=s1s2smg=s_{1}s_{2}\dots s_{m} as in Remark 5.3.11. Since ff is asymptotically fixed by every element of 𝒯\mathcal{T}, Lemma 5.3.9 implies that it is asymptotically fixed by each sis_{i}. In particular, ff is asymptotically fixed by gg. Thus, fb(D,𝒲)Gf\in\mathcal{L}^{\infty}_{b}({}^{*}D,\mathcal{W})^{\sim{}^{*}G}. ∎

6 Vanishing of Ha2(Γ,𝒲)\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})

In this section, we combine the results of the previous sections to prove our main result about uniform stability of lattices in semisimple groups. In §6.1, we build further tools specialized to GG being semisimple groups, which in turn shall be used in §6.2. We begin with some structure properties of semisimple groups with regard to the notions of bounded generation by contracting elements, as discussed in §5.3, and use it prove an asymptotic version of the Mautner property and an asymptotic double ergodicity theorem with coefficients. In §6.2, we study the Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), which is an assumption on GG that would allow us to build on the results of §6.1 to prove our main theorem, namely that Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(G,\mathcal{V})=0. We then conclude the section by studying Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) in more detail in §6.3, and list out a large class of semisimple groups that satisfy it, thus making our main result applicable to them.

6.1 Asymptotic Mautner Property and Ergodicity

In this subsection and henceforth, we shall work with GG being a semisimple group of the form G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected, simply connected, almost KiK_{i}-simple group. Our goal is to use the results developed in the previous sections to prove that Ha2(Γ,𝒲)=Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=\operatorname{H}_{a}^{2}(G,\mathcal{V})=0 (recall that 𝒲\mathcal{W} is the ultraproduct 𝒰𝔲(kn)\prod_{\mathcal{U}}\mathfrak{u}(k_{n}) which is an asymptotic Banach Γ{}^{*}\Gamma-module obtained from the asymptotic homomorphism we start with, while 𝒱=(D,𝒲)\mathcal{V}=\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) is the asymptotic Banach G{}^{*}G-module obtained by the induction procedure in Section 5.2).
Recall Definition 5.3.10 of bounded generation by contracted elements. We shall now see that GG is boundedly generated by 𝔗\mathfrak{T}-contracted elements, for the subgroup family 𝔗\mathfrak{T} being the maximal tori of the radicals of proper parabolic subgroups of GG.

Proposition 6.1.1.

Let G=𝐆(K)G=\mathbf{G}(K) be a connected, simply connected, almost KK-simple KK-isotropic group over local field KK, and let 𝔗\mathfrak{T} be the family of maximal tori of the solvable radicals of proper parabolic subgroups of GG. Then GG is boundedly generated by 𝔗\mathfrak{T}-contracted elements.

Proof.

Let TT be a maximal torus contained in the solvable radical of a proper parabolic subgroup QQ of GG. It is sufficient to show that for some M>0M>0, GG is MM-boundedly generated by TT-contracted elements. This is because the same bound MM works for conjugates of TT, and upto conjugation, the group GG has only finitely many parabolic subgroups.
Let Q=LRu(Q)Q=L\cdot R_{u}(Q) be the Levi decomposition of QQ, where LL is the Levi subgroup and Ru(Q)R_{u}(Q) is the unipotent radical of QQ. Then every element of Ru(Q)R_{u}(Q) is contracted by some element of TT. In fact, the same holds in the case of the opposite parabolic QQ^{-}, that is, every element of Ru(Q)R_{u}(Q^{-}) too is contracted by some element of TT. Hence it is sufficient to show that GG is boundedly generated by ΔRu(Q)(K)Ru(Q)(K)\Delta\coloneqq R_{u}(Q)(K)\cup R_{u}(Q^{-})(K), that is, to show that there exists >0\ell>0 such that G=ΔG=\Delta^{\ell}.
Firstly, note that GG is indeed generated by Δ\Delta ([34, 1.5.4]), and so, there exists k>0k>0 so that the product map ΔkG{\Delta}^{k}\to G is a dominant KK-morphism. From [45, Proposition 3.3], we conclude that ΩΔ\Omega\coloneqq\Delta^{\ell} contains an open neighborhood of the identity. Let TT be a maximal KK-split torus of GG contained in QQQ\cap Q^{-}. For any KK-defined unipotent subgroup UGU\subset G which is normalized by TT, we have

U(K)=tT(K)t(U(K)Ω)t1U(K)=\bigcup_{t\in T(K)}t(U(K)\cap\Omega)t^{-1}

Since T(K)T(K) normalizes Δ\Delta (and hence also Ω\Omega), we see that U(K)ΩU(K)\subseteq\Omega. In particular, for a KK-defined minimal parabolic subgroup PQP\subset Q containing TT, we conclude that Ru(P)(K)ΩR_{u}(P)(K)\subset\Omega.
Let D=TRu(P)D=T\cdot R_{u}(P). Then there exists a compact subset EGE\subset G so that

G=ED(K)=ET(K)Ru(P)(K)G=E\cdot D(K)=E\cdot T(K)\cdot R_{u}(P)(K)

Since Ω\Omega is open and generates GG, there exists s>0s>0 such that EΩsE\subset\Omega^{s}. Hence it is sufficient to show that there exists t>0t>0 such that T(K)ΩtT(K)\subset\Omega^{t}.
There exists N>0N>0 such that for any root α\alpha of TT with coroot α\alpha^{\vee} (recall that for a root α:T(K)K\alpha:T(K)\to K, its dual α:KT(K)\alpha^{\vee}:K\to T(K) is such that α(α)=2\alpha^{\vee}(\alpha)=2), α(t)\alpha^{\vee}(t) for tK×t\in K^{\times} can be written as a product of NN elements of Δ\Delta (for instance, N=6N=6 when G=SLnG=SL_{n}). In particular, this means that every element of T(K)T(K) can be written as a product of NnNn elements of Δ\Delta (where nn is the rank of GG). This concludes the proof. ∎

In fact, Proposition 6.1.1 can be immediately extended to semisimple groups as well, as long as we ensure that the projection of the tori to each factor is non-trivial.

Corollary 6.1.2.

Let GG be a semisimple group of the form G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected, simply connected, almost KiK_{i}-simple KiK_{i}-isotropic group, and let 𝔗\mathfrak{T} be the family of maximal tori of the radicals of parabolic subgroups of GG of the form Q=i=1kQiQ=\prod_{i=1}^{k}Q_{i}, where each QiQ_{i} is a proper parabolic subgroup of 𝐆i(Ki){\mathbf{G}_{i}}(K_{i}). Then GG is boundedly generated by 𝔗\mathfrak{T}-contracted elements.

Combining Corollary 5.3.12 with Corollary 6.1.2, we immediately get the following:

Corollary 6.1.3 (Asymptotic Mautner Property).

Let GG and a parabolic subgroup QGQ\leq G be as in Corollary 6.1.2, and let NGN\leq G be the radical of QQ. Then 𝒱bN=𝒱bG\mathcal{V}_{b}^{\sim{}^{*}N}=\mathcal{V}_{b}^{\sim{}^{*}G}.

Corollary 5.3.12 and Corollary 6.1.2 can also be used to get an internal analogue of ergodicity (and double ergodicity) with coefficients as in [37]. We recall the following classical definitions for comparison:

Definition 6.1.4.

For a locally compact second countable group GG and a regular GG-space SS, the GG-action on SS is said to be ergodic if any measurable GG-invariant function f:S𝐑f:S\to\mathbf{R} is essentially constant. The GG-action on SS is said to be doubly ergodic if the diagonal GG-action on S×SS\times S is ergodic.

There is no difference if 𝐑\mathbf{R} above is replaced by any dual separable Banach space: the GG-action on SS is ergodic (resp., double ergodic) iff for any dual separable Banach space EE, any weak-* measurable GG-invariant map f:SEf:S\to E (resp, f:S×SEf:S\times S\to E) is essentially constant. Indeed the underlying Borel structures are isomorphic. Even if EE is not separable, it suffices that it admits a weak-* measurable (for instance dual linear) injection into a separable dual, since we can then apply ergodicity in the latter space. Here is an illustration of this phenomenon:

Claim 6.1.5.

Let f:GLw(D,𝔲(kn))f\colon G\to L^{\infty}_{w*}(D,\mathfrak{u}(k_{n})) be a GG-invariant weak-* measurable map. Then ff is essentially constant.

Proof.

The GG-action on itself is ergodic, so that this follows from the previous discussion after noticing that Lw(D,𝔲(kn))L^{\infty}_{w*}(D,\mathfrak{u}(k_{n})) admits a dual linear injection into L2(D,𝔲(kn))L^{2}(D,\mathfrak{u}(k_{n})), which is dual and separable. This injection is the dual of the natural map from L2(D,𝔲(kn))L^{2}(D,\mathfrak{u}(k_{n})) to L1(D,𝔲(kn))L^{1}(D,\mathfrak{u}(k_{n})), which is defined because DD has finite measure. ∎

Note that 6.1.5 immediately implies the following (wherein no GG-structure is considered on 𝒱\mathcal{V}):

Claim 6.1.6.

Let Fb(G,𝒱)F\in\mathcal{L}_{b}^{\infty}\left({}^{*}G,\mathcal{V}\right) be G{}^{*}G-invariant (that is, for every xGx\in{}^{*}G and almost every gGg\in{}^{*}G, F(xg)=F(g)F(xg)=F(g)). Then FF is essentially constant. ∎

A classical example of a doubly ergodic action for a semisimple group GG (as in Corollary 6.1.2) is its action on G/PG/P for PP being a minimal parabolic subgroup of GG (which implies that the action of GG on G/QG/Q is doubly ergodic for any parabolic subgroup QQ of GG).

We shall now use Corollary 5.3.12 with Corollary 6.1.2 to get a much stronger double ergodicity result for asymptotically G{}^{*}G-equivariant maps. Classically, this is referred to as double ergodicity with coefficients ([37, Chapter 4.11]):

A GG-action on a regular GG-space SS is said to be ergodic with coefficients in a Banach GG-module EE if any weak-* measurable GG-equivariant map f:SEf:S\to E is essentially constant. Note that this is a much stronger condition than ergodicity; it does not even hold for the transitive action of GG on itself. The action is called doubly ergodic with coefficients if the corresponding condition holds for the diagonal GG-action on S×SS\times S.

In our setting, the goal is to show that for fb((G/P)2,𝒱)Gf\in\mathcal{L}^{\infty}_{b}\left({}^{*}(G/P)^{2},\mathcal{V}\right)^{\sim{}^{*}G}, there exists F𝒱bF\in\mathcal{V}_{b} so that fFinf((G/P)2,𝒱)f-F\in\mathcal{L}^{\infty}_{inf}\left({}^{*}(G/P)^{2},\mathcal{V}\right). In other words, we would like to prove an asymptotic version of double ergodicity of the action of GG on G/PG/P with coefficients being the asymptotic Banach G{}^{*}G-module 𝒱\mathcal{V}. The crucial difference with 6.1.5 is that we involve the GG-structure on 𝒱\mathcal{V} this time.

Remark 6.1.7.

We caution the reader that no (non-trivial) action whatsoever can be ergodic with arbitrary coefficients. Therefore, the nature of the asymptotic GG-module 𝒱\mathcal{V} intervenes in the result. Specifically, the fact that it involves a finite GG-invariant measure on DD is used and it enters the proof through our appeal to the asymptotic Mautner’s lemma (Corollary 6.1.3), which ultimately relies on the L2L^{2}-topology used in Lemma 5.3.9. That L2L^{2}-continuity argument, introduced in §5.3, is not available in the absence of a finite invariant measure (and indeed the statements can fail in that absence).

We begin with a series of short claims that shall be used in the proof of Theorem 6.1.11. The first of these states that it is sufficient to work with fb((G/A),𝒱)Gf\in\mathcal{L}^{\infty}_{b}\left({}^{*}(G/A),\mathcal{V}\right)^{\sim{}^{*}G} for A=PwPw1A=P\cap wPw^{-1} for some wGw\in G.

Claim 6.1.8.

There exists wGw\in G such that for A=PwPw1A=P\cap wPw^{-1}, the map

G/AG/P×G/PG/A\to G/P\times G/P
gA(gP,gwP)gA\mapsto(gP,gwP)

is a measure space isomorphism respecting the action of GG.

Proof.

The result is a consequence of [9, Theorem 3.13 and Corollary 3.15]. ∎

The next claim is a special case of the asymptotic Mautner’s lemma Corollary 6.1.3 for AA.

Claim 6.1.9.

Let AA be as in 6.1.8. Then 𝒱bA=𝒱bG\mathcal{V}_{b}^{\sim{}^{*}A}=\mathcal{V}_{b}^{\sim{}^{*}G}.

Proof.

This is again a consequence of Corollary 6.1.2 and Corollary 5.3.12, since AA contains a maximal torus of GG. ∎

Now we obtain a correction of asymptotically G{}^{*}G-invariant maps to truly G{}^{*}G-invariant maps:

Claim 6.1.10.

Suppose fb(G,𝒱)f\in\mathcal{L}_{b}^{\infty}\left({}^{*}G,\mathcal{V}\right) is such that for every xGx\in{}^{*}G and almost every gGg\in{}^{*}G, f(xg)f(g)𝒱inff(xg)-f(g)\in\mathcal{V}_{inf}. Then there exists Fb(G,𝒱)F\in\mathcal{L}_{b}^{\infty}\left({}^{*}G,\mathcal{V}\right) such that Ffinf(G,𝒱)F-f\in\mathcal{L}_{inf}^{\infty}\left({}^{*}G,\mathcal{V}\right) and for every xGx\in{}^{*}G and almost every gGg\in{}^{*}G, F(xg)=F(g)F(xg)=F(g).

Proof.

The proof is the same as in Lemma 5.1.1, but now using the fact that (G,𝒱)\mathcal{L}^{\infty}\left({}^{*}G,\mathcal{V}\right) is the ultraproduct of relatively injective Banach GG-modules Lw(G,Lw(D,𝔲(kn)))L^{\infty}_{w*}(G,L^{\infty}_{w*}(D,\mathfrak{u}(k_{n}))). ∎

With the above statement at hand, we now prove that:

Theorem 6.1.11 (Double Ergodicity with coefficients).

Let GG be as in Corollary 6.1.2. For fb((G/P)2,𝒱)Gf\in\mathcal{L}^{\infty}_{b}\left({}^{*}(G/P)^{2},\mathcal{V}\right)^{\sim{}^{*}G}, there exists v𝒱bGv\in\mathcal{V}_{b}^{\sim{}^{*}G} so that fvinf((G/P)2,𝒱)f-v\in\mathcal{L}^{\infty}_{inf}\left({}^{*}(G/P)^{2},\mathcal{V}\right).

Proof.

From 6.1.8, we know that this is equivalent to proving that for fb((G/A),𝒱)Gf\in\mathcal{L}^{\infty}_{b}\left({}^{*}(G/A),\mathcal{V}\right)^{\sim{}^{*}G}, there exists v𝒱bv\in\mathcal{V}_{b} so that fvinf((G/A),𝒱)f-v\in\mathcal{L}^{\infty}_{inf}\left({}^{*}(G/A),\mathcal{V}\right). Define fb(G,𝒱)f^{\prime}\in\mathcal{L}^{\infty}_{b}\left({}^{*}G,\mathcal{V}\right) as f(g)πG(g)1f(gA)f^{\prime}(g)\coloneqq\pi_{G}(g)^{-1}f(g\,{}^{*}A). Since πG\pi_{G} is an asymptotic G{}^{*}G-action and ff is asymptotically G{}^{*}G-equivariant, for every xGx\in{}^{*}G and almost every gGg\in{}^{*}G,

f(xg)f(g)𝒱inff^{\prime}(xg)-f^{\prime}(g)\in\mathcal{V}_{inf}

In other words, ff^{\prime}, by construction, is asymptotically G{}^{*}G-invariant. From 6.1.10, there exists Fb(G,𝒱)F\in\mathcal{L}_{b}^{\infty}\left({}^{*}G,\mathcal{V}\right) such that Ffinf(G,𝒱)F-f\in\mathcal{L}_{inf}^{\infty}\left({}^{*}G,\mathcal{V}\right) and for every xGx\in{}^{*}G and almost every gGg\in{}^{*}G, F(xg)=F(g)F(xg)=F(g). Now from 6.1.6, FF is essentially constant, and hence there exists v𝒱bv\in\mathcal{V}_{b} such that for almost every gGg\in{}^{*}G, f(g)v𝒱inff^{\prime}(g)-v\in\mathcal{V}_{inf}. Thus, for almost every gGg\in{}^{*}G,

f(gA)πG(g)v𝒱inff(g{}^{*}A)-\pi_{G}(g)v\in\mathcal{V}_{inf}

We now claim that v𝒱bAv\in\mathcal{V}_{b}^{\sim{}^{*}A}. Then, by 6.1.9, we have v𝒱bGv\in\mathcal{V}_{b}^{\sim{}^{*}G}, allowing us to conclude that for almost every gGg\in{}^{*}G, f(gA)v𝒱inff(g{}^{*}A)-v\in\mathcal{V}_{inf}. Hence it is left to prove that v𝒱bAv\in\mathcal{V}_{b}^{\sim{}^{*}A}.
Define the pullback f^b(G,𝒱)\hat{f}\in\mathcal{L}^{\infty}_{b}\left({}^{*}G,\mathcal{V}\right) by f^(g)f(gA)\hat{f}(g)\coloneqq f(g{}^{*}A), and fix a measurable section s:G/AGs:G/A\to G to define a measure isomorphism

G/A×AGG/A\times A\to G
(gA,a)s(gA)a(gA,a)\mapsto s(gA)a

The above isomorphism can be defined internally to get an internal measure isomorphism (G/A)×AG{}^{*}(G/A)\times{}^{*}A\to{}^{*}G (with the internal section map which we continue denoting ss for simplicity), and we have that for almost every gA(G/A)g{}^{*}A\in{}^{*}(G/A) and almost every aAa\in{}^{*}A,

f^(s(gA)a)πG(s(gA)a)v𝒱inf\hat{f}(s(g{}^{*}A)a)-\pi_{G}(s(g{}^{*}A)a)v\in\mathcal{V}_{inf}

Since f^(s(gA)a)=f^(s(gA))\hat{f}(s(g{}^{*}A)a)=\hat{f}(s(g{}^{*}A)), the above simplifies to the following: for almost every gA(G/A)g{}^{*}A\in{}^{*}(G/A) and almost every aAa\in{}^{*}A

π(s(gA))vπG(s(gA)a)v𝒱inf\pi(s(g{}^{*}A))v-\pi_{G}(s(g{}^{*}A)a)v\in\mathcal{V}_{inf}

Now we fix some gA(G/A)g{}^{*}A\in{}^{*}(G/A) so that for this gAg{}^{*}A, we have that for almost every aAa\in{}^{*}A,

πG(s(gA))vπG(s(gA)a)v𝒱inf\pi_{G}(s(g{}^{*}A))v-\pi_{G}(s(g{}^{*}A)a)v\in\mathcal{V}_{inf}

Since πG\pi_{G} is an asymptotic action of G{}^{*}G, πG(s(gA)a)vπG(s(gA))πG(a)v𝒱inf\pi_{G}(s(g{}^{*}A)a)v-\pi_{G}(s(g{}^{*}A))\pi_{G}(a)v\in\mathcal{V}_{inf}. And so, we conclude that, for almost every aAa\in{}^{*}A,

vπG(a)v𝒱infv-\pi_{G}(a)v\in\mathcal{V}_{inf}

This means that there exists an internal subset 𝒜={An}𝒰\mathcal{A}=\{A^{\prime}_{n}\}_{\mathcal{U}} of A{}^{*}A with AnA^{\prime}_{n} being a conull subset of AA for n𝒰n\in\mathcal{U}, such that v𝒱b𝒜v\in\mathcal{V}_{b}^{\sim\mathcal{A}}. But note that 𝒜𝒜=A\mathcal{A}\cdot\mathcal{A}={}^{*}A (since if AnA_{n}^{\prime} is a co-null subset of the locally compact group AA, then AnAn=AA_{n}^{\prime}A_{n}^{\prime}=A). Hence v𝒱bAv\in\mathcal{V}_{b}^{\sim{}^{*}A} as claimed. ∎

Theorem 6.1.11 tells us that for any fb((G/P)2,𝒱)Gf\in\mathcal{L}^{\infty}_{b}\left({}^{*}(G/P)^{2},\mathcal{V}\right)^{\sim{}^{*}G} the induced map f~~((G/P)2,𝒱)G\tilde{f}\in\tilde{\mathcal{L}}^{\infty}\left({}^{*}(G/P)^{2},\mathcal{V}\right)^{{}^{*}G} is essentially constant. This is the asymptotic analogue of the classical result that the GG-action on G/PG/P is doubly ergodic with coefficients in a suitable GG-module.
We conclude this subsection by showing that Ha2(Q,𝒱)=0\operatorname{H}_{a}^{2}(Q,\mathcal{V})=0 for a proper parabolic subgroup QQ of GG as in Corollary 6.1.2 such that Hb2(Q,𝐑)=0\operatorname{H}_{b}^{2}(Q,\mathbf{R})=0 and Hb3(Q,𝐑)\operatorname{H}_{b}^{3}(Q,\mathbf{R}) is Hausdorff. We begin with the following proposition that is an application of Corollary 6.1.3:

Proposition 6.1.12.

Let GG and a parabolic subgroup QGQ\leq G be as in Corollary 6.1.2. Let ωb((Q)3,𝒱)Q\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} be an asymptotic 22-cocycle for QQ with coefficients in 𝒱\mathcal{V}. Then there exists an asymptotic 22-cocycle ωb((Q)3,𝒱)Q\omega^{\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} for QQ that takes values in 𝒱bG\mathcal{V}_{b}^{\sim{}^{*}G} such that

ω~ω~=d~1α1~\tilde{\omega}-\tilde{\omega}^{\prime}=\tilde{d}^{1}\tilde{\alpha_{1}}

where α1b((Q)2,𝒱)Q\alpha_{1}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{2},\mathcal{V})^{\sim{}^{*}Q}.

Proof.

Let NN be the unipotent radical of QQ, which is a normal amenable closed subgroup of QQ. Since NN is amenable, Theorem 4.3.3 tells us that Ha(Q,𝒱)\operatorname{H}_{a}^{\bullet}(Q,\mathcal{V}) can be computed as the asymptotic cohomology of the asymptotic Q{}^{*}Q-cochain complex

0{0}((Q/N),𝒱){\mathcal{L}^{\infty}({}^{*}(Q/N),\mathcal{V})}(((Q/N))2,𝒱){\mathcal{L}^{\infty}(({}^{*}(Q/N))^{2},\mathcal{V})}(((Q/N))3,𝒱){\mathcal{L}^{\infty}(({}^{*}(Q/N))^{3},\mathcal{V})}{\dots}d0\scriptstyle{d^{0}}d1\scriptstyle{d^{1}}d2\scriptstyle{d^{2}}d3\scriptstyle{d^{3}}

Let k:((Q),𝒱)(((Q/N)),𝒱)k^{\bullet}:\mathcal{L}^{\infty}(({}^{*}Q)^{\bullet},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}(Q/N))^{\bullet},\mathcal{V}) and j:(((Q/N)),𝒱)((Q),𝒱)j^{\bullet}:\mathcal{L}^{\infty}(({}^{*}(Q/N))^{\bullet},\mathcal{V})\to\mathcal{L}^{\infty}(({}^{*}Q)^{\bullet},\mathcal{V}) be the asymptotic Q{}^{*}Q-homotopy equivalences. Consider k2ωb(((Q/N))3,𝒱)Qk^{2}\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}(Q/N))^{3},\mathcal{V})^{\sim{}^{*}Q}. Since NN is normal in QQ, Im(k2ω)𝒱bNIm(k^{2}\omega)\subseteq\mathcal{V}_{b}^{\sim{}^{*}N}. By Corollary 6.1.3, since 𝒱bN=𝒱bG\mathcal{V}_{b}^{\sim{}^{*}N}=\mathcal{V}_{b}^{\sim{}^{*}G}, this implies that Im(k2ω)𝒱bGIm(k^{2}\omega)\subseteq\mathcal{V}_{b}^{\sim{}^{*}G}. The conclusion then follows from (Eq. 4.4). ∎

Thus, the obstacle to the asymptotic 22-cocycle ωb((Q)3,𝒱)Q\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} being an asymptotic 22-coboundary is the asymptotic 22-cocycle ωb((Q)3,𝒱)Q\omega^{\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} that takes values in 𝒱bG\mathcal{V}_{b}^{\sim{}^{*}G}. We shall now see how to handle this using assumptions on QQ.
Consider the asymptotic 22-cocycle ωb((Q)3,𝒱)Q\omega^{\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} for QQ that takes values in 𝒱G\mathcal{V}^{\sim{}^{*}G}. The first step is to correct ω\omega^{\prime} to an element ω′′b((Q)3,𝒲)Q\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{W})^{{}^{*}Q} with Im(ω′′)𝒲bΓIm(\omega^{\prime\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}. Recall the internal map θ:(G,𝒲)(D,𝒲)\theta:\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})\to\mathcal{L}^{\infty}({}^{*}D,\mathcal{W}) as defined in Eq. 5.1.

Lemma 6.1.13.

Given ωb((Q)3,𝒱)Q\omega^{\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} with Im(ω)𝒱GIm(\omega^{\prime})\subseteq\mathcal{V}^{\sim{}^{*}G}, there exists

ω′′b((Q)3,(G,𝒲))Q\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{L}^{\infty}({}^{*}G,\mathcal{W}))^{{}^{*}Q}

with Im(ω′′)((G,𝒲)Γ)G=𝒲bΓIm(\omega^{\prime\prime})\subseteq\left(\mathcal{L}^{\infty}({}^{*}G,\mathcal{W})^{\sim{}^{*}\Gamma}\right)^{{}^{*}G}=\mathcal{W}_{b}^{\sim{}^{*}\Gamma} such that ωθω′′inf((Q)3,𝒱)\omega^{\prime}-\theta\cdot\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{inf}(({}^{*}Q)^{3},\mathcal{V}).

Proof.

This follows from Remark 5.2.4 and Proposition 5.1.6. ∎

Now consider ω′′b((Q)3,𝒲)Q\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{W})^{{}^{*}Q} with Im(ω′′)𝒲bΓIm(\omega^{\prime\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}, and observe that

d3ω′′inf((Q)4,𝒲)d^{3}\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{inf}\left(({}^{*}Q)^{4},\mathcal{W}\right)

That is, ω′′\omega^{\prime\prime} can be thought of as an almost 22-cocycle for QQ with coefficients in 𝒲\mathcal{W} (with a trivial action of Q{}^{*}Q). The underlying idea is that we are now within the domain of classical bounded cohomology of QQ with coefficients in a trivial QQ-module, and can apply the results of Corollary 4.2.14.

Lemma 6.1.14.

Suppose Hb2(Q,𝐑)=0\operatorname{H}_{b}^{2}(Q,\mathbf{R})=0 and Hb3(Q,𝐑)\operatorname{H}_{b}^{3}(Q,\mathbf{R}) is Hausdorff. Then there exists αb((Q)2,𝒲)Q\alpha^{\prime}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}Q)^{2},\mathcal{W}\right)^{{}^{*}Q} with Im(α)𝒲bΓIm(\alpha^{\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma} and d2αω′′inf((Q)3,𝒲)d^{2}\alpha^{\prime}-\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{inf}\left(({}^{*}Q)^{3},\mathcal{W}\right).

Proof.

From Corollary 4.2.14, we know that there exists αb((Q)2,𝒲)Q\alpha^{\prime}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}Q)^{2},\mathcal{W}\right)^{{}^{*}Q} such that d2αω′′inf((Q)3,𝒲)d^{2}\alpha^{\prime}-\omega^{\prime\prime}\in\mathcal{L}^{\infty}_{inf}\left(({}^{*}Q)^{3},\mathcal{W}\right). We only need to show that Im(α)𝒲bΓIm(\alpha^{\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}. That is, we want to show that for an internal cochain αb((Q)2,𝒲)Q\alpha^{\prime}\in\mathcal{L}^{\infty}_{b}\left(({}^{*}Q)^{2},\mathcal{W}\right)^{{}^{*}Q}, if Im(d2α)𝒲bΓIm(d^{2}\alpha^{\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}, then Im(α)𝒲bΓIm(\alpha^{\prime})\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}.
Equivalently, consider the inhomogenous cochain (refer Remark 4.2.7) corresponding to α\alpha^{\prime}, namely fb(Q,𝒲)f\in\mathcal{L}^{\infty}_{b}\left({}^{*}Q,\mathcal{W}\right) defined as follows: for gQg\in{}^{*}Q, f(g)f(g) is the essential value α(x,xg)\alpha^{\prime}(x,xg). Clearly, Im(d2α)=Im(δ1f)Im(d^{2}\alpha^{\prime})=Im(\delta^{1}f) and Im(α)=Im(f)Im(\alpha^{\prime})=Im(f). Hence it is sufficient to show that if fb(Q,𝒲)f\in\mathcal{L}^{\infty}_{b}\left({}^{*}Q,\mathcal{W}\right) such that Im(δ1f)𝒲bΓIm(\delta^{1}f)\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}, then Im(f)𝒲bΓIm(f)\subseteq\mathcal{W}_{b}^{\sim{}^{*}\Gamma}. Note that 𝒲~Γ\tilde{\mathcal{W}}^{{}^{*}\Gamma} is a closed subspace of the real Banach space 𝒲~\tilde{\mathcal{W}}, hence f~~(Q,𝒲~)\tilde{f}\in\tilde{\mathcal{L}}^{\infty}({}^{*}Q,\tilde{\mathcal{W}}) is such that δ~1f~\tilde{\delta}^{1}\tilde{f} is 0 in the quotient Banach space 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma}. We would like to show that f~\tilde{f} is 0 in 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma}.
Consider the image, denoted SS, of Im(f)𝒲bIm(f)\subseteq\mathcal{W}_{b} in the quotient 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma} with the natural map 𝒲b𝒲~/𝒲~Γ\mathcal{W}_{b}\to\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma} denoted ww~w\mapsto\tilde{w}^{\prime}. Let C=sup{v:vS}C=\sup\{\|v\|:v\in S\}. Since fb(Q,𝒲)f\in\mathcal{L}^{\infty}_{b}\left({}^{*}Q,\mathcal{W}\right), we know that C𝐑C\in\mathbf{R}. Let wIm(f)w\in Im(f) be such that w~0.9C\|\tilde{w}^{\prime}\|\geq 0.9C, and consider the ball B0.1C(w~)B_{0.1C}(\tilde{w}^{\prime}) of radius 0.1C0.1C around w~\tilde{w}^{\prime}, and let Y={Yn}𝒰QY=\{Y_{n}\}_{\mathcal{U}}\subseteq{}^{*}Q be an internal subset of Q{}^{*}Q of positive (internal) measure such that the image of f~(Y)\tilde{f}(Y) is in B0.1C(w~)B_{0.1C}(\tilde{w}^{\prime}) (such an internal subset YY of positive measure exists because the image in 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma} of the internal ball B0.1C(w)𝒲bB_{0.1C}(w)\subseteq\mathcal{W}_{b} is contained in B0.1C(w~)B_{0.1C}(\tilde{w}^{\prime}), and f1(B0.1C(w))f^{-1}(B_{0.1C}(w)) is of positive measure).
Since δ~1f~\tilde{\delta}^{1}\tilde{f} is 0 in the quotient Banach space 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma}, we know that there exists an internal subset A={An}𝒰Q×QA=\{A_{n}\}_{\mathcal{U}}\subseteq{}^{*}Q\times{}^{*}Q, where AnA_{n} is co-null in Q×QQ\times Q for n𝒰n\in\mathcal{U}, such that for (g,h)A(g,h)\in A, f~(g)+f~(h)f~(gh)𝒲~Γ\tilde{f}(g)+\tilde{f}(h)-\tilde{f}(gh)\in\tilde{\mathcal{W}}^{{}^{*}\Gamma}. Note that AY×YA\cap Y\times Y, denoted DD, is an internal subset of positive (internal) measure in Q×Q{}^{*}Q\times{}^{*}Q, and so is the product DD{xy:(x,y)D}QD\cdot D\coloneqq\{xy:(x,y)\in D\}\subseteq{}^{*}Q. It follows that the image v𝒲~/𝒲~Γv\in\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma} of f~(g)\tilde{f}(g) has norm v1.6C\|v\|\geq 1.6C, for gDDg\in D\cdot D. This implies that C=0C=0, allowing us to conclude that S𝒲~ΓS\subseteq\tilde{\mathcal{W}}^{{}^{*}\Gamma} and hence f~\tilde{f} is 0 in 𝒲~/𝒲~Γ\tilde{\mathcal{W}}/\tilde{\mathcal{W}}^{{}^{*}\Gamma}. ∎

We now combine the results above to conclude that Ha2(Q,𝒱)=0\operatorname{H}_{a}^{2}(Q,\mathcal{V})=0.

Theorem 6.1.15.

Let GG and a parabolic subgroup QGQ\leq G be as in Corollary 6.1.2, and suppose Hb3(Q,𝐑)\operatorname{H}_{b}^{3}(Q,\mathbf{R}) is Hausdorff and Hb2(Q,𝐑)=0\operatorname{H}_{b}^{2}(Q,\mathbf{R})=0. Then Ha2(Q,𝒱)=0\operatorname{H}_{a}^{2}(Q,\mathcal{V})=0.

Proof.

From Proposition 6.1.12, we know that exists an asymptotic 22-cocycle ω1b((Q)3,𝒱)Q\omega_{1}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{3},\mathcal{V})^{\sim{}^{*}Q} for QQ that takes values in 𝒱G\mathcal{V}^{\sim{}^{*}G} such that ω~ω~=d~1α~1\tilde{\omega}-\tilde{\omega}^{\prime}=\tilde{d}^{1}\tilde{\alpha}_{1} where α1b((Q)2,𝒱)Q\alpha_{1}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{2},\mathcal{V})^{\sim{}^{*}Q}. From Lemma 6.1.13 and Lemma 6.1.14, we conclude that there exists α1b((Q)2,𝒱)Q\alpha_{1}^{\prime}\in\mathcal{L}^{\infty}_{b}(({}^{*}Q)^{2},\mathcal{V})^{\sim{}^{*}Q} such that ω~=d~1α1~\tilde{\omega}^{\prime}=\tilde{d}^{1}\tilde{\alpha_{1}^{\prime}}. We now set α=α1+α1\alpha=\alpha_{1}+\alpha_{1}^{\prime}. ∎

6.2 Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) and the Main Theorem

Recall that the results of §6.1 assumed the existence of a parabolic subgroup Q=i=1kQiQ=\prod_{i=1}^{k}Q_{i} (where each QiQ_{i} is a proper parabolic subgroup of 𝐆i(Ki){\mathbf{G}_{i}}(K_{i})), satisfying the conditions that Hb3(Q,𝐑)\operatorname{H}_{b}^{3}(Q,\mathbf{R}) is Hausdorff and Hb2(Q,𝐑)=0\operatorname{H}_{b}^{2}(Q,\mathbf{R})=0. This allowed us to prove that Ha2(Q,𝒱)\operatorname{H}_{a}^{2}(Q,\mathcal{V}) vanishes. This motivates the following definition:

Definition 6.2.1.

A locally compact group GG has the 2½-property if Hb2(G,𝐑)\operatorname{H}_{b}^{2}(G,\mathbf{R}) vanishes and Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is Hausdorff.

In order to demystify this definition, we should consider that it is a natural strengthening of the vanishing of Hb2(G,𝐑)\operatorname{H}_{b}^{2}(G,\mathbf{R}). Indeed, recall that Hb1(G,𝐑)\operatorname{H}_{b}^{1}(G,\mathbf{R}) always vanishes and that the Hausdorff condition on Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) means that the differential map on two-cochains is an open map. Thus the 2½-property states that the augmented differential complex computing bounded cohomology starts of as an exact sequence up to degree two and retains a weaker consequence of exactness in degree three, namely that the differential is open (see [35] for a detailed discussion of the openness of differentials in chain complexes of Banach spaces).

Proposition 6.2.2.

The 2½-property is preserved under extensions, and hence in particular, under finite direct products.

Proof.

This follows from the Hochschild-Serre spectral sequence as set up in  [37, §12]. Specifically, the Hausdorff assumption allows us to apply Proposition 12.2.2 in [37, §12]. ∎

Furthermore, the results on Hochschild-Serre spectral sequences in [37, §12] imply that a direct product of groups has the 2½-property iff each factor has the 2½-property.
We recall that elementary properties of bounded cohomology allow us to disregard “amenable pieces” when it comes to the 2½-property:

Lemma 6.2.3.
  1. 1.

    Suppose that G~\widetilde{G} is an extension of GG by an amenable kernel, for instance a central extension of GG. Then G~\widetilde{G} has the 2½-property if and only if GG does.

  2. 2.

    Suppose that G1<GG_{1}<G is a closed co-amenable subgroup of GG, for instance a closed normal subgroup with amenable quotient. If G1G_{1} has the 2½-property, then so does GG.

Proof.

The first point follows from the fact that the inflation map Hb(G,𝐑)Hb(G~,𝐑)\operatorname{H}_{b}^{\bullet}(G,\mathbf{R})\to\operatorname{H}_{b}^{\bullet}(\widetilde{G},\mathbf{R}) is an isometric isomorphism, see e.g. [37, 8.5.2]. For the second point, we recall that the restriction map Hb(G,𝐑)Hb(G1,𝐑)\operatorname{H}_{b}^{\bullet}(G,\mathbf{R})\to\operatorname{H}_{b}^{\bullet}(G_{1},\mathbf{R}) is isometrically injective, see e.g. [37, 8.6.6]. ∎

So in Eq. 4.4, we considered a semisimple group G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) (where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected, simply connected, almost KiK_{i}-simple group) and showed that Ha2(Q,𝒱)=0\operatorname{H}_{a}^{2}(Q,\mathcal{V})=0 for a parabolic subgroup Q=i=1kQiQ=\prod_{i=1}^{k}Q_{i} (where each QiQ_{i} is a proper parabolic subgroup of 𝐆i(Ki){\mathbf{G}_{i}}(K_{i})) assuming QQ has the 2½-property. To use this to prove that Ha2(G,𝒱)=0\operatorname{H}_{a}^{2}(G,\mathcal{V})=0, we need the existence of two such parabolic subgroups Q1Q_{1} and Q2Q_{2} both containing PP and boundedly generating GG. This is described in the following definition:

Definition 6.2.4.

Let G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) be a semisimple group (where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected almost KiK_{i}-simple group), and let PP be a minimal parabolic subgroup. Then GG is said to have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) if there exist two parabolic subgroups Q1Q_{1} and Q2Q_{2} of GG satisfying the following properties:

  • Both Q1Q_{1} and Q2Q_{2} are of the form Q1=i=1kQ1,iQ_{1}=\prod_{i=1}^{k}Q_{1,i} and Q2=i=1kQ2,iQ_{2}=\prod_{i=1}^{k}Q_{2,i} with Q1,iQ_{1,i} and Q2,iQ_{2,i} being proper parabolic subgroups of 𝐆i(Ki)\mathbf{G}_{i}(K_{i}) for 1ik1\leq i\leq k.

  • Both Q1Q_{1} and Q2Q_{2} have the 2½-property.

  • The intersection Q1Q2Q_{1}\cap Q_{2} is a parabolic subgroup that contains the minimal parabolic subgroup PP of GG.

  • The group GG is boundedly generated by the union of Q1Q_{1} and Q2Q_{2}.

Observe that the existence of two distinct proper parabolic subgroups immediately implies that if GG has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}), then each of its simple factors has rank at least 22. We shall see explicit examples of groups with the 2½-property and Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) in §6.2.

Theorem 6.2.5.

Let Γ\Gamma be a lattice in a semisimple group G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) (where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected, simply connected, almost KiK_{i}-simple group) that has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}). Then Ha2(Γ,𝒲)=0\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=0.

Proof.

By Theorem 5.2.6, Ha2(Γ,𝒲)=Ha2(G,𝒱)\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})=\operatorname{H}_{a}^{2}(G,\mathcal{V}). Consider an asymptotic 22-cocycle

ωb(((G/P))2,𝒱)G\omega\in\mathcal{L}^{\infty}_{b}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}G}

By Theorem 6.1.15, there exist α1b(((G/P))2,𝒱)Q1\alpha_{1}\in\mathcal{L}_{b}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}Q_{1}} and α2b(((G/P))2,𝒱)Q2\alpha_{2}\in\mathcal{L}_{b}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}Q_{2}} such that

ω~=d~2α~1=d~2α~2\tilde{\omega}=\tilde{d}^{2}\tilde{\alpha}_{1}=\tilde{d}^{2}\tilde{\alpha}_{2}

Since PQ1Q2P\subseteq Q_{1}\cap Q_{2}, note that α1α2b(((G/P))2,𝒱)P\alpha_{1}-\alpha_{2}\in\mathcal{L}_{b}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}P} is an asymptotic 11-cocyle for PP. As PP is amenable, Ha1(P,𝒱)=0\operatorname{H}_{a}^{1}(P,\mathcal{V})=0, and so there exists βb((G/P),𝒱)P\beta\in\mathcal{L}_{b}^{\infty}({}^{*}(G/P),\mathcal{V})^{\sim{}^{*}P} such that

α~1α~2=d~1β~\tilde{\alpha}_{1}-\tilde{\alpha}_{2}=\tilde{d}^{1}\tilde{\beta}

We now use β\beta to define the internal map

β1:((G/P))2𝒱\beta_{1}:({}^{*}(G/P))^{2}\to\mathcal{V}
β1(gP,hP)πG(g)β(g1hP)\beta_{1}(g{}^{*}P,h{}^{*}P)\coloneqq\pi_{G}(g)\beta(g^{-1}h{}^{*}P)

Note that β1b(((G/P))2,𝒱)G\beta_{1}\in\mathcal{L}_{b}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}G}, and by Theorem 6.1.11, there exists F𝒱bGF\in\mathcal{V}_{b}^{\sim{}^{*}G} so that β1Finf(((G/P))2,𝒱)\beta_{1}-F\in\mathcal{L}_{inf}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V}). In particular, the same holds for β\beta as well: βFinf((G/P),𝒱)\beta-F\in\mathcal{L}_{inf}^{\infty}({}^{*}(G/P),\mathcal{V}), implying that

d~1β~=0\tilde{d}^{1}\tilde{\beta}=0

This means that α~1=α~2\tilde{\alpha}_{1}=\tilde{\alpha}_{2}. Setting αα1\alpha\coloneq\alpha_{1}, note that α\alpha is both asymptotically Q1{}^{*}Q_{1}-equivariant and asymptotically Q2{}^{*}Q_{2}-equivariant. Since GG is boundedly generated by elements of Q1Q_{1} and Q2Q_{2}, this implies that αb(((G/P))2,𝒱)G\alpha\in\mathcal{L}_{b}^{\infty}(({}^{*}(G/P))^{2},\mathcal{V})^{\sim{}^{*}G}. ∎

While the above theorem assumes that Γ\Gamma is a lattice in a semisimple group GG that is simply connected, we can extend this as follows:

Theorem 6.2.6.

Let Γ\Gamma be a lattice in a semisimple group G=i=1k𝐆i(Ki)G=\prod^{k}_{i=1}{\mathbf{G}_{i}}(K_{i}) (where for 1ik1\leq i\leq k, KiK_{i} is a local field, and 𝐆i\mathbf{G}_{i} is a connected almost KiK_{i}-simple group) that has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}). Then Γ\Gamma is uniformly 𝔘\mathfrak{U}-stable with a linear estimate.

Proof.

Let G^\hat{G} be the universal cover of GG, which is a semisimple simply connected group, and let Γ^\hat{\Gamma} be the preimage of Γ\Gamma in G^\hat{G}. Note that Γ\Gamma and Γ^\hat{\Gamma} are commensurable, hence Remark 3.0.3 tells us that it is sufficient to prove that Γ^\hat{\Gamma} is is uniformly 𝔘\mathfrak{U}-stable with a linear estimate. Note that since the fundamental group of GG is abelian, from Lemma 6.2.3 it follows that G^\hat{G} has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}). Hence we can now apply Theorem 6.2.5 to Γ^G^\hat{\Gamma}\leq\hat{G} to conclude that Ha2(Γ^,𝒲)=0\operatorname{H}_{a}^{2}(\hat{\Gamma},\mathcal{W})=0, implying that Γ^\hat{\Gamma} is uniformly 𝔘\mathfrak{U}-stable with a linear estimate. ∎

6.3 Groups with Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2})

In this subsection, we shall list classes of groups that satisfy Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) used in the hypothesis of Theorem 6.2.5. Since Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) involves the existence of two proper parabolic subgroups with the 2½-property, we first list classes of groups known to have this property.

Simple Groups with the 2½-property

We shall collect below the necessary statement from the existing literature, and complement them by an additional argument in the non-Archimedan case, to establish that a number of natural semisimple groups have the 2½-property.
Let us first consider simple groups over a non-archimedean field. Here, the vanishing of Hb2(G,𝐑)\operatorname{H}_{b}^{2}(G,\mathbf{R}) was established in [13]. More precisely, this reference establishes the injectivity of the comparison map (and we recall that the case of trivial coefficients 𝐑\mathbf{R} for the semisimple group GG was actually the easy part of this result). On the other hand, the vanishing of the usual cohomology is well-known in this setting, see e.g. [10].
Therefore, what we need to justify is the condition on Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}). It turns out that this vanishes: this result is established by the inductive method introduced in [40], the basis of the induction being provided by [12].

Theorem 6.3.1.

For G=𝐆(K)G=\mathbf{G}(K), where KK is a non-Archimedean local field and 𝐆\mathbf{G} is a connected, simply connected, semisimple KK-group, the bounded cohomology Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) vanishes (and is therefore Hausdorff).

Proof.

The proof is by induction on the KK-rank of 𝐆\mathbf{G}. The case of rank zero corresponds to G=𝐆(K)G=\mathbf{G}(K) being compact, and thus having trivial bounded cohomology in all degrees. The induction really starts with rank one. In that case, GG is an automorphism group of a Bruhat–Tits tree satisfying the assumptions of the main result of [12], which states that Hbn(G,𝐑)\operatorname{H}_{b}^{n}(G,\mathbf{R}) vanishes for all n>0n>0.
We now perform the induction step for 𝐆\mathbf{G} of KK-rank r2r\geq 2. We use (a minor variation of) the spectral sequence introduced in [40, §6.A] as follows. We consider the Tits building TT of 𝐆\mathbf{G} over KK and denote by T(d)T^{(d)} its set of dd-simplices. Thus T(d)T^{(d)} is defined for all dr1d\leq r-1 and is topologized by identifying it with the union of homogeneous spaces G/PIG/P_{I}, where PIP_{I} ranges over standard parabolic subgroups of semisimple rank r1dr-1-d (see  [40] for more details). By convention, we take the augmented simplicial complex, namely we also consider the one-point space of negative simplices T(1)=G/GT^{(-1)}=G/G. In this set-up, an appropriate version of the Solomon-Tits theorem implies that the following sequence of spaces of continuous functions is exact:

0C(T(1))C(T(0))C(T(r1))0\to C(T^{(-1)})\to C(T^{(0)})\to\cdots\to C(T^{(r-1)})

see Theorems 2.7 and 3.9 in [40]. We denote by StG\mathrm{St}_{G} the cokernel of the last map above, which is one version of the Steinberg representation of GG. Finally, the spectral sequence that we consider is the first quadrant hypercohomology spectral sequence obtained by computing the bounded cohomology of GG with coefficients in the following complex of Banach GG-modules

0C(T(1))C(T(0))C(T(r1))StG00\to C(T^{(-1)})\to C(T^{(0)})\to\cdots\to C(T^{(r-1)})\to\mathrm{St}_{G}\to 0 (6.1)

Concretely, the first page of the spectral sequence is by definition E1p,q=Hbq(G,C(T(p1)))E^{p,q}_{1}=\operatorname{H}_{b}^{q}\left(G,C(T^{(p-1)})\right) when prp\leq r, for p=r+1p=r+1 it is E1r+1,q=Hbq(G,StG)E^{r+1,q}_{1}=\operatorname{H}_{b}^{q}(G,\mathrm{St}_{G}), and for p>r+1p>r+1 it is zero. The only difference with [40, §6.A] is that the complex was truncated at T(r1)T^{(r-1)} there, not completing it with StG\mathrm{St}_{G}. A crucial point to allow this definition is that StG\mathrm{St}_{G} is Hausdorff, which follows from an alternative description of the Steinberg representation given by Borel–Serre, see Remark 2.8 in [40].

This spectral sequence abuts to zero since the above complex is exact. By cohomological induction, we have as in Theorem 6.1 of [40] the identifications

E1p,q=Hbq(G,C(T(p1)))Hbq(GI,𝐑)(pr,q)E^{p,q}_{1}=\operatorname{H}_{b}^{q}\left(G,C(T^{(p-1)})\right)\cong\bigoplus\operatorname{H}_{b}^{q}(G_{I},\mathbf{R})\kern 19.91692pt(\forall\,p\leq r,\forall q) (6.2)

where GIG_{I} is the semisimple part of the parabolic subgroup PIP_{I} and the sum is taken over all such subgroups of semisimple rank rpr-p. Notice that the rank is indeed rpr-p and not r1pr-1-p since we started with T(1)T^{(-1)} corresponding to p=0p=0.

Our goal is to prove E10,3=0E^{0,3}_{1}=0 since this is Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}). We have E10,3=E20,3E^{0,3}_{1}=E^{0,3}_{2} since the right hand side is the kernel of the map E10,3E11,3E^{0,3}_{1}\to E^{1,3}_{1} which vanishes by Eq. 6.2 and the inductive hypothesis. Next, E20,3=E30,3E^{0,3}_{2}=E^{0,3}_{3} because the right hand side is the kernel of the map E20,3E22,2E^{0,3}_{2}\to E^{2,2}_{2} and already E12,2E^{2,2}_{1} vanishes by Eq. 6.2 and the general vanishing of Hb2\operatorname{H}_{b}^{2} mentioned earlier.

The next differential to consider is E30,3E33,1E^{0,3}_{3}\to E^{3,1}_{3}. If r3r\geq 3, we can still apply Eq. 6.2 and the general vanishing of Hb1\operatorname{H}_{b}^{1} to conclude E30,3=E40,3E^{0,3}_{3}=E^{0,3}_{4}. If however r=2r=2, then we reach the same conclusion provided we justify that Hb1(G,StG)\operatorname{H}_{b}^{1}(G,\mathrm{St}_{G}) vanishes. To this end, we first observe that the comparison map to ordinary cohomology is always injective in degree one. On the other hand, the ordinary first cohomology of GG with values in the Steinberg representation is known to vanish if (and only if!) the rank of GG is not one, see e.g. Theorem 4.12 p. 205 in [10]. Thus, in either case we have established E30,3=E40,3E^{0,3}_{3}=E^{0,3}_{4}.

The final differential is E40,3E44,0E^{0,3}_{4}\to E^{4,0}_{4}. But we have already second page vanishing of all E2p,0E^{p,0}_{2}. Indeed, by Eq. 6.2 this statement amounts to the acyclicity of the subcomplex of GG-invariants of Eq. 6.1. Ignoring the last term (StG)G(\mathrm{St}_{G})^{G} at first, this comes from the fact that the GG-orbits in the Tits building form by definition a full simplex of dimension r1r-1 (augmented in dimension 1-1), which is hence acyclic. It remains to argue acyclicity of the last term

C(T(r1))G(StG)G0C(T^{(r-1)})^{G}\to(\mathrm{St}_{G})^{G}\to 0

which amounts to showing that (StG)G(\mathrm{St}_{G})^{G} vanishes. This follows e.g. by realizing the Steinberg representation as a space of L2L^{2} harmonic maps on the Bruhat–Tits building, see [32] for a concrete description of this isomorphism (also due to Borel–Serre).

In conclusion, we have shown that E10,3=E40,3=E0,3E^{0,3}_{1}=E^{0,3}_{4}=E^{0,3}_{\infty} holds. Since the spectral sequence converges to zero, this establishes as desired the vanishing of E10,3=Hb3(G,𝐑)E^{0,3}_{1}=\operatorname{H}_{b}^{3}(G,\mathbf{R}). ∎

Next, let us consider simple groups over 𝐑\mathbf{R} or 𝐂\mathbf{C}. Regarding Hb2(G,𝐑)\operatorname{H}_{b}^{2}(G,\mathbf{R}), the injectivity of the comparison map was established for all connected semisimple Lie groups GG in [13]. Therefore, the corresponding vanishing holds in all cases where the usual second bounded cohomology vansihes, which is always the case for G=SLn(𝐂)G=SL_{n}(\mathbf{C}) and is the case for G=SLn(𝐑)G=SL_{n}(\mathbf{R}) if and only if n2n\neq 2. It therefore remains to collect the following results from the existing literature:

Theorem 6.3.2.

Let n2n\geq 2.

  1. 1.

    For G=SLn(𝐑)G=SL_{n}(\mathbf{R}), the bounded cohomology Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) vanishes (and is therefore Hausdorff).

  2. 2.

    For G=SLn(𝐂)G=SL_{n}(\mathbf{C}), the bounded cohomology Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) is one-dimensional and Hausdorff.

Proof.

The case of SLn(𝐑)SL_{n}(\mathbf{R}) was established for n=2n=2 in Theorem 1.5 of [14] and for general nn in Theorem 1.2 of [38].

Regarding G=SLn(𝐂)G=SL_{n}(\mathbf{C}), these two references established that the comparison map is an isomorphism from Hb3(G,𝐑)\operatorname{H}_{b}^{3}(G,\mathbf{R}) to H3(G,𝐑)H^{3}(G,\mathbf{R}), the latter being classically known to be one-dimensional (see [14, Theorem 1.2] and [38, Remark 3.5]). In our context, the only relevant point (and the hard one anyway) is that the comparison map is injective. Indeed, it is a general fact for any group and any degree dd that the injectivity of the comparison map HbdHd\operatorname{H}_{b}^{d}\to H^{d} implies that Hbd\operatorname{H}_{b}^{d} is Hausdorff. This can be seen by combining Theorems 2.3 and 2.8 in [35]. ∎

We now summarize the above results with the following list of simple groups that we know to have the 2½-property:

Theorem 6.3.3.

The following groups GG have the 2½-property:

  1. 1.

    G=𝐆(k)G=\mathbf{G}(k), where kk is a non-Archimedean local field and 𝐆\mathbf{G} is a connected semisimple kk-group.

  2. 2.

    G=SLn(𝐑)G=SL_{n}(\mathbf{R}) for any n2n\neq 2.

  3. 3.

    G=SLn(𝐂)G=SL_{n}(\mathbf{C}) for any nn.

From the 2½-property to Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2})

We can now list out simple groups having Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) using the results discussed earlier in this section. Again, we begin with simple groups, and then extend the results to semisimple groups using Proposition 6.2.2.
Let us first consider the case of a simple group GG over a non-archimedean field, and QGQ\leq G be a parabolic subgroup. Note that, using Theorem 6.3.3, Lemma 6.2.3 and Proposition 6.2.2, we can conclude that QQ has the 2½-property (since, modulo its amenable radical, it is a semisimple group). Thus, since GG has rank at least 22, we can always find two proper parabolic subgroups, both having the 2½-property, generating GG. Hence,

Proposition 6.3.4.

The group G=𝐆(k)G=\mathbf{G}(k), where kk is a non-Archimedean local field and 𝐆\mathbf{G} is a connected, simply connected, semisimple kk-group of rank at least 22, has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

In the complex case, since we know that SLn(𝐂)SL_{n}(\mathbf{C}) has the 2½-property for every nn, we can use Dynkin diagrams to explicitly construct proper parabolic subgroups (which, modulo their amenable radical, would correspond to SLn(𝐂)SL_{n}(\mathbf{C}) that we know has the 2½-property) that together generate the group, to conclude that the simple group has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

Proposition 6.3.5.

Let GG be a simple, simply connected, complex Lie group of rank n2n\geq 2. Then GG has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

Proof.

Recall that simple, simply connected Lie groups over 𝐂\mathbf{C} split, and hence are classified by Dynkin diagrams, so let XX be the Dynkin diagram of GG. Any subdiagram YY of XX corresponds to a parabolic subgroup QGQ\leq G containing a fixed minimal parabolic subgroup PGP\leq G, such that QQ modulo its amenable radical is a semisimple group corresponding to the subdiagram YY. Furthermore, if the subdiagram YY is of type AnA_{n} (for n1n\geq 1), then the parabolic subgroup QQ corresponding to YY, modulo its amenable radical, is SLn+1(𝐂)SL_{n+1}(\mathbf{C}), and so, by Theorem 6.3.3 and Lemma 6.2.3, QQ has the 2½-property.
So constructing two proper parabolic subgroups Q1Q_{1} and Q2Q_{2}, that both contain PP and generate GG, is equivalent to choosing two proper subdiagrams Y1Y_{1} and Y2Y_{2} of XX that the union of the vertex sets of Y1Y_{1} and Y2Y_{2} is the vertex set of XX. Furthermore, if we can ensure both these diagrams are of type AnA_{n}, then this gives us two proper parabolic subgroups with the 2½-property, ensuring that GG has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}). We claim that for any connected Dynkin diagram XX corresponding to a simple complex Lie algebra of rank at least 22, we can find two such subdiagrams Y1Y_{1} and Y2Y_{2} both of type AnA_{n} (for n1n\geq 1). This can be seen by considering each case separately, and is illustrated in the figure below. Observe that for the Dynkin diagram of type F4F_{4}, we construct subdiagrams Y1Y_{1} and Y2Y_{2} both of type A2A_{2}, while for a Dynkin diagram of any other type, we can always choose Y1Y_{1} and Y2Y_{2} to be of type A1A_{1} and Am1A_{m-1} (where mm is the rank of the group GG). ∎

[Uncaptioned image][Uncaptioned image]

We next consider the case of connected, simply connected groups over the reals, where the situation is more involved. Firstly, note that even SL3(𝐑)SL_{3}(\mathbf{R}) (which, by Theorem 6.3.3, has the 2½-property) does not have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) even though it has rank 22, because any proper parabolic subgroup of SL3(𝐑)SL_{3}(\mathbf{R}), modulo its amenable radical, is SL2(𝐑)SL_{2}(\mathbf{R}) for which Hb2(SL2(𝐑),𝐑)0\operatorname{H}_{b}^{2}(SL_{2}(\mathbf{R}),\mathbf{R})\neq 0. However, since SLn(𝐑)SL_{n}(\mathbf{R}) has the 2½-property for n3n\geq 3, we can apply the proof technique of Proposition 6.3.5 in the case of certain split simple real Lie groups to show that:

Proposition 6.3.6.

Let GG be a split simple real Lie group of type AnA_{n}, DnD_{n} (for n3n\geq 3), F4F_{4}, E6E_{6}, E7E_{7} or E8E_{8}.Then GG has Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

Proof.

As in the proof of Proposition 6.3.5, we construct two subdiagrams Y1Y_{1} and Y2Y_{2} of the Satake diagram XX of the split simple real group GG, such that the subdiagrams whose vertex sets together cover the vertex set of XX, such that the simple real Lie groups corresponding to the subdiagrams are both SLn(𝐑)SL_{n}(\mathbf{R}) for some n3n\geq 3. This is illustrated in the figure below, where the subdiagrams are encircled in grey. ∎

[Uncaptioned image]

Note that our method will not work for the split simple real Lie groups of type CnC_{n} or G2G_{2}. In the case of CnC_{n}, for any two two proper subdiagrams covering the vertices of the Satake diagram, at least one of them must either itself be of type CmC_{m}, which corresponds to the simple group Sp(2m,𝐑)Sp(2m,\mathbf{R}) which does not have the 2½-property (as Hb2(Sp2m(𝐑),𝐑)0\operatorname{H}_{b}^{2}(Sp_{2m}(\mathbf{R}),\mathbf{R})\neq 0), or of type A1A_{1}, which corresponds to the simple group SL2(𝐑)SL_{2}(\mathbf{R}). In the case of G2G_{2}, any proper parabolic subgroup is of type A1A_{1}. Thus, split simple real Lie groups of type CnC_{n} or G2G_{2} do not have Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}).

7 Conclusions and Discussion

The current paper presents many questions and directions for futher research. We highlight a few of them now.

  • Prove a complete Ulam-stability result for higher rank lattices in semisimple Lie groups.
    One could hope to extend our main results (specifically Theorem 6.2.5) by considering a larger family of groups for which the Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) is known. However, this has its limitations since there are groups for which Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) is simply not true. An important example of such a group is SL3(𝐑)SL_{3}(\mathbf{R}), where for a maximal parabolic subgroup QSL3(𝐑)Q\leq SL_{3}(\mathbf{R}), Hb2(Q,𝐑)=Hb2(SL2(𝐑),𝐑)0\operatorname{H}_{b}^{2}(Q,\mathbf{R})=\operatorname{H}_{b}^{2}(SL_{2}(\mathbf{R}),\mathbf{R})\neq 0. We can ask if Property-G(𝒬1,𝒬2)G(\mathcal{Q}_{1},\mathcal{Q}_{2}) for the ambient group is even necessary for Ulam stability, or if it just happens to be an artifact of our proof technique.

  • Surjectivity/Injectivity of a comparison map Ha2(Γ,𝒲)Hb2(Γ,𝒲~)\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})\mapsto\operatorname{H}_{b}^{2}(\Gamma,\tilde{\mathcal{W}}).
    There exists a natural forgetful map Ha2(Γ,𝒲)Hb2(Γ,𝒲~)\operatorname{H}_{a}^{2}(\Gamma,\mathcal{W})\mapsto\operatorname{H}_{b}^{2}(\Gamma,\tilde{\mathcal{W}}) (analogous to the comparison map c:Hb2(Γ,W)H2(Γ,W)c:\operatorname{H}_{b}^{2}(\Gamma,W)\to H^{2}(\Gamma,W) for a Γ\Gamma-module WW). It is not immediate if this map is either surjective or injective. Suppose it were injective, then we could truly reduce the question of uniform stability with a linear estimate to the study of the second bounded cohomology group Hb2(Γ,W~)\operatorname{H}_{b}^{2}(\Gamma,\tilde{W}) which is a well-studied notion. For instance, it is known ([13]) that for a lattice Γ\Gamma in a higher rank simple Lie group GG, Hb2(Γ,W)=0\operatorname{H}_{b}^{2}(\Gamma,W)=0 for every dual separable Banach Γ\Gamma-module WW (note, however, that the Banach space ultraproduct W~\tilde{W} is not separable unless it is finite dimensional).

  • Uniform versus non-uniform stability
    All along in this paper, we have dealt with the question of uniform stability as opposed to non-uniform, or pointwise, stability. The connection between pointwise stability and vanishing second cohomology is studied in [19], [33]. We stress that such stability results in the non-uniform setting is far from being known for most lattices. So far such results are known for many lattices in pp-adic Lie groups (with respect to the Frobenius or pp-Schatten norms for p<p<\infty), but almost nothing is known for lattices in real (or complex) Lie groups (see [5]). Moreover, when the family Mn(𝐂)M_{n}(\mathbf{C}) is endowed with the operator norm (i.e. the pp-Schatten norm for p=p=\infty) then it is known that the stability result is not true for most hgh rank lattices. This follows from the results in [18] [17] that show that if H2j(Γ,𝐑)0H^{2j}(\Gamma,\mathbf{R})\neq 0 for some positive integer jj, then Γ\Gamma is not pointwise stable for the operator norm. Note that in our setting of uniform stability, the case of p=p=\infty and p<p<\infty are treated together without any problem.

  • Stability with respect to the (normalized) Hilbert-Schmidt norm.
    In [7] it is shown that a lattice Γ\Gamma that has Property (T) is not pointwise stable for the (normalized) Hilbert-Schmidt norm. In [1] it is shown that if a residually finite group is uniformly stable with respect to the (normalized) Hilbert-Schmidt norm, then it is virtually abelian. In particular, this means that no lattice in a non-compact semisimple Lie group is uniformly stable with respect to the (normalized) Hilbert-Schmidt norm.
    In both cases, the results still leave the possibility that higher rank lattices are flexibly stable (pointwise or uniform) with respect to the (normalized) Hilbert-Schmidt norm. For more on flexible stability, refer [6] and [7].

  • Uniform stability with non-linear estimate.
    Our machinery, whenever it can be applied, works to prove uniform stability with a linear estimate. We do not know of examples of groups that are uniformly stable, but without a linear estimate. It is interesting to compare this with [8] where it is shown that 𝐙2\mathbf{Z}^{2} exhibits pointwise stability (with respect to Hamming metric on Sym(n)Sym(n)) but with a non-linear estimate. However, in the case of uniform stability in our setting, 𝐙2\mathbf{Z}^{2} (being amenable) is uniformly stable for any submultiplicative norm on U(n)U(n). The quantitative aspects of stability are an active line of current research.

  • Stability with respect to non-archimedean metrics.
    A recent monograph of [21] studies stability with respect to pp-adic groups. An interesting feature in this non-archimedean setting is that the ultrametric (strong triangle inequality) forces an equivalence between uniform and pointwise stability (for finitely presented groups). One can ask if further stability results in this setting too can be proved using the framework of asymptotic cohomology.

References

  • [1] Akhtiamov, D., and Dogon, A. On uniform Hilbert-Schmidt stability of groups. Proceedings of the American Mathematical Society (2022).
  • [2] Arkeryd, L. O., Cutland, N. J., and Henson, C. W. Nonstandard analysis: Theory and applications, vol. 493. Springer Science & Business Media, 2012.
  • [3] Arzhantseva, G., and Păunescu, L. Almost commuting permutations are near commuting permutations. Journal of Functional Analysis 269, 3 (2015), 745–757.
  • [4] Avni, N., Lubotzky, A., and Meiri, C. First order rigidity of non-uniform higher rank arithmetic groups. Inventiones mathematicae 217, 1 (2019), 219–240.
  • [5] Bader, U., Lubotzky, A., Sauer, R., and Weinberger, S. Stability and instability of lattices in semisimple Lie groups (in preparation).
  • [6] Becker, O., and Chapman, M. Stability of approximate group actions: Uniform and probabilistic. arXiv preprint arXiv:2005.06652 (2020).
  • [7] Becker, O., and Lubotzky, A. Group stability and property (T). Journal of Functional Analysis 278, 1 (2020), 108298.
  • [8] Becker, O., and Mosheiff, J. Abelian groups are polynomially stable. International Mathematics Research Notices 2021, 20 (2021), 15574–15632.
  • [9] Borel, A., and Tits, J. Groupes réductifs. Publications Mathématiques de l’IHÉS 27 (1965), 55–151.
  • [10] Borel, A., and Wallach, N. Continuous cohomology, discrete subgroups, and representations of reductive groups. No. 67. American Mathematical Soc., 2000.
  • [11] Brown, A., Fisher, D., and Hurtado, S. Zimmer’s conjecture for actions of sl(m,𝐙)sl(m,\mathbf{Z}). Inventiones mathematicae 221, 3 (2020), 1001–1060.
  • [12] Bucher, M., and Monod, N. The bounded cohomology of sl2sl_{2} over local fields and ss-integers. International Mathematics Research Notices 2019, 6 (2019), 1601–1611.
  • [13] Burger, M., and Monod, N. Bounded cohomology of lattices in higher rank Lie groups. Journal of the European Mathematical Society 1, 2 (1999), 199–235.
  • [14] Burger, M., and Monod, N. On and around the bounded cohomology of sl2sl_{2}. Rigidity in dynamics and geometry (Cambridge, 2000) (2002), 19–37.
  • [15] Burger, M., Ozawa, N., and Thom, A. On Ulam stability. Israel Journal of Mathematics 193 (2013), 109–129.
  • [16] Caprace, P.-E., De Cornulier, Y., Monod, N., and Tessera, R. Amenable hyperbolic groups. Journal of the European Mathematical Society 17, 11 (2015), 2903–2947.
  • [17] Connes, A., Gromov, M., and Moscovici, H. Conjecture de Novikov et fibrés presque plats. CR Acad. Sci. Paris Sér. I Math 310, 5 (1990), 273–277.
  • [18] Dadarlat, M. Obstructions to matricial stability of discrete groups and almost flat k-theory. Advances in Mathematics 384 (2021), 107722.
  • [19] De Chiffre, M., Glebsky, L., Lubotzky, A., and Thom, A. Stability, cohomology vanishing, and non-approximable groups. In Forum of Mathematics, Sigma (2020), vol. 8, Cambridge University Press.
  • [20] Elek, G., and Szabó, E. Hyperlinearity, essentially free actions and l2l^{2}-invariants. The sofic property. Mathematische Annalen 332, 2 (2005), 421–441.
  • [21] Fournier-Facio, F. Ultrametric analogues of Ulam stability of groups. arXiv preprint arXiv:2105.00516 (2021).
  • [22] Fournier-Facio, F., Löh, C., and Moraschini, M. Bounded cohomology and binate groups. Journal of the Australian Mathematical Society (2022), 1–36.
  • [23] Fournier-Facio, F., and Rangarajan, B. Ulam stability of lamplighters and Thompson groups. arXiv preprint arXiv:2301.03970 (2023).
  • [24] Frigerio, R. Bounded cohomology of discrete groups, vol. 227. American Mathematical Soc., 2017.
  • [25] Fujiwara, K. The second bounded cohomology of a group acting on a Gromov-hyperbolic space. Proceedings of the London Mathematical Society 76, 1 (1998), 70–94.
  • [26] Gamm, C. ε\varepsilon-representations of groups and Ulam stability. Master’s thesis, Georg-August-Universität Göttingen. 21, 2011.
  • [27] Goldblatt, R. Lectures on the hyperreals: An introduction to nonstandard analysis, vol. 188. Springer Science & Business Media, 2012.
  • [28] Guichardet, A., and Wigner, D. Sur la cohomologie réelle des groupes de Lie simples réels. In Annales scientifiques de l’École Normale Supérieure (1978), vol. 11, pp. 277–292.
  • [29] Heinrich, S. Ultraproducts in Banach space theory. J. Reine Angew. Math. 313 (1980), 77–104.
  • [30] Johnson, B. E. Approximately multiplicative maps between Banach algebras. Journal of the London Mathematical Society 2, 2 (1988), 294–316.
  • [31] Kazhdan, D. On ϵ\epsilon-representations. Israel Journal of Mathematics 43, 4 (1982), 315–323.
  • [32] Klingler, B. Transformation de type Poisson relative aux groupes d’Iwahori. Algebraic groups and Arithmetic (2004), 321–337.
  • [33] Lubotzky, A., and Oppenheim, I. Non pp-norm approximated groups. Journal d’Analyse Mathématique 141, 1 (2020), 305–321.
  • [34] Margulis, G. A. Discrete subgroups of semisimple Lie groups, vol. 17. Springer Science & Business Media, 1991.
  • [35] Matsumoto, S., and Morita, S. Bounded cohomology of certain groups of homeomorphisms. Proceedings of the American Mathematical Society 94, 3 (1985), 539–544.
  • [36] Mitsumatsu, Y. Bounded cohomology and 1\ell^{1}-homology of surfaces. Topology 23, 4 (1984), 465–471.
  • [37] Monod, N. Continuous bounded cohomology of locally compact groups. No. 1758. Springer Science & Business Media, 2001.
  • [38] Monod, N. Stabilization for slnsl_{n} in bounded cohomology. Contemporary Mathematics 347 (2004), 191–202.
  • [39] Monod, N. An invitation to bounded cohomology. In Proceedings of the International Congress of Mathematicians (2006), European Mathematical Society, pp. 1183–1211.
  • [40] Monod, N. On the bounded cohomology of semi-simple groups, s-arithmetic groups and products. J.reine angew.Math [Crelle’s J.] 640 (2010), 167–202.
  • [41] Monod, N. Lamplighters and the bounded cohomology of Thompson’s group. Geometric and Functional Analysis (2022), 1–14.
  • [42] Monod, N., and Nariman, S. Bounded and unbounded cohomology of homeomorphism and diffeomorphism groups. arXiv preprint arXiv:2111.04365 (2021).
  • [43] Monod, N., and Shalom, Y. Cocycle superrigidity and bounded cohomology for negatively curved spaces. Journal of Differential Geometry 67, 3 (2004), 395–455.
  • [44] Monod, N., and Shalom, Y. Orbit equivalence rigidity and bounded cohomology. Annals of mathematics (2006), 825–878.
  • [45] Platonov, V., and Rapinchuk, A. Algebraic groups and number theory. Academic press, 1993.
  • [46] Shtern, A. I. Quasirepresentations of amenable groups: results, errors, and hopes. Russian Journal of Mathematical Physics 20, 2 (2013), 239–253.
  • [47] Specker, E. Additive Gruppen von folgen ganzer Zahlen. Portugaliae Mathematica 9, 3 (1950), 131–140.
  • [48] Turing, A. M. Finite approximations to Lie groups. Ann. of Math. (2) 39, 1 (1938), 105–111.
  • [49] Ulam, S. M. A collection of mathematical problems. Interscience Publishers, 1960.
  • [50] von Neumann, J. Beweis des Ergodensatzes und des H-Theorems in der neuen Mechanik. Z. Phys. 57, 1 (1929), 30–70.
  • [51] Zimmer, R. J. Ergodic theory and semisimple groups, vol. 81. Springer Science & Business Media, 2013.