Characterisations of -pure-injectivity in triangulated categories and applications to endoperfect objects.
Abstract.
We provide various ways to characterise -pure-injective objects in a compactly generated triangulated category. These characterisations mimic analogous well-known results from the model theory of modules. The proof involves two approaches. In the first approach we adapt arguments from the module-theoretic setting. Here the one-sorted language of modules over a fixed ring is replaced with a canonical multi-sorted language, whose sorts are given by compact objects. Throughout we use a variation of the Yoneda embedding, called the resticted Yoneda functor, which associates a multi-sorted structure to each object. The second approach is to translate statements using this functor. In particular, results about -pure-injectives in triangulated categories are deduced from results about -injective objects in Grothendieck categories. Combining the two approaches highlights a connection between sorted pp-definable subgroups and annihilator subobjects of generators in the functor category. Our characterisation motivates the introduction of what we call endoperfect objects, which generalise endofinite objects.
2020 Mathematics Subject Classification:
18E45, 18G80 (primary), 03C60 (secondary)1. Introduction.
The model theory of modules refers to the specification of model theory to the module-theoretic setting. Fundamental work, such as that of Baur [2], placed focus on certain formulas in the language of modules, known as pp-formulas. In particular, module embeddings which reflect solutions to pp-formulas, so-called pure embeddings, became of particular interest. This served as motivation to study modules which are pure-injective: that is, injective with respect to pure embeddings.
In famous work of Ziegler [21], a topological space was defined whose points are indecomposable pure-injective modules. The introduction of the Ziegler spectrum proved to be a groudbreaking moment in this branch of model-theoretic algebra, and interest in understanding pure-injectivity has since grown. Specifically, in work such as that of Huisgen-Zimmerman [11], functional results appeared in which pure-injective and so-called -pure-injective modules were characterised. These characterisations are well documented, for example, by Jensen and Lenzing [14].
A frequently used tool in these characterisations is the relationship between a module and its image in a certain functor category. To explicate, the functor is given by the tensor product, restricted to the full subcategory of finitely presented modules. For example, a module is pure-injective if and only if the corresponding tensor functor is injective. Subsequently one may convert statements about pure-injective modules into statements about injective objects in Grothendieck categories, and translate problems and solutions back and forth.
For example, Garcia and Dung [7] developed the understanding of -injective objects in Grothendieck categories by building on work of Harada [9], which generalised a famous characterisation of -injective modules going back to Faith [6]. These authors showed that, as above, such developments helped simplify arguments about -pure-injective modules.
In this article we attempt to provide, in a utilitarian manner, some analouges to the previously mentioned characterisations. The difference here is that, instead of working in a category of modules, we work in a triangulated category which, in a particular sense, is compactly generated. Never-the-less, the statements we prove and arguments used to prove them are motivated directly from certain module-theoretic counterparts.
The notion of compactness we refer to comes from work of Neeman [19], where the idea was adapted from algebraic topology. Krause [18] provided the definitions of pure monomorphisms, pure-injective objects and the Ziegler spectrum of a compactly generated triangulated category. Garkusha and Prest [8] subsequently introduced a multi-sorted language for this setting, which mimics the role played by the language of modules. They then gave a correspondence between the pp-formulas in this multi-sorted language and coherent functors. In our main result, Theorem 1.1, we use the following notation and assumptions.
-
•
is a compactly generated triangulated category with all small coproducts.
-
•
is the full subcategory of consisting of compact objects, which is assumed to be skeletally small.
-
•
is the category of abelian groups.
-
•
is the category of contravariant additive functors .
-
•
is a set of generators of where each is finitely presented.
-
•
is the restricted Yoenda functor, which takes an object to the restriction of the corepresentable .
Recall that, by the Brown representability theorem, since the category has all small coproducts, it has all small products; see Remark 4.4.
Theorem 1.1.
For any object of the following statements are equivalent.
-
(1)
For any set the coproduct is pure-injective.
-
(2)
The countable coproduct is pure-injective.
-
(3)
For any generator each ascending chain of -annihilator subobjects of must eventually stabilise.
-
(4)
For any set the morphism from to the product , given by the universal property, is a section.
-
(5)
For any object of each descending chain of pp-definable subgroups of of sort must eventually stabilise.
-
(6)
is pure injective, and for any set , is isomorphic to a coproduct of indecomposable pure-injective objects with local endomorphism rings.
The proof of Theorem 1.1 is at the end of the article. The equivalences of (1), (2), (3) and (4) in Theorem 1.1 follow by directly combining work of Garcia and Dung [7] and work of Krause [18]. The equivalence of (5) and (6) with the previous conditions is more involved. For (5) we adapt ideas going back to Faith [6], whilst applying results due to Harada [9] and Garcia and Dung [7]. For (6) we adapt arguments of Huisgen-Zimmerman [11].
The article is organised as follows. In §2 we recall some prerequisite terminology from multi-sorted model theory. In §3 we specify to compactly generated triangulated categories by recalling the canonical multi-sorted language of Garkusha and Prest [8]. In §4 we gather results about products and coproducts of pp-definable subgroups in this context, following ideas of Huisgen-Zimmerman [11]. In §5 we highlight a connection between annihilator subobjects of finitely generated functors and pp-definable subgroups, where the presentation of the functor determines the sort of the subgroup. In §6 we begin combining the results developed in the previous sections with results of Krause [18]. In §7 we complete the proof of Theorem 1.1. In §8 we introduce endoperfect objects, and see applications of Theorem 1.1.
2. Multi-sorted languages, structures and homomorphisms.
There are various module-theoretic characterisations of purity in terms of positive-primitive formulas in the underlying one-sorted language of modules over a ring. Similarly, purity in compactly generated triangulated categories may be discussed in terms of formulas in a multi-sorted language. Although Definitions 2.1, 2.3, 2.5, 2.6 and 2.7 are well-known, we recall them for completeness. We closely follow [5, §2, §7] for consistency.
Definition 2.1.
[5, Definition 34] For a non-empty set , an -sorted predicate language is a tuple where:
-
(1)
each is called a sort;
-
(2)
the symbol denotes a non-empty set of sorted predicate symbols;
-
(3)
the symbol denotes a set of sorted function symbols;
-
(4)
the symbol denotes an arity function ;
-
(5)
the symbol denotes a sort function, taking any -ary (respectively ) to a sequence in of length (respectively ).
When in condition (5) we often write (respectively ). Note that functions with have a sort.
For each sort we introduce a countable set of variables of sort . The terms of each have their own sort, and are defined inductively by stipulating: any variable of sort will be considered a term of sort ; and for any with and for any terms of sort respectively, is considered a term of sort . Note that constant symbols, given by functions with , are also terms.
The atomic formulas with which is equipped are built from the equality between terms of common sort , together with the formulas where , and where each is a term of sort . First-order formulas in are built from: the variables of each sort; the atomic formulas; binary connectives , , and ; negation ; and the quantifiers and .
A positive-primitive or pp formula (with free) has the form
where each is an atomic formula (see, for example, [10, p.50]).
One may build a theory for a multi-sorted language by specficying a set of axioms. For our purposes these axioms are those charaterising objects and morphisms in a fixed category. We explain this idea by means of examples.
Example 2.2.
[14, §6] Let be a unital ring. We recall how the language of -modules may be considered as a predicate language in the sense of Definition 2.1. In this case there is only one sort, which we ignore, and which uniquely determines the function . Let . Let where is binary and is unary.
In Definition 2.3 the notion of a structure is recalled. For the language we have that this notion, together with certain axioms, recovers the defining properties of left -modules. Later we consider homomorphisms.
Definition 2.3.
[5, Definition 35] Fix a set and an -sorted predicate language . An -structure is a tuple
such that:
-
(1)
the symbol denotes a family of sets ;
-
(2)
if then is a subset of ; and
-
(3)
if then is a map .
Denoting the cardinality of any set by , let and, for as above, let be the sum of the cardinalities as runs through .
The so-called one-sorted language from Example 2.2 is trivial in the sense that there is only one possibility for the sort function. In this way, Example 2.4 is a non-trivial example of the multi-sorted languages we recalled in Definition 2.1.
Example 2.4.
[14, §9] Here we recall an example of an -sorted predicate language which is in contrast to Example 2.2. The predicates in this language will be the unary symbols and of sort , and of sort . The functions in this language will be the ternary symbols and where and . After specifying the appropriate axioms, structures are tuples where is a unital ring and is a left -module.
In this way one interprets the symbols and as the additive and multiplicative identities in . Similarly the symbol is interpreted as the additive identity in . In Definition 2.5 the notion of a homomorphism between structures is recalled. In this sense, a homomorphism is given by a pair where is a homomorphism of rings and is a homomorphism of left -modules with the action of on given by .
Definition 2.5.
[5, Definition 3] Fix a non-empty set , an -sorted predicate language and -structures and . By an -homomorphism we mean a family of functions such that:
-
(1)
if then is the set of such that ;
-
(2)
and if then for all we have .
Note that, in the notation of Definition 2.5, [5, Theorem 17] says that a collection of functions defines an -homomorphism if and only if, whenever is an atomic formula with and lies in , we have that
We are now ready to recall the idea of purity coming from model theory.
Definition 2.6.
[5, Definition 36] Fix a set and an -sorted predicate language . By an -embedding we mean an -homomorphism such that:
-
(1)
if is an atomic formula with , then for all , if and only if .
By an -pure embedding we mean an -homomorphism such that:
-
(2)
if is a pp formula with , then for all , if then .
Note that -pure embeddings are -embeddings. Note also that the statement of Definition 2.5(2) is the contrapositive of the definition in [10, p.50], so in this sense, over a ring and in the notation from Example 2.2, an injective left -module homomorphism is pure if and only if it is an -pure embedding.
Definition 2.7.
Fix a non-empty set , an -sorted predicate language and -structures and . We say is an -substructure of if for each and, labelling these inclusions , the family defines an -homomorphism . If, additionally, is an -pure embedding, we say is an -pure substructure of .
3. Purity in the canonical language of a triangulated category.
We now specify the setting of multi-sorted model theory outlined in §2. Throughout the sequel we consider a fixed compactly generated triangulated category; see Assumption 3.3. Before recalling Definition 3.2 we fix some notation.
Notation 3.1.
Let be an additive category. Denote the hom-sets and the identity maps . For any set and any collection of objects in , if the categorical product exists in , we write for the natural morphisms equipping it, in which case the universal property gives unique morphisms such that is the identity on for each . Similarly will denote the morphisms equipping the coproduct if it exists, in which case there exist unique morphisms such that for each .
Fix an object in and consider the covariant functor . Note that both the product and coproduct of the collection exist in the category of abelian groups. We identify with the subgroup of consisting of tuples such that for all but finitely many .
Consequently, if exists in then map from the universal property is given by for each . Similarly if exists in then map from the universal property is given by . In general each of the morphisms are isomorphisms.
Definition 3.2.
[19, Definition 1.1] Let be a triangulated category with suspension functor , and assume all small coproducts in exist. An object of is said to be compact if, for any set and collection of objects in the morphism is an isomorphism. Let be the full triangulated subcategory of consisting of compact objects.
Given a set of compact objects in we say that is compactly generated by if there are no non-zero objects in satisfying for all (or, said another way, any non-zero object gives rise to a non-zero morphism for some ). If is compactly generated by we call a generating set provided for all .
Assumption 3.3.
In the remainder of §3 fix a triangulated category with suspension functor , and we assume:
-
•
that has all small coproducts;
-
•
that is compactly generated by a generating set ; and
-
•
that the subcategory of compact objects is skeletally small.
Definition 3.4 and Remark 3.6 closley follow [8, §3], in which a multi-sorted language associated to the category is introduced.
Definition 3.4.
In what follows let denote a fixed set of objects in given by choosing exactly one representative of each isomorphism class. Such a set exists because we are assuming that is skeletally small.
[8, §3] The canonical language of is given by an -sorted predicate language , defined as follows. The set consists of a symbol with for each . The set consists of: a ternary symbol with for each ; and a unary operation with for each morphism with . Variables of sort will be denoted .
Notation 3.5.
Suppose is any additive category. We write (respectively ) for the category of additive covariant (respectively contravariant) functors where is the category of abelian groups.
For any object of we let denote the object of defined by restriction of to compact objects. We write
to denote the restricted Yoneda functor. That is, takes an object to , and takes a morphism to the natural transformation given by defining, for each compact object , the map by .
Remark 3.6.
[8, §3] Consider the theory given from the set of axioms expressing the positive atomic diagram of the objects in including the specification that all functions are additive. In this way, the category of models for the above theory is equivalent to the category where objects of are regarded as structures for this language.
That is, in the notation of Definition 2.3, we let , we interpret the predicate symbol as the identitly element of , we interpret as the additive group operation on , and we interpret as the map given by .
Lemma 3.7.
Let and be objects in with corresponding -structures and . Then the choice of made in Definition 3.4 defines a bijection between morphisms (that is, natural transformations) in and -homomorphisms .
Proof.
Any object of lies in the same isoclass as some unique , in which case we choose an isomorphism . In this way, any object defines an isomorphism by precomposition with . Recall that here the -structure is defined by setting for each sort . In case we assume, without loss of generality, that . Define the required bijection as follows. Fix an -homomorphism . For any object of define the function by . Converlsey, fixing a natural transformation , let for each .
It suffices to explain why these assignments swap between morphisms in and -homomorphisms. To do so, we explain why the compatability conditions which define these morphisms are in correspondence. To this end, note firstly that the preservation of (the predicate symbol and the function symbols ) is equivalent to saying that each function is a homomorphism of abelian groups. Letting be a morphism in and , for any object of the sorted function symbol is interpreted in by the equation . Thus, by construction, saying that the function symbols are preserved is equivalent to saying that the collection of (for compact) defines a natural transformation. ∎
In what follows we discuss the notion of purity in the context of compactly generated triangulated categories.
Definition 3.8.
[18, Definition 1.1] We say in is a pure monomorphism if is injective for each compact object .
Now we may begin to build results which mimic well-known ideas from the model theory of modules. To consistently compare and contrast our work with the module-theoretic setting, we use a book of Jensen and Lenzing [14]. In this spirit, Lemma 3.9 is analogous to [14, Theorem 6.4(i,ii)], and Lemma 4.2 is analogous to [14, Proposition 6.6]. Similar analogies are found throughout the sequel.
Lemma 3.9.
A morphism is a pure monomorphism if and only if the image under the bijection in Lemma 3.7 is an -pure embedding.
Proof.
By [8, Proposition 3.1] any pp-formula is equivalent to a divisibility formula where is morphism and . By Definition 2.6, is an -pure embedding if and only if, for any morphism with and any pair , if then . Since any compact object is isomorphic to an object in , this is equivalent to the condition which says that, for each compact object , the morphism given by is injective. ∎
4. Products, coproducts, coherent functors and pp-formulas.
Recall, from Definition 2.1, that pp-formulas in are those lying in the closure of the set of equations under conjunction and existential quantification.
Definition 4.1.
[8, §2] Given and an object of with -structure , a pp-definable subgroup of of sort is the set of solutions (in ) to some pp-formula in one free variable of sort .
For any morphism in and any object in recall the map is defined by precomposition. In this case let
If and and are isomorphisms in (as in the proof of Lemma 3.7), then defines a isomorphism in where is the pp-formula where .
Lemma 4.2.
Let and let be a pp-formula in one free variable of sort . If is a pure monomorphism then .
Proof.
We continue, slightly abusing terminology, by reffering to any set of the form (for some ) as a pp-definable subgroup of of sort . Lemma 4.3 is an analgoue of the corresponding result for module categories; see for example [20, Corollary 2.2(i)]. Note that we state it here for convenience, but that it is only used in §8, and not in the proof of Theorem 1.1.
Lemma 4.3.
For any object in and any morphism the pp-definable subgroup is a left -submodule of .
Proof.
By the associativity of the composition of morphisms in the set is closed under postcomposition with endomorphisms of . ∎
Recall that a covariant functor is coherent if there is an exact sequence in of the form . If and are morphisms in with , then for any . So, for the pp-formula in the assignment of objects defines a functor.
Furthermore, by [8, Lemma 4.3] these functors are coherent, and any such coherent functor arises this way. We now recall that the categories we are considering have all small products.
Remark 4.4.
Lemma 4.5.
Let and let be a pp-formula in one free variable of sort . For any set and any collection of objects in the restrictions of and define isomorphisms of abelian groups
Proof.
By the existence of small products and coproducts in and the functorality of , the universal properties give morphisms and . By [8, Lemma 4.3] the functor is coherent, so by the equivalence of statements (1) and (3) from [18, Theorem A] the morphisms and are isomorphisms. It is straightforward to check that and are the respectively restrictions of and from Notation 3.1. ∎
We now adapt some technical results from work of Huisgen-Zimmerman [11], in which a (now well-known) charaterisation of -pure-injective modules was given. Our adaptations, namely Lemmas 4.8 and 7.7, are used in the sequel.
Notation 4.6.
Fix collections and of objects in and let where and exist. By Notation 3.1, and denote the morphisms equipping the product and coproduct, and and denote the morphisms given by universal properties such that and for each and .
Corollary 4.7.
Consider Notation 4.6, let be a morphism with , and let . Then there is an isomorphism
where is given by the universal property, and whose inverse is
Proof.
By Lemma 4.5 the restriction of and define isomorphisms
Letting and , the proof is straightforward. ∎
The proof of Lemma 4.8 follows the proof of the cited result of Huisgen-Zimmerman.
Lemma 4.8.
Proof.
Considering Notation 3.1, for each , we have that is (the inclusion) given by sending (where ) to the sequence , the initial terms of which are . Similarly, we have that is the restriction of for each and each .
We may assume is non-empty since otherwise the statement is automatic. Assume for a contradiction that for any and any finite subset of , there exists with and there exists such that . Following [11, Lemma 4], we firstly claim that there exists: a strictly increasing sequence of integers ; a sequence of pairwise distinct elements of (so, where for all ); and a sequence of -tuples where
for all and all with . We proceed inductively. Choose an element . Let and . By our assumption that the conclusion is false, there exists an integer and an element where . Hence there exists where , and so .
We now iterate this process, yielding sequences , and where is a positive integer, and and such that , , and . Now fix . Since we are considering coproducts of abelian groups, note that we have for all but finitely many . For each define the map in by if for some , and otherwise. Now let
By construction we have for all , and if then
Now, writing where for all , the above gives for all , so we may define
So we have , and therefore
since by definition . Our calculations above likewise show that if then we have . This verifies our initial claim. Now let vary. Since for all , for each observe that there are finitely many with . Hence the sum
is well-defined. Now, for each , by combining everything so far with Corollary 4.7, we have
Now recall that . Since is descending and for all , we have whenever . Together with the above, this shows
for all . Since lies in the coproduct , and therefore must have had finite support over , we have a contradiction, since the set is in bijection with . ∎
5. Annihilator subobjects and pp-definable subgroups.
Recall that, in a category with all small coproducts, a set of objects is called a set of generators provided, for each object , there is an epimorphism . In case is a singleton we say the category has a generator. Recall an additive category is Grothendieck provided: is abelian; has all small coproducts; has a generator; and the direct limit of any short exact sequence in is again exact.
Remark 5.1.
Let be a Grothendieck category. Recall that an object of is finitely presented provided the functor commutes with direct limits. Recall that an object of is finitely generated provided there is an exact sequence in . The categories considered both in work of Garcia and Dung [7] and in work of Harada [9] were Grothendieck categories with a set of finitely generated generators.
Thus, in Definition 5.2 and Lemmas 5.3 and 5.4, we specify various definitions and results from [7] and [9] to (which can be done by Remark 5.1). We now recall a notion introduced by Harada.
Definition 5.2.
[9, §1] Let be a Grothendieck category with a set of finitely generated generators. Let and be objects in . A subobject of is said to be an -annihilator subobject of provided where the intersection is taken over morphisms running through some .
Lemma 5.3 focuses on a particular context of Definition 5.2. That is, we specifiy to the locally coherent category and consider the image of objects under . Note Lemma 5.3 was written only to simplify the proof of Lemma 5.4.
Lemma 5.3.
Let be an epimorphism in where the object of is compact. If and are objects of and respectively, then for any -annihilator subobject of (where ) we have
Proof.
It suffices to assume . Let . Since is onto, for some . Since and each are morphisms in the diagram of abelian groups given by
commutes. By the commutativity of the diagram we have and (hence) . Now suppose . By the above this means for all . Hence lies in the ride hand side of the required equality. The reverse inclusion is straightforward. ∎
Lemma 5.4 is based on a proof of a given by Huisgen-Zimmerman [12, Corollary 7] of a well-known characterisation of -injective modules due to Faith [6, Proposition 3]. We use Lemma 5.4 to simplify the proof of Lemma 6.6, a key result employed in the sequel.
Lemma 5.4.
Let and be objects in and respectively, and let be a sequence in which is exact. Any strictly ascending chain of (-annihilator subobjects of ) gives a strictly descending chain of (pp-definable subgroups of of sort ).
Proof.
Suppose is a strictly ascending chain of -annihilator subobjects of , say where, for each integer , we have for some subset of morphisms in of the form .
We assume without loss of generality. For each there is an object of for which , and we choose . Let be the epimorphism in giving the exact sequence . Since is onto and we have for some morphism . By Lemma 5.3, since we have that for all . Similarly, since there exists such that .
We now follow the proof of [8, Proposition 3.1]. Since is triangulated any morphism in yields a triangle in and an exact sequence in
given by applying the contravaraint functor . In other words, is a pseudocokernel of , and so any morphism in with must satisfy for some .
Let for each . Combining what we have so far, for each we have , so , and so for some morphism , and so . On the other hand, if then which contradicts that , and so . This gives a strict descending chain
A direct application of [8, Proposition 3.1] shows that each finite intersection has the form for some morphism in of the form . So the chain above is, as required, a strictly descending chain of pp-definable subgroups of of sort . ∎
Definition 5.5.
[9, §1] Let be a Grothendieck category with a set of finitely generated generators. Fix an object of . We say that is -injective if, for any set , the coproduct is injective. We say that is fp-injective if, whenever is an exact sequence in where is finitely presented, any morphism extends to a morphism ; see [7, §1].
For the proof of Corollary 5.8 we recall two results: Proposition 5.6, due to Garcia and Dung, characterises -injectivity in the fp-injective setting; and Lemma 5.7, due to Krause, shows that it is sufficient to consider the fp-injective setting.
Proposition 5.6.
[7, Proposition 1.3] Let be an fp-injecitve in a Grothendieck category which has a set of finitely presented generators . Then is -injective if and only if, for each , every ascending chain of -annihilator subobjects of must stabilise.
Lemma 5.7.
[17, Lemma 1.6] For any in the image is fp-injective.
Corollary 5.8.
Let be an object in such that for any compact object of we have that each descending chain of pp-definable subgroups of of sort must stabilise. Then the image of in is -injective.
Proof.
We prove the contrapositive, so we assume is not -injective. Recall, from Remark 5.1, that is locally coherent, and so it is a Grothendieck category with a set of finitely presented generators.
Note is fp-injective by Lemma 5.7, and combining our initial assumption with Proposition 5.6 shows that, for some , there exists a strictly ascending chain of -annihilator subobjects of . Since is finitely presented, there is an exact sequence of the form in where and lie in .
By Lemma 5.4 the aforementioned ascending chain strict ascending chain gives rise to a strictly descending chain of pp-definable subgroups of of sort . ∎
6. -pure-injective objects and canonical morphisms.
Definition 6.1.
Recall Notation 3.1. Let be a set and let be an object of . By the universal properties of the product and coproduct of the collection , there exists a unique summation morphism and a unique canonical morphism for which and for each .
Proposition 6.2.
Let be an object of and let be a set. Then the canonical morphism is a pure monomorphism.
Proof.
Let be a object in which is compact. In general: the morphism is an isomorphism; the canonical morphism is injective; and is the composition .
Since is compact the morphism is an isomorphism. This shows is injective if is compact, and so is a pure monomorphism. ∎
Definition 6.3.
[17, Definition 1.1] An object of is called pure-injective if each pure monomorphism is a section, and is called -pure-injective if, for any set , the coproduct is pure-injective.
At this point it is worth recalling some characterisations of purity due to Krause. Theorem 6.4 is analogous to [14, Theorem 7.1 (ii,v,vi)].
Theorem 6.4.
[18, Theorem 1.8, (1,3,5)] For an object of the following statements are equivalent.
-
(1)
The object of is pure-injective.
-
(2)
The object of is injective.
-
(3)
For any set the morphism factors through the morphism .
Proposition 6.5.
An object of is -pure-injective if and only if, for each set , the canonical morphism is a section.
Proof.
Assume that is -pure-injective and that is a set. By assumption the domain of is pure-injective. Since is a pure monomorphism, this means it is a section. Supposing conversley that is a section for each set , it remains to show that is -pure-injective. Choose a set and let . It suffices to prove is pure-injective. Let be any set and consider the collection . By Theorem 6.4 it suffices to find a map such that . Let .
For each the morphisms satisfy the universal property of the coproduct , and so we assume without loss of generality. Consider the morphisms for each . Since we have by uniqueness. Consequently is the identity on . By the universal property of the product, there is a morphism such that for each . It suffices to let . By the uniqueness of the involved morphisms, it is straightforward to see that .
∎
Lemma 6.6.
If and is a -pure-injective object in then every descending chain of pp-definable subgroups of of sort stabilises.
Proof.
For a contradiction we assume the existence of a strictly descending chain of (pp-definable subgroups of of sort ) for some . Hence there is a collection of compact objects such that for each . By our assumption we may choose elements such that . By Proposition 6.5 the canonical morphism is a section, and so there is some morphism such that is the identity on . After applying the functor (from to the category of abelian groups) this means is the identity on . Let . Fix and let be the formula . Let . Recall is always an isomorphism, and since is compact, is an isomorphism. Define by
Let . The contradiction we will find is that for all , which contradicts that has codomain . Fix . Now let
where the first entries of are .
Note that . Furthermore, since the chain is descending, we have for all and so . By Lemma 4.2(ii), the restrictions of and respectively define isomorphisms and . Similarly restricts to define a morphism . Altogether we have that restricts to a morphism . Let , and so for all .
Recall it suffices to show where . From the above,
and so as otherwise . ∎
We now recall a result of Krause which is used heavily in the sequel.
Corollary 6.7.
[17, Corollary 1.10] The functor gives an equivalence between the full subcategory of consisting of pure-injective objects and the full subcategory of consitsting of injective objects.
Note that, since is additive, by Corollary 6.7 a pure-injective object is indecomposable if and only if is an indecomposable (injective) object.
Remark 6.8.
Corollary 6.9.
The functor gives an equivalence between the full subcategory of consisting of -pure-injective objects and the full subcategory of consitsting of -injective objects.
Proof.
By definition, an object of is -pure-injective if and only if, for every set , the coproduct is pure-injective. By Corollary 6.7 and Remark 6.8, given any such , is pure-injective if and only if is injective.
This shows that induces a functor from the full subcategory of -pure-injectives in and the full subcategory of consisting of -injectives. That this functor is full, faithful and dense, follows by Corollary 6.7 together with the fact that any -pure-injective object of is pure-injective. ∎
Corollary 6.10.
Let be a -pure-injective object of and let .
-
(1)
For any set , the objects and are -pure-injective.
-
(2)
If is a pure monomorphism in then is -pure-injective and is a section.
Proof.
Corollary 6.11.
[7, Corollary 1.6] Let be a Grothendieck category with a set of finitely generated generators. Any -injective object of is a coproduct of indecomposable objects.
Lemma 6.12 is analogous to the implication of (v) by (i) in [14, Theorem 8.1]. The case shows that -pure-injectives decompose into indecomposables.
Lemma 6.12.
If is a -pure-injective object of then for any set the product is a coproduct of indecomposable -pure-injectives.
Proof.
Recall is a Grothendieck category with a set of finitely generated generators. Let so that, by Corollary 6.10(1), is -pure-injective, and so is -injective by Corollary 6.9. By Corollary 6.11 this means where is an indecomposable object of for each .
For each , there is a section , which means is injective, and so by Corollary 6.7 we have for some pure-injective object . Since is indecomposable, we have that is indecomposable. Again, applying Corollary 6.7 gives a section in with . Altogether this means is a pure monomorphism into a -pure-injective object, and so is -pure-injective by Corollary 6.10(2). By Remark 6.8 we have that is injective, and so by Corollary 6.7. ∎
7. Completing the proof of the characterisation.
Before proving Theorem 1.1 we note a consequence of the results gathered so far. Recall, from Definitions 2.3 and 3.2, that the cardinality of the -structure underlying any object of is defined and denoted where is a fixed chosen set of isoclass representatives, one for each class. Corollary 7.1 is analogous to [14, Corollary 8.2(iii)].
Corollary 7.1.
There exists a cardinal such that the -structure underlying any indecomposable pure-injective object of has cardinality at most .
Remark 7.2.
We now note a non-trivial complication in our setting of compactly generated triangulated categories, which is absent in the module-theoretic situation. In the spirit of the results presented so far, it is natural to ask if one may adapt the proof of [14, Corollary 8.2(iii)] to prove Corollary 7.1. This seems straightforward at first glance, since there are multi-sorted versions of the downward Löwenheim-Skolem theorem; see for example [5, Theorem 37].
Recall that, in the language of modules over a fixed ring from Example 2.2, any -structure is an object in the category of -modules, and any elementary embedding of -structures is a pure embedding of -modules. The same correspondence between structures need not be true here. Although objects of correspond to structures over by Remark 3.6, the former need not be given by objects of as need not be essentially surjective. Fortunately, here we may instead use Corollary 7.3, a remark due to Krause, which also shortens the proof.
Corollary 7.3.
(See [17, Corollary 1.10]). There is a set of isomorphism classes of objects in which are pure-injective indecomposable.
Proof of Corollary 7.1.
It suffices to let where runs through and runs through the set from Corollary 7.3. ∎
We now proceed toward proving Theorem 1.1. For this we require Corollary 7.8, and to this end, in Theorem 7.4 we recall well-known results about decompositions of coproducts in abelian categories. For consistency we follow [20].
Theorem 7.4.
Let be an abelian category, and let and be collections of objects in . The following statements hold.
-
(1)
[20, p.82, Theorem 4.A7] Suppose the collections and consist of indecomposable objects with local endomorphism rings. If we have then there is a bijection with for all .
-
(2)
[20, p.82, Theorem 4.A11] Suppose that , and that for all , is indecomposable and is the injective hull of . Suppose there is a section . Then there is a subset with .
To apply Theorem 7.4 we use the following observation of Garkusha and Prest.
Lemma 7.5.
[8, Lemma 2.2] A pure-injective object of is indecomposable if and only if the endomorphism ring is local.
Corollary 7.6.
Let be a collection of indecomposable objects in such that the coproduct is pure injective. If is a summand of then there exists such that .
If additionally where each is indecomposable, then there is a bijection with for all .
Proof.
Let for each . Let and for any subset . Since preserves small coproducts by Remark 6.8, and since each is a summand of , each is a summand of . By Theorem 6.4 and Remark 6.8 is an injective object. Thus each is injective. Since each is indecomposable, each is indecomposable. Similarly, each is pure-injective.
Let . Since is a direct summand of by assumption, is a direct summand of . Hence by Theorem 7.4(2) there exists a subset such that . By Corollary 6.7, since we assume is pure injective, is injective, and hence so too are and . Again, by Corollary 6.7 this gives .
Now suppose also where each is indecomposable and, as above, neccesarilly pure-injective. By Lemma 7.5 each of the objects and has a local endomorphism ring. By Corollary 6.7, for any pure-injective object of there is a ring isomorphism
Since , the second claim follows, as above, by Theorem 7.4(1), using again that preserves coproducts by Remark 6.8. ∎
Lemma 7.7 is analogous to the cited result of Huisgen-Zimmerman.
Lemma 7.7.
[11, Lemma 5] Recall and consider Notation 4.6, where . Let be compact and let be a descending chain of pp-definable subgroups of sort . Suppose additionally that each object is both pure-injective and indecomposable.
Then there exists such that, for each collection where is an indecomposable summand of , we have for all with and all but finitely many .
Proof.
Recall the morphisms defined by the composition
Here the isomorphism is defined in Corollary 4.7, and the maps and are the canonical inclusion and projection respectively. By Lemma 4.8 there exists and a finite subset of such that we have the containment for all with and all .
Choose arbitrary with , and choose arbitrary with for each . Let so that is a summand of . By Corollary 7.6, since is a coproduct of indecomposable pure-injective objects, we have and a bijection with for all .
Corollary 7.8.
Let be a pure-injective object of such that is a coproduct of indecomposable pure-injectives for any set . Then is -pure-injective.
Proof.
Let and for each , so that . By hypothesis we have where each is an indecomposable pure-injective object of . By Lemma 7.7 there exists such that, for each collection where is an indecomposable summand of , we have for all with and all but finitely many . Fixing , for the collection given by for all , we have for all , and for all by Lemma 4.5. ∎
Theorem 7.9.
[9, Theorem 1] (see also [7, Lemma 1.1]). Let be an injective object in a Grothendieck category which has a set of finitely generated generators. Then the following statements are equivalent.
-
(1)
is -injective.
-
(2)
The countable coproduct is injective.
-
(3)
For each , any ascending chain of -annihilator subobjects of eventually stabilises.
Finally, we may now complete the proof of our main result.
of Theorem 1.1.
Taking shows (1) implies (2). That (2) implies (3) follows from Remark 6.8 and Theorem 7.9. That (3) implies (1) follows from Remark 5.1, Proposition 5.6 and Lemma 5.7. The equivalence of (1) and (4) follows from Proposition 6.5. That (1) implies (5) follows from Lemma 6.6, and the converse follows from Corollaries 5.8 and 6.9. The equivalence of (1) and (6) follows from Lemma 6.12 and Corollary 7.8. ∎
8. Applications to endoperfection.
To consider direct applications of Theorem 1.1, recall Lemma 4.3, which says that pp-definable subgroups of an object of a fixed sort define -submodules of . We begin by introducing the definition of an endoperfect object; see Definition 8.1. This concept is motivated by the idea of a perfect module as in Björk [3], and generalises the finite endolength objects defined by Krause [16], recalled in Definition 8.6.
Definition 8.1.
An object of is said to be endoperfect if for all compact objects of the left -module has the descending chain condition on cyclic submodules. That is, is endoperfect provided is a perfect -module in the sense of [3, §1] for each compact object .
As the terminology suggests, any object with a perfect endomorphism is endoperfect by the famous characterisation of perfect rings due to Bass [1, Theorem P]. We now relate this to purity by adapting the notion of endonoerthian modules to our categorical setting. This was motivated by work of Huisgen-Zimmermann and Saorín [13], in which conditions for endomorphism rings are related to -pure-injective modules.
Definition 8.2.
An object of is called endonoetherian (respectively, endoartinian) if is noetherian (respectively, artinian) over for each compact object .
Corollary 8.3.
Any endoperfect endonoetherian object of is -pure-injective, and hence decomposes into a coproduct of indecomposable objects with local endomorphism rings.
Proof.
Let be a compact object of and be a descending chain of pp-definable subgroups of of sort . By Lemma 4.3 this defines a descending chain of -submodules of . Since is endonoetherian each submodule (of the form ) is finitely generated. Since is endoperfect the module has the descending chain condition on cyclic submodules. This means is perfect (over ) in the sense of Björk [3, §1]. By [3, Theorem 2] this means that the descending chain of finitely generated submodules of must terminate.
Lemma 8.4.
If is a pure-injective object of and is a morphism in with compact, then the cyclic -submodule of is an intersection of pp-definable subgroups of of sort .
Proof.
We follow [4, §1.7]. Suppose is a finitely generated subobject of , so that we have an exact sequence in of the form with compact. By construction the quotient is finitely presented, so there is a morphism such that is the composition . Similarly it is straightforward to check that any morphism is given by a morphism in such that .
By [8, Proposition 3.1] this shows that the set of morphisms is the set of where lies in a pp-definable subgroup of of sort . We now follow [4, §4, Lemma]. Since is a locally finitely generated Grothendieck category we have that is the sum over the finitely generated subobjects of . Hence the set of morphisms is the intersection over each of the set of morphisms . Altogether this shows the set of in is given by the set of where lies in an intersection of pp-definable subgroup of of sort .
The pure-injectivity of in is equivalent, by Corollary 6.7, to the injectivity of its image . Thus, considering the canonical embedding given by , any morphism factors through an endomorphism of . Altogether this shows that the cyclic module is the intersection where each pp-formula is given by in the above notation. ∎
Hence we may now provide partial a converse to Corollary 8.3.
Corollary 8.5.
Any -pure-injective object of is endoperfect.
Proof.
By the equivalence of (1) and (5) in Theorem 1.1 we have that for each compact object any descending chain of pp-definable subgroups of of sort must stabilise. By Lemma 8.4 any cyclic -submodule of is an interesction, and hence a finite intersection, of such subgroups. Thus any cyclic -submodule of is a pp-definable subgroup of of sort . So, by definition, and by the equivalence of (1) and (5) in Theorem 1.1, any descending chain of cyclic -submodules stabilises. ∎
The author would be interested in finding, if it exists, an endoperfect object in a compactly generated triangulated category which is not -pure-injective. Of course, by Corollary 8.3 such an example would not be endonoetherian.
Definition 8.6.
[16, Definition 1.1] An object of is called endofinite if has finite length over for all compact objects .
To conclude, we recover a result of Krause. Note endofinite objects are endorartinian, endonoetherian, hence endoperfect and -pure-injective by Corollary 8.3.
Theorem 8.7.
[16, Theorem 1.2(1)] Every endofinite object of decomposes into a coproduct of indecomposable objects with local endomorphism rings.
References
- [1] Hyman Bass, Finitistic dimension and a homological generalization of semi-primary rings, Trans. Amer. Math. Soc. 95 (1960), 466–488. MR 0157984
- [2] Walter Baur, Elimination of quantifiers for modules, Israel J. Math. 25 (1976), no. 1-2, 64–70. MR 0457194
- [3] Jan-Erik Björk, Rings satisfying a minimum condition on principal ideals., Journal für die reine und angewandte Mathematik 1969 (1969), no. 236, 112–119.
- [4] William Crawley-Boevey, Modules of finite length over their endomorphism rings, Representations of algebras and related topics (1992), 127–184.
- [5] Pilar Dellunde, Àngel García-Cerdaña, and Carles Noguera, Löwenheim-Skolem theorems for non-classical first-order algebraizable logics, Log. J. IGPL 24 (2016), no. 3, 321–345. MR 3509921
- [6] Carl Faith, Rings with ascending condition on annihilators, Nagoya Math. J. 27 (1966), 179–191. MR 193107
- [7] Jose Luis Garcia and Nguyen Viet Dung, Some decomposition properties of injective and pure-injective modules, Osaka J. Math. 31 (1994), no. 1, 95–108. MR 1262791
- [8] Grigory Garkusha and Mike Prest, Triangulated categories and the Ziegler spectrum, Algebr. Represent. Theory 8 (2005), no. 4, 499–523. MR 2199207
- [9] Manabu Harada, Perfect categories. IV. Quasi-Frobenius categories, Osaka Math. J. 10 (1973), 585–596. MR 367020
- [10] Wilfrid Hodges, A shorter model theory, Cambridge University Press, Cambridge, 1997. MR 1462612
- [11] Birge Huisgen-Zimmermann, Rings whose right modules are direct sums of indecomposable modules, Proc. Amer. Math. Soc. 77 (1979), no. 2, 191–197. MR 542083
- [12] by same author, Purity, algebraic compactness, direct sum decompositions, and representation type, Infinite length modules (Bielefeld, 1998), Trends Math., Birkhäuser, Basel, 2000, pp. 331–367. MR 1789225
- [13] Birge Huisgen-Zimmermann and Manuel Saorín, Direct sums of representations as modules over their endomorphism rings, Journal of Algebra 250 (2002), no. 1, 67–89.
- [14] Christian Jensen and Helmut Lenzing, Model-theoretic algebra with particular emphasis on fields, rings, modules, Algebra, Logic and Applications, vol. 2, Gordon and Breach Science Publishers, New York, 1989. MR 1057608
- [15] Henning Krause, The spectrum of a locally coherent category, J. Pure Appl. Algebra 114 (1997), no. 3, 259–271. MR 1426488
- [16] by same author, Decomposing thick subcategories of the stable module category, Math. Ann 313 (1999), 95–108.
- [17] by same author, Smashing subcategories and the telescope conjecture—an algebraic approach, Invent. Math. 139 (2000), no. 1, 99–133. MR 1728877
- [18] by same author, Coherent functors in stable homotopy theory, Fund. Math. 173 (2002), no. 1, 33–56. MR 1899046
- [19] Amnon Neeman, The connection between the -theory localization theorem of Thomason, Trobaugh and Yao and the smashing subcategories of Bousfield and Ravenel, Ann. Sci. École Norm. Sup. (4) 25 (1992), no. 5, 547–566. MR 1191736
- [20] Mike Prest, Model theory and modules, London Mathematical Society Lecture Note Series, vol. 130, Cambridge University Press, Cambridge, 1988. MR 933092
- [21] Martin Ziegler, Model theory of modules, Ann. Pure Appl. Logic 26 (1984), no. 2, 149–213. MR 739577