Definable -homology of separable C*-algebras
Abstract.
In this paper we show that the -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 -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 -dimensional connected subspace of .
Key words and phrases:
-homology, -theory, Universal Coefficient Theorem, C*-algebra, definable group2020 Mathematics Subject Classification:
Primary 19K33, 54H05; Secondary 46M20, 46L80Introduction
Given a compact metrizable space , the group classifying extensions of the C*-algebra by the C*-algebra of compact operators was initially considered by Brown, Douglas, and Fillmore in their celebrated work [brown_extensions_1977]. There, they showed that is indeed a group, and that defining, for a compact metrizable space ,
where is the suspension of , 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 -theory [atiyah_theory_1989].
More generally, for an arbitrary separable unital C*-algebra , one can consider a semigroup classifying the essential, unital extensions of by . By Voiculescu’s non-commutative Weyl-von Neumann Theorem [voiculescu_non-commutative_1976, arveson_notes_1977], the trivial element of correspond to the class of trivial essential, unital extensions. The group of invertible elements of corresponds to the essential, unital extensions that are semi-split. Thus, by the Choi–Effros lifting theorem [choi_completely_1976], is a group when is nuclear. One can extend -homology to the category of all separable C*-algebras by setting
where is the suspension of and is the unitization of . This gives a cohomology theory on the category of separable C*-algebras, which can be recognized as the dual of -theory via Paschke duality [paschke_theory_1981, higson_algebra_1995, kaminker_spanier_2017]. Kasparov’s bivariant functor simultaneously generalizes -homology and -theory, where is recovered as and as .
It was already noticed in the seminal work of Brown, Douglas, and Fillmore [brown_extensions_1977, brown_extensions_1973] that the invariant for a compact metrizable space can be endowed with more structure than the purely algebraic group structure. Indeed, one can write as the inverse limit of a tower of compact polyhedra, and endow with the topology induced by the maps for , where is a countable group endowed with the discrete topology. This gives to the structure of a topological group, which is however in general not Hausdorff.
The study of as a topological group for a separable unital C*-algebra 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 -groups.) In [schochet_fine_2001, dadarlat_topology_2005] several natural topologies on , corresponding to different ways to define -homology for separable C*-algebras, are shown to coincide and to turn into a pseudo-Polish group. This means that, if denotes the closure of zero in , then the quotient of by is a Polish group. In [schochet_fine_2002], for a C*-algebra satisfying the Universal Coefficient Theorem (UCT), the topology on is related to the UCT exact sequence, and is shown to be isomorphic to the group classifying pure extensions of by . A characterization of for an arbitrary separable nuclear C*-algebra is obtained in [dadarlat_approximate_2000]. For a separable quasidiagonal C*-algebra satisfying the UCT, is shown to be the subgroup of corresponding to quasidiagonal extensions of by [schochet_fine_2002]; see also [brown_universal_1984] for the commutative case. The quotient of by is the group introduced by Rørdam [rordam_classification_1995]. A universal multicoefficient theorem describing in terms of the -groups of 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 turns out to be trivial, i.e. the closure of zero in is the whole group. For example, the topology on is trivial when is a UHF C*-algebra, despite the fact that is not trivial, and in fact uncountable. Similarly, for every -dimensional solenoid , the topology on is trivial, although is an uncountable group.
In this paper, we take a different approach and consider the group , rather than as a pseudo-Polish topological group, as a definable group. This should be thought of as a group explicitly defined as the quotient of a Polish space by a “well-behaved” equivalence relation , in such a way that the multiplication and inversion operations in are induced by a Borel functions on . 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 and 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 , can be regarded as a definable group. Furthermore, different descriptions of —in terms of extensions, Paschke duality, Fredholm modules, and quasi-homomorphisms—yield naturally definably isomorphic definable groups. For C*-algebras that have a -filtration in the sense of Schochet [schochet_uct_1996], we show that the definable subgroup of is definably isomorphic to . The latter is regarded as a definable group as in [bergfalk_ulam_2020, Section 7]. (In fact, 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 from [bergfalk_ulam_2020, Section 7] where is a torsion-free abelian group without finitely-generated direct summands, we prove that definable -homology provides a finer invariant than the purely algebraic (or topological) groups for a separable C*-algebra , even when one restricts to UHF C*-algebras or commutative unital C*-algebras whose spectrum is a -dimensional subspace of .
Theorem A.
The definable -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 -groups (and trivial -groups).
Theorem B.
The definable -group is a complete invariant for -dimensional solenoids up to homeomorphism. In contrast, there exists an uncountable family of pairwise non homeomorphic -dimensional solenoids with algebraically isomorphic -groups (and trivial -groups).
The historic evolution in the treatment of -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 , its unitary dual is the quotient of the Polish space of unitary irreducible representations of 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 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 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 corresponds to a bijection that is induced by a Borel function . In our terminology from Section 1.4, this corresponds to regarding a unitary dual as a definable set, where an isomorphism of Mackey Borel structures on is a definable bijection . 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 -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 -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 -homology for separable C*-algebras is introduced, and shown to be given by definable groups by considering its description in terms of the -theory of Pashcke dual algebras. The descriptions of -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 -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 -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 -theory, and the rigidity result for definable of torsion-free finite-rank abelian groups from [bergfalk_ulam_2020]. Finally, Section LABEL:Section:spaces considers definable -homology for compact metrizable spaces, and Theorem B is obtained applying again the definable UCT and the rigidity theorem for definable 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 is if and only if it is a Polish space when endowed with the subspace topology. If is a Polish space, then the Borel -algebra of is the -algebra generated by the collection of open sets. By definition, a subset of is Borel if it belongs to the Borel -algebra. If are Polish spaces, then the product is a Polish space when endowed with the product topology. More generally, if is a sequence of Polish spaces, then the product 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 the set of natural numbers including . We regard and any other countable set as a Polish space endowed with the discrete topology. The Baire space is the Polish space obtained as the infinite product of copies of .
A standard Borel space is a set endowed with a -algebra (the Borel -algebra) that comprises the Borel sets with respect to some Polish topology on . A function between standard Borel spaces is Borel if it is measurable with respect to the Borel -algebras. A subset of a standard Borel space is analytic if it is the image of a Borel function for some standard Borel space . This is equivalent to the assertion that there exists a Borel subset such that is the projection of on the first coordinate. A subset of is co-analytic if its complement is analytic. One has that a subset of is Borel if and only if it is both analytic and co-analytic.
Given standard Borel spaces , we let be their product endowed with the product Borel structure, which is also a standard Borel space. If is a sequence of standard Borel spaces, then their disjoint union is a standard Borel space, where a subset of is Borel if and only if is Borel for every . The product 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 be the category that has standard Borel spaces as objects and Borel functions and morphisms.
-
(1)
If is a standard Borel space and is a Borel subset, then is a standard Borel space when endowed with the induced standard Borel structure;
-
(2)
If are standard Borel spaces, is an injective Borel function, and is Borel, then is a Borel subset of ;
-
(3)
If are standard Borel spaces, and is a bijective Borel function, then the inverse function is Borel;
-
(4)
If are standard Borel spaces, and there exist injective Borel functions and , then there exists a Borel bijection ;
-
(5)
The category has finite products, finite coproducts, equalizers, and pullbacks;
-
(6)
A Borel function is monic in if and only if it is injective, and epic in if and only if it is surjective;
-
(7)
An inductive sequence of standard Borel spaces and injective Borel functions has a colimit in .
A Polish group is a topological group whose topology is Polish. If is a Polish group, and is a closed subgroup of , then is a Polish group when endowed with the subspace topology. If furthermore is normal, then is a Polish group when endowed with the quotient topology. If are Polish groups, and is a Borel function, then is continuous. In particular, if is a Polish space, then it has a unique Polish group topology that induces its Borel structure. A subgroup of a Polish group is Polishable if it is Borel and there is a (necessarily unique) Polish group topology on that induces the Borel structure on inherited from . This is equivalent to the assertion that is equal to the range of a continuous group homomorphisms for some Polish group . If is a Polish group, then a Polish -space is a Polish space endowed with a continuous action of . A Borel -space is a standard Borel space endowed with a Borel action of . Given a Borel -space , there exists a Polish topology on such that is a Polish -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 that is also a group, and such that the group operation on and the function , 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 endowed with a topology and a -valued metric such that:
-
(1)
the metric-topology is finer than ;
-
(2)
the metric is lower-semicontinuous with respect to , i.e. for every the set
is -closed in .
A Polish topometric space is a topometric space such that the topology is Polish and the metric is complete. A Polish topometric group is a Polish topometric space that is also a group, and such that endowed with the topology is a Polish group, and the metric on is bi-invariant
1.2. Polish categories
By definition, we let a Polish category be a category enriched over the category of Polish spaces (regarded as a monoidal category with respect to binary products). Thus, for each pair of objects of , is a Polish space, such that for objects , the composition operation is continuous.
Suppose that is a Polish category. For objects of , define be the set of -isomorphisms . While is not necessarily a subset of , and hence not necessarily a Polish space when endowed with the subspace topology, is endowed with a canonical Polish topology, defined as follows. For a net in and , set if and only if in and in . One can then easily show the following.
Lemma 1.2.
Adopt the notations above. Then is a Polish space.
It is clear from the definition that, for every object of , is a Polish group. Furthermore, the canonical (right and left) actions of and on are continuous.
Definition 1.3.
Suppose that and are Polish categories, and is a functor. We say that is continuous if, for every pair of objects of , the map , is continuous. We say that is a topological equivalence if it is continuous, and there exists a continuous functor such that is isomorphic to the identity functor , and is isomorphic to the identity functor .
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 and are Polish categories, and is a functor. The following assertions are equivalent:
-
(1)
is a topological equivalence;
-
(2)
each object of is isomorphic to one of the form for some object of , and for each pair of objects of , the map is a homeomorphism.
1.3. Idealistic equivalence relations
Suppose that is a set. A -filter on is a nonempty family of subsets of that is closed under countable intersections, and such that and if and then . The dual notion is the one of -ideal. Thus, a nonempty family of subsets of is a -ideal if it is closed under countable unions, , and and imply . Clearly, if is a -filter on , then is a -ideal on , and vice-versa. Thus, one can equivalently formulate notions in terms of -filters or in terms of -ideals.
If is a -filter on , then can be thought of as a notion of “largeness” for subsets of . Based on this interpretation, we use the “-filter quantifier” notation “, ” for a subset to express the fact that . If is a unary relation for elements of , “, ” is the assertion that the set of that satisfy belongs to .
Example 1.5.
Suppose that is a Polish space. A subset of 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 form a -ideal . The corresponding dual -filter is the -filter of comeager sets, which are the subsets of whose complement is meager.
Suppose that is a standard Borel space. We consider an equivalence relation on as a subset of , endowed with the product Borel structure. Consistently, we say that is Borel or analytic, respectively, if it is a Borel or analytic subset of . In the following, we will exclusively consider analytic equivalence relations, most of which will in fact be Borel. For an element of we let be its corresponding -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 in Definition 1.6 be the identity function of . In the following definition, for a subset of a product space and , we let be the corresponding vertical section.
Definition 1.6.
An equivalence relation on a standard Borel space is idealistic if there exist a Borel function satisfying for every , and a function that assigns to each -class a -filter of subsets of such that, for every Borel subset of , the set
is Borel.
Idealistic equivalence relations arise naturally as orbit equivalence relations of Polish group actions. Suppose that is a Polish group and is a Polish -space. Let be the corresponding orbit equivalence relation on , obtained by setting, for , if and only if there exists such that . Then is an idealistic equivalence relation, as witnessed by the identity function on and the function where if and only if , . (As in Example 1.5, denotes the -filter of comeager subsets of .) In particular, if is a Polish group, and is a Polishable subgroup of , then the coset equivalence relation of in is Borel and idealistic.
Suppose that is an equivalence relation on a standard Borel space . A Borel selector for is a Borel function such that, for , if and only if . If has a Borel selector, then 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 .
Definition 1.7.
A definable set is a pair where is a standard Borel space and is a Borel and idealistic equivalence relation on . We think of as a presentation of the quotient set . Consistently, we also write the definable set as . A subset of is Borel if there is an -invariant Borel subset of such that .
We now define the notion of morphism between definable sets. Let and be definable sets. A lift of a function is a function such that for every .
Definition 1.8.
Let and be definable sets. A function is Borel-definable if it has a lift 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 , whose morphisms are the definable functions. We regard a standard Borel space as a particular instance of definable set where and is the relation of equality on . This renders the category of standard Borel spaces a full subcategory of the category of definable sets.
If and are definable sets, then their product in is the definable set , being the equivalence relation on defined by setting if and only if and . (It is easy to see that is Borel and idealistic if both and 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 be the category that has definable as objects and definable functions as morphisms.
-
(1)
If is a definable and is a Borel subset, then is itself a definable set;
-
(2)
If are definable sets, is an injective definable function, and a Borel subset, then is a Borel subset of ;
-
(3)
If are definable sets, and is a bijective definable function, then the inverse function is definable;
-
(4)
If are definable sets, and there exist injective definable functions and , then there exists a definable bijection ;
-
(5)
The category has finite products, finite coproducts, equalizers, and pullbacks;
-
(6)
A definable function is monic in if and only if it is injective, and epic in if and only if it is surjective;
-
(7)
An inductive sequence of definable sets and injective definable functions has a colimit in .
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 where is a standard Borel space is an analytic equivalence relation on that is not Borel and idealistic, or has not yet been shown to be Borel and idealistic. In this case, we say that 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 and are semidefinable sets, is a subset and is a function, then is definable if it has a Borel lift , and is Borel if there is a Borel -invariant subset of such that . The category has semidefinable sets as objects and definable functions as morphisms. Notice that, in particular, an isomorphism from to in is a bijection such that both and the inverse function are definable.
Lemma 1.11.
Suppose that is a definable set, is a semidefinable set. If and are isomorphic in , then is a definable set.
Proof.
Suppose that the Borel function and the assignment witness that is idealistic. By assumption, there exists a bijection such that has a Borel lift , and has a Borel lift . For we have that if and only if , whence is Borel. We now show that is idealistic.
Define an assignment from -classes to -filters, by setting if and only if where . Consider also the Borel map . Then it is easy to verify that and the assignment witness that is idealistic. ∎
Lemma 1.12.
Suppose that and are semidefinable sets. Assume that there exists a definable bijection (which is not necessarily an isomorphism in ). If is Borel, then is Borel.
Proof.
By assumption is Borel, and is analytic. Furthermore, has a Borel lift . Since is a bijection, we have that, for ,
This shows that is co-analytic. As is also analytic, we have that 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 be a definable set, be semidefinable set such that is Borel, and be a definable function. If is injective, then the range of a Borel subset of . If is bijective, then the inverse function is definable.
Corollary 1.14.
Suppose that is a definable set, and is a semidefinable set. If is a definable bijection, then is a definable set and is an isomorphism in .
1.5. Definable groups
A definable group can be simply defined as a group in the category in the sense of [mac_lane_categories_1998, Section III.6]. Thus, a definable group is a definable set that is also a group, and such that the group operation is definable, and the function , is also definable. As in the case of sets, we regard a standard Borel group as a particular instance of definable group where is a standard Borel group and is the relation of equality on . Thus, standard Borel groups form a full subcategory of the category of definable groups.
Naturally, a semidefinable group will be a group in , i.e. a semidefinable set that is also a group, and such that the group operation is definable, and the function that maps every element to its inverse is definable.
Lemma 1.15.
If is a semidefinable group, then the equivalence relation is Borel if and only if the identity element of , which is the -class of some element of , is a Borel subset of .
Proof.
Clearly, if is Borel, then is Borel. Conversely, suppose that is Borel. If and are Borel lifts of the group operation in and of the function that maps each element to its inverse, respectively, then we have that if and only if . This shows that is Borel. ∎
Corollary 1.16.
Suppose that is a semidefinable group. If is the orbit equivalence relation of a Borel action of a Polish group on the standard Borel space , then is a definable group.
Proof.
By [becker_descriptive_1996, Theorem 5.2.1] one can assume that is a Polish -space, and is the orbit equivalence relation of a continuous -action on . By [gao_invariant_2009, Proposition 3.1.10], every -class is Borel. Therefore is Borel by Lemma 1.15. Furthermore, 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 is a Polish group and is a Borel Polishable subgroup. Let be the coset equivalence relation of in . The quotient group is the quotient of by the equivalence relation . Since is Polishable, is the orbit equivalence relation of a Borel action of a Polish group on . Thus, 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 be a Banach space. We denote by its unit ball. A seminorm on is bounded if . We say that is contractive if .
Definition 2.1.
A strict Banach space is a Banach space such that is endowed with a topology (called the strict topology) such that, for some sequence of contractive seminorms on , letting be the pseudometric on defined by
one has that:
-
(1)
is a complete metric that induces the strict topology on ;
-
(2)
contains a countable strictly dense subset;
-
(3)
for every .
Example 2.2.
Suppose that is a separable Banach space. Then is a strict Banach space where the strict topology on is the norm topology.
Example 2.3.
Suppose that is a separable Banach space, and is its Banach space dual. Then is a strict Banach space where the strict topology on is the weak*-topology.
Let be a seminormed space, and consider the cone of bounded seminorms on as a complete metric space, with respect to the metric defined by . For a subset , we let be the topology on generated by . We denote by Ball the set of contractive seminorms in . If , , and is a sequence in , then we say that:
-
•
is -Cauchy if for every there exists such that, for , ;
-
•
is -Cauchy if it is -Cauchy for every ;
-
•
is -complete if, for every sequence in Ball, if is -Cauchy, then is -convergent to some element of .
The following lemma is elementary.
Lemma 2.4.
Suppose that is a seminormed space. Let be a topology on . Assume that and are two sets of bounded seminorms on such that the topologies and on coincide with . Then, for a sequence in , is -Cauchy if and only if it is -Cauchy. In this case, we say that is -Cauchy. It follows that is -complete if and only if it is -complete. In this case, we say that is -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 such that is endowed with a topology (called the strict topology) such that, for some separable cone of bounded seminorms on , one has that:
-
(1)
the strict topology on is the -topology, and is strictly complete;
-
(2)
contains a countable strictly dense subset;
-
(3)
for every .
Proposition 2.6.
Suppose that is a strict Banach space. Then is a Polish topometric space when endowed with the strict topology and the norm-distance.
Proof.
By definition, the strict topology on is Polish. Since the strict topology is induced by bounded seminorms on , it is coarser than the norm topology. The function is strictly lower-semicontinuous, being the supremum of strictly continuous functions. Since the norm on is complete, the distance on Ball is complete. ∎
Suppose that is a strict Banach space. We extend the strict topology of Ball to any bounded subset of by declaring the function
to be a homeomorphism with respect to the strict topology, where
Then we have that addition and scalar multiplication on are strictly continuous on bounded sets, and the norm is strictly lower-semicontinuous on bounded sets. In particular is a strictly closed subspace of for . Notice that, if is a norm-closed subspace of such that is strictly closed in Ball, then is a strict Banach space with the induced norm and the induced strict topology.
Definition 2.7.
Let be a strict Banach space. The (standard) Borel structure on is defined by declaring a subset of to be Borel if and only if is Borel for every .
Notice that the Borel structure on is standard, as is Borel isomorphic to the disjoint union of the standard Borel spaces for .
Definition 2.8.
If and are strict Banach spaces. A bounded linear map is contractive if , and strict if it is strictly continuous on bounded sets. A bounded seminorm on 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 is a strict, bijective, and isometric linear map between strict Banach spaces, then the inverse is not necessarily strict, whence is not necessarily an isomorphism in the category of strict Banach spaces. Nonetheless, is a Borel isomorphism, as both and are standard Borel spaces.
Definition 2.9.
Let be a strict Banach space. Define to be the space of bounded, strict seminorms on .
Notice that is a closed subspace of the complete metric space . A sequence in is strictly convergent if and only if it is -Cauchy. A bounded linear map is strict if and only if for every .
Remark 2.10.
Suppose that is a strict Banach space, and is a separable cone of bounded, strict seminorms on that induces the strict topology on . One can consider the globally defined topology on , induced by all the seminorms in . This topology coincides with the strict topology on . However, it is not first countable on the whole of , unless is a separable Banach space and the strict topology is equal to the norm topology. Indeed, if the -topology on is first-countable, then 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, equals the norm topology. In particular, the norm-topology on is equal to the strict topology, and it has a countable dense subset. Hence the norm-topology on 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 and are strict Banach spaces, and . Suppose that is an function. Assume that there exists an sequence of strictly continuous function such that
Then 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 be a strict Banach space. If , then is bounded if and only if, for every , is a bounded subset of .
A natural way to obtain strict Banach spaces is via pairings.
Definition 2.13.
A Banach pairing is a bounded bilinear map , where are Banach spaces. Define the -topology to be the topology on generated by the cone of bounded seminorms for .
The following lemma is an immediate consequence of the definition of strict Banach space.
Lemma 2.14.
Suppose that is a Banach pairing. Assume that:
-
•
are norm-separable Banach spaces;
-
•
for every ,
-
•
is -complete;
-
•
has a countable -dense subset.
Then is a strict Banach space where the strict topology on is the -topology.
Suppose that is a norm-separable Banach space, and is a strict Banach spaces. A linear map is bounded if it maps bounded sets to bounded sets or, equivalently,
This defines a norm on the space of bounded linear maps . We also define the strict topology on Ball to be the topology of pointwise convergence in the strict topology of Ball. Then one can easily show the following.
Proposition 2.15.
Suppose that is a norm-separable Banach space, and is a strict Banach space. Then is a strict Banach space.
2.2. Strict C*-algebras
We now introduce the notion of strict C*-algebra. Given a C*-algebra , we let be the set of its self-adjoint elements. We also denote by the C*-algebra of matrices over , which can be identified with the tensor product . 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 such that, for every , is also a strict Banach space satisfying the following properties:
-
(1)
the *-operation and the multiplication operation on are strictly continuous on bounded sets;
-
(2)
the strict topology on is induced by the inclusion
where is endowed with the strict topology, and is endowed with the product topology.
Example 2.17.
Suppose that is a separable C*-algebra. Then we have that is a strict C*-algebra where, for every , is endowed with the norm-topology.
Suppose that is a strict C*-algebra. Then, for every , is also a strict C*-algebra. If is a strict C*-algebra, then we regard as a standard Borel space with respect to the standard Borel structure induced by the strict topology on as in Definition 2.7. We say that a subset of is Borel if it is Borel with respect to such a Borel structure. We have that the Borel structure on (as a strict C*-algebra) coincides with the product Borel structure.
Definition 2.18.
Suppose that is a strict unital C*-algebra. A strict ideal of is a norm-closed proper two-sided Borel ideal of that is also a strict C*-algebra, and such that the inclusion map is strict.
Remark 2.19.
In order for to be a strict ideal of , we do not require that be strictly closed in nor that the strict topology on be the subspace topology induced by the strict topology of .
Example 2.20.
Suppose that is a strict unital C*-algebra and is a norm-closed and norm-separable proper two-sided ideal of . Then is a strict ideal of .
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 , then we say that is strictly dense in if is dense in with respect to the strict topology.
It follows from the axioms of a strict C*-algebra that, if is a strict C*-algebra, and is a *-polynomial, then defines a function that is strictly continuous on bounded sets. In particular, the sets of normal, self-adjoint, and positive elements of norm at most are strictly closed in . If is a continuous function, then induces by continuous functional calculus and Lemma 2.11 a strictly continuous functions from the strictly closed set of -tuples of pairwise commuting self-adjoint elements in to . Similarly, if is a continuous function, then induces by continuous functional calculus and Lemma 2.11 a strictly continuous function from the strictly closed set of -tuples of pairwise commuting normal elements in to .
Suppose that is a strict C*-algebra. Let be the Borel set of normal elements of . For , the spectrum is a closed subset of . We consider the space of closed subsets of as a standard Borel space endowed with the Effros Borel structure [kechris_classical_1995, Section 12.C]. If is a standard Borel space and is a basis of open subsets of , then a function is Borel if and only if, for every , is Borel. The proof of the following lemma is standard; see [simon_operators_1995, Lemma 1.6].
Lemma 2.21.
Suppose that is a strict C*-algebra. The function , is Borel.
Proof.
It suffices to show that the map , is Borel. Observe that has a basis of open sets of the form where is a continuous function. For such a continuous function , we have that
which is closed in . This concludes the proof. ∎
Suppose that is a strict C*-algebra. Fix and consider the set
Then, for we have that is invertible, , and
It follows from Lemma 2.11 that the function , is strictly continuous.
More generally, suppose that is an open subset of , and is a holomorphic function. Suppose that and is such that . Then admits a Taylor expansion
that converges uniformly for [ahlfors_complex_1978, Chapter 5, Theorem 3 and Chapter 2, Theorem 2]. Fix and set
Then for ,
see [pedersen_analysis_1989, Lemma 4.1.11]. Furthermore, the function , is strictly continuous on by Lemma 2.11, where .
2.3. Multiplier algebras
Suppose that is a separable C*-algebra. A double centralizer for is a pair of bounded linear maps such that and for every . Let be the set of double centralizers for . Then is a C*-algebra with respect to the operations
and the norm
for and . The strict topology on is the topology of pointwise convergence, namely the topology induced by the seminorms
for .
An element can be identified with the multiplier defined by setting and for . This allows one to regard as an essential ideal of . (An ideal of a C*-algebra is essential if is zero or, equivalently, has nonzero intersection with every nonzero ideal of .) If is an approximate unit for [higson_analytic_2000, Definition 1.7.1] then, by definition, strictly converges to in . In particular, is strictly dense in .
If is a strictly Cauchy sequence in , in the sense that is -Cauchy for every , then setting
for defines a double centralizer that is the strict limit of in . For , one can identify with 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 be a separable C*-algebra. Then is a strict unital C*-algebra containing as a strictly dense essential strict ideal where, for every , the strict topology on is as described above, and is identified with .
Example 2.23.
When is the algebra of compact operators on a separable Hilbert space, then and the strict topology on is the strong-* topology [blackadar_operator_2006, Proposition I.8.6.3].
Example 2.24.
One can also regard as the dual space of the Banach space of trace-class operators. This turns into a strict Banach space, where the strict topology on is the weak* topology, which coincides with the weak operator topology [blackadar_operator_2006, Definition I.8.6.2]. As the identity map is strong-*–weak continuous, the strong-* topology and weak operator topology on define the same standard Borel structure on .
One can define as above the strict topology on the whole multiplier algebra to be the topology of pointwise convergence of double multipliers. However, this topology on is not first countable whenever is not unital; see Remark 2.10.
Suppose that is a separable C*-algebra, and is a compact metrizable space. One can then consider the separable C*-algebra of continuous functions . Let also be the C*-algebra of strictly continuous bounded functions . There is an obvious unital *-homomorphism , where acts on by pointwise multiplication. The unital *-homomorphism is in fact a *-isomorphism [akemann_multipliers_1973, Corollary 3.4]. We can thus identify with and regard it as a strict C*-algebra. Observe that, for , the function , is strictly continuous. We let be the C*-algebra of norm-continuous functions , which is a C*-subalgebra of .
Lemma 2.25.
Suppose that is a separable C*-algebra, and is a compact metrizable space. Then is a Borel subset of .
Proof.
Fix a compatible metric on , and a countable dense subset of . Clearly, it suffices to show that is a Borel subset of . Fix, for every , a finite cover of consisting of open sets of diameter less than , and fix elements for . We have that a strictly continuous function is norm-continuous if and only if, for every there exists such that, for every and , . Since is strictly closed and is strictly continuous, we have that is norm-continuous if and only if for every there exists such that, for every and for every , . This shows that the set of norm-continuous functions is Borel. ∎
Corollary 2.26.
Suppose that is a separable C*-algebra. Then the set of norm-continuous paths is a Borel subset of .
Suppose that and are separable C*-algebra. A morphism from to in the sense of [woronowicz_pseudospaces_1980, woronowicz_unbounded_1991, woronowicz_operator_1992, woronowicz_algebras_1995] is a *-homomorphism such that is norm-dense in . (This is called -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 to and strict unital *-homomorphisms ; see [lance_hilbert_1995, Proposition 2.1].
Lemma 2.27.
Let and be separable C*-algebra.
-
•
Suppose that is a strict unital *-homomorphism. Then is a morphism from to .
-
•
Conversely, if is a morphism from to , then extends to a unique strict unital *-homomorphism . If is injective, then is injective.
-
•
If is an approximate unit for , then a *-homomorphism is a morphism from to if and only if strictly converges to .
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 and is meaningful, and it gives a morphism .
Suppose that are separable C*-algebras. A *-homomorphism is quasi-unital [jensen_elements_1991, Definition 1.3.13] (also called strict [lance_hilbert_1995, page 49]) if there exists a projection , called the relative unit of , such that . One has the following generalization of Lemma 2.27; see [lance_hilbert_1995, Corollary 5.7].
Lemma 2.28.
Let and be separable C*-algebra.
-
•
Suppose that is a strict *-homomorphism. Then is a quasi-unital *-homomorphism from to with relative unit .
-
•
Conversely, if is a quasi-unital *-homomorphism from to with relative unit , then extends to a unique strict *-homomorphism with . If is injective, then is injective.
-
•
If is an approximate unit for , then a *-homomorphism is quasi-unital if and only if 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 and , the set Mor of strict unital *-homomorphisms is a Polish space, and composition of morphisms is a continuous function.
Following [woronowicz_algebras_1995] we consider as endowed with the topology of pointwise strict convergence. This is the subspace topology induced by regarding, as in Lemma 2.27, as a subspace of , where is the space of bounded linear maps from to . (Recall that, if is a Banach space and is a strict Banach space, then the space of bounded linear maps is a strict Banach space when is endowed with the topology of pointwise strict convergence; see Proposition 2.15.) As is a subset of , 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 are separable C*-algebras, then the space of isomorphisms 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 of *-isomorphisms 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 , let be the set of all continuous maps . This is endowed with a Polish topology called the compact-open topology, that has as subbasis of open sets the sets of the form
for a compact subset of and an open subset of . This turns the category of locally compact second countable Hausdorff spaces and continuous maps into a Polish category. We let be the set of homeomorphisms . The Polish topology induced on as in Lemma 1.2 was shown in [arens_topologies_1946, Theorem 5], where it is called the -topology, to have as subbasis of open sets the sets of then where are closed sets and at least one between and is compact. For a locally compact second countable Hausdorff space , let be its one-point compactification, obtained by adjoining a point at infinity . Each admits a unique extension to that fixes the point at infinity, in the sense that . By [arens_topologies_1946, Theorem 5], the assignment defines a homeomorphism from onto the closed subset of consisting of the homeomorphisms that fix the point at infinity.
Given a locally compact second countable Hausdorff space , we let be the separable C*-algebra of continuous complex-valued functions on vanishing at infinity. Its multiplier algebra is the algebra of bounded continuous complex-valued functions on . The unit ball of endowed with the strict topology can be identified with the space of continuous functions endowed with the compact-open topology. Every separable commutative C*-algebra is isomorphic to , where is the locally compact second countable Hausdorff space of nonzero homomorphisms (the spectrum of ).
A continuous map induces a strict unital *-homomorphism given by , . 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 , is a homeomorphism, where Mor is endowed as above with the compact-open topology and is endowed with the topology of pointwise strict convergence. Thus, by Lemma 1.4, the assignment 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 is a separable C*-algebra, and is a separable C*-subalgebra. Define then the essential commutant of in to be the C*-algebra
where is the commutator . Define the strict topology on Ball to be the topology generated by the seminorms
for and . If is a approximate unit for that is approximately central for [higson_analytic_2000, Definition 3.2.4], then converges strictly to in Ball.
We have that is strictly complete. Indeed, consider a strictly Cauchy sequence in Ball. Then we have that converges to some in the strict topology of . For every , the sequence is norm-Cauchy in , whence it norm-converges to some element of , which must be equal to . This shows that is the strict limit of in . For , we can identify with , where is the image of under the diagonal embedding . From the above remarks and Lemma 2.14 we thus obtain the following.
Proposition 2.29.
Let be a separable C*-algebra, and let be a separable C*-subalgebra. Let be the corresponding essential commutant. Then is a strict C*-algebra containing as a strictly dense essential strict ideal where, for every , is identified with , and is endowed with the strict topology described above.
Suppose now as above that is a separable C*-algebra, and is a separable C*-subalgebra. Let also be a closed two-sided ideal. Define the essential annihilator
which is a closed two-sided ideal of . The strict topology on is the topology generated by the seminorms
for and . A straightforward argument as above gives the following.
Proposition 2.30.
Let be a separable C*-algebra, let be a separable C*-subalgebra, and be a closed two-sided ideal. Let be the corresponding essential commutant, and be the essential annihilator. Then is a strict ideal of , where for every , is identified with and is endowed with the strict topology described above.
Example 2.31.
Suppose that is a separable unital C*-algebra, is a closed two-sided ideal of , and is a nondegenerate representation of that is ample, in the sense that . We regard as the multiplier algebra of . The Paschke dual as defined in [higson_analytic_2000, Definition 5.1.1] is the essential commutant of inside ; see also [paschke_theory_1981]. The relative dual algebra as defined in [higson_analytic_2000, Definition 5.3.2] is the strict ideal of .
2.5. Homotopy of projections
Suppose that is a strict unital C*-algebra. Recall that a strict ideal of is a proper norm-closed Borel two-sided ideal of that is also a strict C*-algebra and such that the inclusion map is a strict *-homomorphism.
Definition 2.32.
A strict (unital) C*-pair is a pair where is a strict (unital) C*-algebra and is a strict ideal of .
We regard strict unital C*-pairs as objects of a category, where a morphism from to is a strict unital *-homomorphism that maps to .
Every strict unital C*-pair determines a quotient unital C*-algebra . If and are two unital C*-algebras obtained in this way, then we say that a unital *-homomorphism 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 (which is not necessarily a *-homomorphism) such that for every . 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 is considered as the object of this category, we call it a unital C*-algebra with a strict cover, and denote it by , 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 between strict unital C*-pairs induces a definable unital *-homomorphism 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 is a strict unital C*-pair and , we write if . If and , then we set
We let be the identity element of and be the zero element of .
Suppose that is a strict unital C*-pair. A positive element of Ball is a projection if or, equivalently, is a projection in . Define the set to be the Borel set of projections in . The Borel structure on is induced by the Polish topology defined by declaring a net to converge to if and only if strictly in Ball and strictly in Ball. (Recall that the strict topology on might be different from the topology induced by the strict topology on .)
We also say that an element of Ball is a unitary if and or, equivalently, is a unitary in . We let be the Borel set of unitaries in . The Borel structure on is induced by the Polish topology defined by declaring a net to converge to if and only if strictly in , strictly in , and strictly in .
More generally, an element of is called a partial unitary if and is a projection or, equivalently, if is a partial unitary in as in [rordam_introduction_2000, 8.2.12]. We let be the Borel set of partial unitaries in . In a similar fashion one can define the Borel set PI of partial isometries in , consisting of those such that and are 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 -theory for C*-algebras; see [rordam_introduction_2000, higson_analytic_2000, blackadar_theory_1998, wegge-olsen_theory_1993].
Given elements of , subject to a certain relation , we say that an element satisfying a relation can be chosen in a Borel fashion (from ) if there is a Borel function that assign to each -tuple in satisfying an element in such that satisfies . In other words, the set of tuples such that satisfies and satisfies has a Borel uniformization [kechris_classical_1995, Section 18.A].
Suppose that is a unital C*-algebra. Let be the set of projections in . Two projections in are:
-
•
Murray–von Neumann equivalent if there exists such that and , in which case we write ;
-
•
unitary equivalent if there exists such that ;
-
•
homotopic if there is a norm-continuous path in with and .
Lemma 2.33.
Suppose that is a strict unital C*-pair, , and satisfies . Then one can choose in a Borel way a self-adjoint element such that .
Proof.
Consider , and observe that there exists such that and , which can be chosen in a Borel way by strict continuity of the continuous functional calculus. Hence, setting , we have that and . Thus, after replacing with and with , we can assume that .
Let be an holomorphic branch of the logarithm defined on . Considering the holomorphic functional calculus, one can define the element . As
is the uniformly convergent power series expansion in , we have that
In particular, . Define
Then we have that satisfies . ∎
Corollary 2.34.
Suppose that is a strict unital C*-pair, and are unitaries. Then there following assertions are equivalent:
-
(1)
there is a norm-continuous path from to in ;
-
(2)
there exists and such that .
Lemma 2.35.
Suppose that is a strict unital C*-pair, and are such that are projections, , and . Then one can choose in a Borel fashion from such that, setting one has that
and
where does not depend on and .
Proof.
Consider the unitary
Notice that satisfies
and
Consider the norm-continuous path of unitaries
for . Notice that the modulus of continuity of does not depend on and . Fix such that if satisfy , then
Thus, for we have that
By Lemma 2.33 we can choose in a Borel fashion such that . Thus
Consider now and the fact that
Thus
and
Thus by Lemma 2.33 one can choose in a Borel fashion such that
and hence
Proceeding recursively in this way, one can choose in a Borel fashion such that
Then we have that, setting ,
and
This concludes the proof. ∎
Lemma 2.36.
Suppose that is a strict unital C*-pair, and are projections such that . Then one can choose in a Borel fashion from such that, setting , one has that , where does not depend on and .
Proof.
As in the proof of [rordam_introduction_2000, Proposition 2.2.4], consider the norm-continuous path of projections for . Let also , and be the continuous function that is on and on . Then for is a norm-continuous path of projections from to . Notice that the uniform continuity moduli of and do not depend on and .
Thus, there exists (that does depend on and ) such that, for every such that , one has that . Thus, are projections (that depend in a Borel way from by strict continuity of the continuous functional calculus) such that for and is a norm-continuous path from to (whose modulus of continuity does not depend on and ) satisfying
for .
Thus, we can assume without loss of generality that for every . We now proceed as in the proof of [nest_excision_2017, Proposition 2.17]. Define
By definition, we have that . Notice that
and
This implies that
and
for . We have that
Thus, is invertible. Let be its polar decomposition, where is a unitary. Then we have that . Indeed,
Thus
for and in particular .
Notice that is a norm-continuous path, whose modulus of continuity does not depend on and . Therefore, there exists (which do not depend on and ) such that, whenever satisfy , we have . By Lemma 2.33 one can then choose in a Borel way such that, setting , then and hence . This concludes the proof. ∎
Lemma 2.37.
Suppose that is a strict unital C*-pair, and are projections that satisfy . Then one can choose in a Borel fashion from such that, setting , one has that , where does not depend on and .
Proof.
Consider the path
for . This is a norm-continuous path of projections in from to , whose modulus of continuity does not depend on and . Therefore, the conclusion follows from Lemma 2.36. ∎
Lemma 2.38.
Suppose that is a strict unital C*-pair, and are unitaries. Then one can choose in a Borel fashion from and such that , where does not depend on and .
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 -homology. Suppose that is a separable Hilbert space. We regard as the multiplier algebra of the C*-algebra of compact operators on . The corresponding strict topology on is the strong-* topology. Consistently, we consider as a standard Borel space with respect to the induced standard Borel structure. If is a separable Banach space, we consider as a strict Banach space, where is endowed with the topology of pointwise strong-* convergence. We denote by the unitary group of , which is a Polish group when endowed with the strong-* topology.
Suppose that is a separable unital C*-algebra, and is an operator system [paulsen_completely_2002, Chapter 2]. Let be a separable Hilbert space. Arveson’s Extension Theorem asserts that every contractive completely positive (ccp) map [brown_algebras_2008, Section 1.5] admits a contractive completely positive extension [paulsen_completely_2002, Theorem 7.5]. We observe now that can be chosen in a Borel way from . Notice that the space of contractive completely positive maps is closed (hence, compact) in endowed with the topology of pointwise weak* convergence.
Lemma 2.39.
Suppose that is a separable unital C*-algebra, and is an operator system. Let be a separable Hilbert space. Then there exists a Borel function , such that is an extension of .
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 are compact metrizable spaces, and is a Borel subset such that, for every , the vertical section
is a closed nonempty set. Then there exists a Borel function such that for every .
Using this selection theorem, Lemma 2.39 follows immediately from the Arveson Extension Theorem.
Proof of Lemma 2.39.
We consider as a compact metrizable space, endowed with the topology of pointwise weak* convergence. Consider the Borel set of pairs such that . Then by the Arveson Extension Theorem, the vertical sections of are nonempty, and clearly closed. Thus, by Lemma 2.40 there exists a Borel function such that for every . ∎
2.7. The Definable Stinespring Dilation Theorem
Suppose that is a separable unital C*-algebra, and is a separable Hilbert space. Stinespring’s Dilation Theorem asserts that, for every contractive completely positive map , there exists a linear map with and a nondegenerate representation of on such that for every . Notice that the set of nondegenerate representations of on is a subset of , whence Polish with the subspace topology, where is endowed with the strong-* topology. It follows from the proof of the Stinespring Dilation Theorem, where and are explicitly defined in terms of , that they can be chosen in a Borel way from ; see [blackadar_operator_2006, Theorem II.6.9.7]
Lemma 2.41.
Suppose that is a separable unital C*-algebra, and is a separable Hilbert space. Then there exists a Borel function , such that for and for every contractive completely positive map .
2.8. The Definable Voiculescu Theorem
Suppose that is a separable unital C*-algebra, and are two maps. If is a unitary operator, write if for every . If is an isometry, write if for every . A nondegenerate representation of on is ample if, for every , . Notice that the set ARep of ample representations of on is a subset of Ball. Similarly, the set of isometries is a subset of . A formulation of Voiculescu’s Theorem asserts that if is an ample representation, and is a unital completely positive (ucp) map, then there exists an isometry such that ; see [higson_analytic_2000, Theorem 3.4.3, Theorem 3.4.6, Theorem 3.4.7]. We will observe that one can select in a Borel fashion from and .
Lemma 2.42.
Let be a separable unital C*-algebra, and a separable Hilbert space. There exists a Borel function , such that .
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 where is finite-dimensional. The following can be seen a definable version of [higson_analytic_2000, Proposition 3.6.7]. Notice that the set of orthogonal projections is closed subset of . Let 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 of , and regard as a C*-subalgebra of . For every finite subset of and , there exists a Borel map , such that is orthogonal to and for .
One then uses Lemma 2.43 to establish Lemma 2.42 in the case of block-diagonal maps. Recall that is block-diagonal with respect to if is a sequence of pairwise orthogonal finite-rank projections such that and for every (where the convergence is in the strong-* topology). Consider the set of pairs such that is block-diagonal with respect to . The proof of [higson_analytic_2000, Lemma 3.5.2] shows the following.
Lemma 2.44.
There exists a Borel function , such that .
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 , such that .
As a consequence of the definable Voiculescu Theorem, one obtains the following; see [higson_analytic_2000, Theorem 3.4.6].
Lemma 2.46.
Let be a separable unital C*-algebra, and a separable Hilbert space. There exist:
-
•
a Borel map , such that ;
-
•
a Borel map , such that .
2.9. Spectrum
Suppose now that is a strict unital C*-algebra, and is a norm-separable closed two-sided ideal of . One can consider the quotient C*-algebra and, for , the spectrum of in . We also let the resolvent be the complement in of . The following lemma is analogous to [ando_weyl_2015, Theorem 3.16].
Lemma 2.47.
Suppose that is a strict C*-algebra, and a norm-separable closed two-sided ideal of . Suppose that every invertible self-adjoint element of lifts to an invertible self-adjoint element of . If , and is a countable dense subset of , then
Proof.
It suffices to prove that is the union of for . Clearly, for every , so it suffices to prove the other inclusion. Suppose that . We want to show that for some . After replacing with , it suffices to consider the case when . In this case, is invertible in . Therefore, by assumption there exists such that is invertible in . Since the set of invertible elements of is norm-open, there exists such that is invertible in , and hence . ∎
Lemma 2.48.
Suppose that is a strict unital C*-algebra, and a norm-separable closed two-sided ideal of . Suppose that every invertible self-adjoint element of lifts to an invertible self-adjoint element of . Then the function , is Borel.
Proof.
Fix a countable norm-dense subset of . Then by the previous lemma we have that, for ,
As the function , is Borel, this concludes the proof. ∎
Lemma 2.49.
The function , is Borel.
Proof.
An operator induces an invertible element of if and only if it is Fredholm. If is Fredholm and self-adjoint, then it has index , and is an isolated point of the spectrum of that is an eigenvalue with finite multiplicity. Thus, if is the finite-rank projection onto the eigenspace of for , then we have that is invertible and self-adjoint and induces the same element of as . This shows that every invertible self-adjoint element of lifts to an invertible self-adjoint element of . Therefore, the conclusion follows from Proposition 2.48. ∎
Suppose that is a projective. Recall that this means that is a positive operator satisfying . Then it is well-known that there exists a projection . We observe that one can choose such a in a Borel fashion from ; see [andruchow_note_2020, Lemma 3.1].
Lemma 2.50.
Consider the Borel set of projections in . Then there is a Borel function , such that for every .
Proof.
Suppose that . Observe . In particular, is countable, with only accumulation points and . From Lemma 2.49, the maps and are Borel. If then one can set . If , one can set .
Let us consider the case when . By [kechris_classical_1995, Theorem 12.13] there exists a Borel map , such that is an increasing enumeration of . One can then choose in a Borel way such that and then a continuous function such that
One can then set . ∎
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 is a standard Borel space, is a compact metrizable space, and is a Borel subset such that, for every , the vertical section
is a closed nonempty set. Then the assignment , , is Borel, where is endowed with the Effros Borel structure.
As an application, we obtain the following. Let be a separable Hilbert space. We consider the unit ball of as a compact metrizable space endowed with the weak topology. We also consider as a standard Borel space, endowed with the Effros Borel structure.
Lemma 2.52.
The function , , is Borel.
Proof.
By Lemma 2.51, it suffices to show that the set
is Borel. Fix a countable norm-dense subset of . Then we have that, if , then if and only if such that and . Since the norm on is weakly lower-semicontinuous, this shows that is Borel. ∎
Recall that, for an operator , one sets . By strong-* continuity on bounded sets of continuous functional calculus, the function is Borel. Furthermore, there exists a unique partial isometry with such that [pedersen_analysis_1989, Theorem 3.2.17]. The decomposition is then called the polar decomposition of .
Lemma 2.53.
The function , that assigns to an operator the partial isometry in the polar decomposition of is Borel.
Proof.
It suffices to notice that is graph, which is the set of pairs such that is a partial isometry with and , is Borel by Lemma 2.52. ∎
Consider the Borel set of unitaries in . Thus, if and only if . If is the partial isometry in the polar decomposition of , then and is an essential unitary. In fact, one can easily define (in a Borel fashion from ) an isometry or co-isometry such that . One has that is in particular a Fredholm operator. Its index is defined by
Thus, is a Borel function of .
More generally, consider the Borel set of pairs such that is a projection, and . If is the partial isometry in the polar decomposition of , then and the index of regarded as a Fredholm operator on is given by the Borel function
3. -theory of unital C*-algebras with a strict cover
In this section we explain how the and 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 and 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. -group
Suppose that is a unital C*-algebra with a strict cover. Recall that denotes the Polish space of projections in . Similarly, for we have that is a Polish space. We say that an element of for some is a projection over . We define Z be the set of pairs of projections over , which is the disjoint union of Z for endowed with the induced standard Borel structure. Two projections in are Murray–von Neumann equivalent (respectively, unitary equivalent, and homotopic) if and only if and are Murray–von Neumann equivalent (respectively, unitary equivalent, and homotopic) in .
The -group of —see [higson_analytic_2000, Chapter 4]—is defined as a quotient of by an equivalence relation , defined as follows. For , if and only if there exist and such that and are Murray–von Neumann equivalent . By Lemma 2.37, we have the following equivalent description of .
Lemma 3.1.
Suppose that is a unital C*-algebra with a strict cover, and where and . Then if and only if there exist and
such that, setting , one has that
where does not depend on and .
The (commutative) group operation on is induced by the Borel function on , . The neutral element of corresponds to . The function that maps an element to its additive inverse is induced by the Borel function on given by . Thus, is in fact a semidefinable group.
If and are unital C*-algebras with a strict cover, and is a definable unital *-homomorphism, then the induced group homomorphism K is also definable. Thus, the assignment gives a functor from the category of unital C*-algebras with a strict cover to the category of semidefinable abelian groups.
Suppose that is a strict unital C*-pair. We denote by the unitization of , which can be identified with the C*-subalgebra . Since is a proper ideal of , we can write every element of uniquely as where and . More generally, every element of can be written uniquely as where and . As the map , is a unital *-homomorphism, we have that and hence for .
We define to be the set of projections in , which we regard as a Borel subset of . Similarly, the unitary group of is regarded as a Borel subset of . Define also to be the Borel subset of consisting of pairs such that . Finally, let to be the disjoint union of for .
The -group of —see [higson_analytic_2000, Definition 4.2.1]—is defined as a quotient of by an equivalence relation , defined as follows. One has that, for , if and only if there exist and such that and are Murray–von Neumann equivalent. For , we let be the corresponding element of . The (commutative) group operation on is induced by the Borel function on , . The neutral element of corresponds to . The function that maps an element of to its additive inverse is induced by the Borel function on given by . Thus, is a semidefinable group.
If are are strict C*-pairs, and is a strict *-homomorphism, then it induces a strict *-homomorphism . In turn, this induces a definable group homomorphism . This gives a functor from strict unital C*-pairs to semidefinable groups.
Lemma 3.2.
Suppose that is a strict unital C*-pair. Then there is a Borel map , such that and .
Proof.
Suppose that . By definition, we have that for some and , and . Thus, we can define
and
Then we have that
This concludes the proof. ∎
3.2. Relative -group
Suppose now that is a strict unital C*-pair. For , define to be the Borel set of triples where are projections and satisfies and . Define to be the disjoint union of for endowed with the induced standard Borel structure. The elements of are called relative -cycles for ; see [higson_analytic_2000, Definition 4.3.1]. If , then we say that is a relative -cycle of dimension . A relative -cycle for is degenerate if and . Two relative -cycles and of dimension are homotopic if there exists a norm-continuous path of relative -cycles for of dimension with and .
Notice that if is a relative -cycle of dimension , and is a path of unitaries in starting at , then
and
are norm-continuous paths of relative -cycles starting at . If , then
is a norm-continuous path of relative cycles from to . We have the following lemma; see [nest_excision_2017, Proposition 3.4].
Lemma 3.3.
Suppose that is a relative cycle of dimension for . Then is homotopic to the degenerate cycle .
The relative -group is defined to be the quotient of by the equivalence relation defined as follows. For , set if and only if there exists a degenerate relative -cycles such that and are of the same dimension and homotopic. The group operations on are induced by the Borel maps
and
It follows from Lemma 3.3 that is indeed a group. We let be the element of represented by the relative -cycle . The trivial element of is equal to where is any degenerate relative -cycle. Let be the unitization of , which we identify with .
Lemma 3.4.
There is a Borel function , such that , , , and .
Proof.
Notice that is a degenerate relative -cycle of dimension . Consider then
By Lemma 2.38 one can choose in a Borel way such that, setting , one has that , where does not depend on and . Thus, after replacing with
we can assume without loss of generality that .
By Lemma 2.35 one can choose in a Borel fashion from such that, setting , one has that
Thus, after replacing with we can assume without loss of generality that and satisfy and hence .
In this case, we have that , since is a norm-continuous path of relative -cycles from to . This concludes the proof. ∎
Proposition 3.5.
Suppose that is a strict C*-pair. Then is a definable group. The assignment , is a natural definable isomorphism, called the excision isomorphism.
Proof.
By [nest_excision_2017, Theorem 3.9], the excision homomorphism is bijective; see also [higson_analytic_2000, Theorem 4.3.8]. Clearly, it is induced by a Borel function . By Lemma 3.4 the inverse homomorphism is also induced by a Borel function . Thus, is a definable group, and the excision isomorphism is a definable isomorphism. ∎
There is a natural definable homomorphism that is induced by the Borel map Z . We also have a natural definable homomorphism induced by the Borel map . We have the following result; see [higson_analytic_2000, Proposition 4.3.5].
Proposition 3.6.
Suppose that is a strict unital C*-pair. The (natural) sequence of definable groups and definable group homomorphisms
is exact.
Combining the excision isomorphism with the natural definable homomorphism , we obtain a natural definable group homomorphism . This is defined by mapping to regarded as an element of . Combining Proposition 3.5 with Proposition 3.6 we have the following.
Corollary 3.7.
Suppose that is a unital strict C*-algebra and is a proper strict ideal of . Then the natural sequence
is exact.
3.3. group
Suppose that is a strict unital C*-pair. We can then consider the Borel set of elements of that are unitaries . We then let to be the disjoint union of for . The equivalence relation B on is defined by setting B for and if and only if there exist with and such that there is a norm-continuous path from to in the unitary group of the quotient unital C*-algebra . This equivalence relation is analytic by Corollary 2.34. The definable -group is then the semidefinable group obtained as quotient with group operations defined as above. This defines a functor from unital C*-algebras with strict cover to semidefinable groups.
Given a strict unital C*-pair , we also consider the definable -group of . As above, we identify the unitization of with . For we let be the unitary group of . Recall that every element of can be written uniquely as where and . We consider as a Borel subset of . We then set to be the disjoint union of for , and let be the (analytic) equivalence relation on obtained by setting B for and if and only if there exist with and such that there is a norm-continuous path from to in . The definable -group is then the semidefinable group obtained as quotient with group operations defined as above.
Suppose that is a strict unital C*-pair. We have natural definable group homomorphisms
The definable group homomorphism is induced by the inclusion , which gives an inclusion for every . The definable group homomorphism is also induced by the inclusion maps for . 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 is a strict unital C*-pair. The sequence of natural definable homomorphisms
is exact.
3.4. The six-term exact sequence
Suppose that is a strict unital C*-pair. One can define a natural definable group homomorphism called the index map, as follows. An element of is of the form where for some . Then define
and
Then are projections such that and hence . One then defines ; see [higson_analytic_2000, Proposition 4.8.10]. As is obtained in a Borel fashion from , the boundary map is definable.
Equivalently, one can define as follows. Given an element of for some . Consider the partial isometry defined by
and observe that [rordam_introduction_2000, Lemma 9.2.1]. Then and are projections in such that . Therefore, . One has that ; 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 is a strict unital C*-algebra and is a strict ideal of . Then the sequence
is exact.
Suppose that is a strict unital C*-pair. One can consider a natural definable homomorphism called the exponential map. This is defined as follows. Consider an element of of the form for some . Then we have that and are unitary elements of such that . Then one has that ; 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 is a strict unital C*-pair. Then the sequence
is exact.
Suppose that is a strict unital C*-pair. Then as discussed above we have exact sequences
and
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
for the strict unital C*-pair , where the vertical arrows are the index map and the exponential map; see [rordam_introduction_2000, Theorem 12.1.2].
4. Definable -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 -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 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 coincides with the group Ext defined in terms of unital extensions. Using Paschke’s -theoretical description of Ext from [paschke_theory_1981], we show that yields a contravariant functor from separable unital C*-algebras to the category of definable groups.
We also recall the definition of the -homology groups of separable C*-algebras as in [higson_analytic_2000, Chapter 5]. Using their description in terms of , we conclude that they can be endowed with the structure of definable groups, in such a way that the assignments and 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 where is a separable C*-algebra and is a closed two-sided ideal of . A morphism between separable C*-pairs is a *-homomorphisms that maps to . Recall that a C*-algebra is nuclear if the identity map of 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 is nuclear if is nuclear. In this section, we will also introduce the relative definable -homology groups, and the six-term exact sequence in -homology associated with a separable nuclear C*-pair.
4.1. C*-algebra extensions and the Ext group
Let be a separable Hilbert space, and be the algebra of bounded linear operators on . We let be the closed ideal of compact operators, and be the Calkin algebra, which is the quotient of by . Let be the quotient map.
If is a unitary operator, then defines an automorphism Ad given by . As is -invariant, we have an induced automorphism of , still denoted by Ad.
Suppose that is a unital, separable C*-algebra. A unital extension of (by ) is a unital *-homomorphism . A unital extension of is injective or essential if it is an injective *-homomorphism . Two extensions are equivalent if there exists such that Ad. An injective, unital extension is semi-split (or weakly nuclear in the terminology of [elliott_abstract_2001]) if there exists a unital completely positive (ucp) map such that [higson_analytic_2000, Theorem 3.1.5]. An injective unital extension is split or trivial if there exists a unital *-homomorphism such that .
Every unital, essential extension determines an exact sequence
where
and is an essential ideal of . The extension is split if and only if the map is a split epimorphism in the category of unital C*-algebras and unital *-homomorphisms.
Conversely, given an exact sequence
where is a unital *-homomorphism and is an essential ideal of , one can define an essential unital extension as follows. Consider , then define for where is such that . Again, we have that is trivial if and only if is a split epimorphism.
Let be a separable, unital C*-algebra. One defines Ext to be the set of unitary equivalence classes of unital, injective extensions of by ; see [higson_analytic_2000, Definition 2.7.1], and Ext to be the subset of unitary equivalence classes of unital, injective semi-split (or weakly nuclear) extensions of by [blackadar_theory_1998, 15.7.2].
One can define a commutative monoid operation on . The (additively denoted) operation on Ext is induced by the map where 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 is the set of split extensions, which form a single unitary equivalence class [higson_analytic_2000, Theorem 3.4.7]. Furthermore, the set Ext is equal to the set of elements of Ext that have an additive inverse, whence it forms a group [higson_analytic_2000, Definition 2.7.6]. When 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 is semi-split, and Ext. In particular, in this case Ext is itself a group.
Let be a separable unital C*-algebra. We regard Ext as a definable group, as follows. Fix a separable Hilbert space . Let us say that a ucp map is ample if for every . Notice that the set of ample ucp maps is a subset of the space of bounded linear maps of norm at most endowed with the topology of pointwise strong-* convergence. Thus, is a Polish space.
Let be the Borel set of ample ucp maps such that
for . An injective, unital semi-split extension of by definition has a ucp lift, which is an element of , and conversely every element of gives rise to an injective, unital semi-split extension of . Thus, we can regard as the space of representatives of injective, unital semi-split extensions of . We define a Polish topology on that induces the Borel structure on by declaring a net in to converge to if and only if, for every , strong-* converges to , and norm-converges to .
Two elements of represent the same element of Ext if and only if there exists such that for every . This defines an analytic equivalence relation on . We can thus regard Ext as the semidefinable set .
We now observe that the group operations on Ext are definable, and thus this turns into a semidefinable group. We will later show in Proposition LABEL:Proposition:K-Ext that in fact is a definable group.
Proposition 4.1.
Let be a separable unital C*-algebra. The addition operation and the additive inverse operation on are definable functions. Thus, is a semidefinable group.
Proof.
Fix a representation such that . The assertion for addition is clear, as the Borel map , is a lift for the addition operation, where is a fixed surjective linear isometry .
In order to obtain a lift for the function , , one can use the definable Stinespring Dilation Theorem (Lemma 2.41). Thus, if , and and are the nondegenerate representation of and the isometry obtained from in a Borel fashion as in Lemma 2.41, then defining the projection and , , one has that represents , where as above is a fixed surjective linear isometry; see also [higson_analytic_2000, Theorem 3.4.7]. ∎
Suppose that are separable unital C*-algebras. A unital *-homomorphism induces a definable group homomorphism , as follows. If is a representative for an injective, unital extension, then one can consider defined by where is a fixed surjective linear isometry. This defines a Borel function , which induces a definable group homomorphism . Thus, is a contravariant functor from the category of separable unital C*-algebras to the category of semidefinable groups.
4.2. -group and the Voiculescu property
Let be a strict unital C*-algebra. Recall that two projections are Murray–von Neumann (MvN) equivalent if there exists such that and , in which case we write . We say that a projection is ample if is Murray–von Neumann equivalent to , and co-ample if is ample.