Three Viewpoints on Semi-Abelian Homology Julia Goedecke Emmanuel College Department of Pure Mathematics and Mathematical Statistics University of Cambridge A thesis submitted for the degree of Doctor of Philosophy August 2009 I would like to dedicate this thesis to my parents. Although it took them a while to accept that anyone might want to study something without any apparent application to industry, they have always supported me in whatever I wanted to do. I would also like to dedicate this thesis to my collaborator Tim Van der Linden, without whose constant help and encouragement I would probably have given up research long ago. ii Declaration of Originality This dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration except as detailed below and where specifically indicated in the text. Chapter 1 gives the background for the thesis and is known material. Chapter 2 contains more background material in the first two sections, but Section 2.3 presents my own work, generalising known results from the abelian context into the semi-abelian setting. Chapter 3 is based on joint work with Tim Van der Linden, published in the Journal of Homotopy and Related Structures in 2007. I contributed roughly 50% to these results. Chapter 4 again presents known material which is needed in Chapter 5. This chapter is based on another joint paper with Tim Van der Linden, which will be published in the Mathematical Proceedings of the Cambridge Philosophical Society. As before I contributed 50% to the paper; for the thesis I rewrote the results using the concept of an axiomatic class of extensions introduced by Tomas Everaert, which makes many of the statements and proofs easier. The results of Chapter 6 are my own work, developing a suggestion given by Marino Gran. iii iv Acknowledgements I would like to thank my supervisor Peter Johnstone for taking me on as his student even though I did not show much enthusiasm for working with toposes. A very big thank you goes to my collaborator Tim Van der Linden, who is a great pleasure to work with and who supported and advised me throughout my time as a PhD student, and will hopefully carry on doing so in the future. I am also indebted to George Janelidze for pointing Tim and me to his work on satellites as an alternative approach to homology. Marino Gran showed great interest in my work, and his suggestion that I should work on stem extensions resulted in Chapter 6. I am grateful to Tomas Everaert for helpful discussions and comments, and for providing the useful concept of an axiomatic class of extensions, which makes Chapter 5 so much easier. Furthermore I would like to thank Alexander Shannon, arguably the best proof- reader in the world, for finding (hopefully) all the little mistakes in my thesis (though I take full responsibility for any remaining ones). Thank you also to Martin Hyland, Alexander Frolkin and James Griffin for helpful comments and suggestions throughout my time as a PhD student, and to Ignacio Lopez Franco, Enrico Vitale and Burt Totaro for pointing out some mistakes in my examples which nobody else had picked up on. I would also like to thank Emmanuel College and the Department of Pure Mathematics and Mathematical Statistics for their help and support during my time at Cambridge, and the EPSRC for funding the research towards my PhD. All my family and friends have also supported me to get to this point through many ups and downs. In particular I would like to thank Marcus Zibrowius for managing to boost my work morale enough so that I could actually write this thesis. v vi Abstract The main theme of the thesis is to present and compare three different viewpoints on semi-abelian homology, resulting in three ways of defining and calculating homology objects. Any two of these three homology theories coincide whenever they are both de- fined, but having these different approaches available makes it possible to choose the most appropriate one in any given situation, and their respective strengths complement each other to give powerful homological tools. The oldest viewpoint, which is borrowed from the abelian context where it was intro- duced by Barr and Beck, is comonadic homology, generating projective simplicial resolu- tions in a functorial way. This concept only works in monadic semi-abelian categories, such as semi-abelian varieties, including the categories of groups and Lie algebras. Comonadic homology can be viewed not only as a functor in the first entry, giving homology of objects for a particular choice of coefficients, but also as a functor in the second variable, varying the coefficients themselves. As such it has certain universality properties which single it out amongst theories of a similar kind. This is well-known in the setting of abelian categories, but here we extend this result to our semi-abelian context. Fixing the choice of coefficients again, the question naturally arises of how the homol- ogy theory depends on the chosen comonad. Again it is well-known in the abelian case that the theory only depends on the projective class which the comonad generates. We extend this to the semi-abelian setting by proving a comparison theorem for simplicial resolutions. This leads to the result that any two projective simplicial resolutions, the definition of which requires slightly more care in the semi-abelian setting, give rise to the same homology. Thus again the homology theory only depends on the projective class. The second viewpoint uses Hopf formulae to define homology, and works in a non- monadic setting; it only requires a semi-abelian category with enough projectives. Even this slightly weaker setting leads to strong results such as a long exact homology sequence, the Everaert sequence, which is a generalised and extended version of the Stallings- Stammbach sequence known for groups. Hopf formulae use projective presentations of objects, and this is closer to the abelian philosophy of using any projective resolution, rather than a special functorial one generated by a comonad. To define higher Hopf for- mulae for the higher homology objects the use of categorical Galois theory is crucial. This theory allows a choice of Birkhoff subcategory to generate a class of central extensions, which play a big role not only in the definition via Hopf formulae but also in our third viewpoint. This final and new viewpoint we consider is homology via satellites or pointwise Kan extensions. This makes the universal properties of the homology objects apparent, giving a useful new tool in dealing with statements about homology. The driving motivation behind this point of view is the Everaert sequence mentioned above. Janelidze’s theory of generalised satellites enables us to use the universal properties of the Everaert sequence to interpret homology as a pointwise Kan extension, or limit. In the first instance, this allows us to calculate homology step by step, and it removes the need for projective objects from the definition. Furthermore, we show that homology is the limit of the diagram consisting of the kernels of all central extensions of a given object, which forges a strong connection between homology and cohomology. When enough projectives are available, we can interpret homology as calculating fixed points of endomorphisms of a given projective presentation. vii viii Contents Introduction 1 1 The Semi-Abelian Context 7 1.1 Semi-abelian categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1.2 Abelian objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 1.3 Homology in semi-abelian categories . . . . . . . . . . . . . . . . . . . . . . 16 1.4 Categorical Galois theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 1.5 Central extensions in the context of abelianisation . . . . . . . . . . . . . . 24 2 Comonadic Homology 31 2.1 Simplicial objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.2 Comonadic Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.3 Hn(−, E)G as a functor in the variable E . . . . . . . . . . . . . . . . . . . 38 3 A Comparison Theorem for Simplicial Resolutions 47 3.1 Simplicial resolutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.2 The Comparison Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 3.3 Comonads generating the same projective class . . . . . . . . . . . . . . . . 53 4 Homology via Hopf Formulae 61 4.1 Extensions and higher extensions . . . . . . . . . . . . . . . . . . . . . . . . 62 4.2 Strongly (E-)Birkhoff subcategories . . . . . . . . . . . . . . . . . . . . . . . 66 4.3 The Galois structures Γn, centralisation and trivialisation . . . . . . . . . . 68 4.4 Hopf formulae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 5 Homology via Satellites 81 5.1 Satellites and pointwise satellites . . . . . . . . . . . . . . . . . . . . . . . . 82 5.2 Hn+1(−, I)E as a satellite of Hn(−, I1)E1 . . . . . . . . . . . . . . . . . . . . 85 5.3 Hn+1(−, I)E as a satellite of In . . . . . . . . . . . . . . . . . . . . . . . . . 88 5.4 Hn(−, I)E as a right satellite of Hn(−, I1)E1 . . . . . . . . . . . . . . . . . . 90 5.5 Homology without projectives . . . . . . . . . . . . . . . . . . . . . . . . . . 91 5.6 Homology with projectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 6 Stem Extensions in the Context of Abelianisation 111 6.1 Stem extensions and stem covers . . . . . . . . . . . . . . . . . . . . . . . . 111 6.2 Perfect objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 6.3 A natural isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 References 123 Index 127 ix x Introduction The main theme of this thesis is to present and compare three different viewpoints on semi-abelian homology. As a motivation for the newest of these viewpoints, consider a perfect group A. We know that it has a universal central extension, and the kernel of this universal central extension is isomorphic to the second integral homology group H2(A,Z). It would be nice if this kind of property could be extended to the homology of all groups, not only perfect ones. One of the main results in this thesis makes this possible: it shows that the homology H2(A,Z) is the limit of the diagram of kernels of all central extensions of A. This is a natural generalisation of the perfect case, as the limit of a diagram with an initial object is just this initial object. Similarly the higher homology groups form limits of diagrams using kernels of higher central extensions. This result emphasises the close connection of homology with central extensions, and thus makes a connection to cohomology. Semi-abelian categories The results of this thesis are not limited to the category of groups, but take place in the more general setting of semi-abelian categories. Classically, homological algebra is an area which is studied in the context of abelian categories. Many of the diagram lemmas used in homological algebra, such as the Five Lemma and the Snake Lemma, are proved in this context, but they also hold in the category of groups and other settings close to it, such as Lie algebras or crossed modules. Semi-abelian categories give a wider context for homolog- ical algebra which includes these non-abelian examples as well as the traditional abelian categories. Many of the homological algebra results which hold in abelian categories also hold in semi-abelian ones, though not quite all, as can be seen most easily in the category of groups. For example, products and coproducts do not coincide in general semi-abelian categories, and not every monomorphism is a kernel, as for instance not every subgroup is a normal subgroup. We use semi-abelian categories as a framework to study homology in a non-abelian context. Three viewpoints on semi-abelian homology The oldest viewpoint on semi-abelian homology, which is borrowed from the abelian con- text where it was introduced by Barr and Beck, is comonadic homology, generating pro- jective simplicial resolutions in a functorial way. This concept only works in monadic semi-abelian categories, such as semi-abelian varieties, which include the categories of groups and Lie algebras. The second viewpoint uses Hopf formulae to define homology, 1 Three Viewpoints on Semi-Abelian Homology and works in a non-monadic setting; it only requires a semi-abelian category with enough projectives. Even this slightly weaker setting leads to strong results such as a long exact homology sequence. Hopf formulae use projective presentations of objects, and this is closer to the abelian philosophy of using any projective resolution, rather than a special functorial one generated by a comonad. The final viewpoint we consider in this thesis is homology via satellites or pointwise Kan extensions. This makes the universal properties of the homology objects apparent, giving a useful new tool in dealing with statements about homology. Any two of these three definitions of homology coincide whenever they are both defined; but having these different viewpoints available makes it possible to choose the most appropriate one in any given situation, and their respective strengths comple- ment each other to give powerful homological tools. There is another viewpoint we do not consider in this thesis, which uses kernel pairs and Galois groupoids. This viewpoint introduced by Janelidze in [Jan2008] is perhaps closest to our second viewpoint using Hopf formulae. We now describe our three viewpoints in slightly more detail. Viewpoint 1: Comonadic homology Comonadic homology can be viewed not only as a functor in the first entry, giving homol- ogy of objects for a particular choice of coefficients, but also as a functor in the second variable, varying the coefficients themselves. As such it has certain universality properties which single it out amongst theories of a similar kind. This is well-known in the setting of abelian categories, but here we extend this result to our semi-abelian context. Fixing the choice of coefficients again, the question naturally arises of how the homol- ogy theory depends on the chosen comonad. Again it is well-known in the abelian case that the theory only depends on the projective class which the comonad generates. We extend this to the semi-abelian setting by proving a comparison theorem for simplicial resolutions. This leads to the result that any two projective simplicial resolutions, the definition of which requires slightly more care in the semi-abelian setting, give rise to the same homology. Thus again the homology theory only depends on the projective class. Viewpoint 2: Hopf formulae The Hopf formula H2(A,Z) ∼= [P, P ] ∩K[p][K[p], P ] for the second homology group is very well known in the context of integral group homol- ogy. Here p : P −→ A is a projective presentation of the group A with kernel K[p], and [P, P ] and [K[p], P ] are commutator subgroups of P . This formula can be generalised both to a semi-abelian setting and also to higher Hopf formulae, which are then used to define higher homology objects. A good way to summarise these higher Hopf formulae is to say that homology measures the difference between the centralisation and the trivialisation 2 Introduction of a projective presentation of a given object. As hinted above, using this definition any short exact sequence 0 ,2 K  ,2 ,2 B f  ,2 A ,2 0 gives rise to a long exact homology sequence · · · ,2 Hn+1(A, I) δn+1f ,2 K[Hn(f, I1)] γnf ,2 Hn(B, I) Hn(f,I) ,2 Hn(A, I) ,2 · · · · · · ,2 H2(A, I) δ2f ,2 K[H1(f, I1)] γ1f ,2 H1(B, I) H1(f,I) ,2 H1(A, I) ,2 0 which we call the Everaert sequence. This sequence has a different appearance than its abelian counterpart: instead of being functorial in the objects of the short exact sequence, it is functorial in the extension f . It incorporates not only homology of objects, such as Hn(A, I), but also homology of an extension Hn(f, I1), which is a higher-dimensional version of the same theory. The lowest part of the Everaert sequence is the Stallings- Stammbach sequence, which first appeared in the context of groups, but now it can be extended to a full long exact sequence. Its universal properties play a very important role in a large part of this thesis. Birkhoff subcategories and abelian objects The main ingredient of all approaches to semi-abelian homology is a Birkhoff subcategory B of a semi-abelian category A, as the reflector I takes the role of coefficients of the homology theory. Thus semi-abelian homology calculates the derived functors of this reflector, using different approaches to do this. The most common examples such as integral group homology or homology of Lie algebras use the subcategory of abelian objects as the Birkhoff subcategory, which makes it easy to see that all homology objects are abelian. But there are also many examples where the Birkhoff subcategory is not given by the abelian objects, for instance the categories of nilpotent or solvable groups inside the category of groups, the category of Lie algebras inside that of Leibniz algebras, or the category of crossed modules inside that of precrossed modules. In these cases it can still be shown that all homology objects from the second onwards are abelian, but as the first homology just gives the reflection of an object into the Birkhoff subcategory, these will not in general be abelian objects. In the Everaert sequence, this means that at the lower end the objects stop being abelian. But the map δ2f from the last abelian object to the first non-abelian object turns out to be central in the sense of Huq, meaning that its image commutes with everything in its codomain. This nicely connects the abelian part of the sequence with the non-abelian part, and does not seem to have been realised before. 3 Three Viewpoints on Semi-Abelian Homology Viewpoint 3: Homology via satellites The universal properties of the Everaert sequence are the driving motivation for defining homology via satellites, as a pointwise Kan extension or limit. The connecting homo- morphism δ in the Everaert sequence is exactly what makes the Kan extension work. In the first instance, this allows us to calculate homology step by step, and it removes the dependance on projective objects from the definition. For example, the (n+1)st homology Hn+1(−, I) is the left satellite of the nth homology of extensions Hn(−, I1). ExtA cod z    Hn(−,I1) $? ?? ?? ?? A δn+1 +3 Hn+1(−,I) $ ExtA kerz    A For n = 1 this can be reformulated to give the result hinted at in the first paragraph of this introduction: homology is the limit of a diagram of kernels of all central extensions of a given object. H2(A, I) = lim f∈CExtAA K[f ] That is, we consider all central extensions of an object A, and the maps between them. Taking kernels gives us a diagram in the category we are working in, and the second homology H2(A, I) is the limit of this diagram. K[f ]  &- &- B f TTTT TT  &-T TTTT H2(A, I) 40, and de- generacy operators σi : An −→ An+1, for i ∈ [n] and n ∈ N, subject to the simplicial identities ∂i◦∂j = ∂j−1◦∂i if i < j σi◦σj = σj+1◦σi if i ≤ j ∂i◦σj =  σj−1◦∂i if i < j 1 if i = j or i = j + 1 σj◦∂i−1 if i > j + 1. 31 Chapter 2. Comonadic Homology In a semi-abelian category A, we can define the homology of a simplicial object A via the Moore complex of A. 2.1.2 Definition (Moore complex): Let A be a simplicial object in a semi-abelian category A. The Moore complex or normalised chain complex N(A) has as objects N0A = A0, N−nA = 0 and NnA = n−1⋂ i=0 K[∂i : An −→ An−1] = K[(∂i)i∈[n−1] : An −→ Ann−1], for n ≥ 1, and boundary maps dn = ∂n◦ ⋂ iKer ∂i : NnA −→ Nn−1A for n ≥ 1. This gives rise to a functor N: SA −→ ChA from the category of simplicial objects in A to the category of chain complexes in A, called the normalisation functor. The object of n-cycles is ZnA = K[dn] = ⋂n i=0K[∂i : An −→ An−1] for n ≥ 1. We write Z0A = A0. The Moore complex of a simplicial object is always a proper chain complex [EVdL2004b, Theorem 3.6]; thus we can define HnA = HnN(A) for a simplicial object A. In the abelian case, the homology of the Moore complex is the same as the homology of the unnormalised chain complex C(A) of A, where CnA = An and dn = ∂0 − ∂1 + · · ·+ (−1)n∂n. Notice that the Moore complex and thus the homology of a simplicial object only involve the face maps ∂i, and not the degeneracies σi. So we need only consider semi- simplicial maps between simplicial objects, i.e. maps that commute with the ∂i but not necessarily with the σi. This will be used in Chapter 3. The homology objects obtained this way are special objects of A. 2.1.3 Lemma: [EVdL2004b, Theorem 5.5] Let A be a simplicial object in a semi-abelian category A. For any n ≥ 1, the object HnA is an abelian object of A. An important property of the normalisation functor is the following: 2.1.4 Lemma: [EVdL2004b, Proposition 5.6] Let A be a semi-abelian category. The Moore normalisation functor N: SA −→ ChA is exact. Proof. A slightly easier proof than that in [EVdL2004b] can be found in [Eve2007]. This, together with the Snake Lemma, implies the following result. 32 2.1 Simplicial objects 2.1.5 Lemma: [EVdL2004b, Corollary 5.7] A short exact sequence of simplicial objects 0 ,2 A ,2 B ,2 C ,2 0 in A gives rise to a long exact sequence · · · ,2 HnA ,2 HnB ,2 HnC rykkkk kkkk kkkk kkkk k Hn−1A ,2 · · · ,2 H0C ,2 0 of homology objects which depends naturally on the given short exact sequence. Later we will be considering simplicial resolutions, which are really augmented simpli- cial objects. 2.1.6 Definition: An augmented simplicial object is a simplicial object A together with a map ∂0 : A0 −→ A−1 satisfying ∂0◦∂0 = ∂0◦∂1, that is, it has equal composite with the two face maps A1 −→ A0. A contraction of an augmented simplicial object A is a family of maps hn : An −→ An+1, for n ≥ −1, which satisfy ∂0hn = 1An and ∂ihn = hn−1∂i−1 for i > 0. A simplicial object that admits a contraction is called contractible. · · · ,2 ,2,2,2 A2 h2 T] ,2,2,2 A1 h1 U^ ,2,2 A0 h0 U^ ,2 A−1 h−1 U_ When computing the homology of a simplicial object, the following observation is often useful. 2.1.7 Lemma: [EVdL2004b, Proposition 3.9] (cf. [Bou2001]) Given a diagram A ∂1 ,2 ∂0 ,2 B e ,2 C where e◦∂0 = e◦∂1, suppose there is a common splitting t : B −→ A of ∂0 and ∂1, that is, ∂0◦t = 1B = ∂1◦t. Then e is the coequaliser of ∂0 and ∂1 if and only if e is the cokernel of ∂1◦Ker ∂0. A consequence of this result is: 2.1.8 Lemma: [EVdL2004b, Proposition 3.11] A contractible augmented simplicial object A has H0A = A−1 and HnA = 0 for n ≥ 1. 33 Chapter 2. Comonadic Homology A simplicial object in the category of sets is commonly called a simplicial set. A classical property simplicial sets may have is the Kan property. Kan simplicial sets are exactly the fibrant ones (in the usual model structure on SSet) and may be described as follows. 2.1.9 Definition (Kan property): Let S be a simplicial set and n ≥ 1, k ∈ [n] natural numbers. An (n, k)-horn in S is a sequence (si)i∈[n]\{k} of elements of Sn−1 satisfying ∂i(sj) = ∂j−1(si) for all i < j and i, j 6= k. A filler of an (n, k)-horn (si)i is an element s of Sn satisfying ∂i(s) = si for all i 6= k. A simplicial set S is Kan when every horn in S has a filler. This Kan property can be generalised to simplicial objects in a regular category as follows (see [CKP1993]). 2.1.10 Definition (Internal Kan property): Let A be a simplicial object in a regular category A and n ≥ 1, k ∈ [n] natural numbers. An (n, k)-horn in A is a family of maps (bi : B −→ An−1)i∈[n]\{k} satisfying ∂ibj = ∂j−1bi for all i < j and i, j 6= k; we can view this as a map b : B −→ (An−1)n. A filler of an (n, k)-horn b : B −→ (An−1)n is given by a surjection p : Z −→ B and a generalised element z : Z −→ An satisfying ∂iz = bip for i 6= k. This can be viewed as a filler “up to enlargement of domain”. Carboni, Kelly and Pedicchio show in [CKP1993] that every simplicial object of a regular category A is Kan if and only if A is Mal’tsev. Thus when A is regular Mal’tsev, for example semi-abelian, we can apply the internal Kan property for every simplicial object in A. This property is very powerful and can be used in many situations. To demonstrate this, we present a small new result involving the maps used in the Moore complex. Recall that the object NnA is the kernel of the map (∂0, . . . , ∂n−1) : An −→ (An−1)n. 2.1.11 Proposition: Let A be a simplicial object in a semi-abelian category A. Let In denote the image of the map (∂0, . . . , ∂n−1) : An −→ (An−1)n. Then the following is an equaliser diagram: In ,2 (An−1)n (∂0pi1,∂0pi2,...,∂0pin−1,∂1pi2,...,∂n−2pin−1) ,2 (∂0pi0,∂0pi1,...,∂0pin−2,∂1pi1,...,∂n−2pin−2) ,2 (An−2)k where the two parallel maps are constructed to give all the horn conditions ∂ipij = ∂j−1pii for i < j and j ≤ n− 1; so k = 12(n− 1)(n− 2). Proof. The equaliser of these two maps is an (n, n)-horn through which all other (n, n)- horns factor. So we have to show that all (n, n)-horns factor through In. 34 2.1 Simplicial objects Let b : B −→ (An−1)n be an (n, n)-horn. As in a regular Mal’tsev category every simplicial object is Kan [CKP1993, Theorem 4.2], this horn must have a filler (p : Z −→ B, z : Z −→ An). This gives the following diagram: Z z  p  ,2 z′ $> >> >> >> >> B  b %,SSS SSS (An−1)n ,2 ,2 (An−2)k An  ,2 In 2: i 2:lllll Since p is the coequaliser of some pair of maps, and iz′ = bp has equal composite with this pair and i is monic, z′ factors through p and we get a factorisation of the horn B through In as desired. This fact can be used to prove that the normalisation functor is exact, a result which Tim Van der Linden and Tomas Everaert prove in a different way in [EVdL2004b] (see Lemma 2.1.4). 2.1.12 Lemma: Given a short exact sequence of simplicial objects 0 ,2 A f ,2 B g ,2 C ,2 0 in a semi-abelian category A, the induced sequence of chain complexes 0 ,2 NA ,2 NB ,2 NC ,2 0 is also exact. Proof. As mentioned above, this is the same as Lemma 2.1.4, but we give an alternate proof here using Proposition 2.1.11. Given a short exact sequence of simplicial objects as above, we must show that 0 ,2 NnA ,2 NnB ,2 NnC ,2 0 is exact in A for each n ≥ 0. For n = 0 we have N0A = A0, so the result is clear. For n ≥ 1 we use that NnA is the kernel of (∂i)i∈[n−1] : An −→ Ann−1 35 Chapter 2. Comonadic Homology where Ann−1 is the n-fold product of An−1. Then Proposition 2.1.11 tells us that the image IAn of this map is an equaliser. This implies that the image sequence 0 ,2 IAn ,2 I B n ,2 ICn ,2 0 is also exact: we have 0 ,2 An  ,2 fn ,2 _  Bn gn  ,2 _  Cn ,2 _  0 0 ,2 IAn ,2 f ,2   'G GG GG G IBn g  ,2   ICn ,2   0 K[g] 7 7A 7Awwwww v v vv v ]g 0 ,2 Ann−1  ,2 fnn−1 ,2   Bnn−1 gnn−1 ,2   Cnn−1   0 ,2 Akn−2  ,2 fkn−2 ,2 Bkn−2 ,2 Ckn−2 where the first row is exact and the columns are image factorisations. Also, as kernels commute with products, fnn−1 is the kernel of gnn−1. Clearly f factors over the kernel K[g] of g. The kernel property of Ann−1 also induces a map K[g] −→ Ann−1, which has equal composite with the two maps to Akn−2, as fkn−2 is a monomorphism. So this map factors over the equaliser IAn and by the universal properties we see that I A n ∼= K[g]. Now we can use the 3× 3 Lemma: 0  0  0  0 ,2 NnA ,2_   NnB ,2_   NnC_   ,2 0 0 ,2 An  ,2 ,2 _  Bn  ,2 _  Cn ,2 _  0 0 ,2 IAn  ,2 ,2  IBn  ,2  ICn ,2  0 0 0 0 Here all columns and the last two rows are exact, so the first row is also exact, as desired. 36 2.2 Comonadic Homology 2.2 Comonadic Homology When A is a semi-abelian monadic category (e.g., a semi-abelian variety; see [GR2004] for a precise characterisation), there is a canonical forgetful/free comonad G = (G, , δ) on A, which gives rise to a functorial simplicial resolution GA of any object A, that is, an augmented simplicial object over A with face maps ∂i = GiGn−iA : Gn+1A −→ GnA and degeneracies σi = GiδGn−iA : Gn+1A −→ Gn+2A. · · · ,2,2,2,2 G 3A GGA ,2 G2A ,2 G2A ,2 G 2A GA ,2 GA ,2 GA A ,2 A This gives rise to the following Barr-Beck style [BB1969] notion of homology: 2.2.1 Definition: [EVdL2004b] LetB be a Birkhoff subcategory of a semi-abelian monadic category A with reflector I : A −→ B and canonical comonad G. For any object A of A and any n ≥ 0, we define Hn+1(A, I)G = HnNIGA. (A) In fact, comonadic homology can be defined in a more general context than that of a semi-abelian category with a Birkhoff subcategory. 2.2.2 Definition (Comonadic homology): [EVdL2004b] Let C be any category with a comonad G = (G : C −→ C, δ : G =⇒ G2,  : G =⇒ 1C) and let E : C −→ A be a functor to a semi-abelian category A. For n ≥ 1, the object Hn(A,E)G = Hn−1NEGA is called the nth homology object of A (with coefficients in E) relative to the comonad G. This defines a functor Hn(−, E)G : C −→ A, for every n ≥ 1. The dimension shift here is not present in Barr and Beck’s original definition, but was introduced in [EVdL2004b] to make it better adjusted to the non-abelian examples (homology of groups, Lie algebras, crossed modules) which traditionally have a shifted numbering. When E is a contravariant functor, we get comonadic cohomology in a similar way. 2.2.3 Example (Comonads and functors of coefficients): The most common exam- ple of a comonad used for comonadic homology is that of a forgetful/free comonad on a variety of algebras, such as the free group comonad on the category of groups. Using 37 Chapter 2. Comonadic Homology this comonad, the abelianisation functor E = ab: Gp −→ Ab gives rise to integral group homology. The forgetful/free comonad on the category R-Mod of R-modules gives rise to two well-known homology theories. When E = N⊗R− : R-Mod −→ Ab for a fixed module N , we get the Tor groups as homology, that is Hn(M,N ⊗R −)G = TorRn (M,N). The contravariant functor E = HomR(−, N) : R-Mod −→ Ab gives the Ext groups: Hn(M,HomR(−, N))G = ExtnR(M,N) However, if we take the covariant Hom-functor HomR(N,−), we obtain the Eckmann- Hilton homotopy groups (see [BB1969, Example 1.1]). Notice that when taking comonadic homology of the covariant Hom-functor, we are still using projective (in fact free) reso- lutions, and not injective resolutions, which is why we don’t get the Ext groups in this case. There are also other comonads on the category of R-modules: given a ring homomor- phism φ : S −→ R, we can view any R-module as an S-module, and any S-module can be turned into an R-module by tensoring it with R over S. This adjunction gives rise to another comonad on R-Mod. R-Mod U !)LL LLL LLL LL Gφ ,2 R-Mod S-Mod R⊗S(−) 5=rrrrrrrrrr Using this so-called relative comonad, the functors given by tensoring and homing as above give rise to Hochschild’s relative Tor and Ext groups. This comonad will be used in Example 3.3.14. Many more examples of comonads and functors of coefficients exist, see for exam- ple [BB1969]. 2.3 Hn(−, E)G as a functor in the variable E Let G = (G, δ, ) be a comonad on the category C, and E : C −→ A a functor into a semi-abelian category A. We can view the homology functor Hn(−, )G as a functor [C,A] −→ [C,A], taking E to Hn(−, E)G. As such it has certain universal properties, which we discuss in this section. The results in this section are straightforward generalisations from the abelian case discussed in [BB1969], and in most parts the proofs carry over without much change. We still give them here in our own notation for completeness. 38 2.3 Hn(−, E)G as a functor in the variable E 2.3.1 Proposition (G-acyclicity): For n ≥ 1, we have Hn(−, EG)G = 0, and for n = 0 the map λ : H0(−, EG)G ∼= ,2 EG is an isomorphism. Proof. For any object X ∈ C, the augmented simplicial object EGGX is contractible, as we have EGnδX : EGGnX −→ EGGn+1X for n ≥ 0, which satisfies EGGnX◦EGnδX = 1EGGnX and EGGiGn−iX◦EGnδX = EGn−1δX◦EGGiGn−i−1X . Thus using Lemma 2.1.8 the result follows. We define a G-exact sequence to be a sequence of functors 0 −→ E1 −→ E2 −→ E3 −→ 0 such that 0 −→ E1G −→ E2G −→ E3G −→ 0 is exact. We then get 2.3.2 Proposition (G-connectedness): Any G-exact sequence 0 −→ E1 −→ E2 −→ E3 −→ 0 gives rise to a long exact sequence on homology: · · · ,2 Hn(−, E1)G ,2 Hn(−, E2)G ,2 Hn(−, E3)G ∂ ovfffff fffff fffff fffff fff Hn−1(−, E1)G ,2 · · · ,2 H0(−, E3)G ,2 0 Proof. For any object X ∈ C the G-exact sequence gives rise to a short exact sequence of simplicial objects 0 −→ E1GX −→ E2GX −→ E3GX −→ 0 which in turn gives rise to the desired long exact sequence on homology using Lemma 2.1.5. Analogously to [BB1969], we can define a theory of G-left derived functors by the above properties: 39 Chapter 2. Comonadic Homology 2.3.3 Definition: L = (Ln, λ, ∂) is a theory of G-left derived functors if the following are satisfied: (1) Each Ln is a functor Ln : [C,A] −→ [C,A]. (2) λ : L0 −→ 1[C,A] is a natural transformation from L0 to the identity functor on the functor category [C,A]. (3) (G-acyclicity) For a functor of the form EG we have λ : L0(EG) ∼=−→ EG is an isomorphism, Ln(EG) = 0 for n ≥ 1. (4) (G-connectedness) Every G-exact sequence 0 −→ E1 −→ E2 −→ E3 −→ 0 gives rise to a long exact sequence: · · · ,2 Ln(E1) ,2 Ln(E2) ,2 Ln(E3) ∂ qxiiii iiii iiii iiii iii Ln−1(E1) ,2 · · · ,2 L0(E3) ,2 0 where ∂ depends of the given sequence, and LnE3 ∂ ,2  Ln−1E1  LnF3 ∂ ,2 Ln−1F1 commutes for any map of sequences 0 ,2 E1 ,2  E2 ,2  E3 ,2  0 0 ,2 F1 ,2 F2 ,2 F3 ,2 0. We will show that the homology functor above is special amongst these theories of G-left derived functors. To do this, we first prove a result which is needed in the proof of the next theorem. 2.3.4 Lemma: The following is a coequaliser diagram: L0(EG2) L0(EG) ,2 L0(EG) ,2 L0(EG) L0(E) ,2 L0E 40 2.3 Hn(−, E)G as a functor in the variable E Proof. Notice that this diagram is the image under L0 of the lowest part of the augmented simplicial object EG. Similarly to Lemma 2.1.7, it is enough to show that L0(E) is the cokernel of L0(d1) : L0(N1(EG)) −→ L0(EG), which is the composite L0(N1(EG)) L0(Ker (EG)) ,2 L0(EG2) L0(EG) ,2 L0(EG). For this let M be the kernel of E : EG −→ E, and form the following diagram, where the bottom row is G-exact (as EG : EG2 −→ EG is split epic and so regular epic): N1(EG) ν v d1  0 ,2M i ,2 EG E ,2 E ,2 0 The morphism ν is induced by the kernel property of i. We now apply L0 to this diagram and get L0(N1(EG)) L0(ν) t|ppp ppp ppp pp L0(d1)  L0(M) L0(i) ,2 L0(EG) L0(E) ,2 L0(E) ,2 0. Now the bottom row is exact, as the bottom row of the previous diagram was G-exact. Note that we do not require L0 to preserve 0, but we still have L0(E)L0(d1) = 0, as the exact sequence 0 ,2 N1EG ,2 EG2 EG ,2 EG ,2 0 gives rise to the diagram 0 ,2 L0(N1EG) ,2 L0(d1) %,SSS SSS SSS SSS SS L0(EG2) L0(EG) ,2 L0(EG)  L0(EG) ,2 L0(E)  0 L0(EG) L0(E) ,2 L0(E) where the top row is short exact by G-connectedness and G-acyclicity. Thus if we denote the images of L0(i) and L0(d1) by I and I ′ respectively, we get a factorisation I ′ −→ I, which is of course monic. If we can show it is regular epic as well, we have shown that L0(d1) followed by L0(E) is also exact. For this it is sufficient to show that L0(ν) is a regular epi. Let K = K[ν]; then it is sufficient to show that 0 ,2 KG ,2 N1(EGG) νG ,2MG ,2 0 41 Chapter 2. Comonadic Homology is exact, as then G-connectedness implies that L0(K) ,2 L0(N1(EG)) L0(ν) ,2 L0(M) ,2 0 is exact. So we only need to prove that νG is regular epic. In the following diagram both rows are exact, and the left downwards square and right upwards square commute: 0 ,2 N1(EGG) νG  j ,2 EG3 EGG  EG2 ,2 EG2 ,2 0 0 ,2MG h LR iG ,2 EG2 EGδ LR EG ,2 EG ,2 Eδ LR 0 The dotted arrow h is induced by EG2◦EGδ◦iG = 0. Then iGνGh = EGG◦j◦h = EGG◦EGδ◦iG = iG so as iG is monic, νG is split epic, so in particular regular epic. Thus L0(E) is the cokernel of L0(d1), as asserted. This implies that L0(E) is the coequaliser of L0(EG) = h0 and L0(EG) = h1 (we rename them for convenience), in an analogous way to Lemma 2.1.7. Let k = L0(KerEG). In the following diagram, the outer rectangle is a pushout, and both squares commute. L0(N1EG)  L0(KerEG)=k,2 L0(EG2) L0(EG)=h0  L0(EG)=h1 ,2 L0(EG) L0(E)  f  0 ,2 L0(EG) L0(E) ,2 f '. L0(E) D To show that L0(E) is the required coequaliser, it is enough to show that if fh0 = fh1, then fh1k = 0. But this is clear as h0k = 0. Thus f factors through L0(E). We can now prove the following Uniqueness Theorem: 2.3.5 Theorem: Let L be a theory of G-left derived functors. Then there exists a unique family of natural isomorphisms Ln σn ∼= ,2 Hn(−, )G 42 2.3 Hn(−, E)G as a functor in the variable E for n ≥ 0, (i.e. natural isomorphisms LnE −→ Hn(−, E)G which are also natural in E), which are compatible with the augmentation λ and the connecting homomorphism ∂: L0E σ0 ,2 λ &C CC CC CC C H0(−, E)G λ uttt ttt ttt t E LnE3 σn  ∂ ,2 Ln−1E1 σn−1  Hn(−, E3)G ∂ ,2 Hn−1(−, E1)G Proof. The proof from [BB1969] for the abelian case carries over almost completely to the semi-abelian case, but we give it in full in our own notation for convenience. We start by constructing σ0. L0(EG2) L0(EG) ,2 L0(EG) ,2 λ∼=  L0(EG) λ∼=  L0(E) ,2 L0(E) σ0  EG2 EG ,2 EG ,2 EG ,2 H0(−, E)G From Lemma 2.3.4 we know that the top row of this diagram is a coequaliser, as is the bottom row (by definition of homology and Lemma 2.1.7). Thus we get an induced map σ0, which is an isomorphism, since both λ occurring in the diagram are isomorphisms (by the acyclicity property). It is clear that σ0 also commutes with the augmentations λ to E, since L0(E) is a regular epi. Now we construct the other σn inductively. From the construction of σ0 it is clear that it is also natural in E. We take M = K[E : EG −→ E], so that 0 ,2M i ,2 EG E ,2 E ,2 0 is G-exact. For n = 1 we then have L1(EG) = 0 ,2 L1E σ1  ∂L ,2 L0M σ0∼=  L0i ,2 L0(EG) σ0∼=  H1(−, EG)G = 0 ,2 H1(−, E)G ∂H ,2 H0(−,M)G H0(−,i)G ,2 H0(−, EG)G where both rows are exact, by G-connectedness. This induces the map σ1, which is also an isomorphism (by uniqueness of limits, or the Five Lemma). 43 Chapter 2. Comonadic Homology For n > 1 we use the diagram Ln(EG) = 0 ,2 LnE σn  ∂L ∼= ,2 Ln−1M σn−1∼=  ,2 0 = Ln−1(EG) Hn(−, EG)G = 0 ,2 Hn(−, E)G ∂H∼= ,2 Hn−1(−,M)G ,2 0 = Hn−1(−, EG)G Again both rows are exact, so σn = ∂−1H σn−1∂L is also an isomorphism. This defines the natural isomorphisms σn for all n. Now we must verify compatibility with the connecting homomorphisms. Let 0 ,2E1 α ,2E2 β ,2E3 ,20 be G-exact. Let K be the kernel of the composite γ : E2G E2 ,2 E2 β ,2 E3 , and M3 the kernel of E3 as before. Then we get a map of G-exact sequences 0 ,2 K j ,2 κ  E2G γ ,2 βG  E3 ,2 0 0 ,2M3 i3 ,2 E3G E3 ,2 E3 ,2 0 (B) where κ is induced by the kernel property of i3. This gives rise to a diagram 0 ,2 L1E3  ∂′L ,2 ∂L &-UUU UUUU UUUU U L0M3 σ0∼=  ,2 L0(E3G) σ0∼=  L0K L0κ 18iiiiiiiiiiii σ0∼=  0 ,2 H1(−, E3)G ∂′H ,2 ∂H &-TTT TTTT H0(−,M3)G ,2 H0(−, E3G)G H0(−,K)G H0(−,κ)G 18jjjjjjj Here both rows are exact, and the triangles commute by naturality of the connecting homomorphism. The rightmost square and the front right square commute by naturality of σ0. Now since ∂H is the kernel of H0(−, j)G (as H1(−, E2G)G = 0), we get a factorisation L1E3 −→ H1(−, E3)G making the front left square commute. This morphism also makes the back rectangle commute. But that rectangle defines σ1 uniquely, so the front left square must commute with σ1 substituted in. Now letKβ be the kernel of β, then we also have the following map ofG-exact sequences 0 ,2 K j ,2 κ2  E2G γ ,2 E2  E3 ,2 0 0 ,2 Kβ ,2 E2 β ,2 E3 ,2 0 (C) 44 2.3 Hn(−, E)G as a functor in the variable E with the obvious induced κ2. We also have an induced morphism of G-exact sequences 0 ,2 E1 ,2  E2 ,2 E3 ,2 0 0 ,2 Kβ ,2 E2 ,2 E3 ,2 0 which implies that Ln(Kβ) ∼= Ln(E1) and Hn(−,Kβ) ∼= Hn(−, E1) for n ≥ 0, using G- connectedness and the Five Lemma. The map ofG-exact sequences (C) induces the following prism, where we can substitute E1 for Kβ using the above isomorphisms: L1E3 σ1  ∂L ,2 ∂′L %, SSSS SSSS S L0K σ0  L0(κ2)rykk kkkk kkk L0E1 σ0  H1(−, E3)G ∂H ,2 ∂′H %,SS SSS S H0(−,K)G H0(−,κ2)Grzlll lll H0(−, E1)G The triangles again commute by naturality of the connecting homomorphisms, the right front square commutes by naturality of σ0, and the back square is the square whose commutativity we have shown above. Thus the front left square also commutes. The case n ≥ 2 works similarly, again using diagram (B) to get 0 ,2 LnE3 σn  ∂′L ∼= ,2 ∂L ∼= &-UUUU UUUU UUUU Ln−1M3 σn−1∼=  ,2 0 Ln−1K 07hhhhhhhhhhhh σn−1∼=  0 ,2 Hn(−, E3)G ∂′H ∼= ,2 ∂H ∼= &-UUU UUUU U Hn−1(−,M3)G ,2 0 Hn−1(−,K)G 07hhhhhhhh This time it is clear that the front left square commutes with σn substituted in, as ev- erything in sight is an isomorphism (by G-connectedness and G-acyclicity). Then we can again use (C) to get a similar prism to the above, which proves that LnE3 σn  ∂L ,2 Ln−1E1 σn−1  Hn(−, E3)G ∂H ,2 Hn−1(−, E1)G commutes. 45 Chapter 2. Comonadic Homology Thus we have shown that the comonadic homology theories are “the only” such theories of G-left derived functors, up to isomorphism. 46 Chapter 3 A Comparison Theorem for Simplicial Resolutions Introduction In Section 2.3 we studied comonadic homology as a functor in the second variable. This chapter addresses a different question: how does the comonadic homology theory depend on the given comonad G? Barr and Beck showed in [BB1969] that, in the abelian setting, two comonads G and K give rise to the same comonadic homology theory when they generate the same class of projective objects. In this chapter, we prove a suitable extension of their result to the semi-abelian case. That is, we show that given any category C with two comonads G and K, and any functor E : C −→ A to a semi-abelian category A, the functors Hn(−, E)G and Hn(−, E)K : C −→ A are isomorphic for n ∈ N when G and K generate the same Kan projective class. The condition on the projective class ensures that homming from a (G- or K-)projective object into a (G- or K-)simplicial resolution gives a Kan simplicial set. It is exactly the condition needed to prove a comparison theorem for these simplicial resolutions. The examples reveal that this condition is not too strong. In an additive category, homming from any object into any simplicial object results in a Kan simplicial set, and when C is a regular Mal’tsev category, which includes semi-abelian ones, then the condition is fulfilled as soon as the G-projective objects are also regular projectives. For their proof in the abelian case, Barr and Beck make heavy use of additive structure via a free additive completion of the category C. As a semi-abelian category is only additive when it is abelian, we have to use a different approach to extend their result to the semi- abelian setting. We prove a comparison theorem which shows that, given a projective class P, any two P-resolutions are homotopy equivalent and consequently have the same homology. The advantage of our method is that it shows that any P-resolution of a given object will give the same homology, so that it is also possible to use resolutions not coming from a comonad, should this turn out to be more convenient. The only subtlety lies in the definition of a P-resolution of an object A: this is an augmented simplicial object A = (An)n≥−1 where A−1 = A, all other An ∈ P, and for any object P ∈ P the augmented simplicial set Hom(P,A) is Kan and contractible. This is the reason for our condition 47 Chapter 3. A Comparison Theorem for Simplicial Resolutions on the projective class stated above: we must make sure that a G-resolution is also a P-resolution. Section 3.1 sets the scene by explaining the definition of a P-resolution in detail. Sec- tion 3.2 is devoted to the Comparison Theorem 3.2.3: if P is a simplicial object over B with each Pi ∈ P, and A is a simplicial object over A such that all augmented simplicial sets Hom(Pi,A) are contractible and Kan, then any map f : B −→ A extends to a semi- simplicial map f : P −→ A, and any two such extensions are simplicially homotopic. In this section we also relate our comparison theorem to that of Tierney and Vogel [TV1969], which uses a different definition of resolution, in a category with finite limits. The Comparison Theorem is used in Section 3.3 to prove the main result of this chapter, Theorem 3.3.11: under the condition on C mentioned above, any two comonads G and K that generate the same class of projectives induce isomorphic homology theories. We obtain it as an immediate consequence of Corollary 3.3.10 which states that, in a semi- abelian category, simplicially homotopic maps have the same homology: if f ' g then, for any n ∈ N, Hnf = Hng. All results in this chapter are joint work with Tim Van der Linden and also appear in our paper [GVdL2007]. 3.1 Simplicial resolutions To obtain the comonadic homology of a given object, we need to consider simplicial reso- lutions relative to a chosen class of projectives. Here we recall the definition of a projective class and give some examples. 3.1.1 Definition (Projective class): Let C be a category, P an object and e : B −→ A a morphism of C. Then P is called e-projective, and e is called P -epic, if the induced map Hom(P, e) = e◦(·) : Hom(P,B) −→ Hom(P,A) is a surjection. That is, for every map P −→ A, there is a (not necessarily unique) map making the following diagram commute: P y B e ,2 A Let P be a class of objects of C. A morphism e is called P-epic if it is P -epic for every P ∈ P; the class of all P-epimorphisms is denoted by P-epi. Let E be a class of morphisms in C. An object P is called E-projective if it is e-projective for every e in E; the class of 48 3.1 Simplicial resolutions E-projective objects is denoted E-proj. C is said to have enough E-projectives if for every object Y there is a morphism P −→ Y in E with P in E-proj. A projective class on C is a pair (P,E), P a class of objects of C, E a class of morphisms of C, such that P = E-proj, P-epi = E and C has enough E-projectives. Since, given a projective class (P,E), P and E determine each other, we will sometimes abusively write the projective class P or the projective class E. It is easy to see that any retract of a projective object is also projective, as is any coproduct of projectives. 3.1.2 Example: If E is the class of regular epimorphisms, P is called the class of regular projectives. In a variety, the class of regular projectives is generated by the free objects, hence there are enough projectives. The regular projectives in a variety C are also generated by the values of the canonical comonad C, induced by the forgetful functor to Set. More generally, any comonad on a category C generates a projective class: 3.1.3 Definition (Projective class generated by a comonad): Let G = (G, , δ) be a comonad on a category C. An object P in C is called G-projective if it is in the projective class (PG,EG) generated by the objects of the form GA. A map in EG is called a G-epimorphism. The A-component A : GA −→ A of the counit  is always a G-epimorphism. Indeed, any map f : GB −→ A factors over A as Gf◦δB, because A◦Gf◦δB = f◦GB◦δB = f . It is now clear that C has enough projectives of this class, since for any A we have A : GA −→ A. This definition coincides with the definition of G-projectives in [BB1969]. There a G-projective object is an object P which admits a map s : P −→ GP such that P s = 1P . Indeed, if P ∈ P, then the identity on P factors over the P-epimorphism P , which gives the splitting s. A simplicial resolution or a simplicial object in a semi-abelian category A gives rise to simplicial sets, for example by homing into the simplicial object from a fixed object of the category. As we saw in Section 2.1, an important classical property simplicial sets may satisfy is the Kan property defined in 2.1.9. We need the simplicial objects in the category C to satisfy a similar property, but relative to a chosen projective class P on C. For the purposes of this chapter, we will call this the relative Kan property. 3.1.4 Definition (relative Kan property): A simplicial object A is Kan (relative to P) when for every object P ∈ P the simplicial set Hom(P,A) is Kan. 49 Chapter 3. A Comparison Theorem for Simplicial Resolutions 3.1.5 Example: If C is regular with enough regular projectives and P the induced pro- jective class, saying that A is Kan relative to P is the same as saying that the simplicial object A is Kan, in the internal sense of Definition 2.1.10. Every simplicial object of C has this Kan property if and only if C is a Mal’tsev category [CKP1993, Theorem 4.2]. Thus when C is semi-abelian, every simplicial object is Kan with respect to the class of regular projectives. Note, however, that C need not have enough projectives for the internal Kan condition to make sense. The projective objects in the definition of the relative Kan property given here may be replaced by an enlargement of domain as in Definition 2.1.10. In case there are enough projectives, of course the two notions do coincide. 3.1.6 Example: It is well known that the underlying simplicial set of a simplicial group is always Kan. This may be seen as a consequence of the previous example, because the category Gp is a Mal’tsev variety and the forgetful functor U : Gp −→ Set is represented by the group of integers Z. Since C is an arbitrary category (without any extra structure) and A is just semi- abelian (rather than abelian), we have to be careful when considering simplicial resolutions of objects of C. Definition 3.1.7 seems to suit our purposes. 3.1.7 Definition (Simplicial resolution): Let P be a projective class. A P-resolution of A is an augmented simplicial object A = (An)n≥−1 with A−1 = A, where An ∈ P for n ≥ 0, and for every object P ∈ P the augmented simplicial set Hom(P,A) is Kan and contractible. In this chapter, we focus on simplicial resolutions in a category C which are generated by a comonad G on C. For any G-projective object P , the simplicial set Hom(P,GA) is contractible: choose a splitting s for P : GP −→ P ; given a map f : P −→ Gn+1A, define hn(f) = Gf◦s : P −→ Gn+2A. The morphisms hn : Hom(P,Gn+1A) −→ Hom(P,Gn+2A) then satisfy ∂0hn = 1Hom(P,Gn+1A) and ∂ihn = hn−1∂i−1 for i > 0. Thus they give a contraction of the simplicial set Hom(P,GA). Later we assume that the category C and the projective class P generated by G are such that GA is Kan relative to P for any object A, so that GA is a P-resolution of A. In the case when C is a category with finite limits, there exists another definition of simplicial resolution, using simplicial kernels. We give the definition of simplicial kernels here so that we can relate our Comparison Theorem of the next section to that of Tierney and Vogel [TV1969]. 50 3.1 Simplicial resolutions 3.1.8 Definition (Simplicial kernels): [TV1969] Let (fi : B −→ A)0≤i≤n be a sequence of n+ 1 morphisms in the category C. A simplicial kernel of (f0, . . . , fn) is a sequence (ki : K −→ B)0≤i≤n+1 of n+ 2 morphisms in C satisfying fikj = fj−1ki for 0 ≤ i < j ≤ n+ 1, which is universal with respect to this property. In other words, it is the limit for a certain diagram in C. For example, the simplicial kernel of one map is just its kernel pair. If C has finite limits, simplicial kernels always exist. We can then factor any augmented simplicial object through its simplicial kernels as follows: · · · ,2,2,2,2 %B BB BB BB B A2 ,2 ,2 ,2 %B BB BB BB B A1 ,2 ,2 %B BB BB BB B A0 ,2 A−1 K3 9C|||||||| 9C|||||||| 9C|||||||| 9C|||||||| K2 9C|||||||| 9C|||||||| 9C|||||||| K1 9C|||||||| 9C|||||||| Here the Kn+1 are the simplicial kernels of the maps (∂i)i : An −→ An−1. This gives a definition of P-exact simplicial objects: 3.1.9 Definition: [TV1969] Let P be a projective class. An augmented simplicial object A = (An)n≥−1 is called P-exact when the comparison maps to the simplicial kernels and the map ∂0 : A0 −→ A−1 are P-epimorphisms. 3.1.10 Remark: It can be shown that for any P-exact simplicial object A, the simplicial set Hom(P,A) is contractible for any P ∈ P. We will call this property of A relative contractibility. A resolution in the Tierney-Vogel sense is then a P-exact augmented simplicial object A in which all objects An for n ≥ 0 are in the projective class P. For their definition they need the presence of simplicial kernels, so they have to assume for example that the category C has finite limits. In our definition all assumptions are on the comonad G or rather the induced projective class P, and not on the category C. In the next section we will make clear the connections between our definition and theirs. 51 Chapter 3. A Comparison Theorem for Simplicial Resolutions 3.2 The Comparison Theorem Let P be a projective class on C. 3.2.1 Lemma: Let P ∈ P, and let A be an augmented simplicial object for which the aug- mented simplicial set Hom(P,A) is contractible and Kan. Let n ≥ 0. Given a sequence of maps (ai : P −→ An−1)i∈[n] satisfying ∂iaj = ∂j−1ai for i < j, there is a map a : P −→ An with ∂ia = ai. Proof. Define the maps bi+1 = hn−1(ai), where (hn)n≥−1 is the contraction of the simplicial set Hom(P,A). These maps satisfy ∂0bi+1 = ai, and also ∂jbi+1 = ∂i+1bj+1 for i < j ≤ n, since (∂jhn−1)(ai) = hn−2(∂j−1ai), and ∂j−1ai = ∂iaj for i < j. a2 a0 a1 Thus they form an (n + 1, 0)-horn in the simplicial set Hom(P,A), and since we are assuming that this simplicial set is Kan, this horn has a filler b : P −→ An+1. This gives the required map a = ∂0b. 3.2.2 Remark: This lemma shows that in the presence of finite limits our P-resolutions are also simplicial resolutions in the sense of Tierney and Vogel [TV1969]; that is, the com- parison maps to the simplicial kernels are P-epimorphisms. Together with Remark 3.1.10 we see that P-exactness and relative contractibility are equivalent in the situation when we have finite limits and any simplicial object is Kan relative to P. So if C has finite limits, the Comparison Theorem 2.4 from [TV1969] is more general than the one following in this section, but in the absence of finite limits our Comparison Theorem still works. We now prove our Comparison Theorem using the above lemma. 3.2.3 Theorem (Comparison Theorem): Let P be a simplicial object over B with each Pi ∈ P, and let A be a simplicial object over A, for which all the augmented simplicial sets Hom(Pi,A) are contractible and Kan. Then any map f : B −→ A can be extended to a semi-simplicial map f : P −→ A, and any two such extensions are simplicially homotopic. 52 3.3 Comonads generating the same projective class Proof. We construct this semi-simplicial map inductively, using Lemma 3.2.1. · · · ,2,2,2,2 P2 f2  ,2,2,2 P1 f1  ,2,2 P0 f0  ,2 B f−1=f · · · ,2,2,2,2 A2 ,2,2,2 A1 ,2,2 A0 ,2 A Suppose the maps fj : Pj −→ Aj are given for −1 ≤ j < n, and commute appropriately with the ∂i. This gives us n + 1 maps ai : Pn −→ An−1, where i ∈ [n], by composing the ∂i : Pn −→ Pn−1 with fn−1. These maps satisfy ∂iaj = ∂j−1ai for i < j, since ∂ifn−1 = fn−2∂i, and the ∂j in P satisfy the simplicial identities. Thus we can use Lemma 3.2.1 to obtain the map fn : Pn −→ An such that ∂ifn = ai = fn−1∂i. Now suppose f : P −→ A and g : P −→ A are two semi-simplicial maps commuting with f : B −→ A. We construct a homotopy hni : Pn −→ An+1 for n ≥ 0 and 0 ≤ i ≤ n, which satisfies ∂0hn0 = fn, ∂n+1h n n = gn and ∂ih n j =  hn−1j−1 ∂i for i < j ∂ih n i−1 for i = j 6= 0 hn−1j ∂i−1 for i > j + 1. h00 can be constructed using Lemma 3.2.1. Suppose the h k j exist for k < n and commute appropriately with the ∂i. Then hn0 must satisfy ∂0h n 0 = fn, ∂1h n 0 = ∂1h n 1 and ∂ih n 0 = hn−10 ∂i−1 for i > 1. Of these maps, all are known except for ∂1h n 1 . Setting a 0 0 = fn and a0i = h n−1 0 ∂i−1 for i > 1, we form an (n + 1, 1)-horn in Hom(Pn, A). A filler for this horn gives hn0 , and also a 0 1 = ∂1h n 1 , which is needed for the next step. Now suppose h n j are given for j < l, and we have al−1l = ∂lh n l = ∂lh n l−1. Then a l i = h n−1 l−1 ∂i for i < l, all = a l−1 l and a l i = h n−1 l ∂i−1 for i > l+1 form an (n+1, l+1)-horn. A filler for this gives hnl and a l l+1 = a l+1 l+1 for the next step. In the last step we have a n i = h n−1 n−1∂i for i < n, ann = a n−1 n = ∂nh n n−1 and ann+1 = gn. Then we use Lemma 3.2.1 again to get hnn. 3.3 Comonads generating the same projective class In this section we will need an assumption on the category C and the projective class P generated by the comonad G. 3.3.1 Definition (Kan projective class): Let G be a comonad on a category C and let P be the projective class generated G. The projective class P is called a Kan projective class on C when any augmented simplicial object A which is relatively contractible is also Kan relative to P. 53 Chapter 3. A Comparison Theorem for Simplicial Resolutions In particular, when G generates a Kan projective class, the simplicial object GA is Kan relative to P for any object A. Thus if K is a second comonad which generates the same projective class, the simplicial object KA is automatically also Kan relative to P. 3.3.2 Example (Additive categories): If C is an additive category, any projective class P in C is a Kan projective class, since then for any simplicial object A and any object P , the simplicial set Hom(P,A) is actually a simplicial group and thus Kan (cf. Exam- ple 3.1.6). For example, any comonad on the category R-Mod of (left) R-modules generates a Kan projective class, including both the “absolute” comonad generated by the forget- ful/free adjunction to Set and the “relative” comonad induced by a ring homomorphism φ : R −→ S as in Example 2.2.3. 3.3.3 Example (Regular projectives): When C is a regular category and the projective class P is the class of regular projectives, as remarked in Example 3.1.5, saying that a simplicial object A is Kan relative to P is the same as saying A is internally Kan in C. Thus when C is also Mal’tsev, every simplicial object is Kan [CKP1993, Theorem 4.2], and GA is a P-resolution. This includes the forgetful/free comonads on the categories Gp of groups, Rng of non-unital rings, XMod of crossed modules, Comm of commutative rings, K-Alg of associative K-algebras, etc. In fact it includes any variety C with the comonad generated by the forgetful functor to Set, as the following argument shows. Given a comonad G on a category C which comes from an adjunction C U $? ?? ?? ?? G ,2 C D F :D we can determine the class of morphisms of the projective class (P,E) generated by G in the following way: Given an object A and a morphism e : B −→ C in C, the diagram GA f  B e ,2 C corresponds via the adjunction to UA  UB Ue ,2 UC 54 3.3 Comonads generating the same projective class If e is in E, by choosing A = C and f = C , we see that Ue must be split in D, since C corresponds under the adjunction to 1UC . Conversely if Ue is a split epimorphism in D, we can factor any map UA −→ UC over Ue, which implies that we can factor any map f : GA −→ C over e in C, thus e ∈ E. Therefore the class E is exactly the class of morphisms whose images under U are split in D. Thus when C is a variety and U is the forgetful functor to Set, we will always get the class of regular projectives on C. The first step towards our goal is to show that the two simplicial resolutions GA and KA are homotopy equivalent. 3.3.4 Lemma: Let G and K be two comonads on C which generate the same Kan projective class P. Then the simplicial objects GA and KA are homotopy equivalent for any object A. Proof. Our assumptions on C and P imply that for any object A, the simplicial objects GA and KA are both P-resolutions of A. Thus we can use the Comparison Theorem 3.2.3 to get semi-simplicial maps f : GA −→ KA and g : KA −→ GA which commute with the identity on A. · · · ,2 ,2,2,2 G 3A f2  ,2,2,2 G2A f1  ,2,2 GA f0  ,2 A 1A · · · ,2,2,2,2 K 3A g2  ,2,2,2 K2A g1  ,2,2 KA g0  ,2 A 1A · · · ,2,2,2,2 G 3A ,2 ,2 ,2 G2A ,2,2 GA ,2 A Using the second part of the Comparison Theorem we see that both fg and gf are ho- motopic to the identity on KA and GA respectively. Thus GA and KA are homotopy equivalent. 3.3.5 Remark: In this case we don’t actually need the full strength of the second half of Theorem 3.2.3. For any semi-simplicial map f : GA −→ GA which commutes with the identity on A, we can use the homotopy hni = (G i+1fn−i)σi to see that it is homotopic to the identity on GA. Given a functor E : C −→ A, the simplicial objects EGA and EKA are still homotopy equivalent. We now show that, when A is a semi-abelian category, two simplicially ho- motopic semi-simplicial maps induce the same map on homology (see also [VdL2006]). For this we need to define a special simplicial object, so that all the maps that form a simplicial homotopy are taken together to form a single semi-simplicial map. We do this by defining the following limit objects AIn. 55 Chapter 3. A Comparison Theorem for Simplicial Resolutions 3.3.6 Notation: Suppose thatA has finite limits and let A be a simplicial object inA. Put AI0 = A1 and, for n > 0, let A I n be the limit (with projections pr1,. . . , prn+1 : A I n −→ An+1) of the zigzag An+1 ∂1 $? ?? ?? ?? ?? An+1 ∂1z     ∂2 $? ?? ?? ?? ?? · · · z $ An+1 ∂nz     An An An in A. Let 0(A)n, 1(A)n : AIn −→ An and s(A)n : An −→ AIn denote the morphisms respec- tively defined by 0(A)0 = ∂0 0(A)n = ∂0pr1, 1(A)0 = ∂1 1(A)n = ∂n+1prn+1, and s(A)n = (σ0, . . . , σn). 3.3.7 Proposition: Let A be a simplicial object in a finitely complete category A. Then the faces ∂Ii : A I n −→ AIn−1 and degeneracies σIi : AIn −→ AIn+1 given by ∂I0 = ∂0pr2 : A I 1 −→ AI0 ∂I1 = ∂2pr1 : A I 1 −→ AI0 σI0 = (σ1, σ0) : A I 0 −→ AI1 prj∂ I i = ∂i+1prj if j ≤ i∂iprj+1 if j > i : AIn −→ An prkσ I i = σi+1prk if k ≤ i+ 1σiprk−1 if k > i+ 1 : AIn −→ An+2, for i ∈ [n], 1 ≤ j ≤ n and 1 ≤ k ≤ n+2, determine a simplicial object AI . The morphisms mentioned in Notation 3.3.6 above form simplicial morphisms 0(A), 1(A) : AI −→ A and s(A) : A −→ AI such that 0(A)◦s(A) = 1A = 1(A)◦s(A). In other words, (AI , 0(A), 1(A), s(A)) forms a cocylinder on A. Two semi-simplicial maps f, g : B −→ A are simplicially homotopic if and only they are homotopic with respect to the cocylinder (AI , 0(A), 1(A), s(A)): there exists a semi- simplicial map h : B −→ AI satisfying 0(A)◦h = f and 1(A)◦h = g. Using a Kan property argument, we now give a direct proof that homotopic semi- simplicial maps have the same homology. 56 3.3 Comonads generating the same projective class 3.3.8 Proposition: Let A be a simplicial object in a semi-abelian category A; consider 0(A) : AI −→ A. For every n ∈ N, Hn0(A) is an isomorphism. Proof. Recall that in a semi-abelian category every simplicial object is Kan, relative to the class of regular epimorphisms. Using the Kan property, we show that the commutative diagram Nn+1AI Nn+10(A)  ,2 d′n+1  Nn+1A d′n+1  ZnAI Zn0(A)  ,2 ZnA is a generalised regular pushout (see Definition 1.1.19); then it is also a pushout, and Lemma 1.1.17 implies that the induced map HnAI −→ HnA is an isomorphism. Consider morphisms z : Y0 −→ ZnAI and a : Y0 −→ Nn+1A that satisfy d′n+1◦a = Zn0(A)◦z. It is enough to show that there exist a regular epimorphism y : Y −→ Y0 and a morphism h : Y −→ Nn+1AI satisfying d′n+1◦h = z◦y and Nn+10(A)◦h = a◦y: this implies that the comparison map to the pullback is a regular epimorphism, by Lemma 1.1.10 and the fact that the morphisms of a limit cone form a jointly monic family. We first sketch the geometric idea of this in the case n = 0. Consider a = a0 and z = z0 as in Figure 3.1; then (up to enlargement of domain) using the Kan property twice yields the needed (h0, h1) in N1AI . a z 0 Z00(A) ◦ z = d′1 ◦ a 0 h1 h0 Figure 3.1: Using the Kan property twice to obtain (h0, h1) in N1AI . For arbitrary n, write a0 = ⋂ j ker∂j◦a : Y0 −→ An+1, and (z0, . . . , zn) = ⋂ j ker∂j◦z. Note that as z : Y0 −→ ZnAI , we have ∂Ii z = 0 for i ∈ [n], which implies ∂izj−1 = 0 for i < j − 1 and i > j, where 1 ≤ j ≤ n + 1. We also have ∂jzj−1 = ∂jzj for 1 ≤ j ≤ n from the definition of the objects AIn. The map a0 in turn satisfies ∂ia0 = 0 for i ∈ [n], and ∂n+1a0 = ∂0z0. This last equality follows from d′n+1◦a = Zn0(A)◦z. 57 Chapter 3. A Comparison Theorem for Simplicial Resolutions Suppose we have regular epimorphisms yk : Yk −→ Yk−1 for 1 ≤ k ≤ n + 2, and mor- phisms hk−1 : Yk −→ An+2 satisfying ∂i◦hk−1 = 0 for 1 ≤ k ≤ n+2 and i /∈ {k−1, k, n+2}, and ∂n+2hk−1 = zk−1y1 · · · yk for 1 ≤ k ≤ n + 1, and also ∂0h0 = a0y1. We set y = y1◦ · · · ◦yn+2. This gives us the required map h = (h0◦y2◦ · · · ◦yn+2, . . . , hn+1) : Yn+2 −→ Nn+1AI which satisfies d′n+1◦h = z◦y and Nn+10(A)◦h = a◦y. We construct these maps yk and hk inductively. To get h0 we form an (n+ 2, 1)-horn (bi : Y0 −→ An+1)i of A by setting b0 = a0, bn+2 = z0 and bi = 0 for 1 < i < n + 2. A filler for this horn gives y1 : Y1 −→ Y0 and h0 : Y1 −→ An+2 satisfying ∂0h0 = a0y1. Now suppose for 1 ≤ k ≤ n+ 1 we have ak = ∂khk−1 : Yk −→ An+1 with ∂i◦ak = 0 for i ∈ [n], and ∂n+1ak = ∂kzk−1y1y2 · · · yk. Then we can form an (n + 2, k + 1)-horn by setting bk = ak, bn+2 = zky1 · · · yk and bi = 0 for i < k and k + 1 < i < n+ 2, which, when filled, induces yk+1 and hk with the desired properties. 3.3.9 Remark: A homology functor Hn involves an implicit choice of colimits: the coker- nels involved in the construction of the HnA. We may, and from now on we will, assume that these colimits are chosen in such a way that Hn0(A) is an identity instead of just an isomorphism. This gives us the equality in the next corollary. 3.3.10 Corollary: If f ' g then, for any n ∈ N, Hnf = Hng. Proof. Proposition 3.3.8 states that Hn0(A) is an isomorphism; by a careful choice of colimits in the definition of Hn, we may assume that Hn0(A) = 1HnA = Hn1(A). A B h ,2 g .4 f *0 AI 1(A) 4=qqqqqqqqqqqqq 0(A) !*MM MMM MMM MMM MM As(A) lr A If now h is a homotopy f ' g, then Hnf = Hn0(A)◦Hnh = Hn1(A)◦Hnh = Hng. Using the above, we can now prove our Main Theorem. 3.3.11 Theorem: Let G and K be two comonads on C which generate the same Kan projective class P. Let E : C −→ A be a functor into a semi-abelian category. Then the functors Hn(−, E)G and Hn(−, E)K from C to A are isomorphic for all n ≥ 1. 58 3.3 Comonads generating the same projective class Proof. It follows from Lemma 3.3.4 that the simplicial objects EGA and EKA are ho- motopy equivalent. Thus Corollary 3.3.10 implies that Hn(A,E)G ∼= Hn(A,E)K. Given a map f : B −→ A, the two semi-simplicial maps GB ,2 KB Kf ,2 KA and GB Gf ,2 GA ,2 KA are both semi-simplicial extensions of f , so they are homotopic by the Comparison Theo- rem 3.2.3. Again using Corollary 3.3.10, we see that the square HnGB HnGf ,2 ∼=  HnGA ∼=  HnKB HnKf ,2 HnKA commutes, which proves that the isomorphisms are natural. 3.3.12 Remark: In fact, the above isomorphism is also natural in the second variable. If α : E =⇒ F is a natural transformation, then the square Hn(A,E)G ∼= ,2 Hn(A,α)G  Hn(A,E)K Hn(A,α)G  Hn(A,F )G ∼= ,2 Hn(A,F )K also commutes, since EGA ,2 αGA  EKA αKA  FGA ,2 FKA already commutes. 3.3.13 Remark: We could define homology just using a projective class instead of a comonad, since the Comparison Theorem and Corollary 3.3.10 imply that any P-resolution of A will give the same homology. Consider for example the following (Tierney-Vogel) resolution in a category with finite limits: Given an object A, we can find a P-projective object A0 with a P-epimorphism ∂0 : P0 −→ A, since there are enough P-projectives. We call this a presentation of A. Take the kernel pair of ∂0, and take the presentation of the resulting object to get P1. Composition gives two maps ∂0 and ∂1 to P0, and we can take the simplicial kernel of 59 Chapter 3. A Comparison Theorem for Simplicial Resolutions these and the presentation of the resulting object to get P3 and so on. This gives a reso- lution in the Tierney-Vogel sense [TV1969]. When P is a Kan projective class, it is also a P-resolution in our sense and thus gives the same homology. This resolution is often easier to work with than the functorial GA. 3.3.14 Example (Two comonads on R-Mod): Given a ring homomorphism φ : S −→ R, every R-module can also be considered as an S-module by restricting the R-action to S via φ. This gives rise to an adjunction R⊗S (−) a HomR(R,−) between the categories of modules, where R is viewed as an S-module. Now consider two rings S1 and S2 with a surjective ring homomorphism ψ : S1 −→ S2. Let R be another ring, with ring homomorphisms as below which make the diagram commute: S1 φ1 !*MM MMM MMM ψ _  R S2 φ2 4=qqqqqqqq For each i = 1, 2 this gives us a comonad on R-Mod using the adjunction above: R-Mod Ui !)LL LLL LLL LL Gi ,2 R-Mod Si-Mod R⊗Si (−) 5=rrrrrrrrrr We write Ui for the forgetful functor HomR(R,−) from R-modules to Si-modules. As seen in Example 3.3.3, the projective class generated by Gi is given by the class of maps in R-Mod which are split as Si-module maps. An R-module map e : B −→ C is split as an Si-module map if and only if there exists a function f : C −→ B with f(sc+ s′c′) = sf(c)+s′f(c′) for s ∈ I[φi]. Since ψ is a surjection, we have I[φ1] = I[φ2], so any R-module map e has the property that U1(e) is split if and only if U2(e) is split. Thus the comonads G1 and G2 induce the same projective class, and thus give rise to the same homology on R-Mod, for any functor E : R-Mod −→ A to a semi-abelian category A. As mentioned in [BB1969], when using the functors E = N ⊗R − or E = HomR(N,−) for an R-module N , this homology is Hochschild’s S-relative Tor or Ext respectively; so we see we get the same relative Tor or Ext functor for a ring S and any quotient of S. This means we can replace S by the image of φ as a submodule of R. 60 Chapter 4 Homology via Hopf Formulae In the early 1940s, Hopf proved the well-known formula for the second integral homology of a special kind of topological space, which calculates the homology purely in terms of the fundamental group of the space. This led to the definition of group homology independent of a topological space. So today we can write his famous formula as H2A = [P, P ] ∩K[p] [K[p], P ] where p : P −→ A is a projective presentation of a group A. The group commutators used here can be generalised so that this formula makes sense in any semi-abelian category. Furthermore, it can be extended to higher dimensional Hopf formulae which give the higher homology objects HnA. This course was taken by Everaert, Gran and Van der Linden in [EGVdL2008] and further pursued by Everaert in his thesis [Eve2007]. We give the main background and results here, as they will be needed later on in the thesis. Though we will work in semi-abelian categories, most of the results in this section do not need coproducts. In particular, all constructions borrowed from [EGVdL2008] and [Eve2007] which take place in a semi-abelian category still work in pointed exact protomodular ones: though these categories need not have coproducts, they still have cokernels of kernels (see [BB2004, Corollary 4.1.3]). This allows us to apply our results to examples where the category being considered is pointed exact protomodular but lacks coproducts, such as the category of finite groups. In Section 4.1 we introduce the concept of an axiomatically defined class of extensions and the associated higher extensions which are the starting point for the whole theory. Section 4.2 defines strongly E-Birkhoff subcategories, which are a generalisation of the ordinary Birkhoff subcategories we met in Section 1.3. The Galois structures and the induced centralisation and trivialisation functors on which the Hopf formulae crucially build are discussed in Section 4.3. Finally in Section 4.4 we define homology via the Hopf formulae and exhibit the Everaert sequence: a long exact homology sequence which generalises and extends the Stallings-Stammbach sequence known in the case of groups. This sequence and its universal properties play a crucial role in Chapter 5. The content of this chapter is known material, and mainly taken from the work of Everaert, Gran and Van der Linden [EGVdL2008] and Everaert’s thesis [Eve2007]. We will need the results and concepts introduced here in Chapter 5. 61 Chapter 4. Homology via Hopf Formulae 4.1 Extensions and higher extensions The main ingredient for higher Hopf formulae is the concept of higher-dimensional exten- sions. To arrive at this notion, we will first introduce higher-dimensional arrows. Here A always denotes a semi-abelian category, unless stated otherwise. 4.1.1 Definition (Higher-dimensional arrows): The category ArrkA consists of k- dimensional arrows in A: Arr0A = A, Arr1A = ArrA is the category of arrows Fun(2,A) where 2 is generated by a single map ∅ −→ {∅}, and Arrk+1A = ArrArrkA. Thus a dou- ble arrow is a commutative square in A, a 3-arrow is a commutative cube, and a k-arrow is a commutative k-cube. Clearly, ArrkA is also semi-abelian. The functor ker : Arrk+1A −→ ArrkA maps a (k + 1)-arrow a to its kernel K[a], and a morphism (f ′, f) between (k + 1)-arrows b and a to the induced morphism between their kernels. K[b] ker(f ′,f)   ,2Ker b ,2 B′ f ′  b ,2 ⇓ B f  K[a]  ,2 Ker a ,2 A′ a ,2 A Repeating it n times gives a functor kern : Arrk+nA −→ ArrkA which sends a (k+n)-arrow a to the object Kn[a] of ArrkA. We now axiomatically define a class of extensions as in [Eve2007]. The definitions and most of the results and proofs in this section are taken from Tomas Everaert’s thesis [Eve2007]. Given a class of morphisms E in a semi-abelian category A, we write obE for the class of objects A ∈ |A| that occur as domains or codomains of the arrows in E: A ∈ obE if and only if there is at least one f ∈ E with f : A −→ B or f : C −→ A. We also write AE for the full subcategory of A determined by the objects of obE. 4.1.2 Definition (Extensions): [Eve2007] Let E be a class of regular epimorphisms in A with 0 ∈ obE. Then E is called a class of extensions when it satisfies the following properties: (1) E contains all split epimorphisms f : B −→ A with A and B in obE; (2) (a) if f : B −→ A and g : C −→ B are in E, then so is their composite f◦g; (b) if f◦g is in E and B is in obE, then g is in E; 62 4.1 Extensions and higher extensions (3) morphisms f ∈ E are stable under pullback along arrows h : D −→ A in A with D ∈ obE, i.e. in the diagram below, h∗f is again in E whenever D ∈ obE; D ×A B ,2 h∗f  B f  D h ,2 A (4) given a short exact sequence in A as below with B ∈ obE, we have f ∈ E whenever K ∈ obE; 0 ,2K ,2B f ,2A ,20 (5) given a commutative diagram as below with short exact rows in A, whenever both g and k are in E and B ∈ obE, then we also have b ∈ E. 0 ,2 K ,2 k  C g ,2 b  A ,2 0 0 ,2 L ,2 B f ,2 A ,2 0 We call an arrow f ∈ E an E-extension, or just an extension, and write B  ,2A . 4.1.3 Remark: As we have 0 ∈ obE, condition (3) implies that the kernel K[f ] of any extension f is in obE. This gives the converse to (4), so that we have: for any short exact sequence in A as in (4), K ∈ obE if and only if f ∈ E. 4.1.4 Example: The leading example of a class of extensions is the class of all regular epimorphisms in A. In this case we have obE = |A|, all objects of the category A. In fact, it follows from (4) that the class of all regular epimorphisms is the only class E with this property. We also introduce a concept of higher extensions. 4.1.5 Definition (Higher extensions): Given a class of extensions E in a semi-abelian category A, we define the class of n-fold (E-)extensions (called n-extensions when E is understood) inductively as follows: a 0-extension is an object in obE, a 1-extension is an arrow in E, and for n ≥ 1, an (n+ 1)-extension is a morphism (f ′, f) in ArrnA such that all arrows in the induced 63 Chapter 4. Homology via Hopf Formulae diagram B′ f ′  ' b z" r % P  ,2 _  A′ a _  B f  ,2 A (D) are n-extensions. Here P is the pullback of a and f . We will say double (E-)extension for a 2-fold extension. We denote the class of (n+1)-fold E-extensions by En, thus E0 = E and E1 denotes the double E-extensions. To justify the name of (n+ 1)-extension, we will have to show that the class En really is a class of extensions in the sense of 4.1.2. Clearly it is enough to show that E1 is a class of extensions, and the rest follows by induction, as we can then view En as (En−1)1. So we will now concentrate on double extensions. 4.1.6 Remark: Notice that a morphism (f ′, f) : b −→ a in ArrA is a double extension if and only if (b, a) : f ′ −→ f is a double extension. B′ f ′  ,2 b _  A′ a _  B f  ,2 A We can see that, as any extension is a regular epimorphism, a double extension gives a square in A which is a regular pushout (see Section 1.1). In particular, a double extension is a regular epimorphism in ArrA, both viewed as (f ′, f) : a −→ b and as (a, b) : f ′ −→ f . We will now prove a property relating double extensions to extensions, reminiscent of condition (4) in Definition 4.1.2. This will be a key ingredient in proving that E1 is a class of extensions. 4.1.7 Lemma: Given a commutative diagram in A with short exact rows as below, where f ′, f , a and b are extensions, 0 ,2 K ′ k   ,2 ,2 B′ b _  f ′  ,2 A′ ,2 a _  0 0 ,2 K  ,2 ,2 B f  ,2 A ,2 0 the right hand square is a double extension if and only if k is an extension. 64 4.1 Extensions and higher extensions Proof. We can decompose the above diagram as follows, with pullback squares as indicated. 0 ,2 K ′  ,2 ,2 k  B′ f ′  ,2 r  A′ ,2 0 0 ,2 K  ,2 ,2 P  ,2 _  A a _  ,2 0 0 ,2 K  ,2 ,2 B f ,2 A ,2 0 By definition, the right hand square of the original diagram is a double extension if and only if the factorisation r to the pullback is an extension. Notice that by Condition (3) of Definition 4.1.2 the pullback P −→ A of f along a is again an extension and so P is in obE, and by Remark 4.1.3 both K and K ′ are also in obE. If r is an extension, then k is an extension by 4.1.2 (3). Conversely, if k is an extension, then so is r by 4.1.2 (5). 4.1.8 Proposition: Given a class of extensions E in A, the induced class E1 is a class of extensions in ArrA. Proof. We remarked earlier that a double extension is indeed a regular epimorphism in ArrA. Notice that obE1 = E, and clearly 10 ∈ E. We have to show that E1 satisfies conditions (1) to (5) of Definition 4.1.2. To distinguish between these conditions applied to E and to E1, we will denote the conditions referring to E1 by (1)1 to (5)1. A split epimorphism in ArrA is a morphism (f ′, f) : b −→ a such that both f ′ and f are split in A by s′ and s respectively, and (s′, s) : a −→ b is also a morphism in ArrA. Using Remark 4.1.6, to show that (1)1 holds it is enough to show that (b, a) : f ′ −→ f is a double extension whenever a and b are extensions and (f ′, f) is a split epimorphism in ArrA as above. This follows from (1) and Lemma 4.1.7, as the map between the kernels K[b] and K[a] is also a split epimorphism. For the next two conditions we again use Remark 4.1.6, which makes it easy to verify that (2a)1 and (2b)1 follow from Lemma 4.1.7 and (2a) or (2b) respectively. Condition (3)1 follows from (3) and (2a). The key point here is that when pulling back (f ′, f) : b −→ a to say (g′, g) : e −→ d, the square formed by the comparison to the pullbacks E′ ,2  Q  B′  ,2 P is also a pullback square, and this implies that (g′, g) is also an extension. Condition (4)1 follows easily from (4), (2) and Lemma 4.1.7, and finally Condition (5)1 follows from (5) and (3). Again the squares induced by comparison maps to the pullbacks play a crucial role. 65 Chapter 4. Homology via Hopf Formulae Thus the classes En−1 are indeed classes of extensions for any n ≥ 1. We have obEn = En−1. The n-extensions En−1 determine a full subcategory ExtnA of ArrnA. Notice that ExtnA = (ArrnA)En . We sometimes analogously write Ext0A = AE. When we say that a sequence 0 ,2 K[f ]  ,2 ,2 B f  ,2 A ,2 0 is exact in ExtnA, we mean that it is an exact sequence in ArrnA, and the objects are n-extensions. Recall from Remark 4.1.3 that f is an n+1-extension if and only if all three objects are n-extensions. Roughly, the idea behind this definition of k-extensions is the following: suppose we are given a double extension (f ′, f) of an object A of A as in Diagram (D), and let α be any element of A. Then in addition to the existence of elements β of B and α′ of A′ such that f(β) = α and a(α′) = α, there is also an element β′ ∈ B′ such that b(β′) = β and f ′(β′) = α′, whichever β and α′ were chosen. 4.2 Strongly (E-)Birkhoff subcategories Given a class of extensions as in Section 4.1, we will now generalise the notion of Birkhoff subcategory given in Definition 1.3.2. Again this concept is taken from [EGVdL2008] and [Eve2007]. This more general definition will allow us to handle higher extensions at the same time as ordinary Birkhoff subcategories, which makes for clearer statements and proofs later on. 4.2.1 Definition: Given a class of extensions E in a semi-abelian category A, and a reflective subcategory B of AE, we write I : AE −→ B for the reflector and denote the unit of the adjunction by η. We call B a strongly E-Birkhoff subcategory of A if for every (E-)extension f : B −→ A the induced square B ηB  f ,2 A ηA  IB If ,2 IA is a double (E-)extension. 4.2.2 Remark: Notice that this immediately implies that the reflector I takes extensions to extensions, and also that ηA is an extension (and so a regular epimorphism) for each A. This in turn implies that B is closed under subobjects: let A ,2 m ,2B be a mono in A 66 4.2 Strongly (E-)Birkhoff subcategories with B ∈ B. Then as η is a natural transformation we have A ,2 m ,2 ηA _  B ηB IA Im ,2 IB and so ηA is both a mono and a regular epi, and thus an isomorphism. In fact it is well known that a reflective subcategory is closed under subobjects if and only if each ηA is a regular epimorphism, but we only need this one direction. Note that B is also closed under any limits that exist in AE, as it is a reflective subcategory (see for example Proposition 3.5.3 in [Bor1994]). 4.2.3 Example (strongly E-Birkhoff subcategories): When E is the class of regular epimorphisms in A, strongly E-Birkhoff subcategories of A coincide with the usual Birkhoff subcategories (see Lemma 1.3.4), and AE = A. Thus the category Ab of abelian groups is a strongly (regular epi)-Birkhoff subcategory of Gp. Later we will meet special classes of higher extensions, the central n-extensions, which form a strongly E-Birkhoff subcategory of ArrnA when E is the class of n-extensions, so AE = ExtnA. This allows us to state results that work at the same time for a semi-abelian category A with a usual Birkhoff subcategory and for a category of higher extensions ExtnA in A. There is another criterion for a reflective subcategory B to be strongly E-Birkhoff, which uses the kernel of the unit ηA for a given object A. We first introduce some notation. We can view the reflector I as a functor I : AE −→ AE. Then we have another functor J : AE −→ AE, given by JA = K[ηA], which fits into the following short exact sequence of functors. 0 ,2J  ,2 µ ,21AE η  ,2I ,20 4.2.4 Proposition: Let B be a reflective subcategory of AE such that the unit ηA : A −→ IA is a regular epimorphism for any object A ∈ A. Then B is a strongly E-Birkhoff subcategory of A if and only if the functor J : AE −→ AE preserves extensions. Proof. Consider the following diagram with short exact rows: JB Jf   ,2 ,2 B f _  ηB  ,2 IB If _  JA  ,2 ,2 A ηA  ,2 IA By 4.1.7 it follows that if the right hand square is a double extension, then Jf is an extension. Conversely, if Jf is an extension, then 4.1.2(4) implies that ηB and ηA are 67 Chapter 4. Homology via Hopf Formulae extensions, and so by 4.1.2(2) If is as well. Then again by 4.1.7 the right hand square is a double extension for any extension f , and so B is strongly E-Birkhoff. 4.3 The Galois structures Γn, centralisation and trivialisa- tion The strongly E-Birkhoff subcategories give rise to Galois structures Γ as defined in 1.4.1. In fact, we will see that a strongly E-Birkhoff subcategory B induces a whole sequence Γn of Galois structures, each one giving rise to the next. 4.3.1 Proposition: Let E be a class of extensions (in the sense of 4.1.2) in a semi-abelian category A, and B a strongly E-Birkhoff subcategory of A. Let Z be the class of arrows f in B such that f ∈ E, and let I : AE −→ B be the reflector and ⊆ : B −→ AE the inclusion functor. Then (AE,B,E,Z, I,⊆) is a Galois structure. Proof. This is an immediate consequence of the definitions of a class of extensions 4.1.2 and a strongly E-Birkhoff subcategory 4.2.1. 4.3.2 Remark: Notice that we take the extensions E of the Galois structure to be a class of extensions in the sense of 4.1.2, so there is no clash of terminology. This Galois structure Γ = (AE,B,E,Z, I,⊆) gives rise to central extensions as defined in 1.4.3. We denote the full subcategory of ArrA determined by these central extensions by CExtBA. This gives a subcategory of the category ExtA determined by the class of extensions E, which is dependant on the strongly E-Birkhoff subcategory B. When B is clear from the context, we might also write CExtA. We will show that this category of central extensions forms a strongly E1-Birkhoff subcategory of ArrA, thus giving us another Galois structure as above. To show this, we must first construct a reflector I1 : ExtA −→ CExtBA. Recall the short exact sequence of functors AE −→ AE induced by the reflector I = I0: 0 ,2J  ,2 µ ,21A η  ,2I ,20 From this, we build a similar short exact sequence of functors ExtA −→ ExtA as follows. (The construction is made pointwise in ArrA, which has good categorical properties, but the result turns out to be an extension.) 68 4.3 The Galois structures Γn, centralisation and trivialisation 4.3.3 Definition (Centralisation functor): Consider an extension f : B −→ A and its kernel pair (R[f ], pi1, pi2). Write J1[f ] = K[Jpi1] and J1f : J1[f ] −→ 0. J1[f ] = K[Jpi1]_    ,2Ker Jpi1,2 JR[f ] _  µR[f ]  Jpi1 ,2 Jpi2 ,2 JB_  µB  K[f ] = K[pi1]  ,2 Kerpi1 ,2 R[f ] pi1 ,2 pi2 ,2 B This clearly determines a functor J1 : ExtA −→ ExtA. Note that pi2◦Kerpi1 = Ker f , and the left hand square is a pullback. We define the map µ1f : J1f −→ f as in the left hand square below. J1[f ] ,2 µB◦Jpi2◦Ker Jpi1 J1f_  µ1f =⇒ B f _  0 ,2 A B ρ1f  ,2 f _  η1f =⇒ I1[f ] I1f_  A A The composition µB◦Jpi2◦Ker Jpi1 is a normal monomorphism, so we can take cokernels, yielding the right hand square. Since µ1f is the kernel of its cokernel, we obtain the short exact sequence 0 ,2J1  ,2 µ 1 ,21ExtA η1  ,2I1 ,20 of functors ExtA −→ ExtA. 4.3.4 Proposition: The functor I1 : ExtA −→ ExtA corestricts to I1 : ExtA −→ CExtBA. Proof. For a proof see for example [Eve2007, Lemma 1.4.2]. This justifies the name of centralisation functor for I1. 4.3.5 Theorem: Given a strongly E-Birkhoff subcategory B of A, the category CExtBA is a strongly E1-Birkhoff subcategory of ArrA. The reflector is given by I1 : ExtA −→ CExtBA. Proof. For a proof see for example [Eve2007, Theorem 1.4.3]. Now Proposition 4.3.1 implies that Γ1 = (ExtA,CExtBA,E1,Z1, I1,⊆) forms another Galois structure, where Z1 is defined analogously to Z above as those maps in CExtBA that lie in E1. This process may be repeated inductively to obtain Galois structures Γn and functors Jn : ExtnA −→ ExtnA and In : ExtnA −→ CExtnBA. For n ≥ 1 and an n-extension f , we often call the extension Inf the centralisation of f . We now give some examples of centralisation functors I1 for the different Birkhoff subcategories we have met. 69 Chapter 4. Homology via Hopf Formulae 4.3.6 Remark (Property of group commutators): In the next example, we will be using a property of group commutators which is easy to check: for normal subgroups M and N of a group B with N ⊆M , we have[ M N , B N ] = [M,B] N . 4.3.7 Example (centralisation functors): When A = Gp and B = Ab, the category of abelian groups, we saw in Example 1.2.6 that I = ab takes a group G to G/[G,G]. We also saw in Example 1.4.4 that an extension f : B −→ A is central if and only if K[f ] ⊆ ZB, or equivalently if and only if [K[f ], B] = 0. It easily follows, using Remark 4.3.6, that the functor I1 = centr takes an extension f : B −→ A to I1f = centr f : B [K[f ], B] −→ A, that is, J1[f ] = [K[f ], B]. When A = Gp and B = Nilm, the category of nilpotent groups of class at most m, we saw in Example 1.3.3 that I = nilm takes a group G to G/LCmG, for example nil2 takes G to G/[[G,G], G]. Here the centralisation functor I1 takes an extension f : B −→ A to I1f : B lm(K[f ], B, . . . , B) −→ A with lm as in Example 1.4.4. Thus, when m = 2, we have I1f : B/[[K[f ], B], B] −→ A. Again this follows from Example 1.4.4 and Remark 4.3.6. Similarly when A = Gp and B = Solm, the category of solvable groups of class at most m, the reflector I = solm takes a group G to G/DmG, for example sol2 takes G to G/[[G,G], [G,G]], and I1 takes an extension f : B −→ A to I1f : B dm(K[f ], B,B, . . . , B) −→ A. For example, when m = 2, we have I1f : B/[[K[f ], B], [B,B]] −→ A. When A = LeibK and B = LieK for a field K (of characteristic 6= 2), we saw in Example 1.3.3 the reflector lie : g 7−→ g/gAnn, where gAnn is the two-sided ideal generated by elements of the form [x, x]. Given an extension f : b −→ a, let [K[f ], b]lie be the ideal generated by elements of the form ([k, b] + [b, k]) for k ∈ K[f ] and b ∈ b. Notice that [k, b] + [b, k] = [b+ k, b+ k]− [b, b]− [k, k] is an element of bAnn. Then the centralisation functor sends f : b −→ a to I1f : b [K[f ], b]lie −→ a. We see that ([b, b], [b+k, b+k])−([b, b], [b, b])−([0, 0], [k, k]) is an element of R[f ]Ann which is sent to 0 in b by the first projection, so it (or isomorphically, its second projection 70 4.3 The Galois structures Γn, centralisation and trivialisation [k, b] + [b, k]) is an element of [K[f ], b]lie = J1[f ] = K[Jpi1]. Clearly, for any k ∈ K[f ], we have [k, k] = 12([k, k] + [k, k]) ∈ [K[f ], b]lie. Conversely, any element of R[f ]Ann which maps to 0 under the first projection must be made up of elements of the following sort: either the first entry is of the form [b, b] − [b, b], which must have in the second entry [b + k, b + k] − [b + k′, b + k′] = [b, k] + [k, b] + [k, k] − [b, k′] − [k′, b] − [k′, k′] for some k, k′ ∈ K[f ], which is an element of [K[f ], b]lie; or the first entry is of the form [a, a] = 0, accompanied by [b, b] in the second entry, such that f(a) = f(b). But then b − a ∈ K[f ] and we have [b− a, b+ a] + [b+ a, b− a] = 2[b, b] so [b, b] ∈ [K[f ], b]lie. Compare this with the Lie-centre ZLie(b) = {z ∈ b | [b, z] = −[z, b] ∀ b ∈ b} from Example 1.4.4. As mentioned earlier, Theorem 4.3.5 allows us to apply results about strongly E- Birkhoff subcategories at the same time to an ordinary Birkhoff subcategory of a semi- abelian category A and to the higher central extensions viewed as strongly E-Birkhoff subcategories of ArrnA, where E is the class of n-extensions. Thus when we say “B is a strongly E-Birkhoff subcategory of A”, we have one of these two cases in mind. 4.3.8 Remark: Given an n-extension A, for n ≥ 0, the centralisation of the (n + 1)- extension !A : A −→ 0 turns out to be In+1!A : InA −→ 0. The following is also often useful, and quite easy to show using the 3× 3-Lemma and the fact that CExtnBA is strongly E n-Birkhoff. 4.3.9 Lemma: For an (n+ 1)-extension f : B −→ A, we have InIn+1f = Inf : InB −→ InA, i.e., In(In+1[f ]) = InB. Proof. This proof is taken from [EGVdL2008, Lemma 6.2]. Consider the following dia- gram, in which the rows and the middle column are exact sequences (here pi1 and pi2 are 71 Chapter 4. Homology via Hopf Formulae the projections of the kernel pair R[f ] to B, as in Definition 4.3.3): 0  0  0 ,2 J1[f ]_  Jpi2◦kerJpi1  J1[f ] ,2_  µ1f  0  0 ,2 JB  ,2 µB ,2 Jρf_  B ηB  ,2 ρf _  IB ,2 Iρf_  0 0 ,2 JI1[f ]  ,2 µI1[f ] ,2  I1[f ] ηI1[f ] ,2  II1[f ] ,2  0 0 0 0 The top left square commutes by definition of µ1f . Note that this square is a pullback, as µB is a monomorphism. Thus Jpi2◦kerJpi1 is the kernel of Jρf , and the first column is also an exact sequence. Thus by the 3 × 3 Lemma, the last column is also exact, which makes Iρf an isomorphism. This gives II1[f ] II1f  ,2 ∼=  IA IB If  ,2 IA 4.3.10 Remark: Given an n-extension f , the only object of Jnf which is non-zero is domn Jnf , the “initial” object of the n-cube Jnf . This follows easily from the inductive construction of Jnf . Thus we have domn Jnf = Kn[Jnf ] for any n-extension f . This also implies that the only object of the n-cube Inf which differs from f is the initial object domn Inf . The Galois structures Γn also give rise to trivial extensions as defined in 1.4.3. Similarly to the central extensions, the trivial (n+ 1)-extensions of A form a reflective subcategory TExtn+1A of Extn+1A; the reflector Tn+1 : Extn+1A −→ TExtn+1A 72 4.4 Hopf formulae maps an extension f to the pullback Tn+1f : Tn+1[f ] −→ A of Inf along ηnA, the triviali- sation of f . A ηnA $ ?? ?? ?? B f  *0 ηnB # .4 ρnf  ,2 In+1[f ] In+1f $ .5 ηn In+1[f ]  )0  ,2 Tn+1[f ] ????   Tn+1f ?:D  $ ?? ?? ?? InA InB Inf ?:D  Thus we obtain a comparison map rn+1f : In+1[f ] −→ Tn+1[f ], which is a (n+1)-extension by the strong En-Birkhoff property of the reflector In (see 4.2.1) and 4.1.2(2b). This gives an (n+ 2)-extension In+1f −→ Tn+1f . 4.3.11 Remark: Recalling Remark 4.3.10, we see that again the only object of the (n+1)- cube Tn+1f which differs from f is the initial object domn+1 Tn+1f . This implies that the comparison map rn+1f is the identity everywhere except for on this initial object. 4.4 Hopf formulae As in the groups case, the Hopf formulae in the general categorical context use projective presentations, so we must define what exactly we mean by a higher projective presentation. 4.4.1 Definition (Projective presentations): An object of ArrkA is called extension- projective if it is projective with respect to the class of (k + 1)-extensions. A k- extension f : B −→ A is called a (projective) presentation of A when the object B is extension-projective. A k-extension f : B −→ A is called a k-fold presentation, or just k-presentation, when the object B is extension-projective and A is a (k−1)-presentation. (A 0-presentation is an object of AE.) Given an object A of AE, a k-fold presentation p of A is a k-fold presentation with codk p = A, i.e. the “terminal object” of the k-cube p in A is A. We will often denote the “initial object” of a k-presentation p by Pk. We can now state the theorem connecting comonadic homology to the Hopf formulae. 4.4.2 Theorem (Hopf formula): [EGVdL2008, Eve2007] Let E be a class of extensions in a semi-abelian monadic category A, and let B be a strongly E-Birkhoff subcategory of A with reflector I. Given an n-presentation p of an object A of AE with initial object Pn, we have Hn+1(A, I)G ∼= JPn ∩K n[p] Kn[Jnp] . 73 Chapter 4. Homology via Hopf Formulae Proof. The case where B is a straightforward Birkhoff category of A is proved in Theorem 8.1 of [EGVdL2008]. For the axiomatic extensions version see Theorem 3.6.10 in [Eve2007]. Notice that different notation is used there. The functor J is denoted by [·], and Kn[Jn·] by [·]n. The n-fold kernel Kn[p] = ⋂n i=0K[pi] is the intersection of all maps pi with domain Pn in p. In their paper [EG2007], Everaert and Gran define homology in a semi-abelian category with enough projectives via these higher Hopf formulae, which has been shown to give the same result as comonadic homology in a monadic setting. But for this theory no monadicity conditions are required. We will follow this strategy here as well. 4.4.3 Definition (Hopf homology): Let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I. We define Hn+1(A, I)E = JPn ∩Kn[p] Kn[Jnp] with notation and p : P −→ A as in Theorem 4.4.2 above. We also write H1(A, I)E = IA. Notice we have changed the subscript G of the comonadic homology to E to distinguish between these two different definitions of homology. As Theorem 4.4.2 shows, the two definitions coincide as soon as they are both defined, namely in a semi-abelian monadic category. 4.4.4 Example: In the category Gp of groups, the functor J becomes the commutator subgroup functor, that is JA = [A,A]. So in the case n = 1 we recover the well-known formula H2(A,Z) = [P, P ] ∩K[p] [K[p], P ] for integral group homology. We could write the functor J as a commutator in a general semi-abelian category, as it really is a commutator in many examples, but this would give too many different sorts of square brackets in our notation, so we prefer to call it J in most cases. When A = LieK is the category of Lie algebras over a field K, with the Birkhoff subcategory B = AbLie of abelian Lie algebras, the homology defined as in Definition 4.4.3 is the Chevalley-Eilenberg homology of Lie algebras. We can give an alternative formulation of the Hopf formulae, involving centralisation and trivialisation of the n-fold presentation p (cf. [Eve2008, Remark 5.12]). 74 4.4 Hopf formulae 4.4.5 Proposition: Let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I. Given an n-presentation p of an object A of AE with initial object Pn, we have JPn ∩Kn[p] Kn[Jnp] ∼= Kn+1[Inp −→ Tnp] and thus Hn+1(A, I)E ∼= Kn+1[Inp −→ Tnp]. Proof. Notice that as Inp and Tnp coincide in all but their initial object, this (n+ 1)-fold kernel is really just the initial object of the n-cube K[Inp −→ Tnp]. We first prove the case n = 1, then we use the strongly (extension)-Birkhoff prop- erties to apply this case to higher n and use induction to prove the whole statement. For ease of notation let p : P1 = P −→ A have kernel K. Writing I1p : I1[p] −→ A and T1p : T1[p] −→ A, we see that K2[I1p −→ T1p] = K[I1[p] −→ T1[p]] as remarked above. K2[I1p −→ T1p] K[I1[p] −→ T1[p]]  ,2 ,2 _  I1[p]  ,2 I1p _  T1[p] T1p _  0  ,2 ,2 A A So if we manage to prove I1[p] = P K[J1p] and T1[p] = P K ∩ JP , then Noether’s Isomorphism Theorem 1.1.15 will give the required result. By definition I1[p] = P J1[p] 75 Chapter 4. Homology via Hopf Formulae (see 4.3.3), and we have J1[p] = K[J1p] as J1p : J1[p] −→ 0. For T1[p], consider the following diagram, where we are taking kernels to the left: D  ,2 ,2  JP  ,2 _  µP  JA_  µA  K  ,2 ,2 P p  ,2 _  A _  IP Ip  ,2 IA As µA is a mono, the upper left square is a pullback, so D = K ∩ JP . Now consider K ∩ JP  ,2 ,2 JP  ,2_   JA_   JA_   K ∩ JP  ,2 ,2 P τ1p  ,2 _  T1[p] T1p  ,2 _  A _  IP IP Ip  ,2 IA where the three columns are short exact. As the bottom right square is a pullback, the kernels of its two vertical maps coincide. Now the middle square in the top line is also a pullback, since the map IP −→ IP at the bottom is a monomorphism. Thus K ∩ JP is also the kernel of τ1p and thus T1[p] = P/K ∩ JP as required. For higher n we make use of the fact that CExtn−1A is a strongly En−1-Birkhoff sub- category of Arrn−1A. If we write p : P −→ Q with P and Q in Extn−1A, we can apply the above to get K2[Inp −→ Tnp] = Jn−1P ∩K[p]K[Jnp] . Thus taking more kernels gives Kn+1[Inp −→ Tnp] = K n−1[Jn−1P ] ∩Kn[p] Kn[Jnp] (E) as K[B/C] = K[B]/K[C] by the 3 × 3-Lemma, and kernels commute with intersections as both are limits. So we only have to prove that Kn−1[Jn−1P ] ∩ Kn[p] = JPn ∩ Kn[p]. This holds because Kn−1[Jn−1P ] is a subobject of JPn and JPn ∩Kn[p] is a subobject of Kn−1[Jn−1P ], which we will now show. Notice that the (n− 1)-extension P has projective codomain and hence is split. This means that its centralisation and trivialisation coincide (see Lemma 1.4.5). Using (E) for P we see that Kn[In−1P −→ Tn−1P ] = K n−2[Jn−2(domP )] ∩Kn−1[P ] Kn−1[Jn−1P ] = 0 76 4.4 Hopf formulae and so Kn−2[Jn−2(domP )] ∩ Kn[p] ⊆ Kn−2[Jn−2(domP )] ∩ Kn−1[P ] = Kn−1[Jn−1P ] as clearly Kn[p] ⊆ Kn−1[P ]. For n = 2 this immediately gives JP2 ∩K2[p] ⊆ K[J1P ]. Now we use induction. By the induction hypothesis we have JPn ∩Kn−1[P ] ⊆ Kn−2[Jn−2(domP )] and so JPn ∩Kn[p] ⊆ Kn−2[Jn−2(domP )]. We also have Kn−2[Jn−2(domP )] ∩Kn[p] ⊆ Kn−1[Jn−1P ] from above. Thus we can fit all these together to give JPn ∩Kn[p] ⊆ Kn−2[Jn−2(domP )] ∩Kn[p] ⊆ Kn−1[Jn−1P ] as desired. The crucial point here is that the information in the higher homology objects is entirely contained in higher-dimensional versions Ik : ExtkA −→ CExtkBA of the reflector I : A −→ B. One could say that homology measures the difference between the centralisa- tion and the trivialisation of an n-presentation p of A. We now give a long exact homology sequence, which plays a crucial role in our subse- quent theory in Chapter 5. For this we use the homology defined via the Hopf formulae as in 4.4.3 above, so no monadicity conditions are needed. Even so we here only give the proof for the monadic case and refer the reader to [Eve2007] for a proof in full generality. 4.4.6 Theorem (Everaert sequence): [Eve2007, Theorem 2.4.2] Let E be a class of extensions in a semi-abelian category A, and B a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Then any short exact sequence 0 ,2K[f ]  ,2 Ker f ,2B f  ,2A ,20 in AE induces a long exact homology sequence · · · ,2 Hn+1(A, I)E δn+1f ,2 K[Hn(f, I1)E1 ] γnf ,2 Hn(B, I)E Hn(f,I)E,2 Hn(A, I)E ,2 · · · · · · ,2 H2(A, I)E δ2f ,2 K[H1(f, I1)E1 ] γ1f ,2 H1(B, I)E H1(f,I)E ,2 H1(A, I)E ,2 0 (F) in B. This sequence is natural in f . 77 Chapter 4. Homology via Hopf Formulae Proof. A proof of this theorem in its full generality is given in [Eve2007]. However, when we restrict ourselves to the monadic case it becomes relatively easy to understand why the sequence takes this shape. So suppose that we are in a semi-abelian monadic setting, AE = ExtkA with B = CExtkBA, and G is the induced comonad on Ext kA. This comonad produces canonical simplicial resolutions GA and GB of A and B and, by functoriality, also a simplicial resolution Gf of f . The Everaert sequence (F) is the long exact homology sequence (see [EVdL2004b, Corollary 5.7]) obtained from the short exact sequence of simplicial objects 0 ,2K[IkGf ]  ,2 ,2IkGB IkGf ,2IkGA ,20; it remains to be shown that Hn−1K[IkGf ] = K[Hn(f, Ik+1)G] for all n ≥ 1. (Remember the dimension shift in Equation (A).) Now degree-wise, the (k + 1)-extension Ik+1Gf : Ik+1[Gf ] −→ GA is a split epimorphic central extension: it is a centralisation, and GA is degree-wise projec- tive. Via [EGVdL2008, Proposition 4.5], this implies that, degree-wise, it is a trivial exten- sion. This means that Ik+1Gf is the pullback of IkGf along the unit ηkGA : GA −→ IkGA, which in turn implies that K[IkGf ] is the kernel K[Ik+1Gf ] of Ik+1Gf . Since HnGA = 0 for all n ≥ 1, GA being a simplicial resolution, the long exact homology sequence induced by the short exact sequence of simplicial objects 0 ,2K[Ik+1Gf ]  ,2 ,2Ik+1[Gf ] Ik+1Gf ,2GA ,20 gives the needed isomorphism Hn−1K[Ik+1Gf ] ∼= K[Hn(f, Ik+1)G]. Note that in [Eve2007], this sequence has a slightly different appearance: there it contains the objects domHn(f, I1)E1 instead of K[Hn(f, I1)E1 ] for n ≥ 2. But the codomain of Hn(f, I1)E1 is zero (because J1f has zero codomain, hence I1 only changes the domain of an extension), so its domain coincides with its kernel. For us, the sequence in its present, more uniform, shape will be easier to work with. 4.4.7 Corollary: For any n ≥ 2 and any projective presentation p : P −→ A of an object A ∈ |AE|, Hn(p, I1)E1 ∼= (Hn+1(A, I)E −→ 0). Proof. It suffices to note that in the Everaert sequence (F), all Hn+1(P, I)E are zero, because P is projective. This shows how the degree of the homology may be lowered from n+1 to n by raising the degree of the reflector. 78 4.4 Hopf formulae It is well known that the integral group homology objects are abelian groups. The analogous result holds for the homology objects defined via Hopf formulae. 4.4.8 Lemma: Let n ≥ 1. For any object A ∈ AE, the homology object Hn+1(A, I)E is an abelian object. Proof. For a proof see [Eve2007, Proposition 2.3.16]. This result uses Lemma 1.2.7. When the Birkhoff subcategory B is the subcategory of abelian objects AbA, it is easy to see that all objects occurring in the Everaert sequence (F) are abelian objects. However, when B is not made up of abelian objects, then H1(A, I)E = IA will not in general be abelian. In the Everaert sequence, there is a nice bridge between the abelian and the non-abelian objects: the map from the last abelian object H2(A, I)E to the first non-abelian object K[I1f ] is central in the sense of Huq. This is here shown for the first time. 4.4.9 Lemma: Let f : B −→ A be an extension in A. Then δ2f : H2(A, I)E −→ K[H1(f, I1)E1 ] in the Everaert sequence (F) is central in the sense of Huq. Proof. We use some steps leading to the proof of Theorem 2.1 in [Bou2005], which is here quoted as Lemma 1.2.7 (without proof). First of all consider the image Im δ2f : I[δ 2 f ] −→ K[I1f ] = K[H1(f, I1)E1 ]. By Lemma 1.5.8 it is enough to show that Im δ2f is central. As the Everaert sequence is exact, this image is the kernel of γ1f : K[I1f ] −→ IB or equivalently the kernel of the corestriction of γ1f to its image, K[I1f ] −→ K[If ]. Now recalling the definition of I1 and J1 from 4.3.3, we see that K[I1f ] = K[f ] J1[f ] = K[f ] pi2(JR[f ] ∩K[f ]) since J1[f ] = JR[f ] ∩ K[f ] as a normal subobject of R[f ], and its direct image under pi2 gives us a normal subobject of B (note that pi2(J1[f ]) = J1[f ] as µ1f is a normal monomorphism). Similarly K[If ] = K[f ] JB ∩K[f ] = K[f ] pi2(JR[f ]) ∩ pi2(K[f ]) . 79 Chapter 4. Homology via Hopf Formulae Thus, by Noether’s Isomorphism Theorem, we have I[δ2f ] = JB ∩K[f ] J1[f ] = pi2(JR[f ]) ∩ pi2(K[f ]) pi2(JR[f ] ∩K[f ]) which is an abelian object by Lemma 1.2.7. But in fact, following the proof of this result, we see that JB ∩K[f ] J1[f ] = JB J1[f ] ∩ K[f ] J1[f ] = pi2(JR[f ]) pi2(JR[f ] ∩K[f ]) ∩ pi2(K[f ]) pi2(JR[f ] ∩K[f ]) as quotienting out by J1[f ] is a regular epimorphism. Now clearly JR[f ] JR[f ] ∩K[f ] ∩ K[f ] JR[f ] ∩K[f ] = 0 so by [Bou2005, Proposition 2.1] these two subobjects of R[f ]/(JR[f ] ∩ K[f ]) cooperate. Now we take the images under pi2, and [Bou2005, Proposition 1.1] implies that these images JB/J1[f ] and K[f ]/J1[f ] also cooperate, as subobjects of B/J1[f ]. Thus, using Lemma 1.5.7, we see that (JB ∩ K[f ])/J1[f ] and K[f ]/J1[f ] cooperate as subobjects of K[f ]/J1[f ] = K[I1f ], which says exactly that Im δ2f is central. 4.4.10 Remark: Notice that γ1f coincides with the composite in the diagram below, show- ing that the corestriction to the image is in fact K[(I1f, If)] : K[I1f ] −→ K[If ]. K[I1f ] _  γ1f HH H (H HHH  ,2 ,2 I1[f ] ηI1[f ] _  I1f  ,2 A ηA _  K[If ]  ,2 ,2 IB If  ,2 IA Here the map (I1f, If) : ηI1[f ] −→ ηA is a double extension, so its kernel is an extension by Lemma 4.1.7. Thus we have established that K[γ1f ] = JB ∩K[f ] J1[f ] which is reminiscent of the Hopf formula; the only difference is that here f is not a projective presentation and so this expression is not independent of the choice of f . The form of this kernel is not surprising, as when f is a projective presentation, the kernel of γ1f is exactly H2(A, I)E. 80 Chapter 5 Homology via Satellites Introduction Having defined homology via Hopf formulae and introduced the Everaert sequence in the previous chapter, we now use the universal properties of the Everaert sequence to define homology via pointwise Kan extensions or limits. Recall that any short exact sequence 0 ,2K[f ]  ,2 Ker f ,2B f  ,2A ,20 in A gives rise to a long exact homology sequence · · · ,2 Hn+1(A, I)E δn+1f ,2 K[Hn(f, I1)E1 ] γnf ,2 Hn(B, I)E Hn(f,I)E,2 Hn(A, I)E ,2 · · · · · · ,2 H2(A, I)E δ2f ,2 K[H1(f, I1)E1 ] γ1f ,2 H1(B, I)E H1(f,I)E ,2 H1(A, I)E ,2 0 which is natural in f . Janelidze’s theory of generalised satellites now helps us to compute homology objects step by step: the (n + 1)st homology functor Hn+1(−, I) is obtained from Hn(−, I1) as a pointwise right Kan extension, and the connecting homomorphism δ in the Everaert sequence is exactly what makes this work. This approach removes the dependence on projective objects from the definition of homology. In a further step it also cements the connection between homology and central extensions: gluing all the step by step Kan extensions together we see that homology is the limit of the diagram of kernels of all central extensions of a given object. More precisely, given an object A and the category CExtnAA of central n-extensions of A, we have Hn+1(A, I) = lim f∈CExtnAA Kn[f ] for any n ≥ 1. When the category A does have enough projective objects, we can use projective presentations to cut down the size of the diagram of which the homology object is the limit. Given a projective presentation p of A, the (n + 1)st homology Hn+1(A, I) forms the limit of the diagram consisting of the n-fold kernel of p and all maps induced by endomorphisms of p over A. This can be interpreted to say that calculating homology 81 Chapter 5. Homology via Satellites amounts to calculating common fixed points of these maps induced by endomorphisms of a projective presentation of A. The first four sections of this chapter give an analysis of homology in terms of satellites. We start by stating the main definitions in Section 5.1. Then, in Section 5.2, we interpret Hn+1(−, I)E (together with the connecting map δn+1) as a satellite of Hn(−, I1)E1 . In Section 5.3 we prove one of the main results of this chapter: a formula which gives Hn+1 in terms of In. Finally in Section 5.4 we explain how the situation is entirely symmetric, in that the connecting map γn also arises as a pointwise satellite. In the last two sections we discuss the theory obtained by defining homology via satel- lites. Section 5.5 gives this definition of homology without projectives and the result that homology is the limit of the diagram of kernels of central extensions, in Corollary 5.5.10. It also establishes that the homology objects are both objects of the Birkhoff subcategory B and abelian objects of A. In Section 5.6 we investigate the consequences of the new definition when enough projective objects are available. This leads to the interpretation of homology as calculating fixed points of endomorphisms of a projective presentation. Most material in this chapter is based on joint work with Tim Van der Linden and can also be found in our paper [GVdL2008a], though I use some different concepts and proof techniques here that make many statements and proofs easier. 5.1 Satellites and pointwise satellites Modulo a minor terminological change, the following definition is due to Janelidze. 5.1.1 Definition (Satellites): [Jan1976, Definition 2] Let I ′ : A′ −→ B′ be a functor. A left satellite (H, δ) of I ′ (relative to F : A′ −→ A and G : B′ −→ B) is a functor H : A −→ B together with a natural transformation δ : HF =⇒ GI ′ A′ F z    I′ $? ?? ?? ?? A H $ B′ Gz    B δ +3 universal amongst such, i.e., if there is another functor L : A −→ B with a natural trans- formation λ : LF =⇒ GI ′, then there is a unique natural transformation µ : L =⇒ H satis- fying δ◦µF = λ. This means that (H, δ) is the right Kan extension RanFGI ′ of the functor GI ′ along F : A′ −→ A. 82 5.1 Satellites and pointwise satellites This makes it possible to compute derived functors in quite diverse situations. The fol- lowing example, borrowed from [Jan1976], explains how satellites may be used to capture homology in the classical abelian case. 5.1.2 Example: In the abelian context, the (n + 1)st homology functor Hn+1 may be seen as a left satellite of Hn. For instance, let A = B′ and B be categories of modules and G : A −→ B an additive functor. Then G = H0(−, G). Let SESeqA be the category of short exact sequences 0 ,2K  ,2 k ,2B f  ,2A ,20 in A, the functor I ′ : SESeqA −→ A the projection pr1 that maps a sequence (k, f) to the objectK, and F : SESeqA −→ A the projection pr3 that maps (k, f) to A. LetH : A −→ B be the first homology functor H1(−, G). We obtain a satellite diagram SESeqA pr3 z    pr1 $? ?? ?? ?? A H1(−,G) $ A H0(−,G)z    B δ +3 where the natural transformation δ = (δ(k,f))(k,f)∈|SESeqA| consists of the connecting maps from the (classical) long exact homology sequence · · · ,2 H1K H1k ,2 H1B H1f ,2 H1A δ(k,f) ,2 H0K H0k ,2 H0B H0f ,2 H0A ,2 0. The universality of the Kan extension follows from the universality of the long exact homology sequence amongst similar sequences and may for instance be shown as follows. Given any functor L : A −→ B and any natural transformation λ : L◦pr3 =⇒ H0(−, G)◦pr1, we will construct the component at an object A ∈ |A| of the required natural transforma- tion L =⇒ H1(−, G) by using a projective presentation p : P −→ A of A. Let k : K −→ P be the kernel of this projective presentation of A. Since H1P is zero (as P is projective), the exactness of the long homology sequence induced by (k, p) means that δ(k,p) : H1A −→ H0K is the kernel of H0k. Then the string of equalities H0k◦λ(k,p) (1) = λ(1P ,!P )◦L!A (2) = H0(¡P )◦λ(10,10)◦L!A (3) = 0 83 Chapter 5. Homology via Satellites yields the required factorisation LA −→ H1A: (1) expresses the naturality of λ at the upper, downward-pointing morphism of the diagram 0 ,2 K  ,2 k ,2 k  ⇓ P 1P p  ,2 ⇓ A !A  ,2 0 0 ,2 P 1P ⇑ P !P  ,2 ⇑ 0 ,2 0 0 ,2 0 10 ¡P LR 0 ¡P LR 10 0 10 LR ,2 0 (G) in SESeqA, while (2) follows from λ(1P ,!P ) = λ(1P ,!P )◦L10 = H0(¡P )◦λ(10,10), which is the naturality of λ at the lower, upward-pointing morphism; the last equality (3) holds because H00 = 0. Note that, as such, this example does not follow the terminology of Definition 5.1.1. From its point of view one is tempted to call H a left satellite of G (rather than a satellite of I ′), and actually this is how the definition appears in the paper [Jan1976]. But the situation we shall be considering in this thesis demands the change in terminology, and the present example may easily be modified to comply with Definition 5.1.1. Indeed, the functor G may be lifted to a functor SSeqH0(−, G) : SESeqA −→ SSeqB where the latter category consists of short (not necessarily exact) sequences in B. Together with the obvious projection pr1 : SSeqB −→ B (s.t. H0(−, G)◦pr1 = pr1◦SSeqH0(−, G)), this gives us the satellite diagram SESeqA pr3 z    SSeqH0(−,G) $? ?? ?? ?? A H1(−,G) $ SSeqB. pr1z    B δ +3 Whereas such a viewpoint may seem rather far-fetched in the abelian case, it is the only one still available when the context is widened to semi-abelian categories. In fact, even in the abelian setting, this formulation is slightly reminiscent of the universal property for a derived functor (see for example [Wei1997, Section 10.5]), so it is not all that far-fetched after all. In practice, satellites may almost always be computed explicitly using limits—namely, as pointwise Kan extensions. Then the definition given above is strengthened as follows. 84 5.2 Hn+1(−, I)E as a satellite of Hn(−, I1)E1 5.1.3 Notation: Let A be an object of A. We denote by (A ↓ F ) the category of elements of the functor Hom(A,F−) : A′ −→ Set: its objects are pairs (A′, α : A −→ FA′), where A′ is an object of A′ and α is a morphism in A, and its morphisms are defined in the obvious way (cf. [Bor1994, Theorem 3.7.2]). The forgetful functor U : (A ↓ F ) −→ A′ maps a pair (A′, α) to A′. The natural transformation (H, δ) now induces a cone δ on GI ′U : (A ↓ F ) −→ B with vertex HA defined by δ(A′,α : A−→FA′) = δA′◦Hα : HA Hα ,2HFA′ δA′ ,2GI ′A′ = GI ′U(A′, α). 5.1.4 Definition (Pointwise satellites): A left satellite (H, δ) of I ′ relative to the functors F : A′ −→ A and G : B′ −→ B is called pointwise when it is pointwise as a Kan extension, i.e., for every object A of A, the cone (HA, δ) on GI ′U : (A ↓ F ) −→ B is a limit cone. To check that a pair (H, δ) is a pointwise satellite it is not necessary to prove its univer- sality as in Definition 5.1.1, but it suffices to check the limit condition from Definition 5.1.4; see, for example, MacLane [Mac1998, Theorem X.3.1]. 5.2 Hn+1(−, I)E as a satellite of Hn(−, I1)E1 We are now ready to prove the first main result of this chapter: we focus on the universal properties of the Everaert sequence (F), and prove that they allow us to interpret the (n+1)st homology with coefficients in I as a satellite of the nth homology with coefficients in I1. For the whole of this section, let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. 5.2.1 Lemma: For n ≥ 1 and A ∈ |AE|, K[Hn(!A : A −→ 0, I1)E1 ] = Hn(A, I)E. Proof. This follows from the exactness of the Everaert sequence (F) and the fact that all Hn(0, I)E are zero. 85 Chapter 5. Homology via Satellites 5.2.2 Lemma: For all n ≥ 1 and f : B −→ A ∈ |ExtA|, γnf = ker Hn  B f  ⇒ B !B  A !A ,2 0 , I1  E1  : K[Hn(f, I1)E1 ] −→ Hn(B, I)E. Proof. This follows from the previous lemma and the naturality of γn. Indeed, its natu- rality square at the map (1B, !A) is nothing but K[Hn(f, I1)E1 ] kerHn((1B ,!A),I1)E1 ,2 γnf  K[Hn(!B, I1)E1 ] γn!B  Hn(B, I)E Hn(B, I)E; and all kernels may be chosen in such a way that γn!B is an identity. 5.2.3 Proposition: Let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Let n ≥ 1. Then Hn+1(−, I)E : AE −→ AE with the connecting natural transformation ExtA cod z    Hn(−,I1)E1 $? ?? ?? ?? ? AE δn+1 +3 Hn+1(−,I)E $ ExtA kerz    AE (H) is the pointwise left satellite of Hn(−, I1)E1. That is, for any object A of AE, Hn+1(A, I)E = Rancod(ker◦Hn(−, I1)E1)(A) = lim (f,g)∈|(A↓cod)| K[Hn(f, I1)E1 ]. Proof. Let A be an object of AE. Let p : P −→ A be a projective presentation of A. We have to show that (Hn+1(A, I)E, δn+1) is the limit of (A ↓ cod) U ,2ExtA Hn(−,I1)E1 ,2ExtA ker ,2AE. To do so, let (L, λ) be another cone on ker◦Hn(−, I1)E1◦U ; we use the presentation p of A to construct a map of cones l : L −→ Hn+1(A, I)E. First we consider the case n = 1. Recall from Definition 4.4.3 that H1(−, I)E = I and H1(−, I1)E1 = I1. Since p : P −→ A is a projective presentation of A, and thus H2(P, I)E = 0, the lower end of the Everaert sequence (F) of p becomes 86 5.2 Hn+1(−, I)E as a satellite of Hn(−, I1)E1 0 ,2H2(A, I)E  ,2 δ2p ,2K[I1p] γ1p ,2IP. In other words, δ2p is the kernel of γ 1 p . Recalling Diagram (G), consider the following two morphisms in (A ↓ cod): P p  ,2 ⇓ A !A  1A ⇓ A P !P  ,2 ⇑ 0 ⇑ A !Alr 0 ¡P LR 10 0 10 LR A !A lr (I) By Lemma 5.2.2, the naturality of λ at the downward-pointing morphism in Diagram (I) means γ1p◦λ(p,1A) = λ(!P ,!A). This latter morphism is zero, since the naturality of λ at the upward-pointing morphism in (I) means λ(!P ,!A) = I(¡P )◦λ(10,!A), and I0 = 0. Hence there exists a unique morphism l : L −→ H2(A, I)E satisfying λ(p,1A) = δ2p◦l. Higher up in the Everaert sequence (F) of p, for n ≥ 2, Corollary 4.4.7 gives us the isomorphism δn+1p : Hn+1(A, I)E ∼=−→ K[Hn(p, I1)E1 ]. Here we may simply put l = (δn+1p ) −1◦λ(p,1A). It remains to be shown that, in both cases, the constructed map l is a map of cones. Given any object (f : B −→ C, g : A −→ C) of (A ↓ cod), there is a map P p  ,2  ⇓ A g  ⇓ A B f  ,2 C Ag lr as P is projective. Writing h for the image of this morphism under ker◦Hn(−, I1)E1◦U , we see that the diagram L l ,2 λ(p,1A)  λ(f,g) OOOO #+OO OOO OOO OO Hn+1(A, I)E δp oo ooo s{ooo ooo o δ(f,g)  K[Hn(p, I1)E1 ] h ,2 K[Hn(f, I1)E1 ] commutes: λ(f,g) = h◦λ(p,1A) = h◦δp◦l = δ(f,g)◦l. Thus l is indeed a map of cones, and Hn+1(A, I)E is the limit of the given diagram. 5.2.4 Remark: This gives a way to derive Hn+1(−, I)E from Hn(−, I1)E1 for n ≥ 2 in exactly the same way as H2(−, I)E is derived from H1(−, I1)E1 = I1. In other approaches such as [Eve2007, EGVdL2008] the two cases are formally different. 87 Chapter 5. Homology via Satellites 5.2.5 Remark: Notice that we can apply this proposition for higher extensions as well, making use of the fact that CExtkBA is a strongly E k-Birkhoff subcategory of ArrkA for k ≥ 1. This allows us to use induction, as we will see in the next subsection. 5.3 Hn+1(−, I)E as a satellite of In Proposition 5.2.3 gives a way to construct Hn+1(−, I)E from Hn(−, I1)E1 . Here, with The- orem 5.3.2, we obtain a one-step construction of Hn+1(−, I)E out of In. To be able to apply Proposition 5.2.3 repeatedly, we have to show that satellite diagrams like Diagram (H) may be composed in a suitable way (cf. [Jan1976, Theorem 9]). The kernel functor ker : Extk+1A −→ ExtkA that maps an extension f : B −→ A to its kernel K[f ] has a left adjoint, namely the functor ExtkA −→ Extk+1A that sends an object C of ExtkA to the extension !C : C −→ 0. This allows us to use the following result. 5.3.1 Proposition: Suppose that (I ′, δ′) = RanF ′G′I ′′ and (H, δ) = RanFGI ′ as in the diagrams A′ F z    I′ $? ?? ?? ?? ? A H $ δ +3 B′ Gz    B and A′′ F ′ z    I′′ $? ?? ?? ?? ? A′ I′ $ δ′ +3 B′′. G′z    B′ If G is a right adjoint then (H,Gδ′◦δF ′) = RanFF ′GG′I ′′: the two diagrams may be composed to form a single Kan extension diagram A′′ FF ′ z    I′′ $? ?? ?? ?? ? A H $ Gδ′◦δF ′ +3 B′′. GG′z    B If G preserves limits and (I ′, δ′) and (H, δ) are pointwise satellites then (H,Gδ′◦δF ′) is also a pointwise satellite. Proof. We prove the pointwise case. Let A be an object of A, and (C, σ) a cone on the diagram GG′I ′′U : (A ↓ FF ′) −→ B. 88 5.3 Hn+1(−, I)E as a satellite of In For any A′ in A′, the pair (I ′A′, δ′) is the limit of the diagram G′I ′′U ′ : (A′ ↓ F ′) −→ B′. Since G preserves limits, (GI ′A′, Gδ′) is the limit of GG′I ′′U ′ : (A′ ↓ F ′) −→ B. Now for every α : A −→ FA′ the collection (σ(A′′,Fα′◦α))(A′′,α′)∈|(A′↓F ′)| also forms a cone on GG′I ′′U ′; hence there is a unique map µ(A′,α) : C −→ GI ′A′ such that Gδ′A′′◦GI ′α′◦µ(A′,α) = σ(A′′,Fα′◦α). The collection (µ(A′,α))(A′,α)∈|(A↓F )| in turn forms a cone on GI ′U : (A ↓ FF ′) −→ B. Indeed, if (B′, β) is an object of (A ↓ F ) and f ′ : B′ −→ A′ is a map in A′ such that Ff ′◦β = α, then GI ′f ′◦µ(B′,β) = µ(A′,α), because for every (A′′, α′) ∈ |(A′ ↓ F ′)|, Gδ′A′′◦GI ′α′◦GI ′f ′◦µ(B′,β) = σ(A′′,F (α′◦f ′)◦β) = σ(A′′,Fα′◦α) = Gδ′A′′◦GI ′α′◦µ(A′,α), and the Gδ′A′′◦GI ′α′ are jointly monic. This cone gives rise to the required unique map c : C −→ HA. Since it satisfies µ(A′,α) = δA′◦Hα◦c for all (A′, α) ∈ |(A ↓ F )|, we have that Gδ′A′′◦δF ′A′′◦Hα ′′◦c = Gδ′A′′◦δF ′A′′◦HFα ′◦Hα◦c = Gδ′A′′◦GI ′α′◦δA′◦Hα◦c = Gδ′A′′◦GI ′α′◦µ(A′,α) = σ(A′′,Fα′◦α) = σ(A′′,α′′) for all α′′ = Fα′◦α : A −→ FA′ −→ FF ′A′′ in |(A ↓ FF ′)|—and any α′′ allows such a decomposition. 5.3.2 Theorem: Let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Let n ≥ 1. Then Hn+1(−, I)E : AE −→ AE with the connecting natural transformation ∂n+1 = kern−1δ2◦ · · · ◦kerδn◦δn+1 : Hn+1(−, I)E◦ codn =⇒ kern◦In 89 Chapter 5. Homology via Satellites is the pointwise left satellite of In. ExtnA codn z    In $? ?? ?? ?? ? AE Hn+1(−,I)E $ ∂n+1 +3 ExtnA kernz    AE This means, for any object A of A, Hn+1(A, I)E = Rancodn(kern◦In)(A) = lim (f,g)∈|(A↓codn)| Kn[Inf ]. Proof. This follows from gluing diagrams as in Proposition 5.2.3 together using Proposi- tion 5.3.1 and Remark 5.2.5. 5.4 Hn(−, I)E as a right satellite of Hn(−, I1)E1 Proposition 5.2.3 gives an interpretation of the connecting morphisms δnf in the Everaert sequence as left satellites. The connecting morphisms γnf have a dual interpretation: (Hn(−, I)E, γn) is a right satellite (left Kan extension) of Hn(−, I1)E1 . 5.4.1 Proposition: Let E be a class of extensions in a semi-abelian category A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Consider n ≥ 1. Then (Hn(−, I)E, γn), i.e., Hn(−, I)E : AE −→ AE with the connecting natural transformation ExtA Hn(−,I1)E1 z    dom $? ?? ?? ?? ExtA ker $? ?? ?? ?? γn +3 AE Hn(−,I)Ez AE is the pointwise right satellite of Hn(−, I1)E1. Proof. For any A, the category (dom↓A) has a terminal object (!A : A −→ 0, 1A), so the colimit object of the diagram (dom↓A) U ,2 ExtA Hn(−,I1)E1 ,2 ExtA ker ,2 AE 90 5.5 Homology without projectives is K[Hn(!A, I1)E1 ] = Hn(A, I)E. The component of the colimit cocone at (g : B −→ C, f : B −→ A) ∈ |(dom↓A)| is ker Hn  B f ,2 g  ⇒ A !A  C !C ,2 0 , I1  E1  = ker Hn  B f ,2 !B  ⇒ A !A  0 0 , I1  E1  ◦ker Hn  B g  ⇒ B !B  C !C ,2 0 , I1  E1  = Hn(f, I)E◦γng = γn(g,f) by Lemma 5.2.2 and Lemma 5.2.1. 5.5 Homology without projectives In this section we set up a homology theory without projectives by defining homology via pointwise satellites as they appear in Proposition 5.2.3. 5.5.1 Proposition: Let E be a class of extensions in a semi-abelian category A, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Let k ≥ 0, and consider an object A ∈ |AE|. If it exists, write H(2,k) = Rancod(ker◦Ik+1) for the pointwise left satellite of Ik+1 relative to the functors cod and ker. Now suppose H(n,k+1) exists for n ≥ 2, and write H(n+1,k) = Rancod(ker◦H(n,k+1)) for the pointwise left satellite of H(n,k+1) relative to cod and ker, if this exists. Then H(n+1,k) is also the left satellite of In relative to the functors codn and kern. Proof. The proof is the same as the proof of Theorem 5.3.2. 5.5.2 Definition (Homology): Let E be a class of extensions in a semi-abelian category A, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Con- sider an object A ∈ |AE|, and let n ≥ 1. If the functor H(n+1,0) from Proposition 5.5.1 exists, we call it the (n+ 1)st homology functor Hn+1(−, I) : AE −→ AE 91 Chapter 5. Homology via Satellites (with coefficients in I). ExtA cod z    Hn(−,I1) $? ?? ?? ?? AE δn+1 +3 Hn+1(−,I) $ ExtA kerz    AE We also write H1(−, I) = I. 5.5.3 Remark: As CExtkBA is strongly E k-Birkhoff in ArrkA, Definition 5.5.2 also gives us functors Hn+1(−, Ik) : ExtkA −→ ExtkA. These are sometimes needed directly, but whenever possible we will include these higher-dimensional cases in the setting of strongly (extension)-Birkhoff subcategories, to make statements and proofs easier. 5.5.4 Remark: For any object A ∈ |AE|, if H2(A, I) exists, it is the limit object of the diagram (A ↓ cod) U ,2 ExtA I1 ,2 ExtA ker ,2 AE. Similarly, if Hn+1(A, I) exists, it is the limit object of the diagram (A ↓ cod) U ,2 ExtA Hn(−,I1)E1 ,2 ExtA ker ,2 AE or equivalently of (A ↓ codn) U ,2ExtnA In ,2ExtnA kern ,2AE. (J) Potentially, these limits may exist for a given object A even if the homology functors Hn+1(−, I) do not exist in full. 5.5.5 Example (When the reflection is the identity): If B = A then all In are iden- tity functors, and the Hn are zero for n ≥ 2. To see this, we have to prove that the functor 0: AE −→ AE is a pointwise Kan extension of ker : ExtA −→ AE along cod: ExtA −→ AE, for all k ≥ 0. This shows that H2 is zero, which immediately implies that the higher homologies are also zero, being satellites of the zero functor. 92 5.5 Homology without projectives Let A be an object of AE and (L, λ) a cone on ker◦U : (A ↓ cod) −→ AE. Then any map λ(f,g), where (f : B −→ C, g : A −→ C) ∈ |(A ↓ cod)|, fits into the commutative diagram L λ(f,g) ,2 λ(!0,!A)  λ(!B,!A) ?? ?? $? ?? ? K[f ] _  Ker f  0 ,2 B, which means that λ(f,g) is the zero map. If now (L, λ) is a limit cone, this implies that L is zero. The category (A ↓ cod) is rather large, and in a given situation it may be very hard to decide whether the needed limits do indeed exist. Even if they do, they may still be hard to compute. But we may replace the above diagrams with simpler ones, for example using the concept of an initial subcategory. Recall its definition as it occurs in [Mac1998, Section IX.3]: 5.5.6 Definition: An initial functor is a functor F : D −→ C such that for every object C of C, the slice category (F ↓ C) is non-empty and connected. A subcategory D of a category C is called initial when the inclusion of D into C is an initial functor, i.e., for every object C ∈ |C|, the full subcategory (D ↓ C) of (C ↓ C) determined by the maps D −→ C with domain D in D is non-empty and connected. If D is initial in C then limits of diagrams over C may be computed as the limit of their restriction to D. More generally, if F : D −→ C is initial then a diagram G : C −→ E has a limit if and only if GF does, in which case it may be computed as the limit of GF . For any object A of AE, let ExtAA denote the category of extensions of A, the preimage in ExtA of the arrow 1A under the functor cod: ExtA −→ AE. Then the functor U ′ : ExtAA −→ (A ↓ cod) that sends an extension f : B −→ A of A to the pair (f, 1A) is easily seen to be initial: for every object (f : B −→ C, g : A −→ C) of (A ↓ cod) there is the natural morphism U ′f −→ (f, g) B  f  ,2 ⇓ A g  1A ⇓ A B f  ,2 C A,g lr where f is the pullback of f along g; this f is an extension by Definition 4.1.2(3). 93 Chapter 5. Homology via Satellites Also, any other morphism D  h  ,2 ⇓ A g  1A ⇓ A B f  ,2 C A,g lr factors over this morphism U ′f −→ (f, g), by the universal property of a pullback. This means that the limit of ker◦Hn(−, I1)E1◦U may also be computed as the limit of the dia- gram ker◦Hn(−, I1)E1◦UU ′ and moreover, since UU ′ is just the inclusion of the subcategory ExtAA into ExtA, as the limit of ker◦Hn(−, I1)E1 : ExtAA −→ AE. But even now the diagram of shape ExtAA over which the limit is computed may be too large, in the sense that even if A is small-complete, it is still unclear whether the limit of ker◦Hn(−, I1)E1 exists. In the case where A has enough projectives, however, it is possible to further cut down the size of this diagram. In this case Proposition 5.2.3 shows that the limit of this diagram exists and is equal to the homology object defined via the Hopf formulae. But making the diagram smaller gives a new way to calculate this homology. This situation is discussed in Section 5.6. 5.5.7 Notation: Let A ∈ |AE|. Denote by ExtnAA the category of n-extensions of A, defined as the preimage of the arrow 1A under the functor codn : ExtnA −→ AE. This generalises the category ExtAA of extensions of A defined above. Thus the objects are n-extensions with “terminal object” A, when viewed as diagrams in the category A, and the maps are those maps in ExtnA which restrict to the identity on A under codn. Sim- ilarly the category CExtnAA denotes the full subcategory of Ext n AA determined by those n-extensions which are central. The strongly E-Birkhoff subcategory B is understood, and not mentioned in the notation. 5.5.8 Remark: The functor U ′ : ExtnAA −→ (A ↓ codn) which sends an n-extension f of A to (f, 1A) is still initial. This may be shown by induction, using the fact that in a category of n-fold extensions, the (n+ 1)-extensions are pullback-stable (see Definition 4.1.2(3)). 5.5.9 Proposition: Let E be a class of extensions in a semi-abelian category A, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Consider n ≥ 1 and A ∈ |AE|. If it exists, Hn+1(A, I) is also the limit of the diagram kern◦In : ExtnAA −→ AE. Proof. This uses Diagram (J) and the fact that U ′ : ExtnAA −→ (A ↓ codn) is initial. 94 5.5 Homology without projectives 5.5.10 Corollary: For n ≥ 1 and A ∈ |AE|, if it exists, Hn+1(A, I) is the limit of the diagram kern : CExtnAA −→ AE. Proof. The functor In : ExtnAA −→ CExtnAA is initial because, for any central extension f ∈ |CExtnAA|, we have Inf = f , so the slice category (In ↓ f) is non-empty and connected. Since limits commute with kernels, Corollary 5.5.10 also says that Hn+1(A, I) may be computed as the n-fold kernel of a certain n-fold arrow in A, namely, the limit in ArrnA of the inclusion of CExtnAA into Arr nA. Sometimes this n-fold arrow in A itself happens to be an n-fold central extension of A. We say that an n-fold central extension of an object A ∈ |AE| is universal when it is an initial object of CExtnAA. We will see in Lemma 6.2.3 that, when A is a semi-abelian category and I = ab: A −→ AbA is the abelianisation functor, then an object A of A admits a universal central extension p if and only if it is perfect: its abelianisation is zero. In fact, this result holds for any reflector I to a Birkhoff subcategory B of A, not just the abelianisation functor (see [CVdL2009, Theorem 2.9]). In this case, H2(A, I) is the kernel of p. This latter property holds in general, also for higher extensions: 5.5.11 Corollary: Consider n ≥ 1 and A ∈ |AE|. If A has a universal n-fold central extension p then Hn+1(A, I) = Kn[p]. In particular, if A ∈ |A| has a universal central extension p : P −→ A then H2(A, I) = K[p]. Proof. The limit of a functor from a category that has an initial object is the value of the functor at this object. 5.5.12 Example (The homology of zero is zero): If A = 0 then, for any n ≥ 1, the category CExtnAA has an initial object, the zero n-cube. Taking kernels as in Corol- lary 5.5.11 gives Hn+1(0, I) = 0. 5.5.13 Remark: Note that in certain special cases a weakly universal extension can also determine the homology of an object A. When 1A is a weakly universal extension of A, i.e., if every extension f : B −→ A of A is split, we have H2(A, I) = 0. This is be- cause K[I1A] = 0 for any object A, so if 1A is weakly initial, every leg of a cone over ker◦I : ExtAA −→ AE factors over K[I1A] and thus is zero. In particular, we get: 5.5.14 Example (The homology of a projective object is zero): For any projective object P and any n ≥ 1 we have Hn+1(P, I) = 0, since 1P (and also the n-extension only consisting of the maps 1P ) is always weakly initial when P is projective. 95 Chapter 5. Homology via Satellites 5.5.15 Example (Homology of finite groups): For a finite group, we compare its second homology groups with respect to two different adjunctions. On the one hand we have the abelianisation functor ab: Gp −→ AbGp, where Gp is the category of groups, AbGp is the Birkhoff subcategory of abelian groups, and abG = G/[G,G]. This example has been studied in the classical setting in [EVdL2004b] (for lower dimensions) and in [Eve2007, EGVdL2008] (higher dimensions). Here the centralisation functor ab1 takes an extension f : B −→ A to centr f : B/[K[f ], B] −→ A. As mentioned in Example 1.3.6, in this case Definition 5.5.2 gives the classical integral homology of groups. On the other hand, we could focus on finite groups and let A = FinGp be the category of finite groups and B = FinAb = AbFinGp its Birkhoff subcategory of finite abelian groups. Note that FinGp is not semi-abelian and doesn’t have enough projectives, but nevertheless it is pointed, Barr exact and Bourn protomodular. All the results that we apply in this example do not use coproducts, so they are still valid in this context. Here I : A −→ B again sends a group G to finabG = G/[G,G] and I1 : ExtFinGp −→ ExtFinGp sends an extension f : B −→ A to fincentr f : B/[K[f ], B] −→ A. We show that, for any finite group, its second homology groups with respect to the two theories coincide. For perfect groups this is clear. Recall from Corollary 5.5.11 that if a group G has a universal central extension p : P −→ G, then the homology is H2(G, ab) = K[p]; this is the case when G is perfect: abG = 0. So given a finite perfect group G, we know that it has a universal central extension p : P −→ G in the category Gp of all groups, and that H2(G,Z) = H2(G, ab) = K[p]. But we also know that the integral homology of a finite group is a finite group, therefore the group P must also be finite, and the universal central extension p : P −→ G lies in the category FinGp of finite groups. Thus we also have H2(G,finab) = K[p]. So for a finite perfect group G we have H2(G,finab) = H2(G, ab) = H2(G,Z). For a general group, we need a few more steps to prove this equality. Step 1: First we want to show that, for any finite group G, there is a central extension G∗ −→ G with kernel H2(G,Z), such that in the diagram ker : CExtGGp −→ Gp, (K) the leg from the limit H2(G, ab) to this object is an isomorphism. We consider stem extensions: central extensions g : H −→ G with K[g] ≤ [H,H]. This condition implies that abH −→ abG is an isomorphism, or equivalently that the map K[g] −→ abH is zero. So it follows from exactness in (F) that the leg H2(G, ab) −→ K[g] is a surjection when g is a stem extension. To find a stem extension with H2(G,Z) as its kernel, we use the Schur multiplier M(G) of a finite group G introduced in [Sch1904]. Schur proved in [Sch1907] 96 5.5 Homology without projectives that for a finite group G, this multiplier M(G) may be expressed in terms of what is now called the Hopf formula (which, in the infinite case, was only introduced in [Hop1942]), and so we have M(G) ∼= H2(G,Z) (see also, e.g., [Kar1987, Theorem 2.4.6]). In [Sch1904] he showed that, for any finite group G, there is a stem extension f : G∗ −→ G of G with kernel M(G) (see also [Kar1987, Theorem 2.1.4]). Putting these two facts together, we see that H2(G,Z) occurs in the diagram (K) as the kernel of this stem extension f , and that the leg from H2(G, ab) to it must be an isomorphism, being a surjection between finite groups of the same size. From now on we shall assume that this isomorphism is an identity. Step 2: We now consider the diagram of kernels of finite central extensions of G, ker : CExtGFinGp −→ FinGp, (L) which is a small diagram and so has a limit in Gp which we denote by L. We shall show in Step 3 that L ∼= H2(G, ab) and so is actually the limit of (L) in the category FinGp as well, as H2(G, ab) is a finite group. H2(G, ab) forms a cone on (L), using the legs from (K). The induced map of cones to L gives a splitting for the leg p : L −→ H2(G,Z) = K[f ]. As these are all abelian groups, we have L ∼= H2(G,Z) ⊕ E for some abelian group E, and p = pi1 : L −→ H2(G,Z), the first projection. We consider the following central extensions and maps between them: H2(G,Z)  ,2 ,2_   G∗ f  ,2 _  (1G∗ ,0)  G H2(G,Z)⊕ E p _LR  ,2 ,2 pi2 _  G∗ × E pi1 _LR f◦pi1 ,2 f×1E _  G E  ,2 ,2 G× E pi1  ,2G Since the extension pi1 : G× E −→ G is split, the leg from L to E = K[pi1] must be the zero map. So the leg from L to K[f◦pi1] is 1H2(G,Z) ⊕ 0: L ∼= H2(G,Z)⊕ E −→ H2(G,Z)⊕ E, as H2(G,Z)⊕ E is a product. Step 3: Finally we consider a third, even smaller diagram. Let C be the full subcategory of CExtFinGp containing those extensions g of G for which there exists a map f −→ g in CExtGFinGp. We consider the subdiagram ker : C −→ FinGp, (M) the limit of which is H2(G, ab). For any cone D over the diagram (M), the two legs d : D −→ H2(G,Z) = K[f ] and 0: D −→ E = K[pi1] again determine the leg to the prod- 97 Chapter 5. Homology via Satellites uct, (d, 0) : D −→ H2(G,Z)⊕ E = K[f◦pi1]. The leg d also forms the unique cone map D −→ H2(G, ab). Notice that in (M) we also have maps from H2(G,Z)⊕E to any other ob- ject, as p : H2(G,Z)⊕ E −→ H2(G,Z) is part of the diagram. So as (1H2(G,Z)⊕ 0)◦(d, 0) = (d, 0), the map (d, 0) : D −→ L is a cone map and makes L into a limit of (M). So L ∼= H2(G, ab) as promised, and we have H2(G,finab) = H2(G, ab) = H2(G,Z) for any finite group G. It is well known that all integral homology groups of a group are abelian. More generally, both approaches to homology discussed in Chapter 1 are such that the homology objects are abelian objects of the Birkhoff subcategoryB. We now prove that our homology objects Hn+1(A, I) also satisfy these properties. 5.5.16 Lemma: Consider an object A ∈ |AE|. The kernel K[f ] of a central extension f : B −→ A of A is an object of the strongly E-Birkhoff subcategory B. Proof. Let A ∈ |AE|. First consider a trivial extension f : B −→ A. This means f is the pullback of If : IB −→ IA along ηA, so K[f ] is isomorphic to K[If ]. This kernel of the extension If : IB −→ IA is an object of B because B is closed under subobjects (see Remark 4.2.2). Now for a central extension f : B −→ A, recall from Definition 1.4.3 that there exists an extension g such that the pullback f of f along g is trivial. K[f ]  ,2 ,2 B f  ,2 _  A g _  K[f ]  ,2 ,2 B f  ,2 A But then K[f ] = K[f ], which is an object of B as f is trivial. 5.5.17 Remark: The converse implication does not hold, as for example in the category of groups not every extension with abelian kernel is central. 5.5.18 Proposition: Let A be an object of AE and n ≥ 0. Then Hn+1(A, I) is an object of B. Proof. If n = 0 the result is clear as H1(A, I) = IA. For n ≥ 1, we use Lemma 5.5.16 repeatedly to see that the diagram from Corollary 5.5.10 factors over B and becomes the functor kern : CExtnAA −→ B. Since B is closed under limits in A, the limit Hn+1(A, I) of this diagram is still an object of B. 98 5.5 Homology without projectives 5.5.19 Example (When the reflection is zero): If B = 0, the zero subcategory in A, then all homology objects are zero, because they are in B by Proposition 5.5.18. The proofs of the next result Proposition 5.5.22 and its lemma were offered to us by Tomas Everaert. Recall that an object A of a homological category A is abelian if it carries an internal abelian group structure. Such a structure is necessarily unique, and is given by a morphism m : A×A −→ A satisfying m◦(1A, 0) = 1A = m◦(0, 1A), called its addition (see Definition 1.2.1). The abelian objects form a Birkhoff subcategory AbA of A. 5.5.20 Lemma: For any extension f : B −→ A in A, the image of the connecting mor- phism δ2f : H2(A, I) −→ K[H1(f, I1)] = K[I1f ] is an abelian object of A. Proof. We show that I[δ2f ] is a subobject of an abelian object in A, namely the kernel of γ1f . Recall from Remark 4.4.10 that K[γ1f ] = JB ∩K[f ] J1[f ] = pi2(JR[f ]) ∩ pi2(K[f ]) pi2(JR[f ] ∩K[f ]) and thus is abelian by Lemma 1.2.7. Now the composite H2(A, I) δ2f ,2 K[I1f ] γ1f ,2 IB = H1(B, I) is zero, as it is the leg from the limit H2(A, I) to the kernel IB = K[I1!B] of a split central extension. Thus the image of δ2f factors over the kernel of γ 1 f , and I[δ 2 f ] is abelian, as a subobject of the abelian object K[γ1f ] = (JB ∩K[f ])/J1[f ]. 5.5.21 Remark: Notice that we can not assume any more that I[δ2f ] = K[I1f ], as we did in Lemma 4.4.9. As we are using a different definition of homology, we can not assume that the Everaert sequence is exact. It is however still a complex, and the map γ1f does not change, so K[γ1f ] is still the same abelian object. Note that Lemma 1.5.8 still implies that δ2f is central in the sense of Huq, as it factors over the central morphism Ker γ 1 f . Of course, as soon as we are working in a category with enough projectives, the two definitions of homology coincide and the Everaert sequence is exact even when using the definition via limits. 99 Chapter 5. Homology via Satellites 5.5.22 Proposition: Let A be an object of A and n ≥ 1. Then Hn+1(A, I) is an abelian object of A. Proof. It suffices to show that, for any A ∈ |AE|, the object H2(A, I) is abelian in A. We can then use this fact also in higher dimensions when B = CExtkBA, so then the higher homology objects are limits of a diagram of abelian objects, and thus abelian by induction. To show H2(A, I) is abelian, consider the functor H2(−, I)×H2(−, I) : AE −→ AE that sends an object A to the product H2(A, I) × H2(A, I). The previous lemma gives rise to a natural transformation (H2(−, I)×H2(−, I)) ◦ cod =⇒ ker ◦ I1 of functors from ExtA to AE; the component of this natural transformation at an extension f : B −→ A is the composition H2(A, I)×H2(A, I) −→ I[δ2f ]× I[δ2f ] −→ I[δ2f ] −→ K[I1f ]. Here the first arrow is the corestriction of δ2f × δ2f , the second arrow is the addition on the abelian object I[δ2f ], and the last arrow is the inclusion of the image into the codomain of δ2f . The universal property of the Kan extension (H2(−, I), δ2) now yields a natural transformation H2(−, I)×H2(−, I) =⇒ H2(−, I) which is easily seen to define an abelian group structure on all H2(A, I). 5.5.23 Example (Heyting semi-lattices): As mentioned in Example 1.1.4, Heyting semi-lattices form a semi-abelian category (see [Joh2004]). It would be interesting to study homology there, as traditionally this setting is not connected to any homology theories. However, the previous proposition implies that any semi-abelian homology theory in the category of Heyting semi-lattices is necessarily trivial, as the only abelian object is the zero object. We will prove this fact. For the axioms of Heyting semi-lattices see for example [Joh2004] or [Joh2002, A 1.5.11]. Given a morphism f : A −→ B in the category of Heyting semi-lattices which has equal composite with (1A, 0) : A×A −→ A and (0, 1A) : A×A −→ A, we show that f has to be the zero map. This implies that any abelian object is zero. The condition on f can be written as f(>, a) = f(a,>) ∀a ∈ A. We have f(>, a) = f((a⇒ a), (> ⇒ a)) = f((a,>)⇒ (a, a)) = (f(a,>)⇒ f(a, a)) 100 5.6 Homology with projectives so writing b = f(>, a) = f(a,>) and c = f(a, a), we have b = (b⇒ c). But then b = b ∧ (b⇒ c) = b ∧ c and c ∧ b = c ∧ (b⇒ c) = c, so b = c and we have b = (b⇒ b) = >. Therefore, translating back, we have f(>, a) = f(a,>) = > and f(a, a) = > for all a ∈ A. We then see by f(b, a) = (> ⇒ f(b, a)) = (f(b,>)⇒ f(b, a)) = f(>, a) = > that f is the zero map, as claimed. 5.6 Homology with projectives In this section we investigate our new definition of homology in the situation when A does have enough projectives. In this case we know that homology exists, for example via Everaert’s definition using the Hopf formulae, and Proposition 5.2.3 shows that it coincides with the notion introduced in Definition 5.5.2. But by reducing the size of the diagram which defines the homology objects, we obtain a new way to calculate homology. Our main aim is to show Theorem 5.6.6 which states that the (n + 1)st homology of an object A may be computed as a limit over the category Endp of all endomorphisms of an n-presentation p of A. 5.6.1 Notation: For any n-extension f of an object A ∈ |AE|, let Endf , the category of endomorphisms of f over A, be the full subcategory of ExtnAA determined by the object f . Thus maps in Endf are maps from f to itself which restrict to the identity on A under the functor codn. When A has enough projectives we can view Proposition 5.2.3 the other way round: 5.6.2 Theorem (Hopf formula): Let E be a class of extensions in a semi-abelian cate- gory A with enough projectives, and let B be a strongly E-Birkhoff subcategory of A with reflector I : AE −→ B. Let n ≥ 1. Given an n-fold presentation p of an object A ∈ |AE| with initial object Pn, we have Hn+1(A, I) ∼= JPn ∩K n[p] Kn[Jnp] . 101 Chapter 5. Homology via Satellites Proof. This is just Proposition 5.2.3 viewed from the perspective of Definition 5.5.2. 5.6.3 Remark: In [Eve2007, Eve2008] Everaert gives a direct proof that the right hand side of the Hopf formula is a Baer invariant of A: an expression independent of the chosen n-fold presentation p of A (see also [EVdL2004a, Fro¨1963]). More precisely, any morphism p −→ p over A induces the identity on (JPn ∩Kn[p])/Kn[Jnp]. Of course we can still calculate homology as a limit, as defined in Section 5.5. It turns out that in this case, homology may also be computed as a limit over the small subdiagram of shape Êndp, which is a subcategory of (A ↓ codn). 5.6.4 Notation: Let p be an n-presentation of a k-extension A, with initial object Pn. Let ιnPn be the n-cube which has initial object Pn and all other objects zero. The category Êndp is inspired by a higher-dimensional variation of Diagram (I): it is the subcategory of (A ↓ codn) that is generated by the objects (p, 1A), (ιnPn, !A) and (0, !A), all endomorphisms of p over A, and the three maps p f  A !A  1A ⇓ A 1A ιnPn  0  ⇑⇓ A !Alr 0 LR 0 LR A !A lr in (A ↓ codn). Here f is the identity on Pn and obvious everywhere else, and the right side of the diagram displays the maps from A to the “terminal” object of the n-cube depicted on the left. Note that there is an obvious inclusion Endp −→ Êndp sending p to (p, 1A). 5.6.5 Proposition: Consider n ≥ 1 and A ∈ |AE|, and let p be an n-fold presentation of A with initial object Pn. Then Jp ∩Kn[p] Kn[Jnp] = lim ( kern◦In◦U : Êndp −→ AE ) . Proof. The diagram we are considering is Kn[p] Kn[Jnp] = Kn[Inp] kern(Inf)=f¯ ,2Kn[InιnPn] = IPn ,20lr Notice that the initial object of the n-cube Inp is Pn/Kn[Jnp]. We will show that the object (Jp∩Kn[p])/Kn[Jnp] is the kernel of the map f¯ , which will in turn imply that it is indeed the limit of our diagram. 102 5.6 Homology with projectives Forming the cokernel Q of Kn[p] −→ Pn, we construct the following maps: JPn ∩Kn[p]  ,2 ,2_   JPn_    ,2 JQ_   Kn[p] _   ,2 ,2 Pn  ,2 _  Q _  Kn[p] JPn∩Kn[p]  ,2 ,2 IPn  ,2 IQ Here the columns and the first two rows are short exact sequences, and the top left square is a pullback because JQ −→ Q is a mono. The last row clearly composes to give the zero-map, so using the 3×3-Lemma we see that the last row is also a short exact sequence. To avoid fractions in the text we will name the map g : IPn −→ IQ and so refer to its kernel as K[g], and may use Kn[Inp] instead of its explicit description as a fraction. We now wish to show that f¯ factors over K[g]. For this we first see that taking domn of the commutative square p ηnp  ,2 f  Inp Inf  ιnPn ηn (ιnPn) ,2 In(ιnPn) gives Pn  ,2 Pn Kn[Jnp] _  Pn ηPn  ,2 IPn. Also, since domn In(ιnPn) = Kn[In(ιnPn)] = IPn, we see that f¯ factors as Kn[Inp] = Kn[p] Kn[Jnp]  ,2 ,2 f¯  Pn Kn[Jnp] = domn Inp _  IPn IPn . So in the following diagram, all possible squares and triangles commute. Kn[p]  ,2 ,2 _   ' GG GG GG GG Pn  ,2 _   ' FF FF FF FF Q _  Kn[p] Kn[Jnp]  ,2 ,2 f¯ 'F FF FF FF F hw w w w w Pn Kn[Jnp] 8w x xx xx xx x K[g]  ,2 ,2 IPn g  ,2 IQ 103 Chapter 5. Homology via Satellites By considering the composite Kn[p]  ,2Kn[Inp] f¯ ,2IPn g  ,2IQ in this diagram, we see that g◦f¯ = 0 and so f¯ factors over K[g] by the regular epimorphism Kn[p] Kn[Jnp] h  ,2 K n[p] JPn∩Kn[p] = K[g] . By Noether’s Isomorphism Theorem its kernel is K[f¯ ] = K[h] = JPn ∩Kn[p] Kn[Jnp] . Any cone (C, σ) on kern◦In◦U : Êndp −→ AE consists of three maps as shown below. C σ(0,!A)  σ(ιnPn,!A) OOO OOO OO #+OO OOO OOO σ(p,1A) ,2 Kn[Inp] f¯  0 ,2 IPn Thus σ(p,1A) factors over K[f¯ ], which we claim to be the limit of ker n◦In◦U : Êndp −→ AE. It remains to show that K[f¯ ] is itself a cone over this diagram. Given any endomor- phism e of p over A, we write e¯ for the induced endomorphism of Kn[Inp] and en : Pn −→ Pn for its “top” component. To show K[f¯ ] forms a cone over the diagram, we have to prove that e¯◦Ker f¯ = Ker f¯ . But this follows from the fact that (JPn ∩Kn[p])/Kn[Jnp] is a Baer invariant of A (see Remark 5.6.3). Indeed, in the diagram JPn∩Kn[p] Kn[Jnp]  ,2 Ker f¯ ,2 e¯′  Kn[Inp] f¯ ,2 e¯  IPn Ien  JPn∩Kn[p] Kn[Jnp]  ,2 Ker f¯ ,2 Kn[Inp] f¯ ,2 IPn the induced map e¯′ is the identity, and the needed equality follows. Thus JPn ∩Kn[p] Kn[Jnp] forms a cone on kern◦In◦U : Êndp −→ AE and is indeed the limit, as claimed. 5.6.6 Theorem: Consider n ≥ 1 and A ∈ |AE|. If AE has enough projectives and p is an n-fold presentation of A then Hn+1(A, I) = lim ( kern◦In : Endp −→ AE ) . Proof. By Theorem 5.6.2 the (n + 1)st homology of A is (JPn ∩ Kn[p])/Kn[Jnp]. Hence 104 5.6 Homology with projectives by Proposition 5.6.5 it suffices to show that Endp is initial in Êndnp. We must check that the slice categories (Endp ↓ (p, 1A)), (Endp ↓ (ιP , !A)) and (Endp ↓ (10, !A)) are non-empty and connected (here we view Endp as the full subcategory of Êndp determined by (p, 1A)). There is only one possible map from (p, 1A) to (10, !A), and the other two categories fulfil the needed conditions essentially because ((1P , 1Q), (1A, 1A)) is a terminal object of the slice category (Endp ↓ (p, 1A)), and the only maps in Êndp from (p, 1A) to (ιP , !A) are compositions of an endomorphism of (p, 1A) with f . 5.6.7 Remark: This means that computing the homology of an object essentially amounts to finding fixed points of endomorphisms of a projective presentation of this object. The use of this technique will be illustrated in Examples 5.6.9 and 5.6.10. 5.6.8 Remark: We now come back to Remark 5.6.3 and interpret Definition 5.5.2 in terms of Baer invariants. It provides an alternative answer to the following question: “Given a functor I : A −→ A and an object A of A, how can we construct an object Hn+1(A, I) out of the n-extensions of A in a manner which is independent of any particular chosen extension of A?” The classical example is the Hopf formula H2(A, I)G ∼= JP ∩K[p]K[J1p] which expresses H2(A, I)G in terms of a projective presentation p : P −→ A of A. Of course, the very existence of the isomorphism implies that the expression on its right hand side cannot depend on the choice of p. The idea behind Definition 5.5.2 is different but straightforward: simply take the limit of all extensions of A. The independence might now be understood as follows. If p is an n-presentation of A then Hn+1(A, I) is the limit of kern◦In : Endp −→ A, which means that Hn+1(A, I) is the universal object with the property that all endomorphisms of p are mapped to the same automorphism of this object, its identity. Finally we show, as worked out examples, that we can retrieve well-known results in group homology using our new definition. 5.6.9 Example (Finite cyclic groups): We use the methods of our theory to calculate H2(Cn, ab) for any n ∈ N, where Cn is the cyclic group of order n. As Z is projec- tive and abelian, the map p : Z −→ Cn which sends 1 ∈ Z to a generator c ∈ Cn is a projective presentation of Cn, and central. Thus H2(Cn, ab) is the limit of the diagram 105 Chapter 5. Homology via Satellites ker : Endp −→ Gp. Now any endomorphism of p must be Z ·(nk+1)  p  ,2 ⇓ Cn Z p  ,2 Cn i.e., multiplication by (nk + 1) for some k ∈ Z. So H2(Cn, ab) is the limit of the diagram which has nZ as the only object , and maps ·(nk+1): nZ −→ nZ. If λ : H2(Cn, ab) −→ nZ is the leg of the limit cone, we must have λ(x) · (nk + 1) = λ(x) for every element x ∈ H2(Cn, ab) and every k. So we are looking for fixed points of the map ·(nk + 1). But as, in nZ, 0 is the only fixed point of multiplication by (nk + 1) for all k 6= 1, we have λ(x) = 0 for all x ∈ H2(Cn, ab). Thus, as λ is a limit cone and so a monomorphism, H2(Cn, ab) = 0. Notice that as Cn and Z are abelian, they are also nilpotent of any class m ≥ 1 and solvable of any derived length m ≥ 1. Therefore p : Z −→ Cn is also central with respect to Nilm and Solm for any m ≥ 1, and so we have H2(Cn,nilm) = 0 and H2(Cn, solm) = 0 for any m ≥ 1. 5.6.10 Example (Generators and relations): Given a presentation of a group in terms of generators and relations, for example A = 〈a1, . . . , an | ri = 1〉 for some relations ri, the kernel of the free presentation p : Fn −→ A is generated by the relations ri as a normal subgroup of Fn. Here Fn is the free group on n generators. But when we go to the centralisation centr p : Fn [K[p], Fn] −→ A, every element of the kernel commutes with every other element, so now K[centr p] is generated by the relations ri as a subgroup of Fn/[K[p], Fn]. Every endomorphism of p over A must send a generator ai to aiki for some ki ∈ K[f ], and any choice of ki gives such an endomorphism. Thus on centr p we get endomorphisms that send ai ∈ Fn/[K[p], Fn] to ai ∏ j r αij j , for some αij ∈ Z, and again any choice of αij gives an endomorphism. Note that K[centr p] is an abelian group, since it is in the centre of Fn/[K[p], Fn]. From here it is relatively easy to find the fixed points of the induced endomorphism of K[centr p], given 106 5.6 Homology with projectives a specific group in terms of generators and relations. We give as an example Cn × Cn = 〈a, b | an = 1 = bn, aba−1b−1 = 1〉. Here p : F2 −→ Cn × Cn, and K[centr p] is generated by x = an, y = bn and z = aba−1b−1. Note that as aba−1b−1 commutes with everything, we get (aba−1b−1)n = abna−1b−n, and as bn also commutes with everything, we have zn = 1. As described above, any endomorphism of centr p induced by one on p sends a ∈ F2/[K[p], F2] to axα1yα2zα3 and b to bxβ1yβ2zβ3 . On K[centr p] this gives x 7−→ xnα1+1ynα2 y 7−→ xnβ1ynβ2+1 z 7−→ z as the x, y and z commute with everything, and zn = 1. For xl1yl2zl3 to be a fixed point for any of these endomorphisms, we need l1α1 + l2β1 = 0 l1α2 + l2β2 = 0 for any choice of αi and βi, or in other words we need l1α+ l2β = 0 for any choice of α and β. Hence l1 = l2 = 0, and we have fixed points zl3 . Since zn = 1, we get H2(Cn × Cn, ab) = Cn. Note that we can use the diagram over Êndp instead of Endp to see that any fixed point must be of the form aba−1b−1 for some a and b (or a product of such), since the fixed point must be sent to the identity in abFn = Fn/[Fn, Fn]. H2(A, ab) ,2 "*NN NNN NNN NNN  K[centr p]  0 ,2 abFn Comparing this to the Hopf formula H2(A, ab) = [Fn, Fn] ∩K[p] [K[p], Fn] , we see that the calculation using our method is exactly the same as the one using the Hopf 107 Chapter 5. Homology via Satellites formula; the only thing that is different is the interpretation of these elements as fixed points of certain endomorphisms. Note that we of course proved in Proposition 5.6.5 that the limit of the diagram ker◦I1 : Êndp −→ A is the expression of the Hopf formula, so this is exactly what you would expect. In contrast to the situation using abelianisation, we give another example which uses the Birkhoff subcategory of nilpotent groups. 5.6.11 Example (Gp vs. Nil2): We calculate the second homology group of C2×C2 with respect to the Birkhoff subcategory Nil2 of nilpotent groups of class at most 2, using the Hopf formula. We again use the projective presentation p : F2 −→ C2 × C2 with kernel K, but now centralisation gives K [[K,F2],F2]  ,2 ,2 F2 [[K,F2],F2] I1p  ,2 C2 × C2. So the Hopf formula gives H2(C2 × C2,nil2) = K ∩ [[F2, F2], F2][[K,F2], F2] = K [[K,F2], F2] ∩ [[F2, F2], F2] [[K,F2], F2] using Examples 1.3.3 and 4.3.7. If F2 is generated by a and b, a long (and tedious) calculation shows that K[I1p] is generated by the set {x1 = a2, y1 = b2, z = [a, b], x2 = ba2b−1, y2 = ab2a−1, x3 = b−1a2b, y3 = a−1b2a} with many relations, some of which we will list. The main fact we need to get these identities is that any element in [K,F2] commutes with any element of F2. We get that the element z commutes with all other elements of K[I1p], the xi commute amongst each other and so do the yi. We have [x1, y1] = x1y1x−11 y −1 1 = y1x −1 1 y −1 1 x1 = [y1, x −1 1 ] = [x −1 1 , y −1 1 ] = [y −1 1 , x1] = z 4 that is, all cyclic permutations of x1y1x−11 y −1 1 are the same as z 4. Similarly [x1, y2] = [x2, y1] = [x1, y3] = [x3, y1] = [x3, y2] = [x2, y3] = [x2, y2] = [x3, y3] = z4 108 5.6 Homology with projectives with the same cyclic permutations as for [x1, y1] above. Thus we see that z4 commutes with every element of F2, not just elements of the kernel as z does. We also have [a2, b]2 = (x1x−12 ) 2 = [a−2, b−1]2 = (x−11 x3) 2 = z4 and the corresponding expression for y instead of x. Notice that the elements [a2, b] etc. already commute with any element in F2. The following Witt-Hall identities of commutators are very useful: [g, h1h2] = [g, h1]h1[g, h2]h−11 [g1g2, h] = g1[g2, h]g−11 [g1, h] These identities give for example that [a, ba] = [a, b], and also [[a, ab], a] = [[a, b], a] and [[a, ab], b] = [[a, b], b] (using as always that elements of [K,F2] commute with every element of F2). As soon as one each of g, h2 and g2, h lie in the kernel K, these identities become [g, h1h2] = [g, h1][g, h2] [g1g2, h] = [g2, h][g1, h]. We also see that [k, g−1] = [k, g]−1 = [k−1, g] for any k ∈ K and g ∈ F2. Thus we find that the elements [z, a] = [[a, b], a] and [z, b] = [[a, b], b] generate K∩[[F2,F2],F2][[K,F2],F2] , and that [z, a]4 = [z4, a] = 1 and [z, b]4 = [z4, b] = 1. Thus we have H2(C2 × C2,nil2) = C4 × C4. 109 Chapter 5. Homology via Satellites 110 Chapter 6 Stem Extensions in the Context of Abelianisation In this chapter we look at certain special central extensions such as trivial and stem extensions, and establish a bijection between the isomorphism classes of stem extensions of an object A and the subobjects of H2(A, ab). This bijection is induced by the well- known isomorphism Centr(A,K) ∼= Hom(H2A,K) which is usually achieved in two steps, using Centr(A,K) ∼= H2(A,K) ∼= Hom(H2A,K). Here we give an explicit description of the direct isomorphism, without going via the second cohomology group. The Stallings- Stammbach sequence, which is the lower part of the Everaert sequence we have met before, plays a crucial role in these results. 6.1 Stem extensions and stem covers Let us first review some of the material from Chapters 4 and 5 and restate it appropriately for its use in this chapter. As we are dealing with a semi-abelian category A and its Birkhoff subcategory of abelian objects AbA, we will denote the kernel of the unit of abelianisation of A by [A,A] instead of JA: 0 ,2 [A,A]  ,2 ,2A ηA  ,2 abA = H1A ,20 This extends to a functor [−,−] : A −→ A. We also write [K[p], P ] for the object J1[p] = K[J1p] (see Section 4.3). We use homology defined via Hopf formulae, as in Definition 4.4.3, but also use prop- erties of homology exhibited by viewing it as a limit, as in Chapter 5. As we will only need the second homology object, we will repeat the definition here with our slightly changed notation. 6.1.1 Definition ([Eve2007]): Given an object A in A, let K  ,2 ,2P p  ,2A 111 Chapter 6. Stem Extensions in the Context of Abelianisation be a projective presentation of A. We define the second homology object of A by the Hopf Formula H2A = K ∩ [P, P ] [K,P ] . The Stallings-Stammbach sequence [EVdL2004b, EG2007], which is the lowest part of the Everaert sequence (F), plays a big role in our results, but we only use it for central extensions, so we restate it here for this special case. 6.1.2 Proposition: For any central extension 0 ,2K  ,2 ,2B f  ,2A ,20 (N) in A, there is an exact sequence H2B H2f ,2H2A δf ,2K γf ,2H1B H1f ,2H1A ,20 (O) in AbA which is natural in f . Proof. This is just the lowest part of (F) in Theorem 4.4.6. See also [EVdL2004b, EG2007]. Notice that as here f is already central, we get the kernel of f instead of the kernel of I1f = H1(f, I1)E1 . For convenience we also recall Corollary 5.5.10 for the case of H2. 6.1.3 Proposition: Given an object A in A, the second homology object H2A is the limit of the diagram ker : CExtAA −→ A. The legs of the limit cone are the δf : H2A −→ K[f ] as given by the Stallings-Stammbach sequence (O) for a central extension f . Proof. See Proposition 5.2.3 and Corollary 5.5.10. We now define some special central extensions that we will consider throughout the chapter. Trivial extensions have been defined before, in 1.4.3, but we restate the definition here with more equivalent conditions, along with diagrams which clarify the connection to the new definitions. We have also met stem extensions before in the context of groups, in Example 5.5.15. The la´ma´ra extensions, which I have named thus just for the moment, are not particularly important in what follows and are only included for completeness. 6.1.4 Definition: Given a central extension f as in (N), we say it is a 112 6.1 Stem extensions and stem covers (1) trivial extension if one of the following equivalent conditions is satisfied: • δf = 0, • [B,B] −→ [A,A] is an isomorphism, • K ∩ [B,B] = 0; 0 = K ∩ [B,B]  ,2 ,2 _   [B,B] ∼= ,2 _   [A,A] _   K  ,2 ,2 B f  ,2 _  A _  H2A 0 ,2 K  ,2 ,2 H1B  ,2 H1A (2) la´ma´ra extension if δf is a monomorphism; (3) stem extension if one of the following equivalent conditions is satisfied: • δf is a regular epimorphism, • H1B −→ H1A is an isomorphism, • γf = 0, • K ≤ [B,B]; [B,B] _    ,2 [A,A] _   K  ,2 ,2 7B 7B B f  ,2 _  A _  H2A δf  ,2K 0 ,2 H1B ∼= ,2 H1A (4) stem cover if one of the following equivalent conditions is satisfied: • δf is an isomorphism, • H1B −→ H1A is an isomorphism and H2B −→ H2A is the zero map. [B,B]  ,2 _   [A,A] _   K  ,2 ,2 7B 7B B f  ,2 _  A _  H2B 0 ,2 H2A ∼= ,2 K 0 ,2 H1B ∼= ,2 H1A Clearly every stem cover is also a stem extension. 113 Chapter 6. Stem Extensions in the Context of Abelianisation 6.2 Perfect objects From now on we only consider central extensions (N) where A is a perfect object. It is well known that a group has a universal central extension if and only if it is perfect. This holds more generally in any semi-abelian category. 6.2.1 Definition (Universal central extension): A universal central extension (with respect to abelianisation) is a central extension g : C −→ A such that for any other central extension h : D −→ A there is a unique map C −→ D making the following square commute: C g  ,2 ∃!  A D h  ,2 A 6.2.2 Definition: An object A of a semi-abelian category A is called perfect when its abelianisation is zero: AbA = 0. 6.2.3 Lemma: [GVdL2008b, Proposition 4.1] Let A be a semi-abelian category with enough projectives. An object A of A is perfect if and only if A admits a universal central extension. Recall from Corollary 5.5.11 that if A admits a universal central extension g, then the second homology group is the kernel of this central extension: H2(A, ab) = K[g]. In fact, we can say more: 6.2.4 Lemma: Any universal central extension is a stem cover. Proof. Let K  ,2 ,2C g  ,2A be a universal central extension of A. The limit of the dia- gram ker : CExtAA −→ A, is (H2A, δf )f∈CExtAA. As g is initial in CExtAA, the limit H2A is isomorphic to the kernel K of g, and the isomorphism is given by the leg δg. Thus, by definition, g is a stem cover. Notice that we could always make δg into an actual identity by changing the kernel of g to an isomorphic one. When A is a perfect object, the special extensions defined in 6.1.4 take a more specific form. 6.2.5 Lemma: When A is perfect, (1) any central extension (N) has H1B ∼= K/(K ∩ [B,B]), 114 6.2 Perfect objects (2) any trivial extension is isomorphic to a product projection K ×A pi2  ,2A, (3) any la´ma´ra extension satisfies H2A ∼= K ∩ [B,B], (4) any stem extension has perfect domain B. Proof. (1) If H1A = 0, we have the following commuting diagram: K ∩ [B,B]  ,2 ,2 _   [B,B] _    ,2 [A,A] K  ,2 ,2 _  B f  ,2 _  A _ K K∩[B,B] ,2 H1B ,2 0 Here all columns and the top two rows are exact sequences, so using the 3 × 3 Lemma [Bou2001], we conclude that the bottom row is also short exact and the result follows. Notice that the map γf in the Stallings-Stammbach sequence (O) then becomes the canonical quotient K  ,2K/(K ∩ [B,B]). (2) If H1A = 0, the pullback in 6.1.4 becomes a product: K  ,2 ,2 B f  ,2 _  A _  H2A 0 ,2 K ∼= ,2 H1B ,2 0 Thus B = H1B ×A ∼= K ×A and f is a product projection. (3) Combining (1) with the fact that δf is a monomorphism, the Stallings-Stammbach sequence (O) becomes 0 ,2H2A  ,2 ,2K  ,2 KK∩[B,B] ∼= H1B ,20 and so H2A ∼= K ∩ [B,B]. (4) If H1A = 0 and H1B ∼= H1A, clearly H1B = 0 as well. K  ,2 ,2 B f  ,2 _  A _  H2A δf  ,2K ,2 0 ∼= ,2 0 115 Chapter 6. Stem Extensions in the Context of Abelianisation 6.3 A natural isomorphism In this section, we want to define a natural isomorphism θ : Centr(A,−) −→ Hom(H2A,−) of functors AbA −→ Ab. First we must define the abelian group Centr(A,K) for an abelian object K. 6.3.1 Remark: Notice that we will be talking about two different kinds of isomorphism classes of central extensions. Given an object A, we say two central extensions of A are isomorphic if there is an isomorphism K  ,2 ,2 ∼=  B  ,2 ∼=  A K ′  ,2 ,2 B′  ,2 A of central extensions. However, when we fix the kernel K, we say two central extensions of A by K are isomorphic if there is a morphism K  ,2 ,2 B  ,2 ∼=  A K  ,2 ,2 B′  ,2 A where we demand the kernel part to be an actual identity. It follows from the Short Five Lemma that two central extensions of A by K are isomorphic if and only if there is a morphism between them. Let Centr(A,K) denote the set of isomorphism classes of central extensions of A by K, as defined in the remark above. Recall from [Ger1970] how Centr(A,K) can be made into an abelian group using the Baer sum: given two central extensions K  ,2 k B f  ,2A and K  ,2 k ′ B′ f ′  ,2A, let h be their pullback B ×A B′  ,2 _  h (I III III II B′ f ′ _  B f  ,2 A or equivalently B ×A B′ h  ,2  A ∆  B ×B′ f×f ′  ,2 A×A 116 6.3 A natural isomorphism where ∆ is the diagonal. Then the extension f + f ′ is formed as follows, using the multiplication m of the abelian object K: K[m] _  i  K[m] _  (k×k′)◦i  K ×K  ,2k×k ′ ,2 m _  B ×A B′ h  ,2 Coker (k×k′)◦i _  A K  ,2 k×k′ ,2 D f+f ′  ,2 A As h is a central extension with respect to abelianisation, the subobject (k×k′)◦i is normal (see Proposition 1.5.16). So we can form its cokernel, or equivalently the pushout of k×k′ and m. Then k × k′ is a monomorphism as the top square is a pullback. We take its cokernel, the codomain of which is A since the bottom left square is a pushout. Note that this square is also a pullback. This cokernel represents the isomorphism class of f + f ′; it is central because the category of central extensions in A is closed under quotients in the category of extensions in A (see [GVdL2008b] for more details). It is easily checked that this gives a well-defined sum on the isomorphism classes of central extensions of A by K. The zero for this addition is the class of split central extensions, a representative of which is the product projection K  ,2 ,2K ×A pi2  ,2B . Note that any split central extension is trivial [JK1994, Section 4.3], and compare this with 6.2.5(2). Recall from [GVdL2008b] how Centr(A,−) is made into a functor AbA −→ Ab. The main ingredient is the following result: 6.3.2 Lemma: [GVdL2008b, Corollary 3.3] Given a central extension (N) and a map from K to an abelian object K ′, there is a central extension K ′  ,2 ,2B′  ,2A making the diagram below commute. K  ,2 ,2  B  ,2  A K ′  ,2 ,2 B′  ,2 A Proof. See Corollary 3.3 in [GVdL2008b]. For this construction to work it is crucial that the extension B −→ A is central in the sense of Huq, which is the main reason why this isomorphism only works in the setting of abelianisation. This indeed makes Centr(A,−) into a functor AbA −→ Ab (see Propositions 6.1 and 6.2 in [GVdL2008b]). 117 Chapter 6. Stem Extensions in the Context of Abelianisation For a given abelian object K, the set Hom(H2A,K) also forms an abelian group. Two maps α and β are added using the multiplication m of K as follows: α+ β : H2A (α,β) ,2K ×K m ,2K The zero of the addition is the zero map H2A 0 ,2K . Therefore we can view the functor Hom(H2A,−) : A −→ Set also as a functor Hom(H2A,−) : AbA −→ Ab. We can now finally define, for a perfect object A, the required natural transformation. 6.3.3 Definition: Given a perfect object A in A, we define a natural transformation θ : Centr(A,−) −→ Hom(H2A,−) as follows: for each abelian object K in A, θK takes the (isomorphism class of the) central extension K  ,2 k ,2B f  ,2A to the map δf : H2A −→ K defined by the Stallings- Stammbach sequence (O). This is well-defined: if f is isomorphic to K  ,2 k ′ ,2B′ f ′  ,2A , the Stallings-Stammbach sequence gives H2A δf ,2 K ,2 H1B ,2 ∼=  H1A = 0 H2A δf ′ ,2 K ,2 H1B′ ,2 H1A = 0 so δf = δf ′ . Naturality of θ follows from the definition of the functor Centr(A,−) and the fact that δ itself is a natural transformation. For θ to be a natural transformation between functors to Ab, we must show that each θK is a homomorphism of abelian groups. We do this in a separate lemma. 6.3.4 Lemma: For each abelian object K in A, θK is a homomorphism of abelian groups. Proof. Given two central extensions f and f ′ representing two isomorphism classes in Centr(A,K), we consider their pullback B ×A B′ h  ,2A as above in the definition of the sum in Centr(A,K). We again use the limit cone (H2A, δf )f∈CExtAA of the diagram 118 6.3 A natural isomorphism ker : CExtAA −→ A to see that δh = (δf , δf ′) : H2A −→ K ×K: H2A δf vuu uu uu uu u δh  δf ′ (II III III II K_   K ×Kpi1lr pi2  ,2_   K_   B f _  B ×A B′lr  ,2 h _  B′ f ′ _  A A A Remembering that f + f ′ is given by K ×K  ,2k×k ′ ,2 m _  B ×A B′ _  h  ,2 A K  ,2 ,2 D f+f ′  ,2 A the Stallings-Stammbach sequence gives us a commuting diagram H2A δh=(δf ,δf ′ ) ,2 δf+δf ′ !)LL LLL LLL LLL LLL LLL K ×K m _  H2A δf+f ′ ,2 K and the result follows. 6.3.5 Lemma: For each abelian object K in A, θK is injective. Proof. If the central extension f maps to δf = 0, it is a trivial extension by definition, and so as A is perfect, f is (isomorphic to) the product projection K  ,2 ,2K ×A pi2  ,2A (see Lemma 6.2.5). 6.3.6 Lemma: For each abelian object K in A, θK is surjective. Proof. Given a morphism α : H2A −→ K, we use a universal central extension of A and Lemma 6.3.2 to construct a central extension of A by K. Let H2A  ,2 ,2C g  ,2A be a universal central extension of A, which is a stem cover by Lemma 6.2.4. We can choose g such that δg : H2A −→ H2A is the identity on H2A. Then as K is an abelian object, 119 Chapter 6. Stem Extensions in the Context of Abelianisation Lemma 6.3.2 gives us a central extension f and a map from g to f as below: H2A  ,2 ,2 α  C  g  ,2 A K  ,2 ,2 B f  ,2 A Forming the Stallings-Stammbach sequence of f and g, we see that δf = α: H2A δg H2A α  H2A δf ,2 K Thus the isomorphism class of f really is a preimage of α. We can now put together the above results into our main theorem. 6.3.7 Theorem: Given a perfect object A in a semi-abelian category A, there is a natural isomorphism θ : Centr(A,−) −→ Hom(H2A,−) of functors AbA −→ Ab. Fixing an abelian object K, (1) trivial extensions correspond to the zero map H2A 0 ,2K (i.e. there is only one equivalence class of trivial extensions of A by K, one representative of which is the product projection K  ,2 ,2K ×A pi2  ,2A), (2) la´ma´ra extensions correspond to monomorphisms H2A  ,2 ,2K , (3) stem extensions correspond to regular epimorphisms H2A  ,2K , (4) stem covers correspond to isomorphisms H2A ∼= ,2K . Proof. θ is a natural isomorphism by Lemmas 6.3.4, 6.3.5 and 6.3.6. The correspondences (1) to (4) all follow directly from Definition 6.1.4 and the definition of θ. 6.3.8 Corollary: There is only one isomorphism class of stem covers of a perfect object. Proof. If we only look at isomorphisms of central extensions of a perfect object A by a fixed kernel K, as in Centr(A,K), Theorem 6.3.7(4) gives us several isomorphism classes 120 6.3 A natural isomorphism of stem extensions, one for each isomorphism H2A −→ K. But as soon as we vary K, we see that all stem covers of A are isomorphic to each other (cf. Remark 6.3.1). H2A  ,2 ,2 ∼=δf  B g  ,2 ∼=  A K  ,2 ,2 B′ f  ,2 A 6.3.9 Corollary: Any stem cover of a perfect object is a universal central extension. Proof. As any universal central extension is a stem cover by Lemma 6.2.4 and there is only one isomorphism class of stem covers, any stem cover is universal. 6.3.10 Remark: The isomorphism in Theorem 6.3.7 is not new, it exists as a two-step iso- morphism Centr(A,K) ∼= H2(A,K) ∼= Hom(H2A,K) given for example in [GVdL2008b]. We have merely given a direct correspondence without going via the cohomology group. 6.3.11 Corollary: There is a bijective correspondence between the isomorphism classes of stem extensions of a perfect object A and subobjects of H2A. Proof. When we vary K in Theorem 6.3.7, we see that the isomorphism classes of stem extensions of A correspond to regular epimorphisms H2A  ,2K for any K, which in turn correspond to subobjects U of H2A by taking the kernels of these regular epimorphisms. 6.3.12 Lemma: Every stem extension of a perfect object A is the regular image of a stem cover of A. Proof. As A is perfect, it has a universal central extension H2A  ,2 ,2C g  ,2A which is a stem cover. As g is universal, for every stem extension f there is a map H2A  ,2 ,2 δf _  C g  ,2  A K  ,2 ,2 V f  ,2 A where it is easily seen using the Stallings-Stammbach sequence that the induced map H2A −→ K is indeed (isomorphic to) δf . By the Short Five Lemma, which also holds for regular epimorphisms (see [Bou2001]), the map C −→ V in the middle is also a regular epimorphism. 121 Chapter 6. Stem Extensions in the Context of Abelianisation 6.3.13 Remark: The results in Corollaries 6.3.8 and 6.3.11 and in Lemma 6.3.12 are ex- tensions to general semi-abelian categories of the corresponding results on crossed modules in [VC2002]. If C is a semi-abelian category, the category A = Grpd(C) of groupoids in C is also semi-abelian, and its subcategory of abelian objects AbA is RG(AbC), reflective graphs in AbC. A special case of this is the category of crossed modules when C = Gp the category of groups. This gives back the results in [VC2002]. 122 References [Bar1971] M. Barr, Exact categories, Exact categories and categories of sheaves, Lec- ture notes in mathematics, vol. 236, Springer, 1971, pp. 1–120. 8, 9 [BB1969] M. Barr and J. Beck, Homology and standard constructions, Seminar on triples and categorical homology theory, Lecture notes in mathematics, vol. 80, Springer, 1969, pp. 245–335. 31, 37, 38, 39, 43, 47, 49, 60 [BB2004] F. Borceux and D. Bourn, Mal’cev, protomodular, homological and semi- abelian categories, Mathematics and its Applications, vol. 566, Kluwer Aca- demic Publishers, 2004. 7, 10, 11, 12, 15, 16, 26, 61 [BG2002a] D. Bourn and M. Gran, Central extensions in semi-abelian categories, J. Pure Appl. Algebra 175 (2002), 31–44. 15, 29 [BG2002b] D. Bourn and M. Gran, Centrality and normality in protomodular cate- gories, Theory Appl. Categ. 9 (2002), no. 8, 151–165. 24, 28 [BJ2001] F. Borceux and G. Janelidze, Galois theories, Cambridge Studies in Ad- vanced Mathematics, vol. 72, Cambridge University Press, 2001. 20 [Bor1994] F. Borceux, Handbook of categorical algebra, Encyclopedia of Mathematics and its Applications, vol. 50, 51 and 52, Cambridge University Press, 1994. 67, 85 [Bor2004] F. Borceux, A survey of semi-abelian categories, in Janelidze et al. [JPT2004], pp. 27–60. 7, 11, 14, 15 [Bou1991] D. Bourn, Normalization equivalence, kernel equivalence and affine cate- gories, in Carboni et al. [CPR1991], pp. 43–62. 8, 10, 12 [Bou1996] D. Bourn, Mal’cev categories and fibration of pointed objects, Appl. Categ. Struct. 4 (1996), 307–327. 10 [Bou2000] D. Bourn, Normal subobjects and abelian objects in protomodular categories, J. Algebra 228 (2000), 143–164. 15, 27 [Bou2001] D. Bourn, 3 × 3 lemma and protomodularity, J. Algebra 236 (2001), 778– 795. 11, 12, 33, 115, 121 [Bou2002] D. Bourn, Intrinsic centrality and related classifying properties, J. Algebra 256 (2002), 126–145. 14, 24, 26 [Bou2003] D. Bourn, The denormalized 3×3 lemma, J. Pure Appl. Algebra 177 (2003), 113–129. 13 [Bou2005] D. Bourn, On the direct image of intersections in exact homological cate- gories, J. Pure Appl. Algebra 196 (2005), 39–52. 16, 79, 80 123 REFERENCES [CKP1993] A. Carboni, G. M. Kelly, and M. C. Pedicchio, Some remarks on Maltsev and Goursat categories, Appl. Categ. Struct. 1 (1993), 385–421. 10, 13, 34, 35, 50, 54 [CLP1991] A. Carboni, J. Lambek, and M. C. Pedicchio, Diagram chasing in Mal’cev categories, J. Pure Appl. Algebra 69 (1991), no. 3, 271–284. 10 [CPP1992] A. Carboni, M. C. Pedicchio, and N. Pirovano, Internal graphs and inter- nal groupoids in Mal’cev categories, Proceedings of Conf. Category Theory 1991, Montreal, Am. Math. Soc. for the Canad. Math. Soc., Providence, 1992, pp. 97–109. 24, 27 [CPR1991] A. Carboni, M. C. Pedicchio, and G. Rosolini (eds.), Category Theory, Proceedings Como 1990, Lecture notes in mathematics, vol. 1488, Springer, 1991. 123, 125 [CVdL2009] J. M. Casas and T. Van der Linden, A relative theory of universal central extensions, Pre´-Publicac¸o˜es DMUC 09-10 (2009), 1–22, submitted. 22, 95 [EG2006] T. Everaert and M. Gran, Relative commutator associated with varieties of n-nilpotent and of n-solvable groups, Arch. Math. (Brno) 42 (2006), no. 4, 387–396. 23 [EG2007] T. Everaert and M. Gran, On low-dimensional homology in categories, Ho- mology, Homotopy and Appl. 9 (2007), 275–293. 74, 112 [EGVdL2008] T. Everaert, M. Gran, and T. Van der Linden, Higher Hopf formulae for homology via Galois Theory, Adv. Math. 217 (2008), 2231–2267. 61, 66, 71, 73, 74, 78, 87, 96 [EVdL2004a] T. Everaert and T. Van der Linden, Baer invariants in semi-abelian cate- gories I: General theory, Theory Appl. Categ. 12 (2004), no. 1, 1–33. 102 [EVdL2004b] T. Everaert and T. Van der Linden, Baer invariants in semi-abelian cate- gories II: Homology, Theory Appl. Categ. 12 (2004), no. 4, 195–224. 16, 31, 32, 33, 35, 37, 78, 96, 112 [Eve2007] T. Everaert, An approach to non-abelian homology based on Categorical Galois Theory, Ph.D. thesis, Vrije Universiteit Brussel, 2007. 32, 61, 62, 66, 69, 73, 74, 77, 78, 79, 87, 96, 102, 111 [Eve2008] T. Everaert, Higher central extensions and Hopf formulae, to appear in J. Algebra, doi:10.1016/j.jalgebra.2008.12.015, 2008. 74, 102 [Fro¨1963] A. Fro¨hlich, Baer-invariants of algebras, Trans. Amer. Math. Soc. 109 (1963), 221–244. 102 [Ger1970] M. Gerstenhaber, A categorical setting for the Baer extension theory, Appli- cations of Categorical Algebra, New York 1968, Proc. Sympos. Pure Math., vol. XVII, Amer. Math. Soc., Providence, R.I., 1970, pp. 50–64. 116 [GR2004] M. Gran and J. Rosicky´, Semi-abelian monadic categories, Theory Appl. Categ. 13 (2004), no. 6, 106–113. 37 124 REFERENCES [Gra2001] M. Gran, Central extensions and internal groupoids in Maltsev categories, J. Pure Appl. Algebra 155 (2001), 139–166. 17 [GVdL2007] J. Goedecke and T. Van der Linden, A comparison theorem for simplicial resolutions, J. Homotopy and Related Structures 2 (2007), no. 1, 109–126. 48 [GVdL2008a] J. Goedecke and T. Van der Linden, On satellites in semi-abelian categories: Homology without projectives, to appear in Math. Proc. Cambridge Philos. Soc., doi:10.1017/S0305004109990107, 2008. 82 [GVdL2008b] M. Gran and T. Van der Linden, On the second cohomology group in semi- abelian categories, J. Pure Appl. Algebra 212 (2008), 636–651. 28, 29, 114, 117, 121 [Hop1942] H. Hopf, Fundamentalgruppe und zweite Bettische Gruppe, Comment. Math. Helv. 14 (1942), 257–309. 97 [Huq1968] S. A. Huq, Commutator, nilpotency and solvability in categories, Quart. J. Math. Oxford 19 (1968), no. 2, 363–389. 14, 24 [Jan1976] G. Janelidze, On satellites in arbitrary categories, Bulletin of the Academy of Sciences of the Georgian SSR 82 (1976), no. 3, 529–532, in Russian. English translation at arXiv:0809.1504v1. 82, 83, 84, 88 [Jan1990] G. Janelidze, Pure Galois theory in categories, J. Algebra 132 (1990), 270– 286. 21 [Jan1991] G. Janelidze, Precategories and Galois theory, in Carboni et al. [CPR1991], pp. 157–173. 20, 21 [Jan2008] G. Janelidze, Galois groups, abstract commutators and Hopf formula, Appl. Categ. Struct. 16 (2008), 653–668. 2 [JK1994] G. Janelidze and G. M. Kelly, Galois theory and a general notion of central extension, J. Pure Appl. Algebra 97 (1994), 135–161. 17, 19, 20, 21, 24, 117 [JMT2002] G. Janelidze, L. Ma´rki, and W. Tholen, Semi-abelian categories, J. Pure Appl. Algebra 168 (2002), 367–386. 8, 13, 15 [Joh2002] P. T. Johnstone, Sketches of an elephant: A topos theory compendium, Oxford Logic Guides, vol. 43, 44, Oxford University Press, 2002, Volumes 1 and 2. 100 [Joh2004] P. T. Johnstone, A note on the semiabelian variety of Heyting semilattices, in Janelidze et al. [JPT2004], pp. 317–318. 100 [JPT2004] G. Janelidze, B. Pareigis, and W. Tholen (eds.), Galois theory, Hopf alge- bras, and semiabelian categories, Fields Institute Communications Series, vol. 43, A.M.S., 2004. 123, 125 [Kar1987] G. Karpilovsky, The Schur Multiplier, London Math. Soc. Monogr. Ser., vol. 2, Oxford University Press, 1987. 97 125 REFERENCES [Mac1998] S. MacLane, Categories for the working mathematician, second ed., Grad- uate texts in mathematics, vol. 5, Springer, 1998. 85, 93 [Ped1995] M. C. Pedicchio, A categorical approach to commutator theory, J. Algebra 177 (1995), 647–657. 24, 27 [Sch1904] I. Schur, U¨ber die Darstellung der endlichen Gruppen durch gebrochene lineare Substitutionen, J. Reine Angew. Math. 127 (1904), 20–50. 96, 97 [Sch1907] I. Schur, Untersuchungen u¨ber die Darstellung der endlichen Gruppen durch gebrochene lineare Substitutionen, J. Reine Angew. Math. 132 (1907), 85– 137. 96 [Smi1976] J. D. H. Smith, Mal’cev varieties, Lecture notes in mathematics, vol. 554, Springer, 1976. 24 [TV1969] M. Tierney and W. Vogel, Simplicial resolutions and derived functors, Math. Z. 111 (1969), no. 1, 1–14. 48, 50, 51, 52, 60 [VC2002] A.M. Vieites and J.M. Casas, Some results on central extensions of crossed modules, Homology, Homotopy and Applications 4 (2002), no. 1. 122 [VdL2006] T. Van der Linden, Homology and homotopy in semi-abelian categories, Ph.D. thesis, Vrije Universiteit Brussel, 2006, arXiv:math/0607100v1. 14, 55 [VdL2008] T. Van der Linden, Simplicial homotopy in semi-abelian categories, in press, online on arXiv:math/0607076v1, 2008. 14 [Wei1997] Ch. A. Weibel, An introduction to homological algebra, Cambridge studies in advanced mathematics, vol. 38, Cambridge University Press, 1997. 84 126 Index  ,2 ,2 normal mono, 8  ,2 normal epi, 8 0 zero-object, zero map, 7 (A ↓ F ) slice category, 85 AbA subcat. of abelian objects, 15 Ab cat. of abelian groups, 17 ab abelianisation functor, 15 ArrkA higher-dimensional arrows, 62 !A map A −→ 0, 71 Centr(A,K) central ext. of A by K, 116 CExtBA cat. of central extensions, 68 cod codomain functor, 73 Q[f ] cokernel object, 8 Coker f cokernel map, 8 A B cokernel, 8 [G,G] commutator subgroup, 16 G = (G, , δ) comonad, 37 φf,g cooperator, 24 δf connecting homomorphism, 77 DlG lth term of derived series, 18 ∆ diagonal, 15 dom domain functor, 78 Endp cat. of endos of p, 101 Êndp enlarged cat. of endos of p, 102 E class of extensions, 62 obE domains and codomains in E, 62 AE full subcat. determined by obE, 62 En class of higher extensions, 63 ExtA category of extensions, 66 ExtnA category of higher extensions, 66 g Lie or Leibniz algebra, 18 gAnn ideal gen. by all [x, x], 19 Γ Galois structure, 20 γf homom. in Everaert sequence, 77 Gp category of groups, 8 HnC homology of a chain complex, 16 HnA — of a simplicial object, 32 Hn(A, I)G comonadic —, 37 Hn(A, I)E — via Hopf formulae, 74 Hn(A, I) — via satellites/limits, 91 I reflector, 17 I1 centralisation functor, 69 I1[f ] domain of I1f , 69 I[f ] image, 9 Im f image inclusion, 9 ιnPn cube (Pn, 0, . . . , 0), 102 J functor giving kernel of unit, 67 J1f kernel of unit for centralisation, 68 J1[f ] domain of J1f , 68 K[f ] kernel object, 7 Ker f kernel inclusion, 7 ker kernel functor, 62 LCmG term in lower central series, 18 LeibK cat. of Leibniz algebras, 19 LieK cat. of Lie algebras, 19 [n] = {0 . . . , n}, 31 Nilm cat. of nilpotent groups, 18 nilm reflector to Nilm, 18 p projective presentation, 73 Pn initial object of n-cube p, 73 rnf map In[f ] −→ Tn[f ], 73 R-Mod cat. of R-modules, 37 R[f ] kernel pair, 21 (R, pi1, pi2) equivalence relation, 26 dR subdiagonal of equiv. rel., 26 Ran right Kan extension, 82 SESeqA cat. of short exact sequences, 83 A simplicial object, 31 ∂i face operator, 31 σi degeneracy operator, 31 NA Moore complex, 32 ZnA object of n-cycles, 32 S simplicial set, 34 AI cocylinder, 56 0(A) cocylinder map, 56 1(A) cocylinder map, 56 s(A) cocylinder map, 56 Solm cat. of solvable groups, 18 solm reflector to Solm, 18 Tn trivialisation functor, 72 TExtnA cat. of trivial extensions, 72 η unit of reflection, 19 ZG,Zb centre of a group/Lie algebra, 22 ZLie(b) Lie centre of Leibniz algebra, 22 ZNilm(B) nil centre, 22 ZSolm(B) solvable centre, 23 127 INDEX abelian internal — group, 15 abelian category, 1, 8, 15, 37, 47, 54, 60, 83 abelian group, 97, 116, 117 category of —s, see groups, abelian abelian object, 3, 14, 26, 32, 67, 79, 99, 111, 116, 117 abelianisation, 15, 16, 17, 20, 28, 95, 96, 111, 114 addition, see abelian object additive category, 37, 47, 54, 60 additive functor, 83 adjunction, 54, 60, 88, 96 algebras, 54 arrow double —, 62 higher dimensional —, 62 associated normal subobject, 27 augmented simplicial object, 33, 37, 39, 50 Baer invariant, 102, 104, 105 Barr, 8 Barr-Beck, 1, 31, 37, 47 Barr-Beck homology, 37 Beck, see Barr-Beck Birkhoff subcategory, 3, 17, 20, 66, 99, 111 strongly E—, 66, 69, 74–80, 85–105 Bourn, 8, 24 bracket, see Leibniz/Lie algebra category abelian —, 1, 8, 15, 37, 47, 54, 60, 83 additive —, 37, 47, 54, 60 exact —, 8, 61 — with finite limits, 50 homological —, 10 Mal’tsev —, 10, 21, 34 monadic —, 37 pointed —, 7, 61 protomodular —, 8, 61 regular —, 8, 34 semi-abelian —, 1, 8 sub— of abelian objects, 15, 111 unital —, 10, 24 category of elements, 85 category of endomorphisms, 101, 102 category of extensions of A, 93 central equivalence relation, 28 central extension finite —, 97 Galois, 21, 29, 67–69, 81, 94, 95, 98, 111–122 higher —, 71 Huq, 25, 28, 29, 111–122 isomorphism class of —s, 116 split —, 24, 76, 78, 117 universal —, 95, 96, 114, 121 weakly universal —, 95 central extension of A by K, 116 sum, 117 central morphism Huq, 25, 26, 28, 79, 99 central series, 18 centralisation, 68, 69, 70, 73, 75, 85, 88, 95, 96 centralising double relation, 28 centralising equivalence relations, 27 centre — of a group/Lie algebra, 70 — of group/Lie algebra, 22, 25 Lie —, 22 nil —, 22, 70 solvable —, 23 chain complex, 16, 99 Moore complex, 32, 34 normalised —, see Moore complex proper —, 16, 32 unnormalised —, 32 Chevalley-Eilenberg, 19, 74 class of extensions, 62, 68 E1, 65 closed under limits, 67 closed under regular quotients, 17 closed under subobjects, 17, 66 cocone, 90 cocylinder, 56 codomain, 73, 86, 89, 91, 101 coefficients, 37, 38, 85, 91 coequaliser, 40 coequaliser iff cokernel, 33, 40 cohomology, 81 comonadic, 37 second — group, 121 cokernel, 8, 9, 40, 58, 61, 102, 117 cokernel iff coequaliser, 33, 40 colimit, 90 comma category, see slice category 128 INDEX commutative object, 15 commutator, 16, 74, 111 commuting morphisms, 24 comonad, 37, 38, 47, 53, 55 forgetful/free —, 37, 49, 54 projective class of —, 49 relative —, 38, 54, 60 comonadic cohomology, 37 comonadic homology, 2, 19, 37, 38–47, 58– 60 G-acyclicity, 38, 39 G-connectedness, 39 second variable, 38–46, 59 Comparison Theorem, 52, 59 composition of extensions, 62 cone, 85, 86, 92, 95, 97–98, 104, 112, 114 connected equivalence relations, 28 connecting homomorphism, 77, 79, 86, 99, 112, 114, 118 contractibility relative —, 51, 52 contractible simplicial object, 33, 39 homology, 33 contractible simplicial set, 50 contraction, 33 cooperating morphisms, 24, 26, 79 cooperator, 24 coproducts, 61 crossed modules, 8, 17, 21, 54, 122 cycles, see object of n-cycles degeneracy operator, 31, 37 derived functor, 82 derived series, 18 diagonal, 15, 23, 117 diagram (with limit), 85, 86, 90, 92–95, 97– 98, 102, 104, 105, 112, 114 diagram lemmas Five Lemma, 11, 43, 45 Short —, 8, 116, 122 3× 3 Lemma, 11, 71, 103, 115 Noether’s Iso. Theorem, 12, 75, 80, 104 Snake Lemma, 12, 32 direct image, 13, 16, 79, 99 domain, 78, 90, 103 double arrow, 62 double extension, 63, 64, 66, 80 kernel is extension, 64, 68 symmetry, 64, 65 Eckmann-Hilton, 38 Eilenberg, see Chevalley-Eilenberg endomorphism of p over A, 104–108 enlargement of domain, 34, 57 enough projectives, 49, 49, 74–81, 85–91, 101–109, 114 epimorphism normal —, 8 regular —, 8–11, 13, 60, 63, 64, 67, 122 pullback-stable —s, 8 split —, 62, 65 equaliser, 34 equivalence relation, 26 centralising —s, 27 connected —s, 28 normal to an —, 27 Everaert, 5, 31, 36, 61, 74 Everaert sequence, 2, 77, 81, 85, 86, 90, 99, 112 exact category, 8, 61 exact sequence, 9, 99 long —, see homology, — short —, see short exact sequence examples abelianisation functor, 16 Birkhoff subcategories, 17 central extensions Galois, 21 Huq, 25 centralisation functor, 70 class of extensions, 63 comonads, 37 cooperating morphisms, 25 Galois structures, 20 homology finite cyclic groups, 105 finite groups, 96 generators and relations, 106 Heyting semi-lattices, 100 — of projective objects, 95 reflection is identity, 92 — of zero, 95 zero reflection, 98 Hopf formula, 74 Kan projective class, 54 regular projectives, 49 129 INDEX relative Kan property, 50 satellite, abelian, 83 semi-abelian categories, 8 strongly E-Birkhoff subcategories, 67 two comonads on R-Mod, 60 Ext groups, 38, 60 extension central — finite —, 97 Galois, 21, 29, 67–69, 81, 94, 95, 98, 111–122 higher —, 71 Huq, 25, 28, 29, 111–122 isomorphism class of —s, 116 split —, 24, 76, 78, 117 universal —, 95, 96, 114, 121 weakly universal, 95 central — of A by K, 116 sum, 117 class of —s, 62, 68 E1, 65 composition of —s, 62 double —, 63, 64, 66, 80 symmetry, 64, 65 double — has kernel an —, 64, 68 Galois, 20, 68 higher —, 63, 66, 67, 94 higher central —, 71 higher trivial —, 72 kernel of —, 63 la´ma´ra —, 113, 114, 120 n-fold —, see higher — normal —, 21 pullback-stable —s, 62, 65, 94 regular epimorphism, 9 split —, 95, 97 stem —, 96, 113, 114, 120 stem cover, 113, 114, 120, 121 trivial —, 21, 24, 78, 98, 112, 114, 117, 120 higher —, 72 extension preserving functor I, 19 J , 67 extensions of A, 93, 105 higher —, 94 face operator, 31, 37 filler of a horn, see horn filler finite groups, 96 Five Lemma, 11, 43, 45 Short —, 8, 116, 122 fixed points, 81, 105–108 forgetful/free comonad, 37, 49, 54 G-acyclicity, 38, 39 G-connectedness, 39 G-exact sequence, 39, 40–45 G-left derived functors, 39, 42 G-projective object, 49 Galois structure, 20, 28, 68, 69, 72 Gran, 5, 61 groups, 8, 15–17, 19–21, 25, 37, 50, 54, 61, 67, 70, 74, 96, 98, 105, 106 abelian —, 16, 17, 21, 37, 67, 70 integral homology of —, 1, 19, 37, 74, 78, 96, 105–108 nilpotent —, 18, 21, 22, 70 perfect —, 96 solvable —, 18, 21, 22, 70 Heyting semi-lattices, 8, 100 higher central extension, 71 higher dimensional arrow, 62 higher extension, 63, 66, 67, 94 higher presentation, 73, 75 higher trivial extension, 72 Hochschild, 38, 60 Hom functor, 38, 60 homological category, 10 homology abelian object, 32, 79, 99 Barr-Beck —, 37 — of chain complexes, 16 comonadic —, 2, 19, 37, 38–47, 58–60 G-acyclicity, 38, 39 G-connectedness, 39 second variable, 38–46, 59 contractible simpl. obj., 33 Everaert sequence, 2, 77, 81, 85, 86, 90, 99, 112 fixed points, 81, 105–108 — via Hopf formulae, 2, 19, 74, 86, 111– 122 integral group —, 1, 19, 37, 74, 78, 96, 105–108 — via Kan extensions, 4, 19, 91 130 INDEX — via limits, 4, 19, 91, 111 central extensions, 95, 112 equivalent diagrams, 92 with projectives, 104 long exact sequence abelian, 83 Everaert sequence, 2, 77, 81, 85, 86, 90, 99, 112 second variable, 39 simplicial objects, 33, 39 object of B, 98 — of a projective object, 95 — with projectives, 101–109 — without projectives, 91–101 — via satellites, 4, 19, 91 semi-abelian —, 17 — of a simplicial object, 32, 56 — is zero, 92, 95, 98, 100 — of zero, 95 homotopic maps, 52, 58 — have same homology, 58 homotopy, 53, 55 homotopy equivalent, 55 Hopf formula, 2, 61, 73, 74, 80, 96, 101, 102, 105, 111 alternative —, 75–77 Hopf homology, see homology horn, 34, 52, 53, 58 horn filler, 34, 52, 53, 58 Huq central, 25, 26, 28, 79, 99 image, 9, 41, 60, 99 direct —, 13, 16, 79, 99 image factorisation, 9, 13, 41 initial functor, 93, 94 initial object, 95 initial object of n-cube, 72, 73, 75, 102 initial subcategory, 93, 104 integral group homology, 1, 19, 37, 74, 78, 96, 105–108 internal abelian group, 15 internal Kan property, 34, 50 isomorphism class of central extensions, 116 isomorphism class of stem covers, 120 Jacobi identity, 19 Janelidze, 20, 81, 82 jointly epic pair, 10, 15, 24 Kan extension, 81–83, 86, 91, 99 Kan projective class, 53, 55, 58 Kan property, 34, 52, 56 internal —, 34, 50 relative —, 49, 53 Kan simplicial set, 34, 50 kernel, 7, 25, 27–29, 41, 75, 95, 96, 102 kernel functor, 62, 88 kernel of double extension, 64, 68 kernel of extension, 63 kernel pair, 23, 26, 51 la´ma´ra extension, 113, 114, 120 Leibniz algebras, 8, 18, 20, 22, 25, 70 Leibniz identity, 18 Lie algebras, 8, 16, 18–20, 22, 25, 70, 74 Lie-centre, 22, 71 limit, 85, 86, 90, 92–95, 97–98, 102, 104, 105, 112, 114 closed under —s, 67 long exact homology sequence abelian, 83 Everaert sequence, 2, 77, 81, 85, 86, 90, 99, 112 second variable, 39 simplicial objects, 33, 39 Mal’tsev category, 10, 21, 34 modules, 37, 54, 60 monadic category, 37 monomorphism normal —, 8, 9, 13, 27 normal to an equiv. rel., 27 Moore complex, 32, 34 object of n-cycles, 32, 57 multiplication, see abelian object n-fold extension, see higher extension nil-centre, 22, 70 nilpotent groups, 18, 21, 22, 70 3× 3 Lemma, 11, 71, 103, 115 Noether’s Iso. Theorem, 12, 75, 80, 104 normal epimorphism, 8 normal extension, 21 normal monomorphism, 8, 9, 13, 27 to an equivalence relation, 27 normal subobject, 13, 15, 26, 117 associated —, 27 normalisation (of an equiv. rel.), 27 131 INDEX normalisation functor, 32 — is exact, 32, 36 normalised chain complex, see Moore com- plex object of n-cycles, 32, 57 Ω-groups, 8 P-exact simplicial object, 51 P-resolution, 50, 52, 59 perfect group, 96 perfect object, 95, 96, 114, 114–122 pointed category, 7, 61 pointwise satellite, 85, 85–91 presentation, see projective presentation presentation of A, 73, 81, 83, 86, 102, 104, 105, 111 prism, 45 projective class, 49 Kan —, 53, 55, 58 projective class of a comonad, 49 projective presentation, 61, 73, 75, 78, 81, 83, 86, 105, 111 projectives, 49, 73, 81, 95 enough —, 49, 49, 74–81, 85–91, 101– 109, 114 regular —, 49, 54 proper chain complex, 16, 32 proper morphism, 9 protomodular category, 8, 61 pullback, 71, 72, 93, 98 — cancellation, 10, 13 — reflects monos, 10 isomorphic kernels, 12, 64, 76 mono between cokernels, 12, 64, 69, 75– 77, 103, 117 pullback-stable — extensions, 62, 65, 94 — regular epimorphisms, 8 pushout, 57 generalised regular —, 14, 57 isomorphic cokernels, 13, 57 regular —, 13, 19, 64 quotient, see regular quotient R-modules, 37, 54, 60 reflection, see reflector reflective subcategory, 17, 66, 67, 72 reflector, 17, 20, 68, 72, 85, 91, 94, 95, 105 identity, 92 zero, 98 regular category, 8, 34 regular epimorphism, 8–11, 13, 60, 63, 64, 67, 122 pullback-stable —s, 8 regular image, see regular quotient regular projectives, 49, 54 regular pushout, 13, 19, 64 generalised —, 14, 57 regular quotient, 8, 9, 115, 121 closed under —s, 17 relative comonad, 38, 54, 60 relative contractibility, 51, 52 relative Kan property, 49, 53 rings, 8, 54 satellite, 81, 82 composing —s, 88 homology, abelian, 83 homology, semi-abelian, 86 composed, 89 pointwise —, 85, 85–91 Schur multiplier, 96 semi-abelian category, 1, 8 examples, 8 short exact sequence, 63, 77, 83, 103 — of functors, 67–69, 111 Short Five Lemma, 8, 116, 122 simplicial group, 54 simplicial identities, 31 simplicial kernel, 51, 52 simplicial object, 31, 55 augmented —, 33, 37, 39, 50 contractible —, 33, 39 homology, 33 degeneracy operator, 31, 37 face operator, 31, 37 homology, 32 homotopic maps, 52, 58 — have same homology, 58 homotopy equivalent —s, 55 simplicial resolution, 37, 50, 52, 59 simplicial set, 34 contractible —, 50 Kan —, 34, 50 simplicially homotopic, 52, 58 132 INDEX slice category, 85, 93, 95, 104 Smith central, 27 Snake Lemma, 12, 32 solvable centre, 23 solvable groups, 18, 21, 22, 70 split central extension, 24, 76, 78, 117 split epimorphism, 62, 65 split extension, 95, 97 Stallings-Stammbach sequence, 3, 61, 112, 115, 118 stem cover, 113, 114, 120, 121 stem extension, 96, 113, 114, 120 strongly E-Birkhoff, 66, 69, 74–80, 85–105 subdiagonal, 26, 29 subobject closed under —s, 17, 66 sum of central extensions, 117 surjection, 60, 96, 105 symmetry of double extension, 64, 65 tensor, 38, 60 terminal object, 90 terminal object of n-cube, 94, 102 theory of G-left derived functors, 39, 42 Tierney-Vogel, 50, 59 Tor groups, 38, 60 trivial extension, 21, 24, 78, 98, 112, 114, 117, 120 higher —, 72 trivialisation, 72, 75 unital category, 10, 24 universal central extension, 95, 96, 114, 121 unnormalised chain complex, 32 Van der Linden, 5, 31, 36, 48, 61, 82 variety of algebras, 37, 55 Vogel, see Tierney-Vogel weakly universal central extension, 95 zero map, 7 zero-object, 7 zigzag, 56 133