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

Definable K\mathrm{K}-homology of separable C*-algebras

Martino Lupini School of Mathematics and Statistics
Victoria University of Wellington
PO Box 600, 6140 Wellington, New Zealand
[email protected]
Abstract.

In this paper we show that the K\mathrm{K}-homology groups of a separable C*-algebra can be enriched with additional descriptive set-theoretic information, and regarded as definable groups. Using a definable version of the Universal Coefficient Theorem, we prove that the corresponding definable K\mathrm{K}-homology is a finer invariant than the purely algebraic one, even when restricted to the class of UHF C*-algebras, or to the class of unital commutative C*-algebras whose spectrum is a 11-dimensional connected subspace of 3\mathbb{R}^{3}.

Key words and phrases:
K\mathrm{K}-homology, KK\mathrm{KK}-theory, Universal Coefficient Theorem, C*-algebra, definable group
2020 Mathematics Subject Classification:
Primary 19K33, 54H05; Secondary 46M20, 46L80
The author was partially supported by the Marsden Fund Fast-Start Grant VUW1816 from the Royal Society Te Apārangi.

Introduction

Given a compact metrizable space XX, the group Ext(X)\mathrm{Ext}\left(X\right) classifying extensions of the C*-algebra C(X)C\left(X\right) by the C*-algebra K(H)K\left(H\right) of compact operators was initially considered by Brown, Douglas, and Fillmore in their celebrated work [brown_extensions_1977]. There, they showed that Ext()\mathrm{Ext}\left(-\right) is indeed a group, and that defining, for a compact metrizable space XX,

K~p(X):={Ext(X)if p is odd,Ext(ΣX)if p is even;\mathrm{\tilde{K}}_{p}\left(X\right):=\left\{\begin{array}[]{ll}\mathrm{Ext}\left(X\right)&\text{if }p\text{ is odd,}\\ \mathrm{Ext}\left(\Sigma X\right)&\text{if }p\text{ is even;}\end{array}\right.

where ΣX\Sigma X is the suspension of XX, yields a (reduced) homology theory that satisfies all the Eilenberg–Steenrod–Milnor axioms for Steenrod homology, apart from the Dimension Axiom; see also [kaminker_theory_1977]. They furthermore observed, building on a previous insight of Atiyah [atiyah_global_1970], that such a homology theory can be seen as the Spanier–Whitehead dual of topological K\mathrm{K}-theory [atiyah_theory_1989].

More generally, for an arbitrary separable unital C*-algebra AA, one can consider a semigroup Ext(A)\mathrm{Ext}\left(A\right) classifying the essential, unital extensions of AA by K(H)K\left(H\right). By Voiculescu’s non-commutative Weyl-von Neumann Theorem [voiculescu_non-commutative_1976, arveson_notes_1977], the trivial element of Ext(A)\mathrm{Ext}\left(A\right) correspond to the class of trivial essential, unital extensions. The group Ext(A)1\mathrm{Ext}\left(A\right)^{-1} of invertible elements of Ext(A)\mathrm{Ext}\left(A\right) corresponds to the essential, unital extensions that are semi-split. Thus, by the Choi–Effros lifting theorem [choi_completely_1976], Ext(A)\mathrm{Ext}\left(A\right) is a group when AA is nuclear. One can extend K\mathrm{K}-homology to the category of all separable C*-algebras by setting

Kp(A)={Ext(A+)1if p is odd,Ext((SA)+)1if p is even;\mathrm{K}^{p}\left(A\right)=\left\{\begin{array}[]{ll}\mathrm{Ext}\left(A^{+}\right)^{-1}&\text{if }p\text{ is odd,}\\ \mathrm{Ext}((SA)^{+})^{-1}&\text{if }p\text{ is even;}\end{array}\right.

where SXSX is the suspension of AA and A+A^{+} is the unitization of AA. This gives a cohomology theory on the category of separable C*-algebras, which can be recognized as the dual of K\mathrm{K}-theory via Paschke duality [paschke_theory_1981, higson_algebra_1995, kaminker_spanier_2017]. Kasparov’s bivariant functor KK(,)\mathrm{KK}\left(-,-\right) simultaneously generalizes K\mathrm{K}-homology and K\mathrm{K}-theory, where K1(A)\mathrm{K}^{1}\left(A\right) is recovered as KK(A,)\mathrm{KK}\left(A,\mathbb{C}\right) and K0(A)\mathrm{K}^{0}\left(A\right) as KK(A,C0())\mathrm{KK}\left(A,C_{0}\left(\mathbb{R}\right)\right).

It was already noticed in the seminal work of Brown, Douglas, and Fillmore [brown_extensions_1977, brown_extensions_1973] that the invariant Kp(X)\mathrm{K}^{p}\left(X\right) for a compact metrizable space XX can be endowed with more structure than the purely algebraic group structure. Indeed, one can write XX as the inverse limit of a tower (Xn)nω\left(X_{n}\right)_{n\in\omega} of compact polyhedra, and endow Kp(X)\mathrm{K}^{p}\left(X\right) with the topology induced by the maps Kp(X)Kp(Xn)\mathrm{K}^{p}\left(X\right)\rightarrow\mathrm{K}^{p}\left(X_{n}\right) for nωn\in\omega, where Kp(Xn)\mathrm{K}^{p}\left(X_{n}\right) is a countable group endowed with the discrete topology. This gives to Kp(X)\mathrm{K}^{p}\left(X\right) the structure of a topological group, which is however in general not Hausdorff.

The study of Kp(A)\mathrm{K}^{p}\left(A\right) as a topological group for a separable unital C*-algebra AA was later systematically undertaken by Dadarlat [dadarlat_approximate_2000, dadarlat_topology_2005] and Schochet [schochet_fine_2001, schochet_fine_2002, schochet_fine_2005] building on previous work of Salinas [salinas_relative_1992]. (In fact, they consider more generally Kasparov’s KK\mathrm{KK}-groups.) In [schochet_fine_2001, dadarlat_topology_2005] several natural topologies on Kp(A)\mathrm{K}^{p}\left(A\right), corresponding to different ways to define K\mathrm{K}-homology for separable C*-algebras, are shown to coincide and to turn Kp(A)\mathrm{K}^{p}\left(A\right) into a pseudo-Polish group. This means that, if Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) denotes the closure of zero in Kp(A)\mathrm{K}^{p}\left(A\right), then the quotient of Kp(A)\mathrm{K}^{p}\left(A\right) by Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) is a Polish group. In [schochet_fine_2002], for a C*-algebra AA satisfying the Universal Coefficient Theorem (UCT), the topology on Kp(A)\mathrm{K}^{p}\left(A\right) is related to the UCT exact sequence, and Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) is shown to be isomorphic to the group PExt(K1p(A),)\mathrm{PExt}\left(\mathrm{K}_{1-p}\left(A\right),\mathbb{Z}\right) classifying pure extensions of K1p(A)\mathrm{K}_{1-p}\left(A\right) by \mathbb{Z}. A characterization of Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) for an arbitrary separable nuclear C*-algebra AA is obtained in [dadarlat_approximate_2000]. For a separable quasidiagonal C*-algebra satisfying the UCT, K1(A)\mathrm{K}_{\infty}^{1}\left(A\right) is shown to be the subgroup of K1(A)=Ext(A+)\mathrm{K}^{1}\left(A\right)=\mathrm{Ext}\left(A^{+}\right) corresponding to quasidiagonal extensions of A+A^{+} by K(H)K\left(H\right) [schochet_fine_2002]; see also [brown_universal_1984] for the commutative case. The quotient Kwp(A)\mathrm{K}_{\mathrm{w}}^{p}\left(A\right) of Kp(A)\mathrm{K}^{p}\left(A\right) by Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) is the group KLp(A,)\mathrm{KL}_{p}\left(A,\mathbb{C}\right) introduced by Rørdam [rordam_classification_1995]. A universal multicoefficient theorem describing Kwp(A)\mathrm{K}_{\mathrm{w}}^{p}\left(A\right) in terms of the K\mathrm{K}-groups of AA with arbitrary cyclic groups as coefficients is obtained in [dadarlat_classification_2002] for all separable nuclear C*-algebras satisfying the UCT; see [dadarlat_topology_2005, Theorem 5.4].

In many cases of interest, the topology on Kp(A)\mathrm{K}^{p}\left(A\right) turns out to be trivial, i.e. the closure of zero in Kp(A)\mathrm{K}^{p}\left(A\right) is the whole group. For example, the topology on K1(A)\mathrm{K}^{1}\left(A\right) is trivial when AA is a UHF C*-algebra, despite the fact that K1(A)\mathrm{K}^{1}\left(A\right) is not trivial, and in fact uncountable. Similarly, for every 11-dimensional solenoid XX, the topology on K~0(X)\mathrm{\tilde{K}}_{0}\left(X\right) is trivial, although K~0(X)\mathrm{\tilde{K}}_{0}\left(X\right) is an uncountable group.

In this paper, we take a different approach and consider the group Kp(A)\mathrm{K}^{p}\left(A\right), rather than as a pseudo-Polish topological group, as a definable group. This should be thought of as a group GG explicitly defined as the quotient of a Polish space XX by a “well-behaved” equivalence relation EE, in such a way that the multiplication and inversion operations in GG are induced by a Borel functions on XX. This is formally defined in Section 1.5, where the notion of well-behaved equivalence relation is made precise. The definition is devised to ensure that the category of definable groups has good properties, and behaves similarly to the category of standard Borel groups. A morphism in this category is a definable group homomorphism, namely a group homomorphism that lifts to a Borel function between the corresponding Polish spaces.

It has recently become apparent that several homological invariants in algebra and topology can be seen as functors to the category of definable groups. The homological invariants Ext\mathrm{Ext} and lim1\mathrm{lim}^{1} are considered in [bergfalk_ulam_2020], whereas Steenrod homology and Čech cohomology are considered in [bergfalk_cohomology_2020, lupini_definable_2020]. It is shown there that the definable versions of these invariants are finer than the purely algebraic versions.

In this paper, we show that, for an arbitrary separable C*-algebra AA, Kp(A)\mathrm{K}^{p}\left(A\right) can be regarded as a definable group. Furthermore, different descriptions of Kp(A)\mathrm{K}^{p}\left(A\right)—in terms of extensions, Paschke duality, Fredholm modules, and quasi-homomorphisms—yield naturally definably isomorphic definable groups. For C*-algebras that have a KK\mathrm{KK}-filtration in the sense of Schochet [schochet_uct_1996], we show that the definable subgroup Kp(A)\mathrm{K}_{\infty}^{p}\left(A\right) of Kp(A)\mathrm{K}^{p}\left(A\right) is definably isomorphic to PExt(K1p(A),)\mathrm{PExt}\left(\mathrm{K}_{1-p}\left(A\right),\mathbb{Z}\right). The latter is regarded as a definable group as in [bergfalk_ulam_2020, Section 7]. (In fact, PExt(K1p(A),)\mathrm{PExt}\left(\mathrm{K}_{1-p}\left(A\right),\mathbb{Z}\right) is the quotient of a Polish group by a Borel Polishable subgroup, and hence a group with a Polish cover in the parlance of [bergfalk_ulam_2020, Section 7].)

Using this and the rigidity theorem for PExt(Λ,)\mathrm{PExt}\left(\Lambda,\mathbb{Z}\right) from [bergfalk_ulam_2020, Section 7] where Λ\Lambda is a torsion-free abelian group without finitely-generated direct summands, we prove that definable K\mathrm{K}-homology provides a finer invariant than the purely algebraic (or topological) groups Kp(A)\mathrm{K}^{p}\left(A\right) for a separable C*-algebra AA, even when one restricts to UHF C*-algebras or commutative unital C*-algebras whose spectrum is a 11-dimensional subspace of 3\mathbb{R}^{3}.

Theorem A.

The definable K1\mathrm{K}^{1}-group is a complete invariant for UHF C*-algebras up to stable isomorphism. In contrast, there exists an uncountable family of pairwise non stably isomorphic UHF C*-algebras with algebraically isomorphic K1\mathrm{K}^{1}-groups (and trivial K0\mathrm{K}^{0}-groups).

Theorem B.

The definable K~0\mathrm{\tilde{K}}_{0}-group is a complete invariant for 11-dimensional solenoids up to homeomorphism. In contrast, there exists an uncountable family of pairwise non homeomorphic 11-dimensional solenoids with algebraically isomorphic K~0\mathrm{\tilde{K}}_{0}-groups (and trivial K~1\mathrm{\tilde{K}}_{1}-groups).

The historic evolution in the treatment of K\mathrm{K}-homology described above should be compared with the similar evolution in the study of unitary duals of second countable, locally compact groups or, more generally, separable C*-algebras. Given a separable C*-algebra AA, its unitary dual A^\hat{A} is the quotient of the Polish space Irr(A)\mathrm{Irr}(A) of unitary irreducible representations of AA by the relation of unitary equivalence. This includes as a particular instance the case of second countable, locally compact groups, by considering the corresponding universal C*-algebras. While initially A^\hat{A} was considered as a topological space endowed with the quotient topology, it was recognized in the seminal work of Mackey, Glimm, and Effros [mackey_borel_1957, glimm_type_1961, effros_transformation_1965] that a more fruitfuil theory is obtained by considering A^\hat{A} endowed with the quotient Borel structure, called the Mackey Borel structure. This led to the notion of type I C*-algebra, which precisely captures those separable C*-algebras with the property that the Mackey Borel structure is standard. It was soon realized that, in the non type I case, the right notion of “isomorphism” of Macky Borel structures on duals A^,B^\hat{A},\hat{B} corresponds to a bijection A^B^\hat{A}\rightarrow\hat{B} that is induced by a Borel function Irr(A)Irr(B)\mathrm{Irr}\left(A\right)\rightarrow\mathrm{Irr}\left(B\right). In our terminology from Section 1.4, this corresponds to regarding a unitary dual A^\hat{A} as a definable set, where an isomorphism of Mackey Borel structures on A^,B^\hat{A},\hat{B} is a definable bijection A^B^\hat{A}\rightarrow\hat{B}. For example, this approach is taken by Elliott in [elliott_mackey_1977], where he proved that the unitary duals of any two separable AF C*-algebras that are not type I are isomorphic in the category of definable sets. It is a question of Dixmier from 1967 whether the unitary duals of any two non-type I separable C*-algebras are isomorphic in the category of definable sets; see [thomas_descriptive_2015, farah_dichotomy_2012, kerr_turbulence_2010]. This problem was recently considered in the case of groups by Thomas, who showed that the unitary duals of any two countable amenable non-type I groups are isomorphic in the category of definable sets [thomas_descriptive_2015, Theorem 1.10]. Furthermore, the unitary dual of any countable group admits a definable injection to the unitary dual of the free group on two generators [thomas_descriptive_2015, Theorem 1.9].

The work of Mackey, Glimm, and Effors on unitary representations pioneered the application of methods from descriptive set theory to C*-algebras. More recent applications have been obtained by Kechris [kechris_descriptive_1998] and Farah–Toms–Törnquist [farah_turbulence_2014, farah_descriptive_2012], who studied the problem of classifying several classes of C*-algebras from the perspective of Borel complexity theory; see also [lupini_unitary_2014, gardella_conjugacy_2016, elliott_isomorphism_2013].

The rest of this paper is organized as follows. In Section 1 we recall fundamental results from descriptive set theory about Polish spaces and standard Borel spaces, and make precise the notions of definable set, and the corresponding notion of definable group. In Section 2 we introduce the notion of strict C*-algebra, which is a (not necessarily norm-separable) C*-algebra whose unit ball is endowed with a Polish topology induced by bounded seminorms, called the strict topology, such that the C*-algebra operations are strictly continuous on the unit ball. The main example we will consider are multiplier algebras of separable C*-algebras, endowed with their usual strict topology, as well as Paschke dual algebras of separable C*-algebras. In Section 3 we study the K\mathrm{K}-theory of a strict C*-algebra or, more generally, the quotient of a strict C*-algebra by a strict ideal, such as a corona algebra or the commutant of a separable C*-algebra in a corona algebra. We observe that the K\mathrm{K}-theory groups of a strict C*-algebra can be regarded as quotients of a Polish space by an equivalence relation. As such an equivalence relation is not necessarily well-behaved, they are in general only semidefinable groups, although we they will be in fact definable groups in the case of Paschke dual algebras of separable C*-algebras. In Section 4 definable K\mathrm{K}-homology for separable C*-algebras is introduced, and shown to be given by definable groups by considering its description in terms of the K\mathrm{K}-theory of Pashcke dual algebras. The descriptions of K\mathrm{K}-homology due to Cuntz and Kasparov are considered in Section LABEL:Section:Kasparov, where they are shown to yield definably isomorphic groups. In Section LABEL:Section:properties we discuss properties of definable K\mathrm{K}-homology, which can be seen as definable versions of the general properties that an abstract cohomology theory for separable, nuclear C*-algebras in the sense of Schochet satisfies [schochet_topologicalIII_1984]. A definable version of the Universal Coefficient Theorem of Brown [brown_universal_1984], later generalized to KK\mathrm{KK}-groups by Rosenberg and Schochet [rosenberg_kunneth_1987], is considered in Section LABEL:Section:UCT. Theorem A is a consequence of the definable UCT, the classification of AF C*-algebras by K\mathrm{K}-theory, and the rigidity result for definable PExt\mathrm{PExt} of torsion-free finite-rank abelian groups from [bergfalk_ulam_2020]. Finally, Section LABEL:Section:spaces considers definable K\mathrm{K}-homology for compact metrizable spaces, and Theorem B is obtained applying again the definable UCT and the rigidity theorem for definable PExt\mathrm{PExt} from [bergfalk_ulam_2020].

1. Polish spaces and definable groups

In this section we recall some fundamental notions concerning Polish spaces and Polish groups, as well as standard Borel spaces and standard Borel groups, as can be found in [becker_descriptive_1996, kechris_classical_1995, gao_invariant_2009]. We also consider the notion of Polish category, which is a category whose hom-sets are Polish spaces and composition of morphisms is a continuous function, and establish some of its basic properties. Furthermore, we recall the notion of idealistic equivalence relation on a standard Borel space and some of its fundamental properties as established in [kechris_borel_2016, motto_ros_complexity_2012]. We then define precisely the notion of (semi)definable set and (semi)definable group.

1.1. Polish spaces and standard Borel spaces

A Polish space is a second countable topological space whose topology is induced by a complete metric. A subset of a Polish space XX is GδG_{\delta} if and only if it is a Polish space when endowed with the subspace topology. If XX is a Polish space, then the Borel σ\sigma-algebra of XX is the σ\sigma-algebra generated by the collection of open sets. By definition, a subset of XX is Borel if it belongs to the Borel σ\sigma-algebra. If X,YX,Y are Polish spaces, then the product X×YX\times Y is a Polish space when endowed with the product topology. More generally, if (Xn)nω\left(X_{n}\right)_{n\in\omega} is a sequence of Polish spaces, then the product nωXn\prod_{n\in\omega}X_{n} is a Polish space when endowed with the product topology. The class of Polish spaces includes all locally compact second countable Hausdorff spaces. We denote by ω\omega the set of natural numbers including 0. We regard ω\omega and any other countable set as a Polish space endowed with the discrete topology. The Baire space ωω\omega^{\omega} is the Polish space obtained as the infinite product of copies of ω\omega.

A standard Borel space is a set XX endowed with a σ\sigma-algebra (the Borel σ\sigma-algebra) that comprises the Borel sets with respect to some Polish topology on XX. A function between standard Borel spaces is Borel if it is measurable with respect to the Borel σ\sigma-algebras. A subset of a standard Borel space XX is analytic if it is the image of a Borel function f:ZXf:Z\rightarrow X for some standard Borel space ZZ. This is equivalent to the assertion that there exists a Borel subset BX×ωωB\subseteq X\times\omega^{\omega} such that B=projX(A)B=\mathrm{proj}_{X}\left(A\right) is the projection of AA on the first coordinate. A subset of XX is co-analytic if its complement is analytic. One has that a subset of XX is Borel if and only if it is both analytic and co-analytic.

Given standard Borel spaces X,YX,Y, we let X×YX\times Y be their product endowed with the product Borel structure, which is also a standard Borel space. If (Xn)nω\left(X_{n}\right)_{n\in\omega} is a sequence of standard Borel spaces, then their disjoint union XX is a standard Borel space, where a subset AA of XX is Borel if and only if AXnA\cap X_{n} is Borel for every nωn\in\omega. The product nωXn\prod_{n\in\omega}X_{n} is also a standard Borel space when endowed with the product Borel structure. In the following proposition, we collect some well-known properties of the category of standard Borel spaces and Borel functions.

Proposition 1.1.

Let 𝐒𝐁\mathbf{SB} be the category that has standard Borel spaces as objects and Borel functions and morphisms.

  1. (1)

    If XX is a standard Borel space and AXA\subseteq X is a Borel subset, then AA is a standard Borel space when endowed with the induced standard Borel structure;

  2. (2)

    If X,YX,Y are standard Borel spaces, f:XYf:X\rightarrow Y is an injective Borel function, and AXA\subseteq X is Borel, then f(A)f\left(A\right) is a Borel subset of YY;

  3. (3)

    If X,YX,Y are standard Borel spaces, and f:XYf:X\rightarrow Y is a bijective Borel function, then the inverse function f1:YXf^{-1}:Y\rightarrow X is Borel;

  4. (4)

    If X,YX,Y are standard Borel spaces, and there exist injective Borel functions f:XYf:X\rightarrow Y and g:YXg:Y\rightarrow X, then there exists a Borel bijection h:XYh:X\rightarrow Y;

  5. (5)

    The category 𝐒𝐁\mathbf{SB} has finite products, finite coproducts, equalizers, and pullbacks;

  6. (6)

    A Borel function is monic in 𝐒𝐁\mathbf{SB} if and only if it is injective, and epic in 𝐒𝐁\mathbf{SB} if and only if it is surjective;

  7. (7)

    An inductive sequence of standard Borel spaces and injective Borel functions has a colimit in 𝐒𝐁\mathbf{SB}.

A Polish group is a topological group whose topology is Polish. If GG is a Polish group, and HH is a closed subgroup of GG, then HH is a Polish group when endowed with the subspace topology. If furthermore HH is normal, then G/HG/H is a Polish group when endowed with the quotient topology. If G0,G1G_{0},G_{1} are Polish groups, and φ:G0G1\varphi:G_{0}\rightarrow G_{1} is a Borel function, then φ\varphi is continuous. In particular, if GG is a Polish space, then it has a unique Polish group topology that induces its Borel structure. A subgroup HH of a Polish group GG is Polishable if it is Borel and there is a (necessarily unique) Polish group topology on HH that induces the Borel structure on HH inherited from GG. This is equivalent to the assertion that HH is equal to the range of a continuous group homomorphisms G^G\hat{G}\rightarrow G for some Polish group G^\hat{G}. If GG is a Polish group, then a Polish GG-space is a Polish space XX endowed with a continuous action of GG. A Borel GG-space is a standard Borel space XX endowed with a Borel action of GG. Given a Borel GG-space XX, there exists a Polish topology τ\tau on XX such that (X,τ)\left(X,\tau\right) is a Polish GG-space; see [becker_descriptive_1996, Theorem 5.2.1].

A standard Borel group is, simply, a group object in the category of standard Borel spaces [mac_lane_categories_1998, Section III.6]. Explicitly, a standard Borel group is a standard Borel space GG that is also a group, and such that the group operation on GG and the function GGG\rightarrow G, xx1x\mapsto x^{-1} are Borel; see [kechris_classical_1995, Definition 12.23]. Clearly, every Polish group is, in particular, a standard Borel group.

The notion of Polish topometric space was introduced and studied in [ben_yaacov_grey_2015, ben_yaacov_polish_2013, ben_yaacov_topometric_2008, ben_yaacov_continuous_2010]. A topometric space is a Hausdorff space XX endowed with a topology τ\tau and a [0,]\left[0,\infty\right]-valued metric dd such that:

  1. (1)

    the metric-topology is finer than τ\tau;

  2. (2)

    the metric is lower-semicontinuous with respect to τ\tau, i.e. for every r0r\geq 0 the set

    {(a,b)X×X:d(a,b)r}\left\{\left(a,b\right)\in X\times X:d\left(a,b\right)\leq r\right\}

    is τ\tau-closed in X×XX\times X.

A Polish topometric space is a topometric space such that the topology τ\tau is Polish and the metric is complete. A Polish topometric group is a Polish topometric space (G,τ,d)\left(G,\tau,d\right) that is also a group, and such that GG endowed with the topology τ\tau is a Polish group, and the metric dd on GG is bi-invariant

1.2. Polish categories

By definition, we let a Polish category be a category 𝒞\mathcal{C} enriched over the category of Polish spaces (regarded as a monoidal category with respect to binary products). Thus, for each pair of objects a,ba,b of 𝒞\mathcal{C}, 𝒞(a,b)\mathcal{C}\left(a,b\right) is a Polish space, such that for objects a,b,ca,b,c, the composition operation 𝒞(b,c)×𝒞(a,b)𝒞(a,c)\mathcal{C}\left(b,c\right)\times\mathcal{C}\left(a,b\right)\rightarrow\mathcal{C}\left(a,c\right) is continuous.

Suppose that 𝒞\mathcal{C} is a Polish category. For objects a,ba,b of 𝒞\mathcal{C}, define Iso𝒞(a,b)𝒞(a,b)\mathrm{Iso}_{\mathcal{C}}\left(a,b\right)\subseteq\mathcal{C}\left(a,b\right) be the set of 𝒞\mathcal{C}-isomorphisms aba\rightarrow b. While Iso𝒞(a,b)\mathrm{Iso}_{\mathcal{C}}\left(a,b\right) is not necessarily a GδG_{\delta} subset of 𝒞(a,b)\mathcal{C}\left(a,b\right), and hence not necessarily a Polish space when endowed with the subspace topology, Iso𝒞(a,b)\mathrm{Iso}_{\mathcal{C}}\left(a,b\right) is endowed with a canonical Polish topology, defined as follows. For a net (αi)\left(\alpha_{i}\right) in Iso𝒞(a,b)\mathrm{Iso}_{\mathcal{C}}\left(a,b\right) and αIso𝒞(a,b)\alpha\in\mathrm{Iso}_{\mathcal{C}}\left(a,b\right), set αiα\alpha_{i}\rightarrow\alpha if and only if αiα\alpha_{i}\rightarrow\alpha in 𝒞(a,b)\mathcal{C}\left(a,b\right) and αi1α1\alpha_{i}^{-1}\rightarrow\alpha^{-1} in 𝒞(b,a)\mathcal{C}\left(b,a\right). One can then easily show the following.

Lemma 1.2.

Adopt the notations above. Then Iso𝒞(a,b)\mathrm{Iso}_{\mathcal{C}}\left(a,b\right) is a Polish space.

It is clear from the definition that, for every object aa of 𝒞\mathcal{C}, Aut𝒞(a):=Iso𝒞(a,a)\mathrm{\mathrm{Aut}}_{\mathcal{C}}\left(a\right):=\mathrm{Iso}_{\mathcal{C}}\left(a,a\right) is a Polish group. Furthermore, the canonical (right and left) actions of Aut𝒞(a)\mathrm{\mathrm{Aut}}_{\mathcal{C}}\left(a\right) and Aut𝒞(b)\mathrm{\mathrm{Aut}}_{\mathcal{C}}\left(b\right) on 𝒞(a,b)\mathcal{C}\left(a,b\right) are continuous.

Definition 1.3.

Suppose that 𝒞\mathcal{C} and 𝒟\mathcal{D} are Polish categories, and F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} is a functor. We say that FF is continuous if, for every pair of objects a,ba,b of 𝒞\mathcal{C}, the map 𝒞(a,b)𝒟(F(a),F(b))\mathcal{C}\left(a,b\right)\rightarrow\mathcal{D}\left(F\left(a\right),F\left(b\right)\right), fF(f)f\mapsto F\left(f\right) is continuous. We say that FF is a topological equivalence if it is continuous, and there exists a continuous functor G:𝒟𝒞G:\mathcal{D}\rightarrow\mathcal{C} such that GFGF is isomorphic to the identity functor I𝒞I_{\mathcal{C}}, and FGFG is isomorphic to the identity functor I𝒟I_{\mathcal{D}}.

The notion of topological equivalence of categories is the natural analogue of the notion of equivalence of categories in the context of Polish categories; see [mac_lane_categories_1998, Section IV.4]. The same proof as [mac_lane_categories_1998, Section IV.4, Theorem 1] gives the following characterization of topological equivalences.

Lemma 1.4.

Suppose that 𝒞\mathcal{C} and 𝒟\mathcal{D} are Polish categories, and F:𝒞𝒟F:\mathcal{C}\rightarrow\mathcal{D} is a functor. The following assertions are equivalent:

  1. (1)

    FF is a topological equivalence;

  2. (2)

    each object of 𝒟\mathcal{D} is isomorphic to one of the form F(a)F\left(a\right) for some object aa of 𝒞\mathcal{C}, and for each pair of objects c,dc,d of 𝒞\mathcal{C}, the map 𝒞(c,d)𝒞(F(c),F(d))\mathcal{C}\left(c,d\right)\rightarrow\mathcal{C}\left(F\left(c\right),F\left(d\right)\right) is a homeomorphism.

1.3. Idealistic equivalence relations

Suppose that CC is a set. A σ\sigma-filter on CC is a nonempty family \mathcal{F} of subsets of CC that is closed under countable intersections, and such that \varnothing\notin\mathcal{F} and if ABCA\subseteq B\subseteq C and AA\in\mathcal{F} then BB\in\mathcal{F}. The dual notion is the one of σ\sigma-ideal. Thus, a nonempty family \mathcal{I} of subsets of CC is a σ\sigma-ideal if it is closed under countable unions, CC\notin\mathcal{I}, and ABCA\subseteq B\subseteq C and BB\in\mathcal{I} imply AA\in\mathcal{I}. Clearly, if \mathcal{F} is a σ\sigma-filter on CC, then {CA:A}\left\{C\setminus A:A\in\mathcal{F}\right\} is a σ\sigma-ideal on CC, and vice-versa. Thus, one can equivalently formulate notions in terms of σ\sigma-filters or in terms of σ\sigma-ideals.

If \mathcal{F} is a σ\sigma-filter on CC, then \mathcal{F} can be thought of as a notion of “largeness” for subsets of CC. Based on this interpretation, we use the “σ\sigma-filter quantifier” notation “x\mathcal{F}x, xAx\in A” for a subset ACA\subseteq C to express the fact that AA\in\mathcal{F}. If P(x)P(x) is a unary relation for elements of CC, “x\mathcal{F}x, P(x)P(x)” is the assertion that the set of xCx\in C that satisfy P(x)P(x) belongs to \mathcal{F}.

Example 1.5.

Suppose that CC is a Polish space. A subset AA of CC is meager if it is contained in the union of a countable family of closed nowhere dense sets. By the Baire Category Theorem [kechris_classical_1995, Theorem 8.4], meager subsets of CC form a σ\sigma-ideal C\mathcal{I}_{C}. The corresponding dual σ\sigma-filter is the σ\sigma-filter C\mathcal{F}_{C} of comeager sets, which are the subsets of CC whose complement is meager.

Suppose that XX is a standard Borel space. We consider an equivalence relation EE on XX as a subset of X×XX\times X, endowed with the product Borel structure. Consistently, we say that EE is Borel or analytic, respectively, if it is a Borel or analytic subset of X×XX\times X. In the following, we will exclusively consider analytic equivalence relations, most of which will in fact be Borel. For an element xx of XX we let [x]E\left[x\right]_{E} be its corresponding EE-class.

We now recall the notion of idealistic equivalence relation, initially considered in [kechris_countable_1994]; see also [gao_invariant_2009, Definition 5.4.9] and [kechris_borel_2016]. We will consider a slightly more generous definition than the one from [kechris_countable_1994, gao_invariant_2009, kechris_borel_2016]. The more restrictive notion is recovered as a particular case by insisting that the function ss in Definition 1.6 be the identity function of XX. In the following definition, for a subset AA of a product space X×YX\times Y and xXx\in X, we let Ax={yY:(x,y)A}A_{x}=\left\{y\in Y:\left(x,y\right)\in A\right\} be the corresponding vertical section.

Definition 1.6.

An equivalence relation EE on a standard Borel space XX is idealistic if there exist a Borel function s:XXs:X\rightarrow X satisfying s(x)Exs(x)Ex for every xXx\in X, and a function CCC\mapsto\mathcal{F}_{C} that assigns to each EE-class CC a σ\sigma-filter C\mathcal{F}_{C} of subsets of CC such that, for every Borel subset AA of X×XX\times X, the set

As,:={xX:[x]Ex,(s(x),x)A}={xX:As(x)[x]E}.A_{s,\mathcal{F}}:=\left\{x\in X:\mathcal{F}_{[x]_{E}}x^{\prime},\left(s(x),x^{\prime}\right)\in A\right\}=\left\{x\in X:A_{s(x)}\in\mathcal{F}_{\left[x\right]_{E}}\right\}\text{.}

is Borel.

Idealistic equivalence relations arise naturally as orbit equivalence relations of Polish group actions. Suppose that GG is a Polish group and XX is a Polish GG-space. Let EGXE_{G}^{X} be the corresponding orbit equivalence relation on XX, obtained by setting, for x,yXx,y\in X, xEGXyxE_{G}^{X}y if and only if there exists gGg\in G such that gx=yg\cdot x=y. Then EGXE_{G}^{X} is an idealistic equivalence relation, as witnessed by the identity function ss on XX and the function CCC\mapsto\mathcal{F}_{C} where ACA\in\mathcal{F}_{C} if and only if Gg\mathcal{F}_{G}g, gxAg\cdot x\in A. (As in Example 1.5, G\mathcal{F}_{G} denotes the σ\sigma-filter of comeager subsets of GG.) In particular, if GG is a Polish group, and HH is a Polishable subgroup of GG, then the coset equivalence relation EHGE_{H}^{G} of HH in GG is Borel and idealistic.

Suppose that EE is an equivalence relation on a standard Borel space XX. A Borel selector for EE is a Borel function s:XXs:X\rightarrow X such that, for x,yxx,y\in x, xEyxEy if and only if s(x)=s(y)s(x)=s(y). If EE has a Borel selector, then EE is Borel and idealistic; see [gao_invariant_2009, Theorem 5.4.11]. (Precisely, an equivalence relation has a Borel selector if and only if it is Borel, idealistic, and smooth [gao_invariant_2009, Definition 5.4.1].)

1.4. Definable sets

Definable sets are a generalization of standard Borel sets, and can be thought of as sets explicitly presented as the quotient of a standard Borel space by a “well-behaved” equivalence relation EE.

Definition 1.7.

A definable set XX is a pair (X^,E)(\hat{X},E) where X^\hat{X} is a standard Borel space and EE is a Borel and idealistic equivalence relation on X^\hat{X}. We think of (X^,E)(\hat{X},E) as a presentation of the quotient set X=X^/EX=\hat{X}/E. Consistently, we also write the definable set (X^,E)(\hat{X},E) as X^/E\hat{X}/E. A subset ZZ of XX is Borel if there is an EE-invariant Borel subset Z^\hat{Z} of X^\hat{X} such that Z=Z^/EZ=\hat{Z}/E.

We now define the notion of morphism between definable sets. Let X=X^/EX=\hat{X}/E and Y=Y^/FY=\hat{Y}/F be definable sets. A lift of a function f:XYf:X\rightarrow Y is a function f^:X^Y^\hat{f}:\hat{X}\rightarrow\hat{Y} such that f([x]E)=[f^(x)]Ff\left(\left[x\right]_{E}\right)=[\hat{f}(x)]_{F} for every xX^x\in\hat{X}.

Definition 1.8.

Let XX and YY be definable sets. A function f:XYf:X\rightarrow Y is Borel-definable if it has a lift f^:X^Y^\hat{f}:\hat{X}\rightarrow\hat{Y} that is a Borel function.

Remark 1.9.

Since Borel-definability is the only notion of definability we will consider in this paper, we will abbreviate “Borel-definable” to “definable”.

We consider definable sets as objects of a category 𝐃𝐒𝐞𝐭\mathbf{DSet}, whose morphisms are the definable functions. We regard a standard Borel space XX as a particular instance of definable set X=X^/EX=\hat{X}/E where X=X^X=\hat{X} and EE is the relation of equality on XX. This renders the category of standard Borel spaces a full subcategory of the category of definable sets.

If X=X^/EX=\hat{X}/E and Y=Y^/FY=\hat{Y}/F are definable sets, then their product X×YX\times Y in 𝐃𝐒𝐞𝐭\mathbf{DSet} is the definable set X×Y:=(X^×Y^)/(E×F)X\times Y:=(\hat{X}\times\hat{Y})\left/\left(E\times F\right)\right., E×FE\times F being the equivalence relation on X^×Y^\hat{X}\times\hat{Y} defined by setting (x,y)(E×F)(x,y)\left(x,y\right)\left(E\times F\right)\left(x^{\prime},y^{\prime}\right) if and only if xExxEx^{\prime} and yFyyFy^{\prime}. (It is easy to see that E×FE\times F is Borel and idealistic if both EE and FF are Borel and idealistic.)

Many of the good properties of standard Borel spaces, including all the ones listed in Proposition 1.1, generalize to definable sets.

Proposition 1.10.

Let as above 𝐃𝐒𝐞𝐭\mathbf{DSet} be the category that has definable as objects and definable functions as morphisms.

  1. (1)

    If XX is a definable and AXA\subseteq X is a Borel subset, then AA is itself a definable set;

  2. (2)

    If X,YX,Y are definable sets, f:XYf:X\rightarrow Y is an injective definable function, and AXA\subseteq X a Borel subset, then f(A)f\left(A\right) is a Borel subset of YY;

  3. (3)

    If X,YX,Y are definable sets, and f:XYf:X\rightarrow Y is a bijective definable function, then the inverse function f1:YXf^{-1}:Y\rightarrow X is definable;

  4. (4)

    If X,YX,Y are definable sets, and there exist injective definable functions f:XYf:X\rightarrow Y and g:YXg:Y\rightarrow X, then there exists a definable bijection h:XYh:X\rightarrow Y;

  5. (5)

    The category 𝐃𝐒𝐞𝐭\mathbf{DSet} has finite products, finite coproducts, equalizers, and pullbacks;

  6. (6)

    A definable function is monic in 𝐃𝐒𝐞𝐭\mathbf{DSet} if and only if it is injective, and epic in 𝐃𝐒𝐞𝐭\mathbf{DSet} if and only if it is surjective;

  7. (7)

    An inductive sequence of definable sets and injective definable functions has a colimit in 𝐃𝐒𝐞𝐭\mathbf{DSet}.

Proof.

(1) is immediate from the definition. (2) and (3) are consequences of [kechris_borel_2016, Lemma 3.7], after observing that the same proof there applies in the case of the more generous notion of idealistic equivalence relation considered here. (4) is a consequence of (2) and [motto_ros_complexity_2012, Proposition 2.3]. Finally, (5), (6), and (7) are easily verified. ∎

Occasionally we will need to consider quotients X=X^/EX=\hat{X}/E where X^\hat{X} is a standard Borel space EE is an analytic equivalence relation on X^\hat{X} that is not Borel and idealistic, or has not yet been shown to be Borel and idealistic. In this case, we say that X=X^/EX=\hat{X}/E is a semidefinable set. Clearly, every definable set is, in particular, a semidefinable set. The notion of Borel subset and definable function are the same as in the case of definable sets. Thus, if X=X^/EX=\hat{X}/E and Y=Y^/FY=\hat{Y}/F are semidefinable sets, ZXZ\subseteq X is a subset and f:XYf:X\rightarrow Y is a function, then ff is definable if it has a Borel lift f^:X^Y^\hat{f}:\hat{X}\rightarrow\hat{Y}, and ZZ is Borel if there is a Borel EE-invariant subset Z^\hat{Z} of X^\hat{X} such that Z=Z^/EZ=\hat{Z}/E. The category 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭\mathbf{SemiDSet} has semidefinable sets as objects and definable functions as morphisms. Notice that, in particular, an isomorphism from XX to YY in 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭\mathbf{SemiDSet} is a bijection f:XYf:X\rightarrow Y such that both ff and the inverse function f1:YXf^{-1}:Y\rightarrow X are definable.

Lemma 1.11.

Suppose that X=X^/EX=\hat{X}/E is a definable set, Y=Y^/FY=\hat{Y}/F is a semidefinable set. If XX and YY are isomorphic in 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭\mathbf{SemiDSet}, then YY is a definable set.

Proof.

Suppose that the Borel function sX:X^X^s_{X}:\hat{X}\rightarrow\hat{X} and the assignment CCC\mapsto\mathcal{E}_{C} witness that EE is idealistic. By assumption, there exists a bijection f:XYf:X\rightarrow Y such that ff has a Borel lift α:X^Y^\alpha:\hat{X}\rightarrow\hat{Y}, and f1f^{-1} has a Borel lift β:Y^X^\beta:\hat{Y}\rightarrow\hat{X}. For y,yY^y,y^{\prime}\in\hat{Y} we have that yFyyFy^{\prime} if and only if β(y)Eβ(y)\beta(y)E\beta\left(y^{\prime}\right), whence FF is Borel. We now show that FF is idealistic.

Define an assignment DDD\mapsto\mathcal{F}_{D} from FF-classes to σ\sigma-filters, by setting SDS\in\mathcal{F}_{D} if and only if α1(S)C\alpha^{-1}\left(S\right)\in\mathcal{E}_{C} where f(C)=Df\left(C\right)=D. Consider also the Borel map sY:=αsXβ:Y^Y^s_{Y}:=\alpha\circ s_{X}\circ\beta:\hat{Y}\rightarrow\hat{Y}. Then it is easy to verify that sYs_{Y} and the assignment DDD\mapsto\mathcal{F}_{D} witness that FF is idealistic. ∎

Lemma 1.12.

Suppose that X=X^/EX=\hat{X}/E and Y=Y^/FY=\hat{Y}/F are semidefinable sets. Assume that there exists a definable bijection f:XYf:X\rightarrow Y (which is not necessarily an isomorphism in 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭\mathbf{SemiDSet}). If EE is Borel, then FF is Borel.

Proof.

By assumption EX^×X^E\subseteq\hat{X}\times\hat{X} is Borel, and FY^×Y^F\subseteq\hat{Y}\times\hat{Y} is analytic. Furthermore, ff has a Borel lift f^:X^Y^\hat{f}:\hat{X}\rightarrow\hat{Y}. Since ff is a bijection, we have that, for y,yY^y,y^{\prime}\in\hat{Y},

yFyx,xX^((f^(x)Fyf^(x)Fy)xEx).yFy^{\prime}\Leftrightarrow\forall x,x^{\prime}\in\hat{X}\text{, }\left((\hat{f}(x)Fy\wedge\hat{f}\left(x^{\prime}\right)Fy^{\prime})\rightarrow xEx^{\prime}\right)\text{.}

This shows that FF is co-analytic. As FF is also analytic, we have that FF is Borel. ∎

Lemma 3.7 in [kechris_borel_2016] can be stated as the following proposition, which generalizes items (2) and (3) in Proposition 1.10.

Proposition 1.13 (Kechris–Macdonald).

Let X=X^/EX=\hat{X}/E be a definable set, Y=Y^/FY=\hat{Y}/F be semidefinable set such that FF is Borel, and f:XYf:X\rightarrow Y be a definable function. If ff is injective, then the range of ff a Borel subset of YY. If ff is bijective, then the inverse function f1:YXf^{-1}:Y\rightarrow X is definable.

The following result is a consequence of Lemma 1.12 and Proposition 1.13.

Corollary 1.14.

Suppose that X=X^/EX=\hat{X}/E is a definable set, and Y=Y^/FY=\hat{Y}/F is a semidefinable set. If f:XYf:X\rightarrow Y is a definable bijection, then YY is a definable set and ff is an isomorphism in 𝐃𝐒𝐞𝐭\mathbf{DSet}.

Proof.

By Lemma 1.12, FF is Borel. Whence, by Proposition 1.13, ff is an isomorphism in 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭\mathbf{SemiDSet}. Since XX is a definable set, it follows from Lemma 1.11 that YY is also a definable set, and ff is an isomorphism in 𝐃𝐒𝐞𝐭\mathbf{DSet}. ∎

1.5. Definable groups

A definable group can be simply defined as a group in the category 𝐃𝐒𝐞𝐭\mathbf{DSet} in the sense of [mac_lane_categories_1998, Section III.6]. Thus, a definable group is a definable set G=G^/EG=\hat{G}/E that is also a group, and such that the group operation G×GGG\times G\rightarrow G is definable, and the function GGG\rightarrow G, xx1x\mapsto x^{-1} is also definable. As in the case of sets, we regard a standard Borel group as a particular instance of definable group G=G^/EG=\hat{G}/E where G=G^G=\hat{G} is a standard Borel group and EE is the relation of equality on G^\hat{G}. Thus, standard Borel groups form a full subcategory of the category of definable groups.

Naturally, a semidefinable group will be a group in 𝐒𝐞𝐦𝐢𝐃𝐒𝐞𝐭𝐬\mathbf{SemiDSets}, i.e. a semidefinable set G=G^/EG=\hat{G}/E that is also a group, and such that the group operation G×GGG\times G\rightarrow G is definable, and the function GGG\rightarrow G that maps every element to its inverse is definable.

Lemma 1.15.

If G=G^/EG=\hat{G}/E is a semidefinable group, then the equivalence relation EE is Borel if and only if the identity element of GG, which is the EE-class []E\left[\ast\right]_{E} of some element \ast of G^\hat{G}, is a Borel subset of G^\hat{G}.

Proof.

Clearly, if EE is Borel, then []E\left[\ast\right]_{E} is Borel. Conversely, suppose that []E\left[\ast\right]_{E} is Borel. If m:G^×G^G^m:\hat{G}\times\hat{G}\rightarrow\hat{G} and ζ:G^G^\zeta:\hat{G}\rightarrow\hat{G} are Borel lifts of the group operation in XX and of the function that maps each element to its inverse, respectively, then we have that xEyxEy if and only if m(x,ζ(y))[]Em(x,\zeta(y))\in[\ast]_{E}. This shows that EE is Borel. ∎

Corollary 1.16.

Suppose that G=G^/EG=\hat{G}/E is a semidefinable group. If EE is the orbit equivalence relation of a Borel action of a Polish group HH on the standard Borel space G^\hat{G}, then GG is a definable group.

Proof.

By [becker_descriptive_1996, Theorem 5.2.1] one can assume that G^\hat{G} is a Polish HH-space, and EE is the orbit equivalence relation of a continuous HH-action on G^\hat{G}. By [gao_invariant_2009, Proposition 3.1.10], every EE-class is Borel. Therefore EE is Borel by Lemma 1.15. Furthermore, EE is idealistic by [gao_invariant_2009, Proposition 5.4.10]. ∎

Remark 1.17.

A particular instance of definable group is obtained as follows. Suppose that GG is a Polish group and HH is a Borel Polishable subgroup. Let EHGE_{H}^{G} be the coset equivalence relation of HH in GG. The quotient group G/HG/H is the quotient of GG by the equivalence relation EHGE_{H}^{G}. Since HH is Polishable, EHGE_{H}^{G} is the orbit equivalence relation of a Borel action of a Polish group on GG. Thus, G/H=G/EHGG/H=G/E_{H}^{G} is a definable group by Corollary 1.16. The definable groups obtained in this way are called groups with a Polish cover in [bergfalk_ulam_2020].

2. Strict C*-algebras

In this section we introduce the notion of strict Banach space and strict C*-algebra and some of their properties. Briefly, a strict Banach space is a Banach space whose unit ball is endowed with a Polish topology (called the strict topology) that is coarser than the norm-topology and induced by a sequence of bounded seminorms. A suitable semicontinuity requirement relates the norm and the strict topology. A strict C*-algebra is a strict Banach space that is also a C*-algebra with some suitable continuity requirement relating the C*-algebra operations and the strict topology. The name is inspired by the strict topology on the multiplier algebra of a separable C*-algebra, which will be one of the main examples. Other examples are Paschke dual algebras of separable C*-algebras.

2.1. Strict Banach spaces

Let XX be a Banach space. We denote by Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) its unit ball. A seminorm pp on XX is bounded if p:=supxBall(X)p(x)<\left\|p\right\|:=\mathrm{sup}_{x\in\mathrm{\mathrm{Ball}}\left(X\right)}p(x)<\infty. We say that pp is contractive if p1\left\|p\right\|\leq 1.

Definition 2.1.

A strict Banach space is a Banach space 𝔛\mathfrak{X} such that Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is endowed with a topology (called the strict topology) such that, for some sequence (pn)\left(p_{n}\right) of contractive seminorms on 𝔛\mathfrak{X}, letting dd be the pseudometric on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) defined by

d(x,y)=nω2npn(xy),d\left(x,y\right)=\sum_{n\in\omega}2^{-n}p_{n}\left(x-y\right)\text{,}

one has that:

  1. (1)

    dd is a complete metric that induces the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right);

  2. (2)

    Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) contains a countable strictly dense subset;

  3. (3)

    x=supnωpn(x)\left\|x\right\|=\mathrm{sup}_{n\in\omega}p_{n}(x) for every x𝔛x\in\mathfrak{X}.

Example 2.2.

Suppose that XX is a separable Banach space. Then XX is a strict Banach space where the strict topology on Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) is the norm topology.

Example 2.3.

Suppose that YY is a separable Banach space, and YY^{\ast} is its Banach space dual. Then YY^{\ast} is a strict Banach space where the strict topology on Ball(Y)\mathrm{\mathrm{Ball}}\left(Y^{\ast}\right) is the weak*-topology.

Let XX be a seminormed space, and consider the cone 𝒮(X)\mathcal{S}\left(X\right) of bounded seminorms on XX as a complete metric space, with respect to the metric defined by d(p,q)=supxBall(X)|p(x)q(x)|d\left(p,q\right)=\mathrm{sup}_{x\in\mathrm{\mathrm{Ball}}\left(X\right)}\left|p(x)-q(x)\right|. For a subset 𝔖𝒮(X)\mathfrak{S}\subseteq\mathcal{S}\left(X\right), we let σ(X,𝔖)\sigma\left(X,\mathfrak{S}\right) be the topology on Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) generated by 𝔖\mathfrak{S}. We denote by Ball(𝔖)\left(\mathfrak{S}\right) the set of contractive seminorms in 𝔖\mathfrak{S}. If p𝒮(X)p\in\mathcal{S}\left(X\right), 𝔖𝒮(X)\mathfrak{S}\subseteq\mathcal{S}\left(X\right), and (xn)\left(x_{n}\right) is a sequence in Ball(X)\mathrm{\mathrm{Ball}}\left(X\right), then we say that:

  • (xn)nω\left(x_{n}\right)_{n\in\omega} is pp-Cauchy if for every ε>0\varepsilon>0 there exists n0ωn_{0}\in\omega such that, for n,mn0n,m\geq n_{0}, p(xnxm)<εp\left(x_{n}-x_{m}\right)<\varepsilon;

  • (xn)nω\left(x_{n}\right)_{n\in\omega} is 𝔖\mathfrak{S}-Cauchy if it is pp-Cauchy for every p𝔖p\in\mathfrak{S};

  • Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) is 𝔖\mathfrak{S}-complete if, for every sequence (xn)nω\left(x_{n}\right)_{n\in\omega} in Ball(X)\left(X\right), if (xn)nω\left(x_{n}\right)_{n\in\omega} is 𝔖\mathfrak{S}-Cauchy, then (xn)nω\left(x_{n}\right)_{n\in\omega} is σ(X,𝔖)\sigma\left(X,\mathfrak{S}\right)-convergent to some element of Ball(X)\mathrm{\mathrm{Ball}}\left(X\right).

The following lemma is elementary.

Lemma 2.4.

Suppose that XX is a seminormed space. Let τ\tau be a topology on Ball(X)\mathrm{\mathrm{Ball}}\left(X\right). Assume that 𝔖\mathfrak{S} and 𝔗\mathfrak{T} are two sets of bounded seminorms on XX such that the topologies σ(X,𝔖)\sigma\left(X,\mathfrak{S}\right) and σ(X,𝔗)\sigma\left(X,\mathfrak{T}\right) on Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) coincide with τ\tau. Then, for a sequence (xn)\left(x_{n}\right) in Ball(X)\mathrm{\mathrm{Ball}}\left(X\right), (xn)\left(x_{n}\right) is 𝔖\mathfrak{S}-Cauchy if and only if it is 𝔗\mathfrak{T}-Cauchy. In this case, we say that (xn)\left(x_{n}\right) is τ\tau-Cauchy. It follows that Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) is 𝔖\mathfrak{S}-complete if and only if it is 𝔗\mathfrak{T}-complete. In this case, we say that Ball(X)\mathrm{\mathrm{Ball}}\left(X\right) is τ\tau-complete.

In view of Lemma 2.4 one can equivalently define a strict Banach space as follows.

Definition 2.5.

A strict Banach space is a Banach space 𝔛\mathfrak{X} such that Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is endowed with a topology (called the strict topology) such that, for some separable cone 𝔖\mathfrak{S} of bounded seminorms on XX, one has that:

  1. (1)

    the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is the σ(𝔛,𝔖)\sigma\left(\mathfrak{X},\mathfrak{S}\right)-topology, and Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is strictly complete;

  2. (2)

    Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) contains a countable strictly dense subset;

  3. (3)

    x=suppBall(𝔖)p(x)\left\|x\right\|=\mathrm{sup}_{p\in\mathrm{\mathrm{Ball}}\left(\mathfrak{S}\right)}p(x) for every x𝔛x\in\mathfrak{X}.

Proposition 2.6.

Suppose that 𝔛\mathfrak{X} is a strict Banach space. Then Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is a Polish topometric space when endowed with the strict topology and the norm-distance.

Proof.

By definition, the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is Polish. Since the strict topology is induced by bounded seminorms on XX, it is coarser than the norm topology. The function (x,y)xy\left(x,y\right)\mapsto\left\|x-y\right\| is strictly lower-semicontinuous, being the supremum of strictly continuous functions. Since the norm on XX is complete, the distance (x,y)xy\left(x,y\right)\mapsto\left\|x-y\right\| on Ball(𝔛)\left(\mathfrak{X}\right) is complete. ∎

Suppose that 𝔛\mathfrak{X} is a strict Banach space. We extend the strict topology of Ball(𝔛)\left(\mathfrak{X}\right) to any bounded subset of 𝔛\mathfrak{X} by declaring the function

nBall(𝔛)Ball(𝔛)z1nzn\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right)\rightarrow\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right)\text{, }z\mapsto\frac{1}{n}z

to be a homeomorphism with respect to the strict topology, where

nBall(𝔛)={z𝔛:zn}.n\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right)=\left\{z\in\mathfrak{X}:\left\|z\right\|\leq n\right\}\text{.}

Then we have that addition and scalar multiplication on 𝔛\mathfrak{X} are strictly continuous on bounded sets, and the norm is strictly lower-semicontinuous on bounded sets. In particular nBall(X)n\mathrm{\mathrm{Ball}}\left(X\right) is a strictly closed subspace of mBall(X)m\mathrm{\mathrm{Ball}}\left(X\right) for nmn\leq m. Notice that, if 𝔜\mathfrak{Y} is a norm-closed subspace of XX such that Ball(𝔜)\mathrm{\mathrm{Ball}}\left(\mathfrak{Y}\right) is strictly closed in Ball(𝔛)\left(\mathfrak{X}\right), then 𝔜\mathfrak{Y} is a strict Banach space with the induced norm and the induced strict topology.

Definition 2.7.

Let 𝔛\mathfrak{X} be a strict Banach space. The (standard) Borel structure on 𝔛\mathfrak{X} is defined by declaring a subset AA of 𝔛\mathfrak{X} to be Borel if and only if AnBall(𝔛)A\cap n\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is Borel for every n1n\geq 1.

Notice that the Borel structure on 𝔛\mathfrak{X} is standard, as 𝔛\mathfrak{X} is Borel isomorphic to the disjoint union of the standard Borel spaces (n+1)Ball(𝔛)nBall(𝔛)\left(n+1\right)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right)\setminus n\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) for n1n\geq 1.

Definition 2.8.

If 𝔛\mathfrak{X} and 𝔜\mathfrak{Y} are strict Banach spaces. A bounded linear map T:𝔛𝔜T:\mathfrak{X}\rightarrow\mathfrak{Y} is contractive if T1\left\|T\right\|\leq 1, and strict if it is strictly continuous on bounded sets. A bounded seminorm pp on 𝔛\mathfrak{X} is strict if it is strictly continuous on bounded sets.

Clearly, strict Banach spaces form a category where the morphisms are the strict contractive linear maps. Notice that, if TT is a strict, bijective, and isometric linear map T:𝔛𝔜T:\mathfrak{X}\rightarrow\mathfrak{Y} between strict Banach spaces, then the inverse T1:𝔜𝔛T^{-1}:\mathfrak{Y}\rightarrow\mathfrak{X} is not necessarily strict, whence TT is not necessarily an isomorphism in the category of strict Banach spaces. Nonetheless, T:𝔛𝔜T:\mathfrak{X}\rightarrow\mathfrak{Y} is a Borel isomorphism, as both 𝔛\mathfrak{X} and 𝔜\mathfrak{Y} are standard Borel spaces.

Definition 2.9.

Let 𝔛\mathfrak{X} be a strict Banach space. Define 𝒮strict(𝔛)\mathcal{S}_{\mathrm{strict}}\left(\mathfrak{X}\right) to be the space of bounded, strict seminorms on 𝔛\mathfrak{X}.

Notice that 𝒮strict(𝔛)\mathcal{S}_{\mathrm{strict}}\left(\mathfrak{X}\right) is a closed subspace of the complete metric space 𝒮(𝔛)\mathcal{S}\left(\mathfrak{X}\right). A sequence (xn)\left(x_{n}\right) in Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is strictly convergent if and only if it is 𝒮strict(𝔛)\mathcal{S}_{\mathrm{strict}}\left(\mathfrak{X}\right)-Cauchy. A bounded linear map T:𝔛YT:\mathfrak{X}\rightarrow Y is strict if and only if pT𝒮strict(𝔛)p\circ T\in\mathcal{S}_{\mathrm{strict}}\left(\mathfrak{X}\right) for every p𝒮strict(Y)p\in\mathcal{S}_{\mathrm{strict}}\left(Y\right).

Remark 2.10.

Suppose that 𝔛\mathfrak{X} is a strict Banach space, and 𝔖\mathfrak{S} is a separable cone of bounded, strict seminorms on 𝔛\mathfrak{X} that induces the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right). One can consider the globally defined topology σ(𝔛,𝔖)\sigma\left(\mathfrak{X},\mathfrak{S}\right) on 𝔛\mathfrak{X}, induced by all the seminorms in 𝔖\mathfrak{S}. This topology coincides with the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right). However, it is not first countable on the whole of 𝔛\mathfrak{X}, unless 𝔛\mathfrak{X} is a separable Banach space and the strict topology is equal to the norm topology. Indeed, if the σ(𝔛,𝔖)\sigma\left(\mathfrak{X},\mathfrak{S}\right)-topology on 𝔛\mathfrak{X} is first-countable, then (𝔛,σ(𝔛,𝔖))\left(\mathfrak{X},\sigma\left(\mathfrak{X},\mathfrak{S}\right)\right) is a Frechet space. By the Open Mapping Theorem for Frechet spaces [robertson_topological_1964, Theorem 8, page 120], any two comparable Frechet space topologies must be equal. Thus, σ(𝔛,𝔖)\sigma\left(\mathfrak{X},\mathfrak{S}\right) equals the norm topology. In particular, the norm-topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is equal to the strict topology, and it has a countable dense subset. Hence the norm-topology on 𝔛\mathfrak{X} is separable.

For future reference, we record the easily proved observation that a uniform limit of strictly continuous functions is strictly continuous.

Lemma 2.11.

Suppose that 𝔛\mathfrak{X} and 𝔜\mathfrak{Y} are strict Banach spaces, and ABall(𝔛)A\subseteq\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right). Suppose that f:A𝔜f:A\rightarrow\mathfrak{Y} is an function. Assume that there exists an sequence (fn)\left(f_{n}\right) of strictly continuous function fn:A𝔜f_{n}:A\rightarrow\mathfrak{Y} such that

limnsupxAfn(x)f(x)=0.\mathrm{lim}_{n\rightarrow\infty}\mathrm{sup}_{x\in A}\left\|f_{n}(x)-f(x)\right\|=0\mathrm{.}

Then ff is strictly continuous.

A standard Baire Category argument shows that one can characterize bounded subsets in terms of bounded, strict seminorms, as follows.

Lemma 2.12.

Let 𝔛\mathfrak{X} be a strict Banach space. If A𝔛A\subseteq\mathfrak{X}, then AA is bounded if and only if, for every p𝒮strict(𝔛)p\in\mathcal{S}_{\mathrm{strict}}\left(\mathfrak{X}\right), p(A)p\left(A\right) is a bounded subset of \mathbb{R}.

A natural way to obtain strict Banach spaces is via pairings.

Definition 2.13.

A Banach pairing is a bounded bilinear map ,:𝔛×YZ\left\langle\cdot,\cdot\right\rangle:\mathfrak{X}\times Y\rightarrow Z, where 𝔛,Y,Z\mathfrak{X},Y,Z are Banach spaces. Define the σ(𝔛,Y)\sigma\left(\mathfrak{X},Y\right)-topology to be the topology on 𝔛\mathfrak{X} generated by the cone 𝔖Y\mathfrak{S}_{Y} of bounded seminorms xx,yx\mapsto\left\|\left\langle x,y\right\rangle\right\| for yYy\in Y.

The following lemma is an immediate consequence of the definition of strict Banach space.

Lemma 2.14.

Suppose that ,:𝔛×YZ\left\langle\cdot,\cdot\right\rangle:\mathfrak{X}\times Y\rightarrow Z is a Banach pairing. Assume that:

  • Y,ZY,Z are norm-separable Banach spaces;

  • for every x0𝔛x_{0}\in\mathfrak{X},

    x0=supyBall(Y)x0,y;\left\|x_{0}\right\|=\mathrm{sup}_{y\in\mathrm{\mathrm{Ball}}\left(Y\right)}\left\|\left\langle x_{0},y\right\rangle\right\|\text{;}
  • Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is σ(𝔛,Y)\sigma\left(\mathfrak{X},Y\right)-complete;

  • Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) has a countable σ(𝔛,Y)\sigma\left(\mathfrak{X},Y\right)-dense subset.

Then 𝔛\mathfrak{X} is a strict Banach space where the strict topology on Ball(𝔛)\mathrm{\mathrm{Ball}}\left(\mathfrak{X}\right) is the σ(𝔛,Y)\sigma\left(\mathfrak{X},Y\right)-topology.

Suppose that XX is a norm-separable Banach space, and 𝔜\mathfrak{Y} is a strict Banach spaces. A linear map T:X𝔜T:X\rightarrow\mathfrak{Y} is bounded if it maps bounded sets to bounded sets or, equivalently,

T=supxBall(X)T(x)<.\left\|T\right\|=\mathrm{sup}_{x\in\mathrm{\mathrm{Ball}}\left(X\right)}\left\|T(x)\right\|<\infty\text{.}

This defines a norm on the space L(X,𝔜)L\left(X,\mathfrak{Y}\right) of bounded linear maps X𝔜X\rightarrow\mathfrak{Y}. We also define the strict topology on Ball(L(X,𝔜))\left(L\left(X,\mathfrak{Y}\right)\right) to be the topology of pointwise convergence in the strict topology of Ball(𝔜)\left(\mathfrak{Y}\right). Then one can easily show the following.

Proposition 2.15.

Suppose that XX is a norm-separable Banach space, and 𝔜\mathfrak{Y} is a strict Banach space. Then L(X,𝔜)L\left(X,\mathfrak{Y}\right) is a strict Banach space.

2.2. Strict C*-algebras

We now introduce the notion of strict C*-algebra. Given a C*-algebra AA, we let AsaA_{\mathrm{sa}} be the set of its self-adjoint elements. We also denote by Mn(A)M_{n}\left(A\right) the C*-algebra of n×nn\times n matrices over AA, which can be identified with the tensor product Mn()AM_{n}\left(\mathbb{C}\right)\otimes A. We refer to [blackadar_operator_2006, davidson_algebras_1996, murphy_algebras_1990, pedersen_algebras_1979] for fundamental notions and results from the theory of C*-algebras.

Definition 2.16.

A strict C*-algebra is a C*-algebra 𝔄\mathfrak{A} such that, for every n1n\geq 1, Mn(𝔄)M_{n}\left(\mathfrak{A}\right) is also a strict Banach space satisfying the following properties:

  1. (1)

    the *-operation and the multiplication operation on Mn(𝔄)M_{n}\left(\mathfrak{A}\right) are strictly continuous on bounded sets;

  2. (2)

    the strict topology on Ball(Mn(𝔄))\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathfrak{A}\right)\right) is induced by the inclusion

    Ball(Mn(𝔄))Mn(Ball(𝔄)),\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathfrak{A}\right)\right)\subseteq M_{n}\left(\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)\right)\text{,}

    where Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) is endowed with the strict topology, and Mn(Ball(𝔄))M_{n}\left(\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)\right) is endowed with the product topology.

Example 2.17.

Suppose that AA is a separable C*-algebra. Then we have that AA is a strict C*-algebra where, for every n1n\geq 1, Ball(Mn(A))\mathrm{\mathrm{Ball}}\left(M_{n}\left(A\right)\right) is endowed with the norm-topology.

Suppose that 𝔄\mathfrak{A} is a strict C*-algebra. Then, for every n1n\geq 1, Mn(𝔄)M_{n}\left(\mathfrak{A}\right) is also a strict C*-algebra. If 𝔄\mathfrak{A} is a strict C*-algebra, then we regard 𝔄\mathfrak{A} as a standard Borel space with respect to the standard Borel structure induced by the strict topology on Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) as in Definition 2.7. We say that a subset of 𝔄\mathfrak{A} is Borel if it is Borel with respect to such a Borel structure. We have that the Borel structure on Mn(𝔄)M_{n}\left(\mathfrak{A}\right) (as a strict C*-algebra) coincides with the product Borel structure.

Definition 2.18.

Suppose that 𝔄\mathfrak{A} is a strict unital C*-algebra. A strict ideal of 𝔄\mathfrak{A} is a norm-closed proper two-sided Borel ideal 𝔍\mathfrak{J} of 𝔄\mathfrak{A} that is also a strict C*-algebra, and such that the inclusion map 𝔍𝔄\mathfrak{J}\rightarrow\mathfrak{A} is strict.

Remark 2.19.

In order for 𝔍\mathfrak{J} to be a strict ideal of 𝔄\mathfrak{A}, we do not require that Ball(𝔍)\mathrm{\mathrm{Ball}}\left(\mathfrak{J}\right) be strictly closed in Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) nor that the strict topology on Ball(𝔍)\mathrm{\mathrm{Ball}}\left(\mathfrak{J}\right) be the subspace topology induced by the strict topology of Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right).

Example 2.20.

Suppose that 𝔄\mathfrak{A} is a strict unital C*-algebra and J𝔄J\subseteq\mathfrak{A} is a norm-closed and norm-separable proper two-sided ideal of 𝔄\mathfrak{A}. Then JJ is a strict ideal of 𝔄\mathfrak{A}.

We regard strict (unital) C*-algebras as objects of a category with strict (unital) *-homomorphisms as morphisms. (Recall that a bounded linear map is strict if it is strictly continuous on bounded sets.) If 𝔄𝔅\mathfrak{A}\subseteq\mathfrak{B}, then we say that 𝔄\mathfrak{A} is strictly dense in 𝔅\mathfrak{B} if Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) is dense in Ball(𝔅)\mathrm{\mathrm{Ball}}\left(\mathfrak{B}\right) with respect to the strict topology.

It follows from the axioms of a strict C*-algebra that, if 𝔄\mathfrak{A} is a strict C*-algebra, and p(x1,,xn)p\left(x_{1},\ldots,x_{n}\right) is a *-polynomial, then pp defines a function 𝔄n𝔄\mathfrak{A}^{n}\rightarrow\mathfrak{A} that is strictly continuous on bounded sets. In particular, the sets of normal, self-adjoint, and positive elements of norm at most 11 are strictly closed in Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right). If f:[1,1]nnBall()f:[-1,1]^{n}\rightarrow n\mathrm{\mathrm{Ball}}\left(\mathbb{C}\right) is a continuous function, then ff induces by continuous functional calculus and Lemma 2.11 a strictly continuous functions (x1,,xn)f(x1,,xn)\left(x_{1},\ldots,x_{n}\right)\mapsto f\left(x_{1},\ldots,x_{n}\right) from the strictly closed set of nn-tuples of pairwise commuting self-adjoint elements in Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) to nBall(𝔄)n\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right). Similarly, if f:Ball()nkBall()f:\mathrm{\mathrm{Ball}}\left(\mathbb{C}\right)^{n}\rightarrow k\mathrm{\mathrm{Ball}}\left(\mathbb{C}\right) is a continuous function, then ff induces by continuous functional calculus and Lemma 2.11 a strictly continuous function (x1,,xn)f(x1,,xn)\left(x_{1},\ldots,x_{n}\right)\mapsto f\left(x_{1},\ldots,x_{n}\right) from the strictly closed set of nn-tuples of pairwise commuting normal elements in Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) to kBall(𝔄)k\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right).

Suppose that 𝔄\mathfrak{A} is a strict C*-algebra. Let Normal(𝔄)\mathrm{Normal}\left(\mathfrak{A}\right) be the Borel set of normal elements of 𝔄\mathfrak{A}. For aNormal(𝔄)a\in\mathrm{Normal}\left(\mathfrak{A}\right), the spectrum σ(a)\sigma\left(a\right) is a closed subset of \mathbb{C}. We consider the space Closed()\mathrm{Closed}\left(\mathbb{C}\right) of closed subsets of \mathbb{C} as a standard Borel space endowed with the Effros Borel structure [kechris_classical_1995, Section 12.C]. If XX is a standard Borel space and \mathcal{B} is a basis of open subsets of \mathbb{C}, then a function Φ:XClosed()\Phi:X\rightarrow\mathrm{Closed}\left(\mathbb{C}\right) is Borel if and only if, for every UU\in\mathcal{B}, {xX:Φ(x)U}\left\{x\in X:\Phi(x)\cap U\neq\varnothing\right\} is Borel. The proof of the following lemma is standard; see [simon_operators_1995, Lemma 1.6].

Lemma 2.21.

Suppose that 𝔄\mathfrak{A} is a strict C*-algebra. The function Normal(𝔄)Closed()\mathrm{Normal}\left(\mathfrak{A}\right)\rightarrow\mathrm{Closed}\left(\mathbb{C}\right), aσ(a)a\mapsto\sigma\left(a\right) is Borel.

Proof.

It suffices to show that the map Normal(𝔄)Ball(𝔄)Closed()\mathrm{Normal}\left(\mathfrak{A}\right)\cap\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)\rightarrow\mathrm{Closed}\left(\mathbb{C}\right), aσ(a)a\mapsto\sigma\left(a\right) is Borel. Observe that \mathbb{C} has a basis of open sets of the form Uf:={x:f(x)>0}U_{f}:=\left\{x\in\mathbb{C}:f(x)>0\right\} where f:[0,1]f:\mathbb{C}\rightarrow[0,1] is a continuous function. For such a continuous function f:[0,1]f:\mathbb{C}\rightarrow[0,1], we have that

{aNormal(𝔄)Ball(𝔄):σ(a)Uf}={aNormal(𝔄)Ball(𝔄):f(a)0},\left\{a\in\mathrm{Normal}\left(\mathfrak{A}\right)\cap\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right):\sigma\left(a\right)\cap U_{f}\neq\varnothing\right\}=\left\{a\in\mathrm{Normal}\left(\mathfrak{A}\right)\cap\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right):f\left(a\right)\neq 0\right\}\text{,}

which is closed in Normal(𝔄)Ball(𝔄)\mathrm{Normal}\left(\mathfrak{A}\right)\cap\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right). This concludes the proof. ∎

Suppose that 𝔄\mathfrak{A} is a strict C*-algebra. Fix r(0,1)r\in\left(0,1\right) and consider the set

X={x𝔄:1xr}2Ball(𝔄).X=\left\{x\in\mathfrak{A}:\left\|1-x\right\|\leq r\right\}\subseteq 2\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)\text{.}

Then, for xXx\in X we have that xx is invertible, x111r\left\|x^{-1}\right\|\leq\frac{1}{1-r}, and

x1=nωxn.x^{-1}=\sum_{n\in\omega}x^{n}\text{.}

It follows from Lemma 2.11 that the function X11rBall(𝔄)X\rightarrow\frac{1}{1-r}\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right), xx1x\mapsto x^{-1} is strictly continuous.

More generally, suppose that Ω\Omega is an open subset of \mathbb{C}, and f:Ωf:\Omega\rightarrow\mathbb{C} is a holomorphic function. Suppose that 0Ω0\in\Omega and r>0r>0 is such that {z:|z|r}Ω\left\{z\in\mathbb{C}:\left|z\right|\leq r\right\}\subseteq\Omega. Then ff admits a Taylor expansion

f(z)=n=0anznf(z)=\sum_{n=0}^{\infty}a_{n}z^{n}

that converges uniformly for |z|r\left|z\right|\leq r [ahlfors_complex_1978, Chapter 5, Theorem 3 and Chapter 2, Theorem 2]. Fix b0𝔄b_{0}\in\mathfrak{A} and set

X:={x𝔄:xb0r}(1+b0)Ball(𝔄)X:=\left\{x\in\mathfrak{A}:\left\|x-b_{0}\right\|\leq r\right\}\subseteq\left(1+\left\|b_{0}\right\|\right)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)

Then for xXx\in X,

f(xb0):=n=0an(xb0)n𝔄;f\left(x-b_{0}\right):=\sum_{n=0}^{\infty}a_{n}\left(x-b_{0}\right)^{n}\in\mathfrak{A}\text{;}

see [pedersen_analysis_1989, Lemma 4.1.11]. Furthermore, the function XcBall(𝔄)X\rightarrow c\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right), xf(xb0)x\mapsto f\left(x-b_{0}\right) is strictly continuous on XX by Lemma 2.11, where c=sup{|f(z)|:|z|r}c=\mathrm{sup}\left\{\left|f(z)\right|:\left|z\right|\leq r\right\}.

2.3. Multiplier algebras

Suppose that AA is a separable C*-algebra. A double centralizer for AA is a pair (L,R)\left(L,R\right) of bounded linear maps L,R:AAL,R:A\rightarrow A such that L=R\left\|L\right\|=\left\|R\right\| and L(x)y=xR(y)L(x)y=xR(y) for every x,yAx,y\in A. Let M(A)M\left(A\right) be the set of double centralizers for AA. Then M(A)M\left(A\right) is a C*-algebra with respect to the operations

(L1,R1)+(L2,R2)=(L1+L2,R1+R2)\left(L_{1},R_{1}\right)+\left(L_{2},R_{2}\right)=\left(L_{1}+L_{2},R_{1}+R_{2}\right)
(L1,R1)(L2,R2)=(L1L2,R2R1)\left(L_{1},R_{1}\right)\left(L_{2},R_{2}\right)=\left(L_{1}L_{2},R_{2}R_{1}\right)
λ(L,R)=(λL,λR)\lambda\left(L,R\right)=\left(\lambda L,\lambda R\right)
(L,R)=(R,L)\left(L,R\right)^{\ast}=\left(R^{\ast},L^{\ast}\right)

and the norm

(L,R)=L=R\left\|\left(L,R\right)\right\|=\left\|L\right\|=\left\|R\right\|

for (L,R),(L1,R1),(L2,R2)M(A)\left(L,R\right),\left(L_{1},R_{1}\right),\left(L_{2},R_{2}\right)\in M\left(A\right) and λ\lambda\in\mathbb{C}. The strict topology on Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right) is the topology of pointwise convergence, namely the topology induced by the seminorms

pa:(L,R)max{L(a),R(a)}p_{a}:\left(L,R\right)\mapsto\max\{\left\|L\left(a\right)\right\|,\left\|R\left(a\right)\right\|\}

for aAa\in A.

An element aAa\in A can be identified with the multiplier (La,Ra)M(A)\left(L_{a},R_{a}\right)\in M\left(A\right) defined by setting La(x)=axL_{a}(x)=ax and Ra(x)=xaR_{a}(x)=xa for xXx\in X. This allows one to regard AA as an essential ideal of M(A)M\left(A\right). (An ideal JJ of a C*-algebra BB is essential if J:={bB:bJ=0}J^{\bot}:=\left\{b\in B:bJ=0\right\} is zero or, equivalently, JJ has nonzero intersection with every nonzero ideal of BB.) If (vn)nω\left(v_{n}\right)_{n\in\omega} is an approximate unit for AA [higson_analytic_2000, Definition 1.7.1] then, by definition, (vn)\left(v_{n}\right) strictly converges to 11 in Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right). In particular, Ball(A)\mathrm{\mathrm{Ball}}\left(A\right) is strictly dense in Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right).

If (xi)iω\left(x_{i}\right)_{i\in\omega} is a strictly Cauchy sequence in Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right), in the sense that (xi)iω\left(x_{i}\right)_{i\in\omega} is pap_{a}-Cauchy for every aAa\in A, then setting

L(a):=limixiaL\left(a\right):=\mathrm{lim}_{i\rightarrow\infty}x_{i}a
R(a):=limiaxiR\left(a\right):=\mathrm{lim}_{i\rightarrow\infty}ax_{i}

for aAa\in A defines a double centralizer (L,R)Ball(M(A))\left(L,R\right)\in\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right) that is the strict limit of (xi)iω\left(x_{i}\right)_{i\in\omega} in Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right). For n1n\geq 1, one can identify Mn(M(A))M_{n}\left(M\left(A\right)\right) with M(Mn(A))M\left(M_{n}\left(A\right)\right) and consider the corresponding strict topology. From the above remarks and Lemma 2.14, one easily obtains the following; see [farah_combinatorial_2019, Chapter 13] or [wegge-olsen_theory_1993, Chapter 2].

Proposition 2.22.

Let AA be a separable C*-algebra. Then M(A)M\left(A\right) is a strict unital C*-algebra containing AA as a strictly dense essential strict ideal where, for every n1n\geq 1, the strict topology on Ball(M(A))\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right) is as described above, and Mn(M(A))M_{n}\left(M\left(A\right)\right) is identified with M(Mn(A))M\left(M_{n}\left(A\right)\right).

Example 2.23.

When AA is the algebra K(H)K\left(H\right) of compact operators on a separable Hilbert space, then M(A)=B(H)M\left(A\right)=B\left(H\right) and the strict topology on Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) is the strong-* topology [blackadar_operator_2006, Proposition I.8.6.3].

Example 2.24.

One can also regard B(H)B\left(H\right) as the dual space of the Banach space 𝔏1(H)\mathfrak{L}^{1}\left(H\right) of trace-class operators. This turns B(H)B\left(H\right) into a strict Banach space, where the strict topology on Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) is the weak* topology, which coincides with the weak operator topology [blackadar_operator_2006, Definition I.8.6.2]. As the identity map Ball(B(H))Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right)\rightarrow\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) is strong-*–weak continuous, the strong-* topology and weak operator topology on Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) define the same standard Borel structure on B(H)B\left(H\right).

One can define as above the strict topology on the whole multiplier algebra M(A)M\left(A\right) to be the topology of pointwise convergence of double multipliers. However, this topology on M(A)M\left(A\right) is not first countable whenever AA is not unital; see Remark 2.10.

Suppose that AA is a separable C*-algebra, and XX is a compact metrizable space. One can then consider the separable C*-algebra C(X,A)C\left(X,A\right) of continuous functions XAX\rightarrow A. Let also Cβ(X,M(A))C_{\beta}\left(X,M\left(A\right)\right) be the C*-algebra of strictly continuous bounded functions XM(A)X\rightarrow M\left(A\right). There is an obvious unital *-homomorphism Cβ(X,M(A))M(C(X,A))C_{\beta}\left(X,M\left(A\right)\right)\rightarrow M\left(C\left(X,A\right)\right), where Cβ(X,M(A))C_{\beta}\left(X,M\left(A\right)\right) acts on C(X,A)C\left(X,A\right) by pointwise multiplication. The unital *-homomorphism Cβ(X,M(A))M(C(X,A))C_{\beta}\left(X,M\left(A\right)\right)\rightarrow M\left(C\left(X,A\right)\right) is in fact a *-isomorphism [akemann_multipliers_1973, Corollary 3.4]. We can thus identify Cβ(X,M(A))C_{\beta}\left(X,M\left(A\right)\right) with M(C(X,A))M\left(C\left(X,A\right)\right) and regard it as a strict C*-algebra. Observe that, for tXt\in X, the function Ball(Cβ(X,M(A)))Ball(M(A))\mathrm{\mathrm{Ball}}\left(C_{\beta}\left(X,M\left(A\right)\right)\right)\rightarrow\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right), ff(t)f\mapsto f\left(t\right) is strictly continuous. We let C(X,M(A))C\left(X,M\left(A\right)\right) be the C*-algebra of norm-continuous functions XM(A)X\rightarrow M\left(A\right), which is a C*-subalgebra of Cβ(X,M(A))C_{\beta}\left(X,M\left(A\right)\right).

Lemma 2.25.

Suppose that AA is a separable C*-algebra, and XX is a compact metrizable space. Then C(X,M(A))C\left(X,M\left(A\right)\right) is a Borel subset of Cβ(X,M(A))C_{\beta}\left(X,M\left(A\right)\right).

Proof.

Fix a compatible metric dd on XX, and a countable dense subset X0X_{0} of XX. Clearly, it suffices to show that Ball(C(X,M(A)))\mathrm{\mathrm{Ball}}\left(C\left(X,M\left(A\right)\right)\right) is a Borel subset of Ball(Cβ(X,M(A)))\mathrm{\mathrm{Ball}}\left(C_{\beta}\left(X,M\left(A\right)\right)\right). Fix, for every kωk\in\omega, a finite cover {A0k,,Ak1k}\left\{A_{0}^{k},\ldots,A_{\ell_{k}-1}^{k}\right\} of XX consisting of open sets of diameter less than 2k2^{-k}, and fix elements tikAikt_{i}^{k}\in A_{i}^{k} for i<ki<\ell_{k}. We have that a strictly continuous function f:XBall(M(A))f:X\rightarrow\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right) is norm-continuous if and only if, for every nωn\in\omega there exists kωk\in\omega such that, for every i<ki<\ell_{k} and sAiks\in A_{i}^{k}, f(s)f(tik)2k\left\|f\left(s\right)-f\left(t_{i}^{k}\right)\right\|\leq 2^{-k}. Since 2kBall(M(A))2^{-k}\mathrm{\mathrm{Ball}}\left(M\left(A\right)\right) is strictly closed and ff is strictly continuous, we have that ff is norm-continuous if and only if for every nωn\in\omega there exists kωk\in\omega such that, for every i<ki<\ell_{k} and for every sAikX0s\in A_{i}^{k}\cap X_{0}, f(s)f(tik)2k\left\|f\left(s\right)-f\left(t_{i}^{k}\right)\right\|\leq 2^{-k}. This shows that the set of norm-continuous functions is Borel. ∎

Corollary 2.26.

Suppose that AA is a separable C*-algebra. Then the set C([0,1],M(A))C\left([0,1],M\left(A\right)\right) of norm-continuous paths [0,1]M(A)[0,1]\rightarrow M\left(A\right) is a Borel subset of Cβ([0,1],M(A))C_{\beta}\left([0,1],M\left(A\right)\right).

Suppose that AA and BB are separable C*-algebra. A morphism from AA to BB in the sense of [woronowicz_pseudospaces_1980, woronowicz_unbounded_1991, woronowicz_operator_1992, woronowicz_algebras_1995] is a *-homomorphism φ:AM(B)\varphi:A\rightarrow M(B) such that φ(A)B\varphi\left(A\right)B is norm-dense in BB. (This is called SS-morphism in [vallin_algebres_1985, Definition 0.2.7] and a nondegenerate *-homomorphism in [lance_hilbert_1995].) We recall the well-known fact that there is a correspondence between morphisms from AA to BB and strict unital *-homomorphisms M(A)M(B)M\left(A\right)\rightarrow M(B); see [lance_hilbert_1995, Proposition 2.1].

Lemma 2.27.

Let AA and BB be separable C*-algebra.

  • Suppose that ψ:M(A)M(B)\psi:M(A)\rightarrow M(B) is a strict unital *-homomorphism. Then ψ|A\psi|_{A} is a morphism from AA to BB.

  • Conversely, if φ\varphi is a morphism from AA to BB, then φ\varphi extends to a unique strict unital *-homomorphism φ¯:M(A)M(B)\bar{\varphi}:M(A)\rightarrow M(B). If φ\varphi is injective, then φ¯\bar{\varphi} is injective.

  • If (en)\left(e_{n}\right) is an approximate unit for AA, then a *-homomorphism φ:AM(B)\varphi:A\rightarrow M(B) is a morphism from AA to BB if and only if (φ(en))\left(\varphi\left(e_{n}\right)\right) strictly converges to 11.

A further characterization of morphisms is provided in [vallin_algebres_1985, Lemme 0.2.6] and [iorio_hopf_1980, Proposition 1.1]. It follows from Lemma 2.27 that the composition of morphisms ABA\rightarrow B and BCB\rightarrow C is meaningful, and it gives a morphism ACA\rightarrow C.

Suppose that A,BA,B are separable C*-algebras. A *-homomorphism φ:AM(B)\varphi:A\rightarrow M(B) is quasi-unital [jensen_elements_1991, Definition 1.3.13] (also called strict [lance_hilbert_1995, page 49]) if there exists a projection pφM(B)p_{\varphi}\in M(B), called the relative unit of φ\varphi, such that φ(A)B¯=pφB\overline{\varphi(A)B}=p_{\varphi}B. One has the following generalization of Lemma 2.27; see [lance_hilbert_1995, Corollary 5.7].

Lemma 2.28.

Let AA and BB be separable C*-algebra.

  • Suppose that ψ:M(A)M(B)\psi:M(A)\rightarrow M(B) is a strict *-homomorphism. Then ψ|A\psi|_{A} is a quasi-unital *-homomorphism from AA to M(B)M(B) with relative unit ψ(1)\psi\left(1\right).

  • Conversely, if φ\varphi is a quasi-unital *-homomorphism from AA to M(B)M(B) with relative unit pφp_{\varphi}, then φ\varphi extends to a unique strict *-homomorphism φ¯:M(A)M(B)\bar{\varphi}:M(A)\rightarrow M(B) with φ¯=pφ\bar{\varphi}=p_{\varphi}. If φ\varphi is injective, then φ¯\bar{\varphi} is injective.

  • If (en)\left(e_{n}\right) is an approximate unit for AA, then a *-homomorphism φ:AM(B)\varphi:A\rightarrow M(B) is quasi-unital if and only if (φ(en))\left(\varphi\left(e_{n}\right)\right) is strictly Cauchy.

We now observe that the category of multiplier algebras of separable C*-algebras, regarded as a full subcategory of the category of strict unital C*-algebras, can be regarded as a Polish category; see Section 1.2. This means that, for every separable C*-algebras AA and BB, the set Mor(M(A),M(B))\left(M(A),M(B)\right) of strict unital *-homomorphisms M(A)M(B)M(A)\rightarrow M(B) is a Polish space, and composition of morphisms is a continuous function.

Following [woronowicz_algebras_1995] we consider Mor(M(A),M(B))\mathrm{Mor}\left(M(A),M(B)\right) as endowed with the topology of pointwise strict convergence. This is the subspace topology induced by regarding, as in Lemma 2.27, Mor(M(A),M(B))\mathrm{Mor}\left(M(A),M(B)\right) as a subspace of Ball(L(A,M(B)))\mathrm{\mathrm{Ball}}\left(L\left(A,M(B)\right)\right), where L(A,M(B))L\left(A,M(B)\right) is the space of bounded linear maps from AA to M(B)M(B). (Recall that, if XX is a Banach space and 𝔜\mathfrak{Y} is a strict Banach space, then the space L(X,𝔜)L\left(X,\mathfrak{Y}\right) of bounded linear maps X𝔜X\rightarrow\mathfrak{Y} is a strict Banach space when Ball(L(X,𝔜))\mathrm{\mathrm{Ball}}\left(L\left(X,\mathfrak{Y}\right)\right) is endowed with the topology of pointwise strict convergence; see Proposition 2.15.) As Mor(M(A),M(B))\mathrm{Mor}\left(M(A),M(B)\right) is a GδG_{\delta} subset of Ball(L(A,M(B)))\mathrm{\mathrm{Ball}}\left(L\left(A,M(B)\right)\right), it is a Polish space with the induced topology. It is easy to see that this turns the category of muliplier algebras of separable C*-algebras into a Polish category.

If A,BA,B are separable C*-algebras, then the space Iso(M(A),M(B))\mathrm{Iso}\left(M(A),M(B)\right) of isomorphisms M(A)M(B)M(A)\rightarrow M(B) in the category of strict unital C*-algebras endowed with the Polish topology as in Lemma 1.2 can be identified, via the correspondence given by Lemma 2.27, with the space Iso(A,B)\mathrm{Iso}\left(A,B\right) of *-isomorphisms ABA\rightarrow B endowed with the topology of pointwise norm-convergence.

Consider now the category of locally compact second countable Hausdorff spaces, where a morphism is simply a continuous map. Given locally compact second countable Hausdorff spaces X,YX,Y, let Mor(X,Y)\mathrm{\mathrm{Mor}}\left(X,Y\right) be the set of all continuous maps XYX\rightarrow Y. This is endowed with a Polish topology called the compact-open topology, that has as subbasis of open sets the sets of the form

(K,U):={fMor(X,Y):f(K)U}\left(K,U\right):=\left\{f\in\mathrm{\mathrm{Mor}}\left(X,Y\right):f\left(K\right)\subseteq U\right\}

for a compact subset KK of XX and an open subset UU of YY. This turns the category of locally compact second countable Hausdorff spaces and continuous maps into a Polish category. We let Iso(X,Y)Mor(X,Y)\mathrm{Iso}\left(X,Y\right)\subseteq\mathrm{\mathrm{Mor}}\left(X,Y\right) be the set of homeomorphisms XYX\rightarrow Y. The Polish topology induced on Iso(X,Y)\mathrm{Iso}\left(X,Y\right) as in Lemma 1.2 was shown in [arens_topologies_1946, Theorem 5], where it is called the gg-topology, to have as subbasis of open sets the sets of then (K,YL)\left(K,Y\setminus L\right) where K,LK,L are closed sets and at least one between KK and LL is compact. For a locally compact second countable Hausdorff space XX, let X+X^{+} be its one-point compactification, obtained by adjoining a point at infinity X\infty_{X}. Each fIso(X,Y)f\in\mathrm{Iso}\left(X,Y\right) admits a unique extension to f+Iso(X+,Y+)f^{+}\in\mathrm{Iso}(X^{+},Y^{+}) that fixes the point at infinity, in the sense that f+(X)=Yf^{+}\left(\infty_{X}\right)=\infty_{Y}. By [arens_topologies_1946, Theorem 5], the assignment ff+f\mapsto f^{+} defines a homeomorphism from Iso(X,Y)\mathrm{Iso}\left(X,Y\right) onto the closed subset of Iso(X+,Y+)\mathrm{Iso}(X^{+},Y^{+}) consisting of the homeomorphisms that fix the point at infinity.

Given a locally compact second countable Hausdorff space XX, we let C0(X)C_{0}\left(X\right) be the separable C*-algebra of continuous complex-valued functions on XX vanishing at infinity. Its multiplier algebra M(C0(X))M\left(C_{0}\left(X\right)\right) is the algebra Cb(X)C_{b}\left(X\right) of bounded continuous complex-valued functions on XX. The unit ball Ball(Cb(X))\mathrm{\mathrm{Ball}}\left(C_{b}\left(X\right)\right) of Cb(X)=M(C0(X))C_{b}\left(X\right)=M\left(C_{0}\left(X\right)\right) endowed with the strict topology can be identified with the space Mor(X,Ball())\mathrm{\mathrm{Mor}}\left(X,\mathrm{\mathrm{Ball}}\left(\mathbb{C}\right)\right) of continuous functions XBall()X\rightarrow\mathrm{\mathrm{Ball}}\left(\mathbb{C}\right) endowed with the compact-open topology. Every separable commutative C*-algebra AA is isomorphic to C0(A^)C_{0}(\hat{A}), where A^\hat{A} is the locally compact second countable Hausdorff space of nonzero homomorphisms AA\rightarrow\mathbb{C} (the spectrum of AA).

A continuous map f:XYf:X\rightarrow Y induces a strict unital *-homomorphism Cb(Y)Cb(X)C_{b}\left(Y\right)\rightarrow C_{b}\left(X\right) given by φf:Cb(Y)Cb(X)\varphi_{f}:C_{b}\left(Y\right)\rightarrow C_{b}\left(X\right), aafa\mapsto a\circ f. This defines a fully faithful contravariant functor from the category of locally compact second countable Hausdorff spaces to the category of strict unital C*-algebras. In fact, the assignment Mor(X,Y)Mor(Cb(Y),Cb(X))\mathrm{\mathrm{Mor}}\left(X,Y\right)\rightarrow\mathrm{\mathrm{Mor}}\left(C_{b}\left(Y\right),C_{b}\left(X\right)\right), fφff\mapsto\varphi_{f} is a homeomorphism, where Mor(X,Y)\left(X,Y\right) is endowed as above with the compact-open topology and Mor(Cb(X),Cb(Y))\mathrm{\mathrm{Mor}}\left(C_{b}\left(X\right),C_{b}\left(Y\right)\right) is endowed with the topology of pointwise strict convergence. Thus, by Lemma 1.4, the assignment XCb(X)X\rightarrow C_{b}\left(X\right) is a contravariant topological equivalence of categories from the Polish category of locally compact second countable Hausdorff spaces to the Polish category of multiplier algebras of commutative separable C*-algebras; see Definition 1.3.

2.4. Essential commutants and Paschke dual algebras

Suppose that BB is a separable C*-algebra, and CM(B)C\subseteq M(B) is a separable C*-subalgebra. Define then the essential commutant 𝔇(C)\mathfrak{D}\left(C\right) of CC in M(B)M(B) to be the C*-algebra

{xM(B):cB,[x,c]B},\left\{x\in M(B):\forall c\in B,\left[x,c\right]\in B\right\}\text{,}

where [x,c]\left[x,c\right] is the commutator xccxxc-cx. Define the strict topology on Ball(𝔇(C))\left(\mathfrak{D}\left(C\right)\right) to be the topology generated by the seminorms

xmax{xb,bx,[x,c]}x\mapsto\max\left\{\left\|xb\right\|,\left\|bx\right\|,\left\|\left[x,c\right]\right\|\right\}

for cCc\in C and bBb\in B. If (vn)nω\left(v_{n}\right)_{n\in\omega} is a approximate unit for BB that is approximately central for CC [higson_analytic_2000, Definition 3.2.4], then (vn)nω\left(v_{n}\right)_{n\in\omega} converges strictly to 11 in Ball(𝔇(B))\left(\mathfrak{D}(B)\right).

We have that 𝔇(C)\mathfrak{D}\left(C\right) is strictly complete. Indeed, consider a strictly Cauchy sequence (xi)iω\left(x_{i}\right)_{i\in\omega} in Ball(𝔇(C))\left(\mathfrak{D}\left(C\right)\right). Then we have that (xi)iω\left(x_{i}\right)_{i\in\omega} converges to some xBall(M(B))x\in\mathrm{\mathrm{Ball}}\left(M(B)\right) in the strict topology of M(B)M(B). For every cCc\in C, the sequence ([xi,c])iω\left(\left[x_{i},c\right]\right)_{i\in\omega} is norm-Cauchy in BB, whence it norm-converges to some element of BB, which must be equal to [x,c]\left[x,c\right]. This shows that xBall(𝔇(C))x\in\mathrm{\mathrm{Ball}}\left(\mathfrak{D}\left(C\right)\right) is the strict limit of (xi)iω\left(x_{i}\right)_{i\in\omega} in Ball(𝔇(C))\mathrm{\mathrm{Ball}}\left(\mathfrak{D}\left(C\right)\right). For n1n\geq 1, we can identify Mn(𝔇(C))M_{n}\left(\mathfrak{D}\left(C\right)\right) with 𝔇(Δn(C))M(Mn(B))\mathfrak{D}\left(\Delta_{n}\left(C\right)\right)\subseteq M\left(M_{n}(B)\right), where Δn(C)Mn(B)\Delta_{n}\left(C\right)\subseteq M_{n}(B) is the image of CC under the diagonal embedding Δn:BMn(B)\Delta_{n}:B\rightarrow M_{n}(B). From the above remarks and Lemma 2.14 we thus obtain the following.

Proposition 2.29.

Let BB be a separable C*-algebra, and let CM(B)C\subseteq M(B) be a separable C*-subalgebra. Let 𝔇(C)\mathfrak{D}\left(C\right) be the corresponding essential commutant. Then 𝔇(C)\mathfrak{D}\left(C\right) is a strict C*-algebra containing BB as a strictly dense essential strict ideal where, for every n1n\geq 1, Mn(𝔇(C))M_{n}\left(\mathfrak{D}\left(C\right)\right) is identified with 𝔇(Δn(C))\mathfrak{D}\left(\Delta_{n}\left(C\right)\right), and Ball(𝔇(Δn(C)))\mathrm{\mathrm{Ball}}\left(\mathfrak{D}\left(\Delta_{n}\left(C\right)\right)\right) is endowed with the strict topology described above.

Suppose now as above that BB is a separable C*-algebra, and CM(B)C\subseteq M(B) is a separable C*-subalgebra. Let also ICI\subseteq C be a closed two-sided ideal. Define the essential annihilator

𝔇(C//I)={x𝔇(C):aI,xaB},\mathfrak{D}\left(C//I\right)=\left\{x\in\mathfrak{D}\left(C\right):\forall a\in I,xa\in B\right\}\text{,}

which is a closed two-sided ideal of 𝔇(C)\mathfrak{D}\left(C\right). The strict topology on Ball(𝔇(C//I))\mathrm{\mathrm{Ball}}\left(\mathfrak{D}\left(C//I\right)\right) is the topology generated by the seminorms

xmax{xb,bx,[x,c]}x\mapsto\mathrm{\max}\left\{\left\|xb\right\|,\left\|bx\right\|,\left\|\left[x,c\right]\right\|\right\}

for bBIb\in B\cup I and cCc\in C. A straightforward argument as above gives the following.

Proposition 2.30.

Let BB be a separable C*-algebra, let CM(B)C\subseteq M(B) be a separable C*-subalgebra, and ICI\subseteq C be a closed two-sided ideal. Let 𝔇(C)\mathfrak{D}\left(C\right) be the corresponding essential commutant, and 𝔇(C//I)\mathfrak{D}\left(C//I\right) be the essential annihilator. Then 𝔇(C//I)\mathfrak{D}\left(C//I\right) is a strict ideal of 𝔇(C)\mathfrak{D}\left(C\right), where for every n1n\geq 1, Mn(𝔇(C//I))M_{n}\left(\mathfrak{D}\left(C//I\right)\right) is identified with 𝔇(Δn(C)//Δn(I))\mathfrak{D}\left(\Delta_{n}\left(C\right)//\Delta_{n}\left(I\right)\right) and Ball(𝔇(Δn(C)//Δn(I)))\mathrm{\mathrm{Ball}}\left(\mathfrak{D}\left(\Delta_{n}\left(C\right)//\Delta_{n}\left(I\right)\right)\right) is endowed with the strict topology described above.

Example 2.31.

Suppose that AA is a separable unital C*-algebra, JJ is a closed two-sided ideal of AA, and ρ:AB(H)\rho:A\rightarrow B\left(H\right) is a nondegenerate representation of AA that is ample, in the sense that ρ(A)K(H)={0}\rho(A)\cap K\left(H\right)=\left\{0\right\}. We regard B(H)B\left(H\right) as the multiplier algebra of K(H)K\left(H\right). The Paschke dual 𝔇ρ(A)\mathfrak{D}_{\rho}(A) as defined in [higson_analytic_2000, Definition 5.1.1] is the essential commutant 𝔇(ρ(A))\mathfrak{D}\left(\rho(A)\right) of ρ(A)\rho(A) inside B(H)B\left(H\right); see also [paschke_theory_1981]. The relative dual algebra 𝔇ρ(A//J)\mathfrak{D}_{\rho}\left(A//J\right) as defined in [higson_analytic_2000, Definition 5.3.2] is the strict ideal 𝔇(ρ(A)//ρ(J))\mathfrak{D}\left(\rho(A)//\rho\left(J\right)\right) of 𝔇ρ(A)=𝔇(ρ(A))\mathfrak{D}_{\rho}(A)=\mathfrak{D}\left(\rho(A)\right).

2.5. Homotopy of projections

Suppose that 𝔄\mathfrak{A} is a strict unital C*-algebra. Recall that a strict ideal of 𝔄\mathfrak{A} is a proper norm-closed Borel two-sided ideal 𝔍\mathfrak{J} of 𝔄\mathfrak{A} that is also a strict C*-algebra and such that the inclusion map 𝔍𝔄\mathfrak{J}\rightarrow\mathfrak{A} is a strict *-homomorphism.

Definition 2.32.

A strict (unital) C*-pair is a pair (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) where 𝔄\mathfrak{A} is a strict (unital) C*-algebra and 𝔍\mathfrak{J} is a strict ideal of 𝔄\mathfrak{A}.

We regard strict unital C*-pairs as objects of a category, where a morphism from (𝔄,)\left(\mathfrak{A},\mathfrak{I}\right) to (𝔅,𝔍)\left(\mathfrak{B},\mathfrak{J}\right) is a strict unital *-homomorphism φ:𝔄𝔅\varphi:\mathfrak{A}\rightarrow\mathfrak{B} that maps \mathfrak{I} to 𝔍\mathfrak{J}.

Every strict unital C*-pair (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) determines a quotient unital C*-algebra 𝔄/𝔍\mathfrak{A/J}. If 𝔄/\mathfrak{A/I} and 𝔅/𝔍\mathfrak{B/J} are two unital C*-algebras obtained in this way, then we say that a unital *-homomorphism φ:𝔄/𝔅/𝔍\varphi:\mathfrak{A/I}\rightarrow\mathfrak{B/J} is definable if it has a Borel lift  (or a Borel representation in the terminology of [farah_automorphisms_2011, ghasemi_isomorphisms_2015]). This is a Borel function f:𝔄𝔅f:\mathfrak{A}\rightarrow\mathfrak{B} (which is not necessarily a *-homomorphism) such that φ(a+)=f(a)+𝔍\varphi\left(a+\mathfrak{I}\right)=f\left(a\right)+\mathfrak{J} for every aAa\in A. The notion of definable unital *-homomorphisms determines a category, whose objects are strict unital C*-pairs and whose morphisms are the definable unital *-homomorphism. When the strict unital C*-pair (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is considered as the object of this category, we call it a unital C*-algebra with a strict cover, and denote it by 𝔄/𝔍\mathfrak{A/J}, as we think of it as a unital C*-algebra explicitly presented as the quotient of a strict unital C*-algebra by a strict ideal. The category of unital C*-algebras with a strict cover thus has unital C*-algebras with strict cover as objects and definable unital *-homomorphisms as morphisms. The notion of a unital C*-algebra with a strict cover is the analogue in the context of C*-algebras to the notion of group with a Polish cover considered in [bergfalk_ulam_2020]; see Remark 1.17.

Notice that every strict unital *-homomorphism (𝔄,)(𝔅,𝔍)\left(\mathfrak{A},\mathfrak{I}\right)\rightarrow\left(\mathfrak{B},\mathfrak{J}\right) between strict unital C*-pairs induces a definable unital *-homomorphism 𝔄/𝔍𝔅/𝔍\mathfrak{A/J}\rightarrow\mathfrak{B/J} between the corresponding unital C*-algebras with a strict cover. This allows one to regard the category of strict unital C*-pairs as a subcategory of the category of unital C*-algebras with a strict cover. These categories have the same objects, but different morphisms.

If (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair and a,b𝔄a,b\in\mathfrak{A}, we write abmod𝔍a\equiv b\mathrm{\ \mathrm{mod}}\ \mathfrak{J} if ab𝔍a-b\in\mathfrak{J}. If aMn(𝔄)a\in M_{n}\left(\mathfrak{A}\right) and bMk(𝔄)b\in M_{k}\left(\mathfrak{A}\right), then we set

ab=[a00b]Mn+k(𝔄).a\oplus b=\begin{bmatrix}a&0\\ 0&b\end{bmatrix}\in M_{n+k}\left(\mathfrak{A}\right)\text{.}

We let 1n1_{n} be the identity element of Mn(𝔄)M_{n}\left(\mathfrak{A}\right) and 0n0_{n} be the zero element of Mn(𝔄)M_{n}\left(\mathfrak{A}\right).

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. A positive element of Ball(𝔄)\left(\mathfrak{A}\right) is a projection mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} if p2pmod𝔍p^{2}\equiv p\mathrm{\ \mathrm{mod}}\ \mathfrak{J} or, equivalently, p+𝔍p+\mathfrak{J} is a projection in 𝔄/𝔍\mathfrak{A/J}. Define the set Proj(𝔄/𝔍)Ball(𝔄)\mathrm{Proj}\left(\mathfrak{A/J}\right)\subseteq\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) to be the Borel set of projections mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} in 𝔄\mathfrak{A}. The Borel structure on Proj(𝔄/𝔍)\mathrm{Proj}\left(\mathfrak{A/J}\right) is induced by the Polish topology defined by declaring a net (pi)iI\left(p_{i}\right)_{i\in I} to converge to pp if and only if pipp_{i}\rightarrow p strictly in Ball(𝔄)\left(\mathfrak{A}\right) and pi2pip2pp_{i}^{2}-p_{i}\rightarrow p^{2}-p strictly in 22Ball(𝔍)\left(\mathfrak{J}\right). (Recall that the strict topology on Ball(𝔍)\mathrm{\mathrm{Ball}}\left(\mathfrak{J}\right) might be different from the topology induced by the strict topology on Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right).)

We also say that an element uu of Ball(𝔄)\left(\mathfrak{A}\right) is a unitary mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} if uu1mod𝔍uu^{\ast}\equiv 1\mathrm{\ \mathrm{mod}}\ \mathfrak{J} and uu1mod𝔍u^{\ast}u\equiv 1\mathrm{\ \mathrm{mod}}\ \mathfrak{J} or, equivalently, u+𝔍u+\mathfrak{J} is a unitary in 𝔄/𝔍\mathfrak{A/J}. We let U(𝔄/𝔍)U\left(\mathfrak{A/J}\right) be the Borel set of unitaries mod𝔍\mathrm{\mathrm{\mathrm{mod}}\ }\mathfrak{J} in 𝔄\mathfrak{A}. The Borel structure on U(𝔄/𝔍)U\left(\mathfrak{A/J}\right) is induced by the Polish topology defined by declaring a net (ui)iI\left(u_{i}\right)_{i\in I} to converge to uu if and only if uiuu_{i}\rightarrow u strictly in Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right), uiui1uu1u_{i}u_{i}^{\ast}-1\rightarrow uu^{\ast}-1 strictly in 2Ball(𝔍)2\mathrm{\mathrm{Ball}}\left(\mathfrak{J}\right), and uiui1uu1u_{i}^{\ast}u_{i}-1\rightarrow u^{\ast}u-1 strictly in 2Ball(𝔍)2\mathrm{\mathrm{Ball}}\left(\mathfrak{J}\right).

More generally, an element vv of Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) is called a partial unitary mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} if uuuuuu^{\ast}\equiv uu^{\ast} mod𝔍\mathrm{\ \mathrm{mod}}\ \mathfrak{J} and uuuu^{\ast} is a mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projection or, equivalently, if v+𝔍v+\mathfrak{J} is a partial unitary in 𝔄/𝔍\mathfrak{A/J} as in [rordam_introduction_2000, 8.2.12]. We let PU(𝔄/𝔍)\mathrm{PU}\left(\mathfrak{A/J}\right) be the Borel set of mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} partial unitaries in 𝔄\mathfrak{A}. In a similar fashion one can define the Borel set PI(𝔄/𝔍)\left(\mathfrak{A/J}\right) of mod𝔍\mathrm{\mathrm{\mathrm{\mathrm{mod}}\ }}\mathfrak{J} partial isometries in 𝔄\mathfrak{A}, consisting of those vBall(𝔄)v\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) such that vvvv^{\ast} and vvv^{\ast}v are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections.

In the rest of this section we record some lemmas about unitaries and projections modulo a strict ideal in a strict unital C*-algebra. The content of these lemmas can be summarized as the assertion that a homotopy between projections and unitaries in a unital C*-algebra with a strict cover is witnessed by unitary elements in the path-component of the identity of the unitary group that can be chosen in a Borel fashion. The proofs follow standard arguments from the literature on K\mathrm{K}-theory for C*-algebras; see [rordam_introduction_2000, higson_analytic_2000, blackadar_theory_1998, wegge-olsen_theory_1993].

Given elements y1,,yny_{1},\ldots,y_{n} of Ball(𝔄)\mathrm{\mathrm{\mathrm{\mathrm{Ball}}}}\left(\mathfrak{A}\right), subject to a certain relation P(y1,,yn)P\left(y_{1},\ldots,y_{n}\right), we say that an element zBall(𝔄)z\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) satisfying a relation R(y1,,yn,z)R\left(y_{1},\ldots,y_{n},z\right) can be chosen in a Borel fashion (from y1,,yny_{1},\ldots,y_{n}) if there is a Borel function (y1,,yn)z(y1,,y)\left(y_{1},\ldots,y_{n}\right)\mapsto z\left(y_{1},\ldots,y\right) that assign to each nn-tuple (y1,,yn)\left(y_{1},\ldots,y_{n}\right) in Ball(𝔄)\mathrm{\mathrm{\mathrm{\mathrm{Ball}}}}\left(\mathfrak{A}\right) satisfying PP an element z(y1,,yn)z\left(y_{1},\ldots,y_{n}\right) in Ball(𝔄)\mathrm{\mathrm{\mathrm{\mathrm{Ball}}}}\left(\mathfrak{A}\right) such that (y1,,yn,z(y1,,y))\left(y_{1},\ldots,y_{n},z\left(y_{1},\ldots,y\right)\right) satisfies RR. In other words, the set of tuples (y1,,yn,z)Ball(𝔄)n×Ball(𝔄)\left(y_{1},\ldots,y_{n},z\right)\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)^{n}\times\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) such that (y1,,yn)\left(y_{1},\ldots,y_{n}\right) satisfies PP and (y1,,yn,z)\left(y_{1},\ldots,y_{n},z\right) satisfies RR has a Borel uniformization [kechris_classical_1995, Section 18.A].

Suppose that AA is a unital C*-algebra. Let Proj(A)\mathrm{Proj}(A) be the set of projections in AA. Two projections p,qp,q in AA are:

  • Murray–von Neumann equivalent if there exists vAv\in A such that vv=pv^{\ast}v=p and vv=qvv^{\ast}=q, in which case we write pMvNqp\sim_{\mathrm{MvN}}q;

  • unitary equivalent if there exists uU(A)u\in U(A) such that uqu=pu^{\ast}qu=p;

  • homotopic if there is a norm-continuous path (pt)t[0,1]\left(p_{t}\right)_{t\in\left[0,1\right]} in Proj(Mn(A))\mathrm{Proj}\left(M_{n}(A)\right) with p0=pp_{0}=p and p1=qp_{1}=q.

Lemma 2.33.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, dd\in\mathbb{N}, and udBall(𝔄)u\in d\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) satisfies u1+𝔍1/2\left\|u-1+\mathfrak{J}\right\|\leq 1/2. Then one can choose in a Borel way a self-adjoint element yBall(𝔄)y\in\mathrm{\mathrm{\mathrm{Ball}}}\left(\mathfrak{A}\right) such that eiyumod𝔍e^{iy}\equiv u\mathrm{\ \mathrm{mod}}\ \mathfrak{J}.

Proof.

Consider u1+𝔍𝔄/𝔍u-1+\mathfrak{J}\in\mathfrak{A/J}, and observe that there exists a𝔄a\in\mathfrak{A} such that a1/2\left\|a\right\|\leq 1/2 and a+𝔍=u1+𝔍a+\mathfrak{J}=u-1+\mathfrak{J}, which can be chosen in a Borel way by strict continuity of the continuous functional calculus. Hence, setting u~:=a+1(d+1)Ball(𝔄)\tilde{u}:=a+1\in\left(d+1\right)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right), we have that u~umod𝔍\tilde{u}\equiv u\mathrm{\mathrm{\ \mathrm{mod}}\ }\mathfrak{J} and u~11/2\left\|\tilde{u}-1\right\|\leq 1/2. Thus, after replacing dd with d+1d+1 and uu with u~\tilde{u}, we can assume that u11/2\left\|u-1\right\|\leq 1/2.

Let log:D\log:D\rightarrow\mathbb{C} be an holomorphic branch of the logarithm defined on {z:|z1|<1}\left\{z\in\mathbb{C}:\left|z-1\right|<1\right\}. Considering the holomorphic functional calculus, one can define the element log(u)𝔄\mathrm{\log}(u)\in\mathfrak{A}. As

log(z)=n=0(1z)nn\mathrm{\log}(z)=\sum_{n=0}^{\infty}\frac{\left(1-z\right)^{n}}{n}

is the uniformly convergent power series expansion in {z:|z|1/2}\left\{z\in\mathbb{C}:\left|z\right|\leq 1/2\right\}, we have that

log(u)=n=0(1u)nn.\mathrm{\log}(u)=\sum_{n=0}^{\infty}\frac{\left(1-u\right)^{n}}{n}\text{.}

In particular, log(u)1\left\|\mathrm{\log}(u)\right\|\leq 1. Define

y:=log(u)+log(u)2Ball(𝔄sa).y:=\frac{\mathrm{\log}(u)+\log(u)^{\ast}}{2}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right)\text{.}

Then we have that ylog(u)mod𝔍y\equiv\log(u)\mathrm{\ \mathrm{mod}}\ \mathfrak{J} satisfies exp(iy)umod𝔍\exp\left(iy\right)\equiv u\mathrm{\ \mathrm{mod}}\ \mathfrak{J}. ∎

Corollary 2.34.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, and u,w𝔄u,w\in\mathfrak{A} are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} unitaries. Then there following assertions are equivalent:

  1. (1)

    there is a norm-continuous path from u+𝔍u+\mathfrak{J} to w+𝔍w+\mathfrak{J} in 𝔄/𝔍\mathfrak{A/J};

  2. (2)

    there exists 1\ell\geq 1 and y1,,yBall(𝔄sa)y_{1},\ldots,y_{\ell}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) such that eiy1eiyuwmod𝔍e^{iy_{1}}\cdots e^{iy_{\ell}}u\equiv w\mathrm{\ \mathrm{mod}}\ \mathfrak{J}.

Lemma 2.35.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, and p,q,xBall(𝔄)p,q,x\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) are such that p,qp,q are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections, xxpmod𝔍x^{\ast}x\equiv p\mathrm{\ \mathrm{mod}}\ \mathfrak{J}, and xxqmod𝔍xx^{\ast}\equiv q\mathrm{\ \mathrm{mod}}\ \mathfrak{J}. Then one can choose Y1,,YBall(M2(𝔄)sa)Y_{1},\ldots,Y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{2}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) in a Borel fashion from p,q,xp,q,x such that, setting U:=eiY1eiY,U:=e^{iY_{1}}\cdots e^{iY_{\ell}}\mathrm{,} one has that

U(q0)U(p0)modM2(𝔍)U^{\ast}\left(q\oplus 0\right)U\equiv\left(p\oplus 0\right)\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)

and

(q0)U(p0)x0modM2(𝔍),\left(q\oplus 0\right)U\left(p\oplus 0\right)\equiv x\oplus 0\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)\text{,}

where 1\ell\geq 1 does not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,q,xp,q,x.

Proof.

Consider the modM2(𝔍)\mathrm{\mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right) unitary

X:=[x1q1px]M2(𝔄).X:=\begin{bmatrix}x&1-q\\ 1-p&x^{\ast}\end{bmatrix}\in M_{2}\left(\mathfrak{A}\right)\text{.}

Notice that XX satisfies

X(q0d)X(p0d)modM2(𝔍)X^{\ast}\left(q\oplus 0_{d}\right)X\equiv\left(p\oplus 0_{d}\right)\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)

and

(q0d)X(p0d)x0dmodM2(𝔍).\left(q\oplus 0_{d}\right)X\left(p\oplus 0_{d}\right)\equiv x\oplus 0_{d}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)\text{.}

Consider the norm-continuous path of modM2(𝔍)\mathrm{\mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right) unitaries

Xt:=[cos(πt2)x1(1sin(πt2))q(1sin(πt2))p1cos(πt2)x]X_{t}:=\begin{bmatrix}\mathrm{cos}\left(\frac{\pi t}{2}\right)x&1-\left(1-\mathrm{sin}\left(\frac{\pi t}{2}\right)\right)q\\ \left(1-\mathrm{sin}\left(\frac{\pi t}{2}\right)\right)p-1&\mathrm{cos}\left(\frac{\pi t}{2}\right)x^{\ast}\end{bmatrix}

for t[0,1]t\in[0,1]. Notice that the modulus of continuity of (Xt)t[0,1]\left(X_{t}\right)_{t\in\left[0,1\right]} does not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and (p,q,x)\left(p,q,x\right). Fix 1\ell\geq 1 such that if t,s[0,1]t,s\in\left[0,1\right] satisfy |st|1/\left|s-t\right|\leq 1/\ell, then

XtXs1/2.\left\|X_{t}-X_{s}\right\|\leq 1/2\text{.}

Thus, for i{1,2,,}i\in\left\{1,2,\ldots,\ell\right\} we have that

Xi/X(i+1)/1/2.\left\|X_{i/\ell}-X_{(i+1)/\ell}\right\|\leq 1/2\text{.}

By Lemma 2.33 we can choose in a Borel fashion Y1Ball(𝔄sa)Y_{1}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) such that eiY1X1/modM2(𝔍)e^{iY_{1}}\equiv X_{1/\ell}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right). Thus

R1:=exp(iY1)X1/M2(𝔍).R_{1}:=\mathrm{\exp}\left(iY_{1}\right)-X_{1/\ell}\in M_{2}\left(\mathfrak{J}\right).

Consider now X2/X_{2/\ell} and the fact that

X2/X1/1/2.\left\|X_{2/\ell}-X_{1/\ell}\right\|\leq 1/2\text{.}

Thus

exp(iY1)(X2/+R1)1/2.\left\|\mathrm{\exp}\left(iY_{1}\right)-\left(X_{2/\ell}+R_{1}\right)\right\|\leq 1/2\text{.}

and

1exp(iY1)(X2/+R1)1/2.\left\|1-\mathrm{\exp}\left(-iY_{1}\right)\left(X_{2/\ell}+R_{1}\right)\right\|\leq 1/2\text{.}

Thus by Lemma 2.33 one can choose in a Borel fashion Y2Ball(𝔄sa)Y_{2}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) such that

exp(iY2)exp(iY1)(X2/+R1)exp(iY1)X2/modM2(𝔍)\exp\left(iY_{2}\right)\equiv\mathrm{\exp}\left(-iY_{1}\right)\left(X_{2/\ell}+R_{1}\right)\equiv\exp\left(-iY_{1}\right)X_{2/\ell}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)

and hence

exp(iY1)exp(iY2)X2/modM2(𝔍).\exp\left(iY_{1}\right)\exp\left(iY_{2}\right)\equiv X_{2/\ell}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)\text{.}

Proceeding recursively in this way, one can choose Y1,,YBall(𝔄sa)Y_{1},\ldots,Y_{\ell}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) in a Borel fashion such that

exp(iY1)exp(iY)XmodM2(𝔍).\mathrm{\exp}\left(iY_{1}\right)\cdots\mathrm{\exp}\left(iY_{\ell}\right)\equiv X\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)\text{.}

Then we have that, setting U:=exp(iY1)exp(iY)U:=\mathrm{\exp}\left(iY_{1}\right)\cdots\mathrm{\exp}\left(iY_{\ell}\right),

U(q0d)UX(q0d)Xp0dmodM2(𝔍)U^{\ast}\left(q\oplus 0_{d}\right)U\equiv X^{\ast}\left(q\oplus 0_{d}\right)X\equiv p\oplus 0_{d}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)

and

(q0d)U(p0d)(q0d)X(p0d)x0dmodM2(𝔍).\left(q\oplus 0_{d}\right)U\left(p\oplus 0_{d}\right)\equiv\left(q\oplus 0_{d}\right)X\left(p\oplus 0_{d}\right)\equiv x\oplus 0_{d}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right)\text{.}

This concludes the proof. ∎

Lemma 2.36.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, and p,q𝔄sap,q\in\mathfrak{A}_{\mathrm{sa}} are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections such that pq1/2\left\|p-q\right\|\leq 1/2. Then one can choose y1,,yBall(𝔄sa)y_{1},\ldots,y_{\ell}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) in a Borel fashion from p,qp,q such that, setting u:=eiy1eiynu:=e^{iy_{1}}\cdots e^{iy_{n}}, one has that uqupmod𝔍u^{\ast}qu\equiv p\mathrm{\ \mathrm{mod}}\ \mathfrak{J}, where 1\ell\geq 1 does not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q.

Proof.

As in the proof of [rordam_introduction_2000, Proposition 2.2.4], consider the norm-continuous path of mod𝔍\mathrm{\mathrm{\mathrm{mod}}\ }\mathfrak{J} projections at:=(1t)p+tqa_{t}:=\left(1-t\right)p+tq for t[0,1]t\in\left[0,1\right]. Let also K=[1/4,1/4][3/4,5/4]K=[-1/4,1/4]\cup[3/4,5/4]\subseteq\mathbb{R}, and f:Kf:K\rightarrow\mathbb{C} be the continuous function that is 0 on [1/4,1/4][-1/4,1/4] and 11 on [3/4,5/4][3/4,5/4]. Then pt:=f(at)p_{t}:=f\left(a_{t}\right) for t[0,1]t\in\left[0,1\right] is a norm-continuous path of mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections from pp to qq. Notice that the uniform continuity moduli of tatt\mapsto a_{t} and tptt\mapsto p_{t} do not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q.

Thus, there exists kωk\in\omega (that does depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q) such that, for every t,s[0,1]t,s\in\left[0,1\right] such that |ts|1/k\left|t-s\right|\leq 1/k, one has that ptps1/6\left\|p_{t}-p_{s}\right\|\leq 1/6. Thus, p0=p,p1/k,p2/k,,p1=qp_{0}=p,p_{1/k},p_{2/k},\ldots,p_{1}=q are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections (that depend in a Borel way from p,qp,q by strict continuity of the continuous functional calculus) such that p(i+1)/kpi/k1/6\left\|p_{(i+1)/k}-p_{i/k}\right\|\leq 1/6 for i{0,1,,k1}i\in\left\{0,1,\ldots,k-1\right\} and (pi(1s)+(1+i)sk)s[0,1](p_{\frac{i\left(1-s\right)+\left(1+i\right)s}{k}})_{s\in[0,1]} is a norm-continuous path from pi/kp_{i/k} to p(i+1)/kp_{(i+1)/k} (whose modulus of continuity does not depend on 𝔄\mathfrak{A} and p,q𝔄p,q\in\mathfrak{A}) satisfying

pi(1s)+(1+i)skpi/k1/6\left\|p_{\frac{i\left(1-s\right)+\left(1+i\right)s}{k}}-p_{i/k}\right\|\leq 1/6

for s[0,1]s\in\left[0,1\right].

Thus, we can assume without loss of generality that ptp1/6\left\|p_{t}-p\right\|\leq 1/6 for every t[0,1]t\in\left[0,1\right]. We now proceed as in the proof of [nest_excision_2017, Proposition 2.17]. Define

xt:=(2p1)(ptp)+1x_{t}:=\left(2p-1\right)\left(p_{t}-p\right)+1

By definition, we have that x0=1x_{0}=1. Notice that

xtppt+(p1)(pt1)mod𝔍x_{t}\equiv pp_{t}+\left(p-1\right)\left(p_{t}-1\right)\mathrm{\ \mathrm{mod}}\ \mathfrak{J}
pxtpptxtptmod𝔍px_{t}\equiv pp_{t}\equiv x_{t}p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}
ptxtxtxtpxtxtxtptmod𝔍p_{t}x_{t}^{\ast}x_{t}\equiv x_{t}^{\ast}px_{t}\equiv x_{t}^{\ast}x_{t}p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}

and

pxtxtxtptxtxtxtpmod𝔍.px_{t}x_{t}^{\ast}\equiv x_{t}p_{t}x_{t}^{\ast}\equiv x_{t}x_{t}^{\ast}p\mathrm{\ \mathrm{mod}}\ \mathfrak{J}\text{.}

This implies that

p|xt||xt|pmod𝔍p\left|x_{t}^{\ast}\right|\equiv\left|x_{t}^{\ast}\right|p\mathrm{\ \mathrm{mod}}\ \mathfrak{J}

and

pt|xt||xt|ptmod𝔍p_{t}\left|x_{t}\right|\equiv\left|x_{t}\right|p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}

for t[0,1]t\in\left[0,1\right]. We have that

xt1=(2p1)(ptp)2p1ptp1/2.\left\|x_{t}-1\right\|=\left\|\left(2p-1\right)\left(p_{t}-p\right)\right\|\leq\left\|2p-1\right\|\left\|p_{t}-p\right\|\leq 1/2\text{.}

Thus, xtx_{t} is invertible. Let xt:=ut|xt|x_{t}:=u_{t}\left|x_{t}\right| be its polar decomposition, where utu_{t} is a unitary. Then we have that pututptmod𝔍pu_{t}\equiv u_{t}p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}. Indeed,

putpxt|xt|1xtpt|xt|1xt|xt|1ptutptmod𝔍.pu_{t}\equiv px_{t}\left|x_{t}\right|^{-1}\equiv x_{t}p_{t}\left|x_{t}\right|^{-1}\equiv x_{t}\left|x_{t}\right|^{-1}p_{t}\equiv u_{t}p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}\text{.}

Thus

utputptmod𝔍u_{t}^{\ast}pu_{t}\equiv p_{t}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}

for t[0,1]t\in\left[0,1\right] and in particular u1pu1qmod𝔍u_{1}^{\ast}pu_{1}\equiv q\mathrm{\ \mathrm{mod}}\ \mathfrak{J}.

Notice that (ut)t[0,1]\left(u_{t}\right)_{t\in\left[0,1\right]} is a norm-continuous path, whose modulus of continuity does not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q. Therefore, there exists k1k\geq 1 (which do not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q) such that, whenever s,t[0,1]s,t\in\left[0,1\right] satisfy |st|1/k\left|s-t\right|\leq 1/k, we have utus1/2\left\|u_{t}-u_{s}\right\|\leq 1/2. By Lemma 2.33 one can then choose in a Borel way y1,,ykBall(𝔄sa)y_{1},\ldots,y_{k}\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}_{\mathrm{sa}}\right) such that, setting u:=exp(iy1)exp(iyk)u:=\exp\left(iy_{1}\right)\cdots\exp\left(iy_{k}\right), then uu1mod𝔍u\equiv u_{1}\mathrm{\ \mathrm{mod}}\ \mathfrak{J} and hence upuqmod𝔍u^{\ast}pu\equiv q\mathrm{\ \mathrm{mod}}\ \mathfrak{J}. This concludes the proof. ∎

Lemma 2.37.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, and p,q𝔄p,q\in\mathfrak{A} are mod𝔍\mathrm{\mathrm{\mathrm{mod}}\ }\mathfrak{J} projections that satisfy pqqp0mod𝔍pq\equiv qp\equiv 0\mathrm{\ \mathrm{mod}}\ \mathfrak{J}. Then one can choose Y1,,YBall(M2(𝔄)sa)Y_{1},\ldots,Y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{2}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) in a Borel fashion from p,qp,q such that, setting U:=eiY1eiYU:=e^{iY_{1}}\cdots e^{iY_{\ell}}, one has that U(pq)U(p+q)0modM2(𝔍)U^{\ast}\left(p\oplus q\right)U\equiv\left(p+q\right)\oplus 0\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right), where 1\ell\geq 1 does not depend on (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) and p,qp,q.

Proof.

Consider the path

rt:=[1000]p+[cos2(πt2)cos(πt2)sin(πt2)cos(πt2)sin(πt2)sin2(πt2)]qr_{t}:=\begin{bmatrix}1&0\\ 0&0\end{bmatrix}p+\begin{bmatrix}\mathrm{cos}^{2}\left(\frac{\pi t}{2}\right)&\mathrm{cos}\left(\frac{\pi t}{2}\right)\mathrm{sin}\left(\frac{\pi t}{2}\right)\\ \mathrm{cos}\left(\frac{\pi t}{2}\right)\mathrm{sin}\left(\frac{\pi t}{2}\right)&\mathrm{sin}^{2}\left(\frac{\pi t}{2}\right)\end{bmatrix}q

for t[0,1]t\in\left[0,1\right]. This is a norm-continuous path of modM2(𝔍)\mathrm{\mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right) projections in M2(𝔄)M_{2}\left(\mathfrak{A}\right) from (p+q)0\left(p+q\right)\oplus 0 to pqp\oplus q, whose modulus of continuity does not depend on 𝔄\mathfrak{A} and p,qp,q. Therefore, the conclusion follows from Lemma 2.36. ∎

Lemma 2.38.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair, and u,v𝔄u,v\in\mathfrak{A} are mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} unitaries. Then one can choose y1,,yBall(M2(𝔄)sa)y_{1},\ldots,y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{2}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) in a Borel fashion from uu and vv such that (uv)eiy1eiy(uv1)modM2(𝔍)\left(u\oplus v\right)\equiv e^{iy_{1}}\cdots e^{iy_{\ell}}\left(uv\oplus 1\right)\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right), where 1\ell\geq 1 does not depend on 𝔄,𝔍\mathfrak{A},\mathfrak{J} and u,vu,v.

Proof.

Fix a unitary path (Wt)t[0,1]\left(W_{t}\right)_{t\in\left[0,1\right]} in U(M2())U\left(M_{2}\left(\mathbb{C}\right)\right) from 11 to

[0110].\begin{bmatrix}0&1\\ 1&0\end{bmatrix}\text{.}

Fix 1\ell\geq 1 such that, for s,t[0,1]s,t\in\left[0,1\right], WsWt1/2\left\|W_{s}-W_{t}\right\|\leq 1/2. Then

ut:=(u1)Wt(v1)Wt(uv1)u_{t}:=\left(u\oplus 1\right)W_{t}^{\ast}\left(v\oplus 1\right)W_{t}\left(uv\oplus 1\right)^{\ast}

is a path of modM2(𝔍)\mathrm{\mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right) unitaries from 11 to (uv)(uv1)\left(u\oplus v\right)\left(uv\oplus 1\right)^{\ast} to 11. Then, as in the proof of Lemma 2.35, using Lemma 2.33 one can recursively choose in a Borel fashion y1,,yBall(M2(𝔄)sa)y_{1},\ldots,y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{2}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) such that eiykuk/u(k+1)/modM2(𝔍)e^{iy_{k}}\equiv u_{k/\ell}^{\ast}u_{(k+1)/\ell}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right) for k{0,1,,1}k\in\left\{0,1,\ldots,\ell-1\right\} and hence (uv)(uv1)eiy1eiymodM2(𝔍)\left(u\oplus v\right)\left(uv\oplus 1\right)^{\ast}\equiv e^{iy_{1}}\cdots e^{iy_{\ell}}\mathrm{\ \mathrm{mod}}\ M_{2}\left(\mathfrak{J}\right). ∎

2.6. The Definable Arveson Extension Theorem

In the rest of this section, we present definable versions of some fundamental results in operator algebras, to be used in the development of definable K\mathrm{K}-homology. Suppose that HH is a separable Hilbert space. We regard B(H)B\left(H\right) as the multiplier algebra of the C*-algebra K(H)K\left(H\right) of compact operators on HH. The corresponding strict topology on Ball(B(H))\mathrm{\mathrm{\mathrm{\mathrm{Ball}}}}\left(B\left(H\right)\right) is the strong-* topology. Consistently, we consider B(H)B\left(H\right) as a standard Borel space with respect to the induced standard Borel structure. If ZZ is a separable Banach space, we consider L(Z,B(H))L\left(Z,B\left(H\right)\right) as a strict Banach space, where Ball(L(Z,B(H)))\mathrm{\mathrm{Ball}}\left(L\left(Z,B\left(H\right)\right)\right) is endowed with the topology of pointwise strong-* convergence. We denote by U(H)U\left(H\right) the unitary group of B(H)B\left(H\right), which is a Polish group when endowed with the strong-* topology.

Suppose that AA is a separable unital C*-algebra, and XAX\subseteq A is an operator system [paulsen_completely_2002, Chapter 2]. Let HH be a separable Hilbert space. Arveson’s Extension Theorem asserts that every contractive completely positive (ccp) map ϕ:XB(H)\phi:X\rightarrow B\left(H\right) [brown_algebras_2008, Section 1.5] admits a contractive completely positive extension ϕ^:AB(H)\hat{\phi}:A\rightarrow B\left(H\right) [paulsen_completely_2002, Theorem 7.5]. We observe now that ϕ^\hat{\phi} can be chosen in a Borel way from ϕ\phi. Notice that the space CCP(X,B(H))\mathrm{CCP}\left(X,B\left(H\right)\right) of contractive completely positive maps is closed (hence, compact) in Ball(L(X,B(H)))\mathrm{\mathrm{Ball}}\left(L\left(X,B\left(H\right)\right)\right) endowed with the topology of pointwise weak* convergence.

Lemma 2.39.

Suppose that AA is a separable unital C*-algebra, and XAX\subseteq A is an operator system. Let HH be a separable Hilbert space. Then there exists a Borel function CCP(X,B(H))CCP(A,B(H))\mathrm{CCP}\left(X,B\left(H\right)\right)\rightarrow\mathrm{CCP}\left(A,B\left(H\right)\right), ϕϕ^\phi\mapsto\hat{\phi} such that ϕ^\hat{\phi} is an extension of ϕ\phi.

Towards the proof of Lemma 2.39, we recall the following particular case of the selection theorem for relations with compact sections [kechris_classical_1995, Theorem 28.8].

Lemma 2.40.

Suppose that X,YX,Y are compact metrizable spaces, and AX×YA\subseteq X\times Y is a Borel subset such that, for every xXx\in X, the vertical section

Ax={yY:(x,y)A}A_{x}=\left\{y\in Y:\left(x,y\right)\in A\right\}

is a closed nonempty set. Then there exists a Borel function f:XYf:X\rightarrow Y such that (x,f(x))A\left(x,f(x)\right)\in A for every xXx\in X.

Using this selection theorem, Lemma 2.39 follows immediately from the Arveson Extension Theorem.

Proof of Lemma 2.39.

We consider CCP(X,B(H))\mathrm{CCP}\left(X,B\left(H\right)\right) as a compact metrizable space, endowed with the topology of pointwise weak* convergence. Consider the Borel set ACCP(X,B(H))×CCP(A,B(H))A\subseteq\mathrm{CCP}\left(X,B\left(H\right)\right)\times\mathrm{CCP}\left(A,B\left(H\right)\right) of pairs (ϕ,ψ)\left(\phi,\psi\right) such that ψ|X=ϕ\psi|_{X}=\phi. Then by the Arveson Extension Theorem, the vertical sections of AA are nonempty, and clearly closed. Thus, by Lemma 2.40 there exists a Borel function f:CCP(X,B(H))CCP(A,B(H))f:\mathrm{CCP}\left(X,B\left(H\right)\right)\rightarrow\mathrm{CCP}\left(A,B\left(H\right)\right) such that f(ϕ)|X=ϕf\left(\phi\right)|_{X}=\phi for every ϕCCP(X,B(H))\phi\in\mathrm{CCP}\left(X,B\left(H\right)\right). ∎

2.7. The Definable Stinespring Dilation Theorem

Suppose that AA is a separable unital C*-algebra, and HH is a separable Hilbert space. Stinespring’s Dilation Theorem asserts that, for every contractive completely positive map ϕ:AB(H)\phi:A\rightarrow B\left(H\right), there exists a linear map Vϕ:HHV_{\phi}:H\rightarrow H with Vϕ2=ϕ\left\|V_{\phi}\right\|^{2}=\left\|\phi\right\| and a nondegenerate representation πϕ\pi_{\phi} of AA on HH such that ϕ(a)=Vϕπϕ(a)Vϕ\phi\left(a\right)=V_{\phi}^{\ast}\pi_{\phi}\left(a\right)V_{\phi} for every aAa\in A. Notice that the set Rep(A,H)\mathrm{Rep}\left(A,H\right) of nondegenerate representations of AA on HH is a GδG_{\delta} subset of Ball(L(A,B(H)))\mathrm{\mathrm{Ball}}\left(L\left(A,B\left(H\right)\right)\right), whence Polish with the subspace topology, where Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) is endowed with the strong-* topology. It follows from the proof of the Stinespring Dilation Theorem, where VV and π\pi are explicitly defined in terms of ϕ\phi, that they can be chosen in a Borel way from ϕ\phi; see [blackadar_operator_2006, Theorem II.6.9.7]

Lemma 2.41.

Suppose that AA is a separable unital C*-algebra, and HH is a separable Hilbert space. Then there exists a Borel function CCP(A,B(H))Ball(B(H))×Rep(A,B(H))\mathrm{CCP}\left(A,B\left(H\right)\right)\rightarrow\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right)\times\mathrm{Rep}\left(A,B\left(H\right)\right), ϕ(Vϕ,πϕ)\phi\mapsto\left(V_{\phi},\pi_{\phi}\right) such that ϕ(a)=Vϕπϕ(a)Vϕ\phi\left(a\right)=V_{\phi}^{\ast}\pi_{\phi}\left(a\right)V_{\phi} for aAa\in A and Vϕ2=ϕ\left\|V_{\phi}\right\|^{2}=\left\|\phi\right\| for every contractive completely positive map ϕ:AB(H)\phi:A\rightarrow B\left(H\right).

2.8. The Definable Voiculescu Theorem

Suppose that AA is a separable unital C*-algebra, and ρ,ρ:AB(H)\rho,\rho^{\prime}:A\rightarrow B\left(H\right) are two maps. If UU(H)U\in U\left(H\right) is a unitary operator, write ρUρ\rho^{\prime}\thickapprox_{U}\rho if ρ(a)Uρ(a)UmodK(H)\rho^{\prime}\left(a\right)\equiv U^{\ast}\rho\left(a\right)U\mathrm{\ \mathrm{mod}}\ K\left(H\right) for every aAa\in A. If V:HHV:H\rightarrow H is an isometry, write ρVρ\rho^{\prime}\lesssim_{V}\rho if ρ(a)Vρ(a)VmodK(H)\rho^{\prime}\left(a\right)\equiv V^{\ast}\rho\left(a\right)V\mathrm{\ \mathrm{mod}}\ K\left(H\right) for every aAa\in A. A nondegenerate representation ρ\rho of AA on B(H)B\left(H\right) is ample if, for every aB(H)a\in B\left(H\right), ρ(a)K(H)a=0\rho\left(a\right)\in K\left(H\right)\Rightarrow a=0. Notice that the set ARep(A,H)\left(A,H\right) of ample representations of AA on B(H)B\left(H\right) is a GδG_{\delta} subset of Ball(L(A,B(H)))\left(L\left(A,B\left(H\right)\right)\right). Similarly, the set Iso(H)\mathrm{Iso}\left(H\right) of isometries HHH\rightarrow H is a GδG_{\delta} subset of Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right). A formulation of Voiculescu’s Theorem asserts that if ρ:AB(H)\rho:A\rightarrow B\left(H\right) is an ample representation, and σ:AB(H)\sigma:A\rightarrow B\left(H\right) is a unital completely positive (ucp) map, then there exists an isometry V:HHV:H\rightarrow H such that σVρ\sigma\lesssim_{V}\rho; see [higson_analytic_2000, Theorem 3.4.3, Theorem 3.4.6, Theorem 3.4.7]. We will observe that one can select VV in a Borel fashion from ρ\rho and σ\sigma.

Lemma 2.42.

Let AA be a separable unital C*-algebra, and HH a separable Hilbert space. There exists a Borel function UCP(A,B(H))×ARep(A,H)Iso(H)\mathrm{\mathrm{UCP}}\left(A,B\left(H\right)\right)\times\mathrm{ARep}\left(A,H\right)\rightarrow\mathrm{Iso}\left(H\right), (σ,ρ)Vσ,ρ\left(\sigma,\rho\right)\mapsto V_{\sigma,\rho} such that σVσ,ρρ\sigma\lesssim_{V_{\sigma,\rho}}\rho.

Towards obtaining a proof of Lemma 2.42, we argue as in the proof of Voiculescu’s theorem as expounded in [higson_analytic_2000, Chapter 3]. First, one considers the case of ucp maps AB(H)A\rightarrow B\left(H\right) where HH is finite-dimensional. The following can be seen a definable version of [higson_analytic_2000, Proposition 3.6.7]. Notice that the set Proj(H)\mathrm{Proj}\left(H\right) of orthogonal projections HHH\rightarrow H is closed subset of Ball(B(H))\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right). Let Projfd(H)\mathrm{Proj}_{\mathrm{fd}}\left(H\right) be the Borel subset of finite-dimensional projections. The following lemma is a consequence of [higson_analytic_2000, Proposition 3.6.7] itself and the Luzin–Novikov Uniformization Theorem for Borel relations with countable sections [kechris_classical_1995, Theorem 18.10].

Lemma 2.43.

Fix a finite-dimensional subspace H0H_{0} of HH, and regard B(H0)B\left(H_{0}\right) as a C*-subalgebra of B(H)B\left(H\right). For every finite subset FF of AA and ε>0\varepsilon>0, there exists a Borel map UCP(A,B(H0))×ARep(A,H)×Projfd(H)Ball(H)\mathrm{\mathrm{UCP}}\left(A,B\left(H_{0}\right)\right)\times\mathrm{ARep}\left(A,H\right)\times\mathrm{Proj}_{\mathrm{fd}}\left(H\right)\rightarrow\mathrm{\mathrm{Ball}}\left(H\right), (σ,ρ,P)V\left(\sigma,\rho,P\right)\mapsto V such that Ran(V)\mathrm{Ran}(V) is orthogonal to P(H)P\left(H\right) and σ(a)Vρ(a)V<ε\left\|\sigma\left(a\right)-V^{\ast}\rho\left(a\right)V\right\|<\varepsilon for aFa\in F.

One then uses Lemma 2.43 to establish Lemma 2.42 in the case of block-diagonal maps. Recall that σ\sigma is block-diagonal with respect to (Pn)nω\left(P_{n}\right)_{n\in\omega} if (Pn)nω\left(P_{n}\right)_{n\in\omega} is a sequence of pairwise orthogonal finite-rank projections PnB(H)P_{n}\in B\left(H\right) such that nPn=I\sum_{n}P_{n}=I and σ(a)=nPnσ(a)Pn\sigma\left(a\right)=\sum_{n}P_{n}\sigma\left(a\right)P_{n} for every aAa\in A (where the convergence is in the strong-* topology). Consider the set BlockUCP(A,B(H))\mathrm{BlockUCP}\left(A,B\left(H\right)\right) of pairs (σ,(Pn)nω)UCP(A,B(H))×Projfd(H)ω\left(\sigma,\left(P_{n}\right)_{n\in\omega}\right)\in\mathrm{\mathrm{UCP}}\left(A,B\left(H\right)\right)\times\mathrm{Proj}_{\mathrm{fd}}\left(H\right)^{\omega} such that σ\sigma is block-diagonal with respect to (Pn)nω\left(P_{n}\right)_{n\in\omega}. The proof of [higson_analytic_2000, Lemma 3.5.2] shows the following.

Lemma 2.44.

There exists a Borel function BlockUCP(A,B(H))×ARep(A,H)Iso(H)\mathrm{BlockUCP}\left(A,B\left(H\right)\right)\times\mathrm{ARep}\left(A,H\right)\rightarrow\mathrm{Iso}\left(H\right),(σ,(Pn)nω,ρ)V\left(\sigma,\left(P_{n}\right)_{n\in\omega},\rho\right)\mapsto V such that σVρ\sigma\lesssim_{V}\rho.

Finally, one shows that the general case of Voiculescu’s theorem can be reduced to the block-diagonal case, as in [higson_analytic_2000, Theorem 3.5.5].

Lemma 2.45.

There exists a Borel function UCP(A,B(H))BlockUCP(A,B(H))×Iso(H)\mathrm{\mathrm{UCP}}\left(A,B\left(H\right)\right)\rightarrow\mathrm{BlockUCP}\left(A,B\left(H\right)\right)\times\mathrm{Iso}\left(H\right), σ(σ,(Pn)nω,V)\sigma\mapsto\left(\sigma^{\prime},\left(P_{n}\right)_{n\in\omega},V^{\prime}\right) such that σVσ\sigma\lesssim_{V^{\prime}}\sigma^{\prime}.

Lemma 2.42 is then obtained by combining Lemma 2.44 and Lemma 2.45.

As a consequence of the definable Voiculescu Theorem, one obtains the following; see [higson_analytic_2000, Theorem 3.4.6].

Lemma 2.46.

Let AA be a separable unital C*-algebra, and HH a separable Hilbert space. There exist:

  • a Borel map Rep(A,H)×ARep(A,H)U(H)\mathrm{Rep}\left(A,H\right)\times\mathrm{ARep}\left(A,H\right)\rightarrow U\left(H\right), (ρ,ρ)Uρ,ρ\left(\rho^{\prime},\rho\right)\mapsto U_{\rho^{\prime},\rho} such that ρρUρ,ρρ\rho^{\prime}\oplus\rho\thickapprox_{U_{\rho^{\prime},\rho}}\rho;

  • a Borel map ARep(A,H)×ARep(A,H)U(H)\mathrm{ARep}\left(A,H\right)\times\mathrm{ARep}\left(A,H\right)\rightarrow U\left(H\right), (ρ,ρ)Wρ,ρ\left(\rho^{\prime},\rho\right)\mapsto W_{\rho^{\prime},\rho} such that ρWρ,ρρ\rho\thickapprox_{W_{\rho^{\prime},\rho}}\rho^{\prime}.

2.9. Spectrum

Suppose now that 𝔄\mathfrak{A} is a strict unital C*-algebra, and JJ is a norm-separable closed two-sided ideal of 𝔄\mathfrak{A}. One can consider the quotient C*-algebra 𝔄/J\mathfrak{A}/J and, for a𝔄a\in\mathfrak{A}, the spectrum σ𝔄/J(a)\sigma_{\mathfrak{A}/J}\left(a\right) of a+Ja+J in 𝔄/J\mathfrak{A}/J. We also let the resolvent ρ𝔄/J(a)\rho_{\mathfrak{A}/J}\left(a\right) be the complement in \mathbb{C} of σ𝔄/J(a)\sigma_{\mathfrak{A}/J}\left(a\right). The following lemma is analogous to [ando_weyl_2015, Theorem 3.16].

Lemma 2.47.

Suppose that 𝔄\mathfrak{A} is a strict C*-algebra, and JJ a norm-separable closed two-sided ideal of 𝔄\mathfrak{A}. Suppose that every invertible self-adjoint element of 𝔄/J\mathfrak{A}/J lifts to an invertible self-adjoint element of 𝔄\mathfrak{A}. If a𝔄saa\in\mathfrak{A}_{\mathrm{sa}}, and J0J_{0} is a countable dense subset of J𝔄saJ\cap\mathfrak{A}_{\mathrm{sa}}, then

σ𝔄/J(a)=dJ0σ(a+d).\sigma_{\mathfrak{A}/J}\left(a\right)=\bigcap_{d\in J_{0}}\sigma\left(a+d\right)\text{.}
Proof.

It suffices to prove that ρ𝔄/J(a)\rho_{\mathfrak{A}/J}\left(a\right) is the union of ρ(a+d)\rho\left(a+d\right) for dJ0d\in J_{0}. Clearly, ρ(a+d)ρ𝔄/J(a)\rho\left(a+d\right)\subseteq\rho_{\mathfrak{A}/J}\left(a\right) for every dJ0d\in J_{0}, so it suffices to prove the other inclusion. Suppose that λρ𝔄/J(a)\lambda\in\rho_{\mathfrak{A}/J}\left(a\right)\cap\mathbb{R}. We want to show that λρ(a+d)\lambda\in\rho\left(a+d\right) for some dJ0d\in J_{0}. After replacing aa with aλa-\lambda, it suffices to consider the case when λ=0\lambda=0. In this case, a+Ja+J is invertible in 𝔄/J\mathfrak{A}/J. Therefore, by assumption there exists dJ𝔄sad\in J\cap\mathfrak{A}_{\mathrm{sa}} such that a+da+d is invertible in 𝔄\mathfrak{A}. Since the set of invertible elements of 𝔄\mathfrak{A} is norm-open, there exists d0J0d_{0}\in J_{0} such that a+d0a+d_{0} is invertible in 𝔄\mathfrak{A}, and hence 0ρ(a+d0)0\in\rho\left(a+d_{0}\right). ∎

Lemma 2.48.

Suppose that 𝔄\mathfrak{A} is a strict unital C*-algebra, and JJ a norm-separable closed two-sided ideal of 𝔄\mathfrak{A}. Suppose that every invertible self-adjoint element of 𝔄/J\mathfrak{A}/J lifts to an invertible self-adjoint element of 𝔄\mathfrak{A}. Then the function 𝔄saClosed()\mathfrak{A}_{\mathrm{sa}}\rightarrow\mathrm{Closed}\left(\mathbb{R}\right), aσ𝔄/J(a)a\mapsto\sigma_{\mathfrak{A}/J}\left(a\right) is Borel.

Proof.

Fix a countable norm-dense subset J0J_{0} of J𝔄saJ\cap\mathfrak{A}_{\mathrm{sa}}. Then by the previous lemma we have that, for a𝔄saa\in\mathfrak{A}_{\mathrm{sa}},

σ𝔄/J(a)=dJ0σ(a+d).\sigma_{\mathfrak{A}/J}\left(a\right)=\bigcap_{d\in J_{0}}\sigma\left(a+d\right)\text{.}

As the function Closed()ωClosed()\mathrm{Closed}\left(\mathbb{R}\right)^{\omega}\rightarrow\mathrm{Closed}\left(\mathbb{R}\right), (Fn)nωFn\left(F_{n}\right)\mapsto\bigcap_{n\in\omega}F_{n} is Borel, this concludes the proof. ∎

Lemma 2.49.

The function B(H)saK()B\left(H\right)_{\mathrm{sa}}\rightarrow K\left(\mathbb{R}\right), Tσess(T)=σQ(H)(T)T\mapsto\sigma_{\mathrm{ess}}\left(T\right)=\sigma_{Q\left(H\right)}\left(T\right) is Borel.

Proof.

An operator TB(H)T\in B\left(H\right) induces an invertible element of Q(H)Q\left(H\right) if and only if it is Fredholm. If TT is Fredholm and self-adjoint, then it has index 0, and 0 is an isolated point of the spectrum of TT that is an eigenvalue with finite multiplicity. Thus, if PP is the finite-rank projection onto the eigenspace of 0 for TT, then we have that T+PT+P is invertible and self-adjoint and induces the same element of Q(H)Q\left(H\right) as TT. This shows that every invertible self-adjoint element of Q(H)Q\left(H\right) lifts to an invertible self-adjoint element of B(H)B\left(H\right). Therefore, the conclusion follows from Proposition 2.48. ∎

Suppose that TBall(B(H))T\in\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right) is a modK(H)\mathrm{\mathrm{\ \mathrm{mod}}\ }K\left(H\right) projective. Recall that this means that TT is a positive operator satisfying T2TmodK(H)T^{2}\equiv T\mathrm{\ \mathrm{mod}}\ K\left(H\right). Then it is well-known that there exists a projection PTmodK(H)P\equiv T\mathrm{\ \mathrm{mod}}\ K\left(H\right). We observe that one can choose such a PP in a Borel fashion from TT; see [andruchow_note_2020, Lemma 3.1].

Lemma 2.50.

Consider the Borel set Proj(B(H)/K(H))\mathrm{Proj}\left(B\left(H\right)/K\left(H\right)\right) of modK(H)\mathrm{\mathrm{mod}}\ K\left(H\right) projections in B(H)B\left(H\right). Then there is a Borel function Proj(B(H)/K(H))Proj(B(H))\mathrm{Proj}\left(B\left(H\right)/K\left(H\right)\right)\rightarrow\mathrm{\mathrm{Proj}}\left(B\left(H\right)\right), TPTT\mapsto P_{T} such that TPTmodK(H)T\equiv P_{T}\mathrm{\ \mathrm{mod}}\ K\left(H\right) for every TProj(B(H)/K(H))T\in\mathrm{Proj}\left(B\left(H\right)/K\left(H\right)\right).

Proof.

Suppose that TProj(B(H)/K(H))T\in\mathrm{Proj}\left(B\left(H\right)/K\left(H\right)\right). Observe σess(T){0,1}\sigma_{\mathrm{ess}}\left(T\right)\subseteq\left\{0,1\right\}. In particular, σess(T)\sigma_{\mathrm{ess}}\left(T\right) is countable, with only accumulation points 0 and 11. From Lemma 2.49, the maps Tσess(T)T\mapsto\sigma_{\mathrm{ess}}\left(T\right) and Tσ(T)T\mapsto\sigma\left(T\right) are Borel. If σess(T)={0}\sigma_{\mathrm{ess}}\left(T\right)=\left\{0\right\} then one can set PT=0P_{T}=0. If σess(T)={1}\sigma_{\mathrm{ess}}\left(T\right)=\left\{1\right\}, one can set PT=1P_{T}=1.

Let us consider the case when {0,1}=σess(T)\left\{0,1\right\}=\sigma_{\mathrm{ess}}(T). By [kechris_classical_1995, Theorem 12.13] there exists a Borel map Proj(B(H)/K(H))[0,1]ω\mathrm{Proj}\left(B\left(H\right)/K\left(H\right)\right)\rightarrow\left[0,1\right]^{\omega}, T(tn)T\mapsto\left(t_{n}\right) such that (tn)\left(t_{n}\right) is an increasing enumeration of σ(T)\sigma\left(T\right). One can then choose in a Borel way n0ωn_{0}\in\omega such that tn0<tn0+1t_{n_{0}}<t_{n_{0}+1} and then a continuous function f:[0,1][0,1]f:\left[0,1\right]\rightarrow\left[0,1\right] such that

f(ti)={0if in0,1if in0+1.f\left(t_{i}\right)=\left\{\begin{array}[]{ll}0&\text{if }i\leq n_{0}\text{,}\\ 1&\text{if }i\geq n_{0}+1\text{.}\end{array}\right.

One can then set PT=f(T)P_{T}=f\left(T\right). ∎

2.10. Polar decompositions

We now observe that the polar decomposition of an operator is given by a Borel function. We will use the following version of the selection theorem for relations with compact sections from [kechris_classical_1995, Theorem 28.8].

Lemma 2.51.

Suppose that XX is a standard Borel space, YY is a compact metrizable space, and AX×YA\subseteq X\times Y is a Borel subset such that, for every xXx\in X, the vertical section

Ax={yY:(x,y)A}A_{x}=\left\{y\in Y:\left(x,y\right)\in A\right\}

is a closed nonempty set. Then the assignment XClosed(Y)X\rightarrow\mathrm{Closed}\left(Y\right), xAxx\mapsto A_{x}, is Borel, where Closed(Y)\mathrm{Closed}\left(Y\right) is endowed with the Effros Borel structure.

As an application, we obtain the following. Let HH be a separable Hilbert space. We consider the unit ball Ball(H)\mathrm{\mathrm{Ball}}\left(H\right) of HH as a compact metrizable space endowed with the weak topology. We also consider Closed(Ball(H))\mathrm{Closed}\left(\mathrm{\mathrm{Ball}}\left(H\right)\right) as a standard Borel space, endowed with the Effros Borel structure.

Lemma 2.52.

The function B(H)Closed(Ball(H))B\left(H\right)\rightarrow\mathrm{Closed}\left(\mathrm{\mathrm{Ball}}\left(H\right)\right), TKer(T)Ball(H)T\mapsto\mathrm{\mathrm{Ker}}\left(T\right)\cap\mathrm{Ball}\left(H\right), is Borel.

Proof.

By Lemma 2.51, it suffices to show that the set

A={(T,x)B(H)×Ball(H):Tx=0}A=\left\{\left(T,x\right)\in B\left(H\right)\times\mathrm{\mathrm{Ball}}\left(H\right):Tx=0\right\}

is Borel. Fix a countable norm-dense subset {xn:nω}\left\{x_{n}:n\in\omega\right\} of Ball(H)\mathrm{\mathrm{Ball}}\left(H\right). Then we have that, if (T,x)Ball(B(H))×Ball(H)\left(T,x\right)\in\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right)\times\mathrm{\mathrm{Ball}}\left(H\right), then (T,x)A\left(T,x\right)\in A if and only if kω\forall k\in\omega nω\exists n\in\omega such that xxn<2k\left\|x-x_{n}\right\|<2^{-k} and Txk<2k\left\|Tx_{k}\right\|<2^{-k}. Since the norm on Ball(H)\mathrm{\mathrm{Ball}}\left(H\right) is weakly lower-semicontinuous, this shows that AA is Borel. ∎

Recall that, for an operator TB(H)T\in B\left(H\right), one sets |T|:=(TT)1/2\left|T\right|:=\left(T^{\ast}T\right)^{1/2}. By strong-* continuity on bounded sets of continuous functional calculus, the function T|T|T\mapsto\left|T\right| is Borel. Furthermore, there exists a unique partial isometry UU with Ker(U)=Ker(T)\mathrm{\mathrm{Ker}}\left(U\right)=\mathrm{\mathrm{\mathrm{Ker}}}\left(T\right) such that T=U|T|T=U\left|T\right| [pedersen_analysis_1989, Theorem 3.2.17]. The decomposition T=U|T|T=U\left|T\right| is then called the polar decomposition of TT.

Lemma 2.53.

The function B(H)B(H)B\left(H\right)\rightarrow B\left(H\right), TUT\mapsto U that assigns to an operator the partial isometry UU in the polar decomposition of TT is Borel.

Proof.

It suffices to notice that is graph, which is the set of pairs (T,U)\left(T,U\right) such that UU is a partial isometry with Ker(U)=Ker(T)\mathrm{\mathrm{Ker}}\left(U\right)=\mathrm{\mathrm{\mathrm{Ker}}}\left(T\right) and T=U|T|T=U\left|T\right|, is Borel by Lemma 2.52. ∎

Consider the Borel set U(B(H)/K(H))U\left(B\left(H\right)/K\left(H\right)\right) of modK(H)\mathrm{\mathrm{mod}}\ K\left(H\right) unitaries in B(H)B\left(H\right). Thus, TU(B(H)/K(H))T\in U\left(B\left(H\right)/K\left(H\right)\right) if and only if TTTTImodK(H)T^{\ast}T\equiv TT^{\ast}\equiv I\mathrm{\ \mathrm{mod}}\ K\left(H\right). If UU is the partial isometry in the polar decomposition of TT, then UTmodK(H)U\equiv T\mathrm{\ \mathrm{mod}}\ K\left(H\right) and UU is an essential unitary. In fact, one can easily define (in a Borel fashion from TT) an isometry or co-isometry VV such that TVmodK(H)T\equiv V\mathrm{\ \mathrm{mod}}\ K\left(H\right). One has that TT is in particular a Fredholm operator. Its index is defined by

index(T)=rank(1VV)rank(1VV).\mathrm{index}\left(T\right)=\mathrm{\mathrm{rank}}\left(1-V^{\ast}V\right)-\mathrm{\mathrm{rank}}\left(1-VV^{\ast}\right)\text{.}

Thus, index(T)\mathrm{index}\left(T\right) is a Borel function of TU(B(H)/K(H))T\in U\left(B\left(H\right)/K\left(H\right)\right).

More generally, consider the Borel set of pairs (P,T)Ball(B(H))2\left(P,T\right)\in\mathrm{\mathrm{Ball}}\left(B\left(H\right)\right)^{2} such that PP is a projection, PT=TP=TPT=TP=T and TTTTPTT^{\ast}\equiv T^{\ast}T\equiv P. If VV is the partial isometry in the polar decomposition of TT, then VTmodK(H)V\equiv T\mathrm{\ \mathrm{mod}}\ K\left(H\right) and the index of PTPPTP regarded as a Fredholm operator on PHPH is given by the Borel function

index(PTP)=rank(PVV)rank(PVV).\mathrm{index}\left(PTP\right)=\mathrm{\mathrm{rank}}\left(P-V^{\ast}V\right)-\mathrm{\mathrm{rank}}\left(P-VV^{\ast}\right)\text{.}

3. K\mathrm{K}-theory of unital C*-algebras with a strict cover

In this section we explain how the K0\mathrm{K}_{0} and K1\mathrm{K}_{1} groups of a unital C*-algebra with a strict cover can be regarded as semidefinable groups. We also recall the definition of the index map and the exponential map between the K0\mathrm{K}_{0} and K1\mathrm{K}_{1} groups, and observe that they are definable homomorphisms. Finally, we consider the six-term exact sequence associated with a strict unital C*-pair, and observe that the connective maps are all definable group homomorphisms.

3.1. K0\mathrm{K}_{0}-group

Suppose that 𝔄/𝔍\mathfrak{A/J} is a unital C*-algebra with a strict cover. Recall that Proj(𝔄/𝔍)\mathrm{Proj}\left(\mathfrak{A/J}\right) denotes the Polish space of mod𝔍\mathrm{\mathrm{\mathrm{mod}}\ }\mathfrak{J} projections in 𝔄\mathfrak{A}. Similarly, for n1n\geq 1 we have that Proj(Mn(𝔄/𝔍)):=Proj(Mn(𝔄)/Mn(𝔍))\mathrm{Proj}\left(M_{n}\left(\mathfrak{A/J}\right)\right):=\mathrm{Proj}\left(M_{n}\left(\mathfrak{A}\right)/M_{n}\left(\mathfrak{J}\right)\right) is a Polish space. We say that an element of Proj(Mn(𝔄/𝔍))\mathrm{Proj}\left(M_{n}\left(\mathfrak{A/J}\right)\right) for some n1n\geq 1 is a mod𝔍\mathrm{\mathrm{\mathrm{mod}}\ }\mathfrak{J} projection over 𝔄\mathfrak{A}. We define Z0(𝔄/𝔍){}_{0}\left(\mathfrak{A/J}\right) be the set of pairs of mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections over 𝔄\mathfrak{A}, which is the disjoint union of Z0(n)(𝔄/𝔍):=Proj(Mn(𝔄/𝔍))×Proj(Mn(𝔄/𝔍)){}_{0}^{\left(n\right)}\left(\mathfrak{A/J}\right):=\mathrm{Proj}\left(M_{n}\left(\mathfrak{A/J}\right)\right)\times\mathrm{Proj}\left(M_{n}\left(\mathfrak{A/J}\right)\right) for n1n\geq 1 endowed with the induced standard Borel structure. Two mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} projections p,qp,q in 𝔄\mathfrak{A} are Murray–von Neumann equivalent (respectively, unitary equivalent, and homotopic) mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J} if and only if p+𝔍p+\mathfrak{J} and q+𝔍q+\mathfrak{J} are Murray–von Neumann equivalent (respectively, unitary equivalent, and homotopic) in 𝔄/𝔍\mathfrak{A/J}.

The K0\mathrm{K}_{0}-group K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A/J}\right) of 𝔄/𝔍\mathfrak{A/J}—see [higson_analytic_2000, Chapter 4]—is defined as a quotient of Z0(𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A/J}\right) by an equivalence relation B0(𝔄/𝔍)\mathrm{B}_{0}\left(\mathfrak{A/J}\right), defined as follows. For (p,p),(q,q)Z0(𝔄/𝔍)\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\in\mathrm{Z}_{0}\left(\mathfrak{A/J}\right), (p,p)B0(𝔄/𝔍)(q,q)\left(p,p^{\prime}\right)\mathrm{B}_{0}\left(\mathfrak{A/J}\right)\left(q,q^{\prime}\right) if and only if there exist m,nωm,n\in\omega and rProj(Mm(𝔄/𝔍))r\in\mathrm{Proj}\left(M_{m}\left(\mathfrak{A/J}\right)\right) such that pqr0np\oplus q^{\prime}\oplus r\oplus 0_{n} and qpr0nq\oplus p^{\prime}\oplus r\oplus 0_{n} are Murray–von Neumann equivalent mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J}. By Lemma 2.37, we have the following equivalent description of B0(𝔄/𝔍)\mathrm{B}_{0}\left(\mathfrak{A/J}\right).

Lemma 3.1.

Suppose that 𝔄/𝔍\mathfrak{A/J} is a unital C*-algebra with a strict cover, and (p,p),(q,q)Z0(𝔄/𝔍)\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\in\mathrm{Z}_{0}\left(\mathfrak{A/J}\right) where p,pMd(𝔄/𝔍)p,p^{\prime}\in M_{d}\left(\mathfrak{A/J}\right) and q,qMk(𝔄/𝔍)q,q^{\prime}\in M_{k}\left(\mathfrak{A/J}\right). Then (p,p)B0(𝔄/𝔍)(q,q)\left(p,p^{\prime}\right)\mathrm{B}_{0}\left(\mathfrak{A/J}\right)\left(q,q^{\prime}\right) if and only if there exist m,nωm,n\in\omega and

y1,,yBall(Md+k+m+n(𝔄)sa)y_{1},\ldots,y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{d+k+m+n}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right)

such that, setting u:=eiy1eiyu:=e^{iy_{1}}\cdots e^{iy_{\ell}}, one has that

u(pq1m0n)uqp1m0nmod𝔍,u\left(p\oplus q^{\prime}\oplus 1_{m}\oplus 0_{n}\right)u^{\ast}\equiv q\oplus p^{\prime}\oplus 1_{m}\oplus 0_{n}\mathrm{\ \mathrm{mod}}\ \mathfrak{J}\text{,}

where 1\ell\geq 1 does not depend on 𝔄/𝔍\mathfrak{A/J} and (p,p),(q,q)Z0(𝔄/𝔍)\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\in\mathrm{Z}_{0}\left(\mathfrak{A/J}\right).

The (commutative) group operation on K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A/J}\right) is induced by the Borel function on Z0(𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A/J}\right), ((p,p),(q,q))(pq,pq)\left(\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\right)\mapsto\left(p\oplus q,p^{\prime}\oplus q^{\prime}\right). The neutral element of K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A/J}\right) corresponds to (0,0)Z0(𝔄/𝔍)\left(0,0\right)\in\mathrm{Z}_{0}\left(\mathfrak{A/J}\right). The function that maps an element to its additive inverse is induced by the Borel function on Z0(𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A/J}\right) given by (p,p)(p,p)\left(p,p^{\prime}\right)\mapsto\left(p^{\prime},p\right). Thus, K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A/J}\right) is in fact a semidefinable group.

If 𝔄/\mathfrak{A/I} and 𝔅/𝔍\mathfrak{B/J} are unital C*-algebras with a strict cover, and φ:𝔄/𝔅/𝔍\varphi:\mathfrak{A/I}\rightarrow\mathfrak{B/J} is a definable unital *-homomorphism, then the induced group homomorphism K0(𝔄/)K0(𝔅/𝔍){}_{0}\left(\mathfrak{A/I}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{B/J}\right) is also definable. Thus, the assignment 𝔄/𝔍K0(𝔄/𝔍)\mathfrak{A/J}\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right) gives a functor from the category of unital C*-algebras with a strict cover to the category of semidefinable abelian groups.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. We denote by 𝔍+\mathfrak{J}^{+} the unitization of 𝔍\mathfrak{J}, which can be identified with the C*-subalgebra 𝔍+=span{𝔍,1}𝔄\mathfrak{J}^{+}=\mathrm{\mathrm{span}}\left\{\mathfrak{J},1\right\}\subseteq\mathfrak{A}. Since 𝔍\mathfrak{J} is a proper ideal of 𝔄\mathfrak{A}, we can write every element of 𝔍+\mathfrak{J}^{+} uniquely as a+λ1a+\lambda 1 where a𝔍a\in\mathfrak{J} and λ\lambda\in\mathbb{C}. More generally, every element of Mn(𝔍+)M_{n}\left(\mathfrak{J}^{+}\right) can be written uniquely as a+α1a+\alpha 1 where aMn(𝔍)a\in M_{n}\left(\mathfrak{J}\right) and αMn()\alpha\in M_{n}\left(\mathbb{C}\right). As the map Mn(𝔍+)Mn()M_{n}\left(\mathfrak{J}^{+}\right)\rightarrow M_{n}\left(\mathbb{C}\right), a+α1αa+\alpha 1\mapsto\alpha is a unital *-homomorphism, we have that αa+α1\left\|\alpha\right\|\leq\left\|a+\alpha 1\right\| and hence a2a+α1\left\|a\right\|\leq 2\left\|a+\alpha 1\right\| for a+α1Mn(𝔍+)a+\alpha 1\in M_{n}\left(\mathfrak{J}^{+}\right).

We define Proj(Mn(𝔍+))\mathrm{Proj}(M_{n}(\mathfrak{J}^{+})) to be the set of projections in Mn(𝔍+)M_{n}(\mathfrak{J}^{+}), which we regard as a Borel subset of 2Ball(Mn(𝔍))×Ball(Mn())2\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathfrak{J}\right)\right)\times\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathbb{C}\right)\right). Similarly, the unitary group U(Mn(𝔍+))U(M_{n}(\mathfrak{J}^{+})) of Mn(𝔍+)M_{n}(\mathfrak{J}^{+}) is regarded as a Borel subset of 2Ball(Mn(𝔍))×Ball(Mn())2\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathfrak{J}\right)\right)\times\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathbb{C}\right)\right). Define also Z0(n)(𝔍)\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{J}\right) to be the Borel subset of Proj(Mn(𝔍+))×Proj(Mn(𝔍+))\mathrm{Proj}(M_{n}(\mathfrak{J}^{+}))\times\mathrm{Proj}(M_{n}(\mathfrak{J}^{+})) consisting of pairs (p,p)\left(p,p^{\prime}\right) such that ppmodMn(𝔍)p\equiv p^{\prime}\mathrm{\ \mathrm{mod}}\ M_{n}\left(\mathfrak{J}\right). Finally, let Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right) to be the disjoint union of Z0(n)(𝔍)\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{J}\right) for n1n\geq 1.

The K0\mathrm{K}_{0}-group K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right) of 𝔍\mathfrak{J}—see [higson_analytic_2000, Definition 4.2.1]—is defined as a quotient of Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right) by an equivalence relation B0(𝔍)\mathrm{B}_{0}\left(\mathfrak{J}\right), defined as follows. One has that, for (p,p),(q,q)Z0(𝔍)\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\in\mathrm{Z}_{0}\left(\mathfrak{J}\right), (p,p)B0(𝔍)(q,q)\left(p,p^{\prime}\right)\mathrm{B}_{0}\left(\mathfrak{J}\right)\left(q,q^{\prime}\right) if and only if there exist m,nωm,n\in\omega and xProj(Mm(𝔍+))x\in\mathrm{Proj}(M_{m}(\mathfrak{J}^{+})) such that pqx0np\oplus q^{\prime}\oplus x\oplus 0_{n} and qpx0nq\oplus p^{\prime}\oplus x\oplus 0_{n^{\prime}} are Murray–von Neumann equivalent. For (p,p)Z0(𝔍)\left(p,p^{\prime}\right)\in\mathrm{Z}_{0}\left(\mathfrak{J}\right), we let [p][p]\left[p\right]-\left[p^{\prime}\right] be the corresponding element of K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right). The (commutative) group operation on K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right) is induced by the Borel function on Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right), ((p,p),(q,q))(pq,pq)\left(\left(p,p^{\prime}\right),\left(q,q^{\prime}\right)\right)\mapsto\left(p\oplus q,p^{\prime}\oplus q^{\prime}\right). The neutral element of K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right) corresponds to (0,0)Z0(𝔍)\left(0,0\right)\in\mathrm{Z}_{0}\left(\mathfrak{J}\right). The function that maps an element of K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right) to its additive inverse is induced by the Borel function on Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right) given by (p,p)(p,p)\left(p,p^{\prime}\right)\mapsto\left(p^{\prime},p\right). Thus, K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right) is a semidefinable group.

If (𝔄,)\left(\mathfrak{A},\mathfrak{I}\right) are (𝔅,𝔍)\left(\mathfrak{B},\mathfrak{J}\right) are strict C*-pairs, and φ:(𝔄,)(𝔅,𝔍)\varphi:\left(\mathfrak{A},\mathfrak{I}\right)\rightarrow\left(\mathfrak{B},\mathfrak{J}\right) is a strict *-homomorphism, then it induces a strict *-homomorphism φ|:𝔍\varphi|_{\mathfrak{I}}:\mathfrak{I}\rightarrow\mathfrak{J}. In turn, this induces a definable group homomorphism K0(𝔍)K0()\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{I}\right). This gives a functor (𝔄,)K0()\left(\mathfrak{A},\mathfrak{I}\right)\mapsto\mathrm{K}_{0}\left(\mathfrak{I}\right) from strict unital C*-pairs to semidefinable groups.

Lemma 3.2.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. Then there is a Borel map Z0(𝔍)Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{J}\right), (P,P)(p,p)\left(P,P^{\prime}\right)\mapsto\left(p,p^{\prime}\right) such that [P][P]=[p][p]\left[P\right]-\left[P^{\prime}\right]=\left[p\right]-\left[p^{\prime}\right] and pMn()p^{\prime}\in M_{n}\left(\mathbb{C}\right).

Proof.

Suppose that (P,P)Z0(d)(𝔍)\left(P,P^{\prime}\right)\in\mathrm{Z}_{0}^{\left(d\right)}\left(\mathfrak{J}\right). By definition, we have that for some x,xMd(𝔍)x,x^{\prime}\in M_{d}\left(\mathfrak{J}\right) and αMd()\alpha\in M_{d}\left(\mathbb{C}\right), P=x+αP=x+\alpha and P=x+αP^{\prime}=x^{\prime}+\alpha. Thus, we can define

p:=[P001dP]M2d(𝔍+)p:=\begin{bmatrix}P&0\\ 0&1_{d}-P^{\prime}\end{bmatrix}\in M_{2d}\left(\mathfrak{J}^{+}\right)

and

p:=[α001dα]M2d().p^{\prime}:=\begin{bmatrix}\alpha&0\\ 0&1_{d}-\alpha\end{bmatrix}\in M_{2d}\left(\mathbb{C}\right)\text{.}

Then we have that

[p][p]=[P]+[1dP]+[α][1dα]=[P][P].\left[p\right]-\left[p^{\prime}\right]=\left[P\right]+\left[1_{d}-P^{\prime}\right]+\left[\alpha\right]-\left[1_{d}-\alpha\right]=\left[P\right]-\left[P^{\prime}\right]\text{.}

This concludes the proof. ∎

3.2. Relative K0\mathrm{K}_{0}-group

Suppose now that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. For n1n\geq 1, define Z0(n)(𝔄,𝔄/𝔍)\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{A},\mathfrak{A/J}\right) to be the Borel set of triples (p,q,x)Ball(𝔄)3\left(p,q,x\right)\in\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right)^{3} where p,qp,q are projections and xBall(Mn(𝔄))x\in\mathrm{\mathrm{Ball}}\left(M_{n}\left(\mathfrak{A}\right)\right) satisfies xxpmodMn(𝔍)x^{\ast}x\equiv p\mathrm{\ \mathrm{mod}}\ M_{n}\left(\mathfrak{J}\right) and xxqmodMn(𝔍)xx^{\ast}\equiv q\mathrm{\ \mathrm{mod}}\ M_{n}\left(\mathfrak{J}\right). Define Z0(𝔄,𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) to be the disjoint union of K0(n)(𝔄,𝔄/𝔍)\mathrm{K}_{0}^{\left(n\right)}\left(\mathfrak{A},\mathfrak{A/J}\right) for n1n\geq 1 endowed with the induced standard Borel structure. The elements of Z0(𝔄,𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) are called relative K\mathrm{K}-cycles for (𝔄,𝔄/𝔍)\left(\mathfrak{A},\mathfrak{A/J}\right); see [higson_analytic_2000, Definition 4.3.1]. If (p,q,x)Z0(n)(𝔄,𝔄/𝔍)\left(p,q,x\right)\in\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{A},\mathfrak{A/J}\right), then we say that (p,q,x)\left(p,q,x\right) is a relative K\mathrm{K}-cycle of dimension nn. A relative K\mathrm{K}-cycle (p,q,x)\left(p,q,x\right) for (𝔄,𝔄/𝔍)\left(\mathfrak{A},\mathfrak{A/J}\right) is degenerate if xx=px^{\ast}x=p and xx=qxx^{\ast}=q. Two relative K\mathrm{K}-cycles (p,q,x)\left(p,q,x\right) and (p,q,x)\left(p^{\prime},q^{\prime},x^{\prime}\right) of dimension nn are homotopic if there exists a norm-continuous path ((pt,qt,xt))t[0,1]\left(\left(p_{t},q_{t},x_{t}\right)\right)_{t\in\left[0,1\right]} of relative K\mathrm{K}-cycles for (𝔄,𝔄/𝔍)\left(\mathfrak{A},\mathfrak{A/J}\right) of dimension nn with (p,q,x)=(p0,q0,x0)\left(p,q,x\right)=\left(p_{0},q_{0},x_{0}\right) and (p,q,x)=(p1,q1,x1)\left(p^{\prime},q^{\prime},x^{\prime}\right)=\left(p_{1},q_{1},x_{1}\right).

Notice that if (p,q,x)\left(p,q,x\right) is a relative K\mathrm{K}-cycle of dimension dd, and (ut)t[0,1]\left(u_{t}\right)_{t\in\left[0,1\right]} is a path of unitaries in Md(𝔄)M_{d}\left(\mathfrak{A}\right) starting at 11, then

(utput,utqut,utxut)\left(u_{t}^{\ast}pu_{t},u_{t}^{\ast}qu_{t},u_{t}^{\ast}xu_{t}\right)

and

(p,utqut,utx)\left(p,u_{t}^{\ast}qu_{t},u_{t}^{\ast}x\right)

are norm-continuous paths of relative K\mathrm{K}-cycles starting at (p,q,x)\left(p,q,x\right). If pqxmodMd(𝔍)p\equiv q\equiv x\mathrm{\ \mathrm{mod}}\ M_{d}\left(\mathfrak{J}\right),  then

(p,q,tp+(1t)q)\left(p,q,tp+\left(1-t\right)q\right)

is a norm-continuous path of relative cycles from (p,q,x)\left(p,q,x\right) to (p,q,p)\left(p,q,p\right). We have the following lemma; see [nest_excision_2017, Proposition 3.4].

Lemma 3.3.

Suppose that (p,q,x)\left(p,q,x\right) is a relative cycle of dimension nn for (𝔄,𝔄/𝔍)\left(\mathfrak{A},\mathfrak{A/J}\right). Then r0:=(pq,qp,xx)r_{0}:=\left(p\oplus q,q\oplus p,x\oplus x^{\ast}\right) is homotopic to the degenerate cycle (pq,pq,pq)\left(p\oplus q,p\oplus q,p\oplus q\right).

The relative K0\mathrm{K}_{0}-group K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is defined to be the quotient of Z0(𝔄,𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) by the equivalence relation B0(𝔄,𝔄/𝔍)\mathrm{B}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) defined as follows. For (p,q,x),(p,q,x)\left(p,q,x\right),\left(p^{\prime},q^{\prime},x^{\prime}\right), set (p,q,x)B0(𝔄,𝔄/𝔍)(p,q,x)\left(p,q,x\right)\mathrm{B}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\left(p^{\prime},q^{\prime},x^{\prime}\right) if and only if there exists a degenerate relative K\mathrm{K}-cycles (p0,q0,x0),(p0,q0,x0)\left(p_{0},q_{0},x_{0}\right),\left(p_{0}^{\prime},q_{0}^{\prime},x_{0}^{\prime}\right) such that (pp0,qq0,xx0)\left(p\oplus p_{0},q\oplus q_{0},x\oplus x_{0}\right) and (pp0,qq0,xx0)\left(p^{\prime}\oplus p_{0}^{\prime},q^{\prime}\oplus q_{0}^{\prime},x^{\prime}\oplus x_{0}^{\prime}\right) are of the same dimension and homotopic. The group operations on K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) are induced by the Borel maps

((p,q,x),(p,q,x))(pp,qq,xx)\left(\left(p,q,x\right),\left(p^{\prime},q^{\prime},x^{\prime}\right)\right)\mapsto\left(p\oplus p^{\prime},q\oplus q^{\prime},x\oplus x^{\prime}\right)

and

(p,q,x)(q,p,x).\left(p,q,x\right)\mapsto\left(q,p,x^{\ast}\right)\text{.}

It follows from Lemma 3.3 that K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is indeed a group. We let [p,q,x]\left[p,q,x\right] be the element of K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) represented by the relative K\mathrm{K}-cycle (p,q,x)\left(p,q,x\right). The trivial element of K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is equal to [p,q,x]\left[p,q,x\right] where (p,q,x)\left(p,q,x\right) is any degenerate relative K\mathrm{K}-cycle. Let 𝔍+\mathfrak{J}^{+} be the unitization of 𝔍\mathfrak{J}, which we identify with span(𝔍,1)𝔄\mathrm{\mathrm{span}}\left(\mathfrak{J},1\right)\subseteq\mathfrak{A}.

Lemma 3.4.

There is a Borel function Z0(𝔄,𝔄/𝔍)Z0(𝔄,𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right), (P,Q,X)(p,q,p)\left(P,Q,X\right)\mapsto\left(p,q,p\right) such that pMn()p\in M_{n}\left(\mathbb{C}\right), qMn(𝔍+)q\in M_{n}\left(\mathfrak{J}^{+}\right), pqMn(𝔍)p\equiv q\in M_{n}\left(\mathfrak{J}\right), and [P,Q,X]=[p,q,p]K0(𝔄,𝔄/𝔍)[P,Q,X]=[p,q,p]\in\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right).

Proof.

Notice that (1P,1P,1P)\left(1-P,1-P,1-P\right) is a degenerate relative K\mathrm{K}-cycle of dimension dd. Consider then

(P(1P),Q(1P),X(1P)).\left(P\oplus\left(1-P\right),Q\oplus\left(1-P\right),X\oplus\left(1-P\right)\right).

By Lemma 2.38 one can choose Y1,,YBall(M2d(𝔄)sa)Y_{1},\ldots,Y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{2d}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) in a Borel way such that, setting U:=eiY1eiYU:=e^{iY_{1}}\cdots e^{iY_{\ell}}, one has that U(P(1P)U)=1d0dU\left(P\oplus\left(1-P\right)U^{\ast}\right)=1_{d}\oplus 0_{d}, where 1\ell\geq 1 does not depend on 𝔄,𝔍\mathfrak{A},\mathfrak{J} and (P,Q,X)\left(P,Q,X\right). Thus, after replacing (P,Q,X)\left(P,Q,X\right) with

(U(P(1P))U,U(Q(1P))U,U(X(1P))U),\left(U\left(P\oplus\left(1-P\right)\right)U^{\ast},U\left(Q\oplus\left(1-P\right)\right)U^{\ast},U\left(X\oplus\left(1-P\right)\right)U^{\ast}\right),

we can assume without loss of generality that P=1d0dP=1_{d}\oplus 0_{d}.

By Lemma 2.35 one can choose Y1,,YBall(M4d(𝔄)sa)Y_{1},\ldots,Y_{\ell}\in\mathrm{\mathrm{Ball}}\left(M_{4d}\left(\mathfrak{A}\right)_{\mathrm{sa}}\right) in a Borel fashion from (P,Q,X)\left(P,Q,X\right) such that, setting U:=eiY1eiYU:=e^{iY_{1}}\cdots e^{iY_{\ell}}, one has that

U(Q02d)U(P02d)modM2d(𝔍).U^{\ast}\left(Q\oplus 0_{2d}\right)U\equiv\left(P\oplus 0_{2d}\right)\mathrm{\ \mathrm{mod}}\ M_{2d}\left(\mathfrak{J}\right)\text{.}

Thus, after replacing (P,Q,X)\left(P,Q,X\right) with (P02d,U(Q02d)U,U(X02d))\left(P\oplus 0_{2d},U^{\ast}\left(Q\oplus 0_{2d}\right)U,U^{\ast}\left(X\oplus 0_{2d}\right)\right) we can assume without loss of generality that P=1d03dM4d()P=1_{d}\oplus 0_{3d}\in M_{4d}\left(\mathbb{C}\right) and QM4d(𝔄)Q\in M_{4d}\left(\mathfrak{A}\right) satisfy PQmodM4d(𝔍)P\equiv Q\mathrm{\ \mathrm{mod}}\ M_{4d}\left(\mathfrak{J}\right) and hence QM4n(𝔍+)Q\in M_{4n}\left(\mathfrak{J}^{+}\right).

In this case, we have that [P,Q,X]=[P,Q,P]\left[P,Q,X\right]=\left[P,Q,P\right], since (Pt,Qt,tP+(1t)X)t[0,1]\left(P_{t},Q_{t},tP+\left(1-t\right)X\right)_{t\in\left[0,1\right]} is a norm-continuous path of relative K\mathrm{K}-cycles from (P,Q,X)\left(P,Q,X\right) to (P,Q,P)\left(P,Q,P\right). This concludes the proof. ∎

Proposition 3.5.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict C*-pair. Then K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is a definable group. The assignment K0(𝔍)K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right), [P][Q][P,Q,P][P]-\left[Q\right]\mapsto[P,Q,P] is a natural definable isomorphism, called the excision isomorphism.

Proof.

By [nest_excision_2017, Theorem 3.9], the excision homomorphism K0(𝔍)K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is bijective; see also [higson_analytic_2000, Theorem 4.3.8]. Clearly, it is induced by a Borel function Z0(𝔍)Z0(𝔄,𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right). By Lemma 3.4 the inverse homomorphism K0(𝔄,𝔄/𝔍)K0(𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{J}\right) is also induced by a Borel function Z0(𝔄,𝔄/𝔍)Z0(𝔍)\mathrm{Z}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{J}\right). Thus, K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) is a definable group, and the excision isomorphism is a definable isomorphism. ∎

There is a natural definable homomorphism K0(𝔄,𝔄/𝔍)K0(𝔄)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right) that is induced by the Borel map Z0(𝔄,𝔄/𝔍)Z0(𝔄){}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{A}\right) (p,q,x)(p,q)\left(p,q,x\right)\mapsto\left(p,q\right). We also have a natural definable homomorphism K0(𝔄)K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right) induced by the Borel map Z0(𝔄)Z0(𝔄/𝔍)\mathrm{Z}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{Z}_{0}\left(\mathfrak{A/J}\right). We have the following result; see [higson_analytic_2000, Proposition 4.3.5].

Proposition 3.6.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. The (natural) sequence of definable groups and definable group homomorphisms

K0(𝔄,𝔄/𝔍)K0(𝔄)K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right)

is exact.

Combining the excision isomorphism K0(𝔍)K0(𝔄,𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right) with the natural definable homomorphism K0(𝔄,𝔄/𝔍)K0(𝔄)\mathrm{K}_{0}\left(\mathfrak{A},\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right), we obtain a natural definable group homomorphism K0(𝔍)K0(𝔄)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right). This is defined by mapping (p,q)Z0(n)(𝔍)\left(p,q\right)\in\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{J}\right) to (p,q)\left(p,q\right) regarded as an element of Z0(n)(𝔄)\mathrm{Z}_{0}^{\left(n\right)}\left(\mathfrak{A}\right). Combining Proposition 3.5 with Proposition 3.6 we have the following.

Corollary 3.7.

Suppose that 𝔄\mathfrak{A} is a unital strict C*-algebra and 𝔍\mathfrak{J} is a proper strict ideal of 𝔄\mathfrak{A}. Then the natural sequence

K0(𝔍)K0(𝔄)K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right)

is exact.

3.3. K1\mathrm{K}_{1} group

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. We can then consider the Borel set U(𝔄/𝔍)U\left(\mathfrak{A/J}\right) of elements of Ball(𝔄)\mathrm{\mathrm{Ball}}\left(\mathfrak{A}\right) that are unitaries mod𝔍\mathrm{\mathrm{mod}}\ \mathfrak{J}. We then let Z1(𝔄/𝔍)\mathrm{Z}_{1}\left(\mathfrak{A/J}\right) to be the disjoint union of U(Mn(𝔄)/Mn(𝔍))U\left(M_{n}\left(\mathfrak{A}\right)/M_{n}\left(\mathfrak{J}\right)\right) for n1n\geq 1. The equivalence relation B1(𝔄/𝔍){}_{1}\left(\mathfrak{A/J}\right) on Z1(𝔄/𝔍)\mathrm{Z}_{1}\left(\mathfrak{A/J}\right) is defined by setting uuB1(𝔄/𝔍)u{}_{1}\left(\mathfrak{A/J}\right)u^{\prime} for uU(Mn(𝔄)/Mn(𝔍))u\in U\left(M_{n}\left(\mathfrak{A}\right)/M_{n}\left(\mathfrak{J}\right)\right) and uU(Mn(𝔄)/Mn(𝔍))u^{\prime}\in U\left(M_{n^{\prime}}\left(\mathfrak{A}\right)/M_{n^{\prime}}\left(\mathfrak{J}\right)\right) if and only if there exist k,kωk,k^{\prime}\in\omega with n+k=n+kn+k=n^{\prime}+k^{\prime} and such that there is a norm-continuous path from u1k+Mn+k(𝔍)u\oplus 1_{k}+M_{n+k}\left(\mathfrak{J}\right) to u1k+Mn+k(𝔍)u^{\prime}\oplus 1_{k^{\prime}}+M_{n+k}\left(\mathfrak{J}\right) in the unitary group of the quotient unital C*-algebra Mn+k(𝔄)/Mn+k(𝔍)M_{n+k}\left(\mathfrak{A}\right)/M_{n+k}\left(\mathfrak{J}\right). This equivalence relation is analytic by Corollary 2.34. The definable K1\mathrm{K}_{1}-group K1(𝔄/𝔍)\mathrm{K}_{1}\left(\mathfrak{A/J}\right) is then the semidefinable group obtained as quotient Z1(𝔄/𝔍)/B1(𝔄/𝔍)\mathrm{Z}_{1}\left(\mathfrak{A/J}\right)/\mathrm{B}_{1}\left(\mathfrak{A/J}\right) with group operations defined as above. This defines a functor 𝔄/𝔍K1(𝔄/𝔍)\mathfrak{A/J}\mapsto\mathrm{K}_{1}\left(\mathfrak{A/J}\right) from unital C*-algebras with strict cover to semidefinable groups.

Given a strict unital C*-pair (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right), we also consider the definable K1\mathrm{K}_{1}-group K1(𝔍)\mathrm{K}_{1}\left(\mathfrak{J}\right) of 𝔍\mathfrak{J}. As above, we identify the unitization 𝔍+\mathfrak{J}^{+} of 𝔍\mathfrak{J} with span{𝔍,1}𝔄\mathrm{\mathrm{span}}\left\{\mathfrak{J},1\right\}\subseteq\mathfrak{A}. For n1n\geq 1 we let U(Md(𝔍+))U\left(M_{d}\left(\mathfrak{J}^{+}\right)\right) be the unitary group of Md(𝔍+)M_{d}\left(\mathfrak{J}^{+}\right). Recall that every element of Md(𝔍+)M_{d}\left(\mathfrak{J}^{+}\right) can be written uniquely as x+α1x+\alpha 1 where xMd(𝔍)x\in M_{d}\left(\mathfrak{J}\right) and αMd()\alpha\in M_{d}\left(\mathbb{C}\right). We consider U(Md(𝔍+))U\left(M_{d}\left(\mathfrak{J}^{+}\right)\right) as a Borel subset of Ball(Md(𝔍))×Ball(Md(()))\mathrm{\mathrm{Ball}}\left(M_{d}\left(\mathfrak{J}\right)\right)\times\mathrm{\mathrm{Ball}}\left(M_{d}\left(\left(\mathbb{C}\right)\right)\right). We then set Z1(𝔍)\mathrm{Z}_{1}\left(\mathfrak{J}\right) to be the disjoint union of U(Md(𝔍+))U\left(M_{d}\left(\mathfrak{J}^{+}\right)\right) for d1d\geq 1, and let B1(𝔍)\mathrm{B}_{1}\left(\mathfrak{J}\right) be the (analytic) equivalence relation on Z1(𝔍)\mathrm{Z}_{1}\left(\mathfrak{J}\right) obtained by setting uuB1(𝔍)u{}_{1}\left(\mathfrak{J}\right)u^{\prime} for uU(Mn(𝔍+))u\in U\left(M_{n}\left(\mathfrak{J}^{+}\right)\right) and uU(Mn(𝔍+))u^{\prime}\in U\left(M_{n^{\prime}}\left(\mathfrak{J}^{+}\right)\right) if and only if there exist k,kωk,k^{\prime}\in\omega with n+k=n+kn+k=n^{\prime}+k^{\prime} and such that there is a norm-continuous path from u1ku\oplus 1_{k} to u1ku^{\prime}\oplus 1_{k^{\prime}} in U(Mn+k(𝔍+))U\left(M_{n+k}\left(\mathfrak{J}^{+}\right)\right). The definable K1\mathrm{K}_{1}-group K1(𝔍)\mathrm{K}_{1}\left(\mathfrak{J}\right) is then the semidefinable group obtained as quotient Z1(𝔍)/B1(𝔍)\mathrm{Z}_{1}\left(\mathfrak{J}\right)/\mathrm{B}_{1}\left(\mathfrak{J}\right) with group operations defined as above.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. We have natural definable group homomorphisms

K1(𝔍)K1(𝔄)K1(𝔄/𝔍).\mathrm{K}_{1}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A/J}\right)\text{.}

The definable group homomorphism K1(𝔍)K1(𝔄)\mathrm{K}_{1}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A}\right) is induced by the inclusion 𝔍+𝔄\mathfrak{J}^{+}\subseteq\mathfrak{A}, which gives an inclusion U(Md(𝔍+))U(Md(𝔄))U\left(M_{d}\left(\mathfrak{J}^{+}\right)\right)\rightarrow U\left(M_{d}\left(\mathfrak{A}\right)\right) for every d1d\geq 1. The definable group homomorphism K1(𝔄)K1(𝔄/𝔍)\mathrm{K}_{1}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A/J}\right) is also induced by the inclusion maps U(Md(𝔄))U(Md(𝔄)/Md(𝔍))U\left(M_{d}\left(\mathfrak{A}\right)\right)\rightarrow U\left(M_{d}\left(\mathfrak{A}\right)/M_{d}\left(\mathfrak{J}\right)\right) for d1d\geq 1. We have the following result, which can be easily verified directly, and also follows from Corollary 3.7 via the Bott isomorphism theorem [higson_analytic_2000, Theorem 4.9.1].

Proposition 3.8.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. The sequence of natural definable homomorphisms

K1(𝔍)K1(𝔄)K1(𝔄/𝔍)\mathrm{K}_{1}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A/J}\right)

is exact.

3.4. The six-term exact sequence

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. One can define a natural definable group homomorphism 1:K1(𝔄/𝔍)K0(𝔍)\partial_{1}:\mathrm{K}_{1}\left(\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{J}\right) called the index map, as follows. An element of K1(𝔄/𝔍)\mathrm{K}_{1}\left(\mathfrak{A/J}\right) is of the form [u]\left[u\right] where uU(Md(𝔄)/Md(𝔍))u\in U\left(M_{d}\left(\mathfrak{A}\right)/M_{d}\left(\mathfrak{J}\right)\right) for some d1d\geq 1. Then define

P:=[uuu(1uu)1/2u(1uu)1/21uu]M2d(𝔍+)P:=\begin{bmatrix}uu^{\ast}&u\left(1-u^{\ast}u\right)^{1/2}\\ u^{\ast}\left(1-uu^{\ast}\right)^{1/2}&1-u^{\ast}u\end{bmatrix}\in M_{2d}\left(\mathfrak{J}^{+}\right)

and

Q:=1d0dM2d(𝔍+).Q:=1_{d}\oplus 0_{d}\in M_{2d}\left(\mathfrak{J}^{+}\right)\text{.}

Then P,QP,Q are projections such that PQmodM2d(𝔍)P\equiv Q\mathrm{\ \mathrm{mod}}\ M_{2d}\left(\mathfrak{J}\right) and hence (P,Q)Z0(𝔍)\left(P,Q\right)\in\mathrm{Z}_{0}\left(\mathfrak{J}\right). One then defines 1([u])=[P][Q]\partial_{1}\left(\left[u\right]\right)=\left[P\right]-\left[Q\right]; see [higson_analytic_2000, Proposition 4.8.10]. As (P,Q)\left(P,Q\right) is obtained in a Borel fashion from uu, the boundary map 1:K1(𝔄/𝔍)K0(𝔍)\partial_{1}:\mathrm{K}_{1}\left(\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{J}\right) is definable.

Equivalently, one can define 1\partial_{1} as follows. Given an element [u]\left[u\right] of K1(𝔄/𝔍)\mathrm{K}_{1}\left(\mathfrak{A/J}\right) for some uU(Md(𝔄)/Md(𝔍))u\in U\left(M_{d}\left(\mathfrak{A}\right)/M_{d}\left(\mathfrak{J}\right)\right). Consider the partial isometry vM2d(𝔄)v\in M_{2d}\left(\mathfrak{A}\right) defined by

v=[u0(1uu)1/20]v=\begin{bmatrix}u&0\\ \left(1-u^{\ast}u\right)^{1/2}&0\end{bmatrix}

and observe that vu0mod𝔍v\equiv u\oplus 0\mathrm{\ \mathrm{mod}}\ \mathfrak{J} [rordam_introduction_2000, Lemma 9.2.1]. Then 12dvv1_{2d}-v^{\ast}v and 12dvv1_{2d}-vv^{\ast} are projections in M2d(𝔍+)M_{2d}\left(\mathfrak{J}^{+}\right) such that 12dvv12dvv0d1dMd()1_{2d}-v^{\ast}v\equiv 1_{2d}-vv^{\ast}\equiv 0_{d}\oplus 1_{d}\in M_{d}\left(\mathbb{C}\right). Therefore, (12dvv,12dvv)Z0(𝔍)\left(1_{2d}-v^{\ast}v,1_{2d}-vv^{\ast}\right)\in\mathrm{Z}_{0}\left(\mathfrak{J}\right). One has that 1[u]=[12dvv][12dvv]K0(𝔍)\partial_{1}[u]=[1_{2d}-v^{\ast}v]-[1_{2d}-vv^{\ast}]\in\mathrm{K}_{0}\left(\mathfrak{J}\right); see [rordam_introduction_2000, Proposition 9.2.3]. Then we have the following; see [rordam_introduction_2000, Lemma 9.3.1 and Lemma 9.3.2].

Proposition 3.9.

Suppose that 𝔄\mathfrak{A} is a strict unital C*-algebra and 𝔍\mathfrak{J} is a strict ideal of 𝔄\mathfrak{A}. Then the sequence

K1(𝔄)K1(𝔄/𝔍)1K0(𝔍)K0(𝔄)\mathrm{K}_{1}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A/J}\right)\overset{\partial_{1}}{\rightarrow}\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right)

is exact.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. One can consider a natural definable homomorphism 0:K0(𝔄/𝔍)K1(𝔍)\partial_{0}:\mathrm{K}_{0}\left(\mathfrak{A/J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{J}\right) called the exponential map. This is defined as follows. Consider an element of K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{A/J}\right) of the form [p][q][p]-[q] for some p,qProj(Mn(𝔄)/Mn(𝔍))p,q\in\mathrm{Proj}\left(M_{n}\left(\mathfrak{A}\right)\mathfrak{/}M_{n}\left(\mathfrak{J}\right)\right). Then we have that exp(2πip)\exp\left(2\pi ip\right) and exp(2πiq)\exp\left(2\pi iq\right) are unitary elements of Mn(𝔍+)M_{n}\left(\mathfrak{J}^{+}\right) such that exp(2πip)exp(2πiq)modMn(𝔍)\exp\left(2\pi ip\right)\equiv\exp\left(2\pi iq\right)\mathrm{\ \mathrm{mod}}\ M_{n}\left(\mathfrak{J}\right). Then one has that 0([p][q])=[exp(2πip)][exp(2πiq)]K1(𝔍)\partial_{0}\left([p]-[q]\right)=[\exp\left(2\pi ip\right)]-[\exp\left(2\pi iq\right)]\in\mathrm{K}_{1}\left(\mathfrak{J}\right); see [rordam_introduction_2000, Proposition 12.2.2] and [higson_analytic_2000, Section 4.9]. From Proposition 3.9 one can obtain via the Bott isomorphism theorem [higson_analytic_2000, Theorem 4.9.1] the following.

Proposition 3.10.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. Then the sequence

K0(𝔄)K0(𝔄/𝔍)0K1(𝔍)K1(𝔄)\mathrm{K}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right)\overset{\partial_{0}}{\rightarrow}\mathrm{K}_{1}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A}\right)

is exact.

Suppose that (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right) is a strict unital C*-pair. Then as discussed above we have exact sequences

K0(𝔍)K0(𝔄)K0(𝔄/𝔍)\mathrm{K}_{0}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{0}\left(\mathfrak{A/J}\right)

and

K1(𝔍)K1(𝔄)K1(𝔄/𝔍).\mathrm{K}_{1}\left(\mathfrak{J}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A}\right)\rightarrow\mathrm{K}_{1}\left(\mathfrak{A/J}\right)\text{.}

These are joined together by the index and exponential maps. From Proposition 3.10, Proposition 3.9, Corollary 3.7 and Proposition 3.8, one obtains the six-term exact sequence

K1(𝔍){\mathrm{K}_{1}\left(\mathfrak{J}\right)}K1(𝔄){\mathrm{K}_{1}\left(\mathfrak{A}\right)}K1(𝔄/𝔍){\mathrm{K}_{1}\left(\mathfrak{A/J}\right)}K0(𝔄/𝔍){\mathrm{K}_{0}\left(\mathfrak{A/J}\right)}K0(𝔄){\mathrm{K}_{0}\left(\mathfrak{A}\right)}K0(𝔍){\mathrm{K}_{0}\left(\mathfrak{J}\right)}0\scriptstyle{\partial_{0}}1\scriptstyle{\partial_{1}}

for the strict unital C*-pair (𝔄,𝔍)\left(\mathfrak{A},\mathfrak{J}\right), where the vertical arrows are the index map and the exponential map; see [rordam_introduction_2000, Theorem 12.1.2].

4. Definable K\mathrm{K}-homology of separable C*-algebras

In this section we recall the definition of the Ext invariant for separable unital C*-algebras, and its description due to Paschke in terms of the K\mathrm{K}-theory of Paschke dual algebras as defined in [higson_algebra_1995, higson_analytic_2000] or, equivalently, of commutants in the Calkin algebra. Following [higson_analytic_2000, Chapter 3], we consider the group Ext()1\left(-\right)^{-1} defined in terms of unital semi-split extensions. In the case of separable unital nuclear C*-algebras, every unital extension is semi-split, and the group Ext()1\left(-\right)^{-1} coincides with the group Ext()\left(-\right) defined in terms of unital extensions. Using Paschke’s K\mathrm{K}-theoretical description of Ext from [paschke_theory_1981], we show that Ext()1\mathrm{Ext}\left(-\right)^{-1} yields a contravariant functor from separable unital C*-algebras to the category of definable groups.

We also recall the definition of the K\mathrm{K}-homology groups of separable C*-algebras as in [higson_analytic_2000, Chapter 5]. Using their description in terms of Ext\mathrm{Ext}, we conclude that they can be endowed with the structure of definable groups, in such a way that the assignments AK0(A)A\mapsto\mathrm{K}^{0}\left(A\right) and AK1(A)A\mapsto\mathrm{K}^{1}\left(A\right) are functors from the category of separable C*-algebras to the category of definable groups.

We define a separable C*-pair to be a pair (A,I)\left(A,I\right) where AA is a separable C*-algebra and II is a closed two-sided ideal of AA. A morphism (A,I)(B,J)\left(A,I\right)\rightarrow\left(B,J\right) between separable C*-pairs is a *-homomorphisms ABA\rightarrow B that maps II to JJ. Recall that a C*-algebra AA is nuclear if the identity map of AA is the pointwise limit of contractive completely positive maps that factor through finite-dimensional C*-algebras; see [higson_analytic_2000, Section 3.3]. We say that a separable C*-pair (A,I)\left(A,I\right) is nuclear if AA is nuclear. In this section, we will also introduce the relative definable K\mathrm{K}-homology groups, and the six-term exact sequence in K\mathrm{K}-homology associated with a separable nuclear C*-pair.

4.1. C*-algebra extensions and the Ext group

Let HH be a separable Hilbert space, and B(H)B\left(H\right) be the algebra of bounded linear operators on HH. We let K(H)B(H)K\left(H\right)\subseteq B\left(H\right) be the closed ideal of compact operators, and Q(H)Q\left(H\right) be the Calkin algebra, which is the quotient of B(H)B\left(H\right) by K(H)K\left(H\right). Let π:B(H)Q(H)\pi:B\left(H\right)\rightarrow Q\left(H\right) be the quotient map.

If UU(H)U\in U\left(H\right) is a unitary operator, then UU defines an automorphism Ad(U):B(H)B(H)\left(U\right):B\left(H\right)\rightarrow B\left(H\right) given by TUTUT\mapsto U^{\ast}TU. As K(H)K\left(H\right) is Ad(U)\mathrm{Ad}\left(U\right)-invariant, we have an induced automorphism of Q(H)Q\left(H\right), still denoted by Ad(U)\left(U\right).

Suppose that AA is a unital, separable C*-algebra. A unital extension of AA (by K(H)K\left(H\right)) is a unital *-homomorphism φ:AQ(H)\varphi:A\rightarrow Q\left(H\right). A unital extension of AA is injective or essential if it is an injective *-homomorphism AQ(H)A\rightarrow Q\left(H\right). Two extensions φ,φ:AQ(H)\varphi,\varphi^{\prime}:A\rightarrow Q\left(H\right) are equivalent if there exists UU(H)U\in U\left(H\right) such that Ad(U)φ=φ\left(U\right)\circ\varphi^{\prime}=\varphi. An injective, unital extension φ:AQ(H)\varphi:A\rightarrow Q\left(H\right) is semi-split (or weakly nuclear in the terminology of [elliott_abstract_2001]) if there exists a unital completely positive (ucp) map σ:AB(H)\sigma:A\rightarrow B\left(H\right) such that φ=πσ\varphi=\pi\circ\sigma [higson_analytic_2000, Theorem 3.1.5]. An injective unital extension φ:AQ(H)\varphi:A\rightarrow Q\left(H\right) is split or trivial if there exists a unital *-homomorphism φ~:AB(H)\tilde{\varphi}:A\rightarrow B\left(H\right) such that φ=πφ~\varphi=\pi\circ\tilde{\varphi}.

Every unital, essential extension φ:AQ(H)\varphi:A\rightarrow Q\left(H\right) determines an exact sequence

0K(H)EφA00\rightarrow K\left(H\right)\rightarrow E_{\varphi}\rightarrow A\rightarrow 0

where

Eφ={(x,y)AB(H):φ(x)=π(y)}.E_{\varphi}=\left\{\left(x,y\right)\in A\oplus B\left(H\right):\varphi(x)=\pi(y)\right\}\text{.}

and K(H)K\left(H\right) is an essential ideal of EφE_{\varphi}. The extension is split if and only if the map EφAE_{\varphi}\rightarrow A is a split epimorphism in the category of unital C*-algebras and unital *-homomorphisms.

Conversely, given an exact sequence

0K(H)EpA00\rightarrow K\left(H\right)\rightarrow E\overset{p}{\rightarrow}A\rightarrow 0

where p:EAp:E\rightarrow A is a unital *-homomorphism and K(H)K\left(H\right) is an essential ideal of EE, one can define an essential unital extension φ:AQ(H)\varphi:A\rightarrow Q\left(H\right) as follows. Consider K(H)EB(H)K\left(H\right)\subseteq E\subseteq B\left(H\right), then define φ(a)=π(a^)Q(H)\varphi\left(a\right)=\pi\left(\widehat{a}\right)\in Q\left(H\right) for aAa\in A where a^E\widehat{a}\in E is such that p(a^)=ap\left(\widehat{a}\right)=a. Again, we have that φ\varphi is trivial if and only if p:EAp:E\rightarrow A is a split epimorphism.

Let AA be a separable, unital C*-algebra. One defines Ext(A)\left(A\right) to be the set of unitary equivalence classes of unital, injective extensions of AA by K(H)K\left(H\right); see [higson_analytic_2000, Definition 2.7.1], and Extnuc(A)=Ext(A)1{}_{\text{nuc}}\left(A\right)=\mathrm{Ext}\left(A\right)^{-1} to be the subset of unitary equivalence classes of unital, injective semi-split (or weakly nuclear) extensions of AA by K(H)K\left(H\right) [blackadar_theory_1998, 15.7.2].

One can define a commutative monoid operation on Ext(A)\mathrm{Ext}\left(A\right). The (additively denoted) operation on Ext(A)\left(A\right) is induced by the map (φ,φ)Ad(V)(φφ)\left(\varphi,\varphi^{\prime}\right)\mapsto\mathrm{Ad}\left(V\right)\circ(\varphi\oplus\varphi^{\prime}) where V:HHHV:H\rightarrow H\oplus H is a surjective linear isometry; see [higson_analytic_2000, Proposition 2.7.2]. By Voiculescu’s Theorem [higson_analytic_2000, Theorem 3.4.3], one has that the neutral element of Ext(A)\mathrm{Ext}\left(A\right) is the set of split extensions, which form a single unitary equivalence class [higson_analytic_2000, Theorem 3.4.7]. Furthermore, the set Ext(A)1\left(A\right)^{-1} is equal to the set of elements of Ext(A)\left(A\right) that have an additive inverse, whence it forms a group [higson_analytic_2000, Definition 2.7.6]. When AA is a nuclear unital separable C*-algebra, by the Choi–Effros lifting theorem [higson_analytic_2000, Theorem 3.3.6], one has that every extension of AA is semi-split, and Ext(A)=Ext(A)1\left(A\right)=\mathrm{Ext}\left(A\right)^{-1}. In particular, in this case Ext(A)\left(A\right) is itself a group.

Let AA be a separable unital C*-algebra. We regard Ext(A)1\left(A\right)^{-1} as a definable group, as follows. Fix a separable Hilbert space HH. Let us say that a ucp map ϕ:AB(H)\phi:A\rightarrow B\left(H\right) is ample if (πϕ)(x)=x\left\|\left(\pi\circ\phi\right)(x)\right\|=\left\|x\right\| for every xAx\in A. Notice that the set AUCP(A,B(H))\mathrm{A\mathrm{UCP}}\left(A,B\left(H\right)\right) of ample ucp maps AB(H)A\rightarrow B\left(H\right) is a GδG_{\delta} subset of the space Ball(L(A,B(H)))\mathrm{\mathrm{Ball}}\left(L\left(A,B\left(H\right)\right)\right) of bounded linear maps of norm at most 11 endowed with the topology of pointwise strong-* convergence. Thus, AUCP(A,B(H))\mathrm{A\mathrm{UCP}}\left(A,B\left(H\right)\right) is a Polish space.

Let (A)AUCP(A,B(H))\mathcal{E}\left(A\right)\subseteq\mathrm{A\mathrm{UCP}}\left(A,B\left(H\right)\right) be the Borel set of ample ucp maps φ:AB(H)\varphi:A\rightarrow B\left(H\right) such that

φ(xy)φ(x)φ(y)modK(H)\varphi\left(xy\right)\equiv\varphi(x)\varphi(y)\mathrm{\ \mathrm{mod}}\ K\left(H\right)

for x,yAx,y\in A. An injective, unital semi-split extension of AA by definition has a ucp lift, which is an element of (A)\mathcal{E}\left(A\right), and conversely every element of (A)\mathcal{E}\left(A\right) gives rise to an injective, unital semi-split extension of AA. Thus, we can regard (A)\mathcal{E}\left(A\right) as the space of representatives of injective, unital semi-split extensions of AA. We define a Polish topology on (A)\mathcal{E}\left(A\right) that induces the Borel structure on (A)\mathcal{E}\left(A\right) by declaring a net (φi)iI\left(\varphi_{i}\right)_{i\in I} in (A)\mathcal{E}\left(A\right) to converge to φ\varphi if and only if, for every x,yXx,y\in X, (φi(x))iω\left(\varphi_{i}(x)\right)_{i\in\omega} strong-* converges to φ(x)\varphi(x), and (φi(xy)φi(x)φi(y))iω\left(\varphi_{i}\left(xy\right)-\varphi_{i}(x)\varphi_{i}(y)\right)_{i\in\omega} norm-converges to φ(xy)φ(x)φ(y)\varphi\left(xy\right)-\varphi(x)\varphi(y).

Two elements φ,φ\varphi,\varphi^{\prime} of (A)\mathcal{E}\left(A\right) represent the same element [φ]\left[\varphi\right] of Ext(A)1\left(A\right)^{-1} if and only if there exists UU(H)U\in U\left(H\right) such that Uφ(a)Uφ(a)modK(H)U^{\ast}\varphi\left(a\right)U\equiv\varphi^{\prime}\left(a\right)\mathrm{\ \mathrm{mod}}\ K\left(H\right) for every aAa\in A. This defines an analytic equivalence relation \thickapprox on (A)\mathcal{E}\left(A\right). We can thus regard Ext(A)1\left(A\right)^{-1} as the semidefinable set (A)/\mathcal{E}\left(A\right)\left/\thickapprox\right..

We now observe that the group operations on Ext(A)1\left(A\right)^{-1} are definable, and thus this turns Ext(A)1\mathrm{Ext}\left(A\right)^{-1} into a semidefinable group. We will later show in Proposition LABEL:Proposition:K-Ext that in fact Ext(A)1\mathrm{Ext}\left(A\right)^{-1} is a definable group.

Proposition 4.1.

Let AA be a separable unital C*-algebra. The addition operation (x,y)x+y\left(x,y\right)\mapsto x+y and the additive inverse operation xxx\mapsto-x on Ext(A)1\mathrm{Ext}\left(A\right)^{-1} are definable functions. Thus, (A)/=Ext(A)1\mathcal{E}\left(A\right)\left/\thickapprox\right.=\mathrm{Ext}\left(A\right)^{-1} is a semidefinable group.

Proof.

Fix a representation AB(H)A\subseteq B\left(H\right) such that AK(H)={0}A\cap K\left(H\right)=\left\{0\right\}. The assertion for addition is clear, as the Borel map (A)×(A)(A)\mathcal{E}\left(A\right)\times\mathcal{E}\left(A\right)\rightarrow\mathcal{E}\left(A\right), (φ,φ)Ad(W)(φφ)\left(\varphi,\varphi^{\prime}\right)\mapsto\mathrm{Ad}\left(W\right)\circ\left(\varphi\oplus\varphi^{\prime}\right) is a lift for the addition operation, where WW is a fixed surjective linear isometry HHHH\rightarrow H\oplus H.

In order to obtain a lift for the function Ext(A)1Ext(A)1\mathrm{Ext}\left(A\right)^{-1}\rightarrow\mathrm{Ext}\left(A\right)^{-1}, xxx\mapsto-x, one can use the definable Stinespring Dilation Theorem (Lemma 2.41). Thus, if φ(A)\varphi\in\mathcal{E}\left(A\right), and π\pi and VV are the nondegenerate representation π\pi of AA and the isometry V:HHV:H\rightarrow H obtained from φ\varphi in a Borel fashion as in Lemma 2.41, then defining the projection P:=IVVB(H)P:=I-VV^{\ast}\in B\left(H\right) and φ:AB(H)\varphi^{\prime}:A\rightarrow B\left(H\right), aW(Pπ(a)Pa)Wa\mapsto W^{\ast}\left(P\pi\left(a\right)P\oplus a\right)W, one has that φ(A)\varphi^{\prime}\in\mathcal{E}\left(A\right) represents [φ]-\left[\varphi\right], where as above W:HHHW:H\rightarrow H\oplus H is a fixed surjective linear isometry; see also [higson_analytic_2000, Theorem 3.4.7]. ∎

Suppose that A,BB(H)A,B\subseteq B\left(H\right) are separable unital C*-algebras. A unital *-homomorphism α:AB\alpha:A\rightarrow B induces a definable group homomorphism Ext(B)1Ext(A)1\mathrm{Ext}\left(B\right)^{-1}\rightarrow\mathrm{Ext}\left(A\right)^{-1}, as follows. If φ(B)\varphi\in\mathcal{E}\left(B\right) is a representative for an injective, unital extension, then one can consider α(φ)(A)\alpha^{\ast}\left(\varphi\right)\in\mathcal{E}\left(A\right) defined by aW((φα)(a)a)Wa\mapsto W^{\ast}\left(\left(\varphi\circ\alpha\right)\left(a\right)\oplus a\right)W where W:HHHW:H\rightarrow H\oplus H is a fixed surjective linear isometry. This defines a Borel function α:(B)(A)\alpha^{\ast}:\mathcal{E}\left(B\right)\rightarrow\mathcal{E}\left(A\right), which induces a definable group homomorphism α:Ext(B)1Ext(A)1\alpha^{\ast}:\mathrm{Ext}\left(B\right)^{-1}\rightarrow\mathrm{Ext}\left(A\right)^{-1}. Thus, Ext()1\mathrm{Ext}\left(-\right)^{-1} is a contravariant functor from the category of separable unital C*-algebras to the category of semidefinable groups.

4.2. K0\mathrm{K}_{0}-group and the Voiculescu property

Let 𝔄\mathfrak{A} be a strict unital C*-algebra. Recall that two projections p,q𝔄p,q\in\mathfrak{A} are Murray–von Neumann (MvN) equivalent if there exists v𝔄v\in\mathfrak{A} such that vv=pv^{\ast}v=p and vv=qvv^{\ast}=q, in which case we write pMvNqp\sim_{\mathrm{MvN}}q. We say that a projection p𝔄p\in\mathfrak{A} is ample if p0p\oplus 0 is Murray–von Neumann equivalent to p1p\oplus 1, and co-ample if 1p1-p is ample.