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

Characterisations of Σ\Sigma-pure-injectivity in triangulated categories and applications to endoperfect objects.

Raphael Bennett-Tennenhaus R. Bennett-Tennenhaus
Faculty of Mathematics
Bielefeld University
Universität sstraße 25
33615 Bielefeld
Germany
[email protected]
Abstract.

We provide various ways to characterise Σ\Sigma-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 Σ\Sigma-pure-injectives in triangulated categories are deduced from results about Σ\Sigma-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 Σ\Sigma-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 Σ\Sigma-injective objects in Grothendieck categories by building on work of Harada [9], which generalised a famous characterisation of Σ\Sigma-injective modules going back to Faith [6]. These authors showed that, as above, such developments helped simplify arguments about Σ\Sigma-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.

  • 𝒯\mathcal{T} is a compactly generated triangulated category with all small coproducts.

  • 𝒯c\mathcal{T}^{c} is the full subcategory of 𝒯\mathcal{T} consisting of compact objects, which is assumed to be skeletally small.

  • 𝐀𝐛\mathbf{Ab} is the category of abelian groups.

  • 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is the category of contravariant additive functors 𝒯c𝐀𝐛\mathcal{T}^{c}\to\mathbf{Ab}.

  • 𝙶\mathtt{G} is a set of generators of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} where each 𝒢𝙶\mathscr{G}\in\mathtt{G} is finitely presented.

  • 𝐘:𝒯𝐌𝐨𝐝-𝒯c\mathbf{Y}\colon\mathcal{T}\to\mathbf{Mod}\text{-}\mathcal{T}^{c} is the restricted Yoenda functor, which takes an object MM to the restriction of the corepresentable 𝒯(,M)\mathcal{T}(-,M).

Recall that, by the Brown representability theorem, since the category 𝒯\mathcal{T} has all small coproducts, it has all small products; see Remark 4.4.

Theorem 1.1.

For any object MM of 𝒯\mathcal{T} the following statements are equivalent.

  1. (1)

    For any set 𝙸\mathtt{I} the coproduct M(𝙸)=𝙸MM^{(\mathtt{I})}=\coprod_{\mathtt{I}}M is pure-injective.

  2. (2)

    The countable coproduct M()M^{(\mathbb{N})} is pure-injective.

  3. (3)

    For any generator 𝒢𝙶\mathscr{G}\in\mathtt{G} each ascending chain of 𝐘(M)\mathbf{Y}(M)-annihilator subobjects of 𝒢\mathscr{G} must eventually stabilise.

  4. (4)

    For any set 𝙸\mathtt{I} the morphism from M(𝙸)M^{(\mathtt{I})} to the product M𝙸=𝙸MM^{\mathtt{I}}=\prod_{\mathtt{I}}M, given by the universal property, is a section.

  5. (5)

    For any object XX of 𝒯c\mathcal{T}^{c} each descending chain of pp-definable subgroups of MM of sort XX must eventually stabilise.

  6. (6)

    MM is pure injective, and for any set 𝙸\mathtt{I}, M𝙸M^{\mathtt{I}} 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 𝚂\mathtt{S}, an 𝚂\mathtt{S}-sorted predicate language 𝔏\mathfrak{L} is a tuple pred𝚂,func𝚂,ar𝚂,sort𝚂\langle\mathrm{pred}_{\mathtt{S}},\,\mathrm{func}_{\mathtt{S}},\,\mathrm{ar}_{\mathtt{S}},\,\mathrm{sort}_{\mathtt{S}}\rangle where:

  1. (1)

    each s𝚂s\in\mathtt{S} is called a sort;

  2. (2)

    the symbol pred𝚂\mathrm{pred}_{\mathtt{S}} denotes a non-empty set of sorted predicate symbols;

  3. (3)

    the symbol func𝚂\mathrm{func}_{\mathtt{S}} denotes a set of sorted function symbols;

  4. (4)

    the symbol ar𝚂\mathrm{ar}_{\mathtt{S}} denotes an arity function pred𝚂func𝚂={0,1,2,}\mathrm{pred}_{\mathtt{S}}\sqcup\mathrm{func}_{\mathtt{S}}\to\mathbb{N}=\{0,1,2,\dots\};

  5. (5)

    the symbol sort𝚂\mathrm{sort}_{\mathtt{S}} denotes a sort function, taking any nn-ary Rpred𝚂R\in\mathrm{pred}_{\mathtt{S}} (respectively Ffunc𝚂F\in\mathrm{func}_{\mathtt{S}}) to a sequence in 𝚂\mathtt{S} of length nn (respectively n+1n+1).

When n>0n>0 in condition (5) we often write sort𝚂(R)=(s1,,sn)\mathrm{sort}_{\mathtt{S}}(R)=(s_{1},\dots,s_{n}) (respectively sort𝚂(F)=(s1,,sn,s)\mathrm{sort}_{\mathtt{S}}(F)=(s_{1},\dots,s_{n},s)). Note that functions FF with ar𝚂(F)=0\mathrm{ar}_{\mathtt{S}}(F)=0 have a sort.

For each sort ss we introduce a countable set 𝒱s\mathcal{V}_{s} of variables of sort ss. The terms of 𝔏\mathfrak{L} each have their own sort, and are defined inductively by stipulating: any variable xx of sort ss will be considered a term of sort ss; and for any Ffunc𝚂F\in\mathrm{func}_{\mathtt{S}} with sort𝚂(F)=(s1,,sn,s)\mathrm{sort}_{\mathtt{S}}(F)=(s_{1},\dots,s_{n},s) and for any terms t1,,tnt_{1},\dots,t_{n} of sort s1,,sns_{1},\dots,s_{n} respectively, F(t1,,tn)F(t_{1},\dots,t_{n}) is considered a term of sort ss. Note that constant symbols, given by functions FF with ar𝚂(F)=0\mathrm{ar}_{\mathtt{S}}(F)=0, are also terms.

The atomic formulas with which 𝔏\mathfrak{L} is equipped are built from the equality t=stt=_{s}t^{\prime} between terms t,tt,t^{\prime} of common sort ss, together with the formulas R(t1,,tn)R(t_{1},\dots,t_{n}) where Rpred𝚂R\in\mathrm{pred}_{\mathtt{S}}, sort𝚂(R)=(s1,,sn)\mathrm{sort}_{\mathtt{S}}(R)=(s_{1},\dots,s_{n}) and where each tit_{i} is a term of sort sis_{i}. First-order formulas φ\varphi in 𝔏\mathfrak{L} are built from: the variables of each sort; the atomic formulas; binary connectives \wedge, \vee, and \implies; negation ¬\neg; and the quantifiers \forall and \exists.

A positive-primitive or pp formula φ(x1,,xn)\varphi(x_{1},\dots,x_{n}) (with xix_{i} free) has the form

wn+1,,wm:j=1kψj(x1,,xn,wn+1,,wm),\begin{array}[]{c}\exists\,w_{n+1},\dots,w_{m}\colon\bigwedge_{j=1}^{k}\psi_{j}(x_{1},\dots,x_{n},w_{n+1},\dots,w_{m}),\end{array}

where each ψj\psi_{j} is an atomic formula (see, for example, [10, p.50]).

One may build a theory for a multi-sorted language 𝔏\mathfrak{L} 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 AA be a unital ring. We recall how the language 𝔏A\mathfrak{L}_{A} of AA-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 sortA\mathrm{sort}_{A}. Let predA={0}\mathrm{pred}_{A}=\{0\}. Let funcA={+}{a×aA}\mathrm{func}_{A}=\{+\}\cup\{a\times-\mid a\in A\} where ++ is binary and a×a\times- is unary.

In Definition 2.3 the notion of a structure is recalled. For the language 𝔏A\mathfrak{L}_{A} we have that this notion, together with certain axioms, recovers the defining properties of left AA-modules. Later we consider homomorphisms.

Definition 2.3.

[5, Definition 35] Fix a set 𝚂\mathtt{S}\neq\emptyset and an 𝚂\mathtt{S}-sorted predicate language 𝔏\mathfrak{L}. An 𝔏\mathfrak{L}-structure is a tuple

𝖬=𝚂(𝖬),(R(𝖬)Rpred𝚂),(F(𝖬)Ffunc𝚂)\mathsf{M}=\langle\mathtt{S}(\mathsf{M}),(R(\mathsf{M})\mid R\in\mathrm{pred}_{\mathtt{S}}),(F(\mathsf{M})\mid F\in\mathrm{func}_{\mathtt{S}})\rangle

such that:

  1. (1)

    the symbol 𝚂(𝖬)\mathtt{S}(\mathsf{M}) denotes a family of sets {s(𝖬)s𝚂}\{s(\mathsf{M})\mid s\in\mathtt{S}\};

  2. (2)

    if sort𝚂(R)=(s1,,sn)\mathrm{sort}_{\mathtt{S}}(R)=(s_{1},\dots,s_{n}) then R(𝖬)R(\mathsf{M}) is a subset of s1(𝖬)××sn(𝖬)s_{1}(\mathsf{M})\times\dots\times s_{n}(\mathsf{M}); and

  3. (3)

    if sort𝚂(F)=(s1,,sn,s)\mathrm{sort}_{\mathtt{S}}(F)=(s_{1},\dots,s_{n},s) then F(𝖬)F(\mathsf{M}) is a map s1(𝖬)××sn(𝖬)s(𝖬)s_{1}(\mathsf{M})\times\dots\times s_{n}(\mathsf{M})\to s(\mathsf{M}).

Denoting the cardinality of any set 𝚇\mathtt{X} by |𝚇||\mathtt{X}|, let |𝔏|=|pred𝚂func𝚂||\mathfrak{L}|=|\mathrm{pred}_{\mathtt{S}}\sqcup\mathrm{func}_{\mathtt{S}}| and, for 𝖬\mathsf{M} as above, let |𝖬||\mathsf{M}| be the sum of the cardinalities |s(𝖬)||s(\mathsf{M})| as ss runs through 𝚂\mathtt{S}.

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 {𝚛,𝚖}\{\mathtt{r},\mathtt{m}\}-sorted predicate language which is in contrast to Example 2.2. The predicates in this language will be the unary symbols 0𝚛0_{\mathtt{r}} and 1𝚛1_{\mathtt{r}} of sort 𝚛\mathtt{r}, and 0𝚖0_{\mathtt{m}} of sort 𝚖\mathtt{m}. The functions in this language will be the ternary symbols ++ and ×\times where sort𝚛,𝚖(+)=(𝚖,𝚖,𝚖)\mathrm{sort}_{\mathtt{r},\mathtt{m}}(+)=(\mathtt{m},\mathtt{m},\mathtt{m}) and sort𝚛,𝚖(×)=(𝚛,𝚖,𝚖)\mathrm{sort}_{\mathtt{r},\mathtt{m}}(\times)=(\mathtt{r},\mathtt{m},\mathtt{m}). After specifying the appropriate axioms, structures 𝖬𝖠{}_{\mathsf{A}}\mathsf{M} are tuples (A,M)(A,M) where AA is a unital ring and MM is a left AA-module.

In this way one interprets the symbols 0𝚛0_{\mathtt{r}} and 1𝚛1_{\mathtt{r}} as the additive and multiplicative identities in AA. Similarly the symbol 0𝚖0_{\mathtt{m}} is interpreted as the additive identity in MM. In Definition 2.5 the notion of a homomorphism between structures is recalled. In this sense, a homomorphism (A,M)(B,N)(A,M)\to(B,N) is given by a pair (f,l)(f,l) where f:ABf\colon A\to B is a homomorphism of rings and l:MNl\colon M\to N is a homomorphism of left AA-modules with the action of AA on NN given by ff.

Definition 2.5.

[5, Definition 3] Fix a non-empty set 𝚂\mathtt{S}, an 𝚂\mathtt{S}-sorted predicate language 𝔏\mathfrak{L} and 𝔏\mathfrak{L}-structures 𝖫\mathsf{L} and 𝖬\mathsf{M}. By an 𝔏\mathfrak{L}-homomorphism 𝗁:𝖫𝖬\mathsf{h}\colon\mathsf{L}\to\mathsf{M} we mean a family {𝗁ss𝚂}\{\mathsf{h}_{s}\mid s\in\mathtt{S}\} of functions 𝗁s:s(𝖫)s(𝖬)\mathsf{h}_{s}\colon s(\mathsf{L})\to s(\mathsf{M}) such that:

  1. (1)

    if sort𝚂(R)=(s1,,sn)\mathrm{sort}_{\mathtt{S}}(R)=(s_{1},\dots,s_{n}) then R(𝖬)R(\mathsf{M}) is the set of (𝗁s1(a1),,𝗁sn(an))(\mathsf{h}_{s_{1}}(a_{1}),\dots,\mathsf{h}_{s_{n}}(a_{n})) such that (a1,,an)R(𝖫)(a_{1},\dots,a_{n})\in R(\mathsf{L});

  2. (2)

    and if sort𝚂(F)=(s1,,sn,s)\mathrm{sort}_{\mathtt{S}}(F)=(s_{1},\dots,s_{n},s) then for all (a1,,an)s1(𝖫)××sn(𝖫)(a_{1},\dots,a_{n})\in s_{1}(\mathsf{L})\times\dots\times s_{n}(\mathsf{L}) we have 𝗁s(F(𝖫)(a1,,an))=F(𝖬)(𝗁s1(a1),,𝗁sn(an))\mathsf{h}_{s}(F(\mathsf{L})(a_{1},\dots,a_{n}))=F(\mathsf{M})(\mathsf{h}_{s_{1}}(a_{1}),\dots,\mathsf{h}_{s_{n}}(a_{n})).

Note that, in the notation of Definition 2.5, [5, Theorem 17] says that a collection of functions 𝗁s:s(𝖫)s(𝖬)\mathsf{h}_{s}:s(\mathsf{L})\to s(\mathsf{M}) defines an 𝔏\mathfrak{L}-homomorphism if and only if, whenever φ(x1,,xn)\varphi(x_{1},\dots,x_{n}) is an atomic formula with sort𝚂(xi)=si\mathrm{sort}_{\mathtt{S}}(x_{i})=s_{i} and (a1,,an)(a_{1},\dots,a_{n}) lies in s1(𝖫)××sn(𝖫)s_{1}(\mathsf{L})\times\dots\times s_{n}(\mathsf{L}), we have that

𝖫φ(a1,,an)𝖬φ(𝗁s1(a1),,𝗁sn(an)).\mathsf{L}\models\varphi(a_{1},\dots,a_{n})\implies\mathsf{M}\models\varphi(\mathsf{h}_{s_{1}}(a_{1}),\dots,\mathsf{h}_{s_{n}}(a_{n})).

We are now ready to recall the idea of purity coming from model theory.

Definition 2.6.

[5, Definition 36] Fix a set 𝚂\mathtt{S}\neq\emptyset and an 𝚂\mathtt{S}-sorted predicate language 𝔏\mathfrak{L}. By an 𝔏\mathfrak{L}-embedding we mean an 𝔏\mathfrak{L}-homomorphism 𝗁:𝖫𝖬\mathsf{h}\colon\mathsf{L}\to\mathsf{M} such that:

  1. (1)

    if φ(x1,,xn)\varphi(x_{1},\dots,x_{n}) is an atomic formula with sort𝚂(xi)=si\mathrm{sort}_{\mathtt{S}}(x_{i})=s_{i}, then for all (a1,,an)s1(𝖫)××sn(𝖫)(a_{1},\dots,a_{n})\in s_{1}(\mathsf{L})\times\dots\times s_{n}(\mathsf{L}), 𝖫φ(a1,,an)\mathsf{L}\models\varphi(a_{1},\dots,a_{n}) if and only if 𝖬φ(𝗁s1(a1),,𝗁sn(an))\mathsf{M}\models\varphi(\mathsf{h}_{s_{1}}(a_{1}),\dots,\mathsf{h}_{s_{n}}(a_{n})).

By an 𝔏\mathfrak{L}-pure embedding we mean an 𝔏\mathfrak{L}-homomorphism 𝗁:𝖫𝖬\mathsf{h}\colon\mathsf{L}\to\mathsf{M} such that:

  1. (2)

    if φ(x1,,xn)\varphi(x_{1},\dots,x_{n}) is a pp formula with sort𝚂(xi)=si\mathrm{sort}_{\mathtt{S}}(x_{i})=s_{i}, then for all (a1,,an)s1(𝖫)××sn(𝖫)(a_{1},\dots,a_{n})\in s_{1}(\mathsf{L})\times\dots\times s_{n}(\mathsf{L}), if 𝖬φ(𝗁s1(a1),,𝗁sn(an))\mathsf{M}\models\varphi(\mathsf{h}_{s_{1}}(a_{1}),\dots,\mathsf{h}_{s_{n}}(a_{n})) then 𝖫φ(a1,,an)\mathsf{L}\models\varphi(a_{1},\dots,a_{n}).

Note that 𝔏\mathfrak{L}-pure embeddings are 𝔏\mathfrak{L}-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 AA and in the notation from Example 2.2, an injective left AA-module homomorphism is pure if and only if it is an 𝔏A\mathfrak{L}_{A}-pure embedding.

Definition 2.7.

Fix a non-empty set 𝚂\mathtt{S}, an 𝚂\mathtt{S}-sorted predicate language 𝔏\mathfrak{L} and 𝔏\mathfrak{L}-structures 𝖫\mathsf{L} and 𝖬\mathsf{M}. We say 𝖫\mathsf{L} is an 𝔏\mathfrak{L}-substructure of 𝖬\mathsf{M} if s(𝖫)s(𝖬)s(\mathsf{L})\subseteq s(\mathsf{M}) for each s𝚂s\in\mathtt{S} and, labelling these inclusions 𝗂s\mathsf{i}_{s}, the family {𝗂ss𝚂}\{\mathsf{i}_{s}\mid s\in\mathtt{S}\} defines an 𝔏\mathfrak{L}-homomorphism 𝗂:𝖫𝖬\mathsf{i}\colon\mathsf{L}\to\mathsf{M}. If, additionally, 𝗂:𝖫𝖬\mathsf{i}\colon\mathsf{L}\to\mathsf{M} is an 𝔏\mathfrak{L}-pure embedding, we say 𝖫\mathsf{L} is an 𝔏\mathfrak{L}-pure substructure of 𝖬\mathsf{M}.

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 𝒜\mathcal{A} be an additive category. Denote the hom-sets 𝒜(X,Y)\mathcal{A}(X,Y) and the identity maps 1X1_{X}. For any set 𝙸\mathtt{I} and any collection B={Bii𝙸}\mathrm{B}=\{B_{i}\mid i\in\mathtt{I}\} of objects in 𝒜\mathcal{A}, if the categorical product iBi\prod_{i}B_{i} exists in 𝒜\mathcal{A}, we write pj,B:iBiBjp_{j,\mathrm{B}}\colon\prod_{i}B_{i}\to B_{j} for the natural morphisms equipping it, in which case the universal property gives unique morphisms vj,B:BjiBiv_{j,\mathrm{B}}\colon B_{j}\to\prod_{i}B_{i} such that pj,Bvj,Bp_{j,\mathrm{B}}v_{j,\mathrm{B}} is the identity 1j1_{j} on BjB_{j} for each jj. Similarly uj,B:BjiBiu_{j,\mathrm{B}}\colon B_{j}\to\coprod_{i}B_{i} will denote the morphisms equipping the coproduct iBi\coprod_{i}B_{i} if it exists, in which case there exist unique morphisms qj,B:iBiBjq_{j,\mathrm{B}}\colon\coprod_{i}B_{i}\to B_{j} such that qj,Buj,B=1jq_{j,\mathrm{B}}u_{j,\mathrm{B}}=1_{j} for each jj.

Fix an object AA in 𝒜\mathcal{A} and consider the covariant functor 𝒜(A,)\mathcal{A}(A,-). Note that both the product and coproduct of the collection 𝒜(A,B)={𝒜(A,Bi)i𝙸}\mathcal{A}(A,\mathrm{B})=\{\mathcal{A}(A,B_{i})\mid i\in\mathtt{I}\} exist in the category 𝐀𝐛\mathbf{Ab} of abelian groups. We identify i𝙸𝒜(A,Bi)\coprod_{i\in\mathtt{I}}\mathcal{A}(A,B_{i}) with the subgroup of i𝙸𝒜(A,Bi)\prod_{i\in\mathtt{I}}\mathcal{A}(A,B_{i}) consisting of tuples (gii𝙸)(g_{i}\mid i\in\mathtt{I}) such that gi=0g_{i}=0 for all but finitely many i𝙸i\in\mathtt{I}.

Consequently, if iBi\prod_{i}B_{i} exists in 𝒜\mathcal{A} then map λA,B:𝒜(A,iBi)i𝒜(A,Bi)\lambda_{A,\mathrm{B}}\colon\mathcal{A}(A,\prod_{i}B_{i})\to\prod_{i}\mathcal{A}(A,B_{i}) from the universal property is given by f(pi,Bfi𝙸)f\mapsto(p_{i,\mathrm{B}}f\mid i\in\mathtt{I}) for each f𝒜(A,iBi)f\in\mathcal{A}(A,\prod_{i}B_{i}). Similarly if iBi\coprod_{i}B_{i} exists in 𝒜\mathcal{A} then map γA,B:i𝒜(A,Bi)𝒜(A,iBi)\gamma_{A,\mathrm{B}}\colon\coprod_{i}\mathcal{A}(A,B_{i})\to\mathcal{A}(A,\coprod_{i}B_{i}) from the universal property is given by γA,B(gii𝙸)=iui,Bgi\gamma_{A,\mathrm{B}}(g_{i}\mid i\in\mathtt{I})=\sum_{i}u_{i,\mathrm{B}}g_{i}. In general each of the morphisms λA,B\lambda_{A,\mathrm{B}} are isomorphisms.

Definition 3.2.

[19, Definition 1.1] Let 𝒯\mathcal{T} be a triangulated category with suspension functor Σ\Sigma, and assume all small coproducts in 𝒯\mathcal{T} exist. An object XX of 𝒯\mathcal{T} is said to be compact if, for any set 𝙸\mathtt{I} and collection M={Mii𝙸}\mathrm{M}=\{M_{i}\mid i\in\mathtt{I}\} of objects in 𝒯\mathcal{T} the morphism γX,M\gamma_{X,\mathrm{M}} is an isomorphism. Let 𝒯c\mathcal{T}^{c} be the full triangulated subcategory of 𝒯\mathcal{T} consisting of compact objects.

Given a set 𝒢\mathcal{G} of compact objects in 𝒯\mathcal{T} we say that 𝒯\mathcal{T} is compactly generated by 𝒢\mathcal{G} if there are no non-zero objects MM in 𝒯\mathcal{T} satisfying 𝒯(X,M)=0\mathcal{T}(X,M)=0 for all X𝒢X\in\mathcal{G} (or, said another way, any non-zero object MM gives rise to a non-zero morphism XMX\to M for some X𝒢X\in\mathcal{G}). If 𝒯\mathcal{T} is compactly generated by 𝒢\mathcal{G} we call 𝒢\mathcal{G} a generating set provided ΣX𝒢\Sigma X\in\mathcal{G} for all X𝒢X\in\mathcal{G}.

Assumption 3.3.

In the remainder of §3 fix a triangulated category 𝒯\mathcal{T} with suspension functor Σ\Sigma, and we assume:

  • that 𝒯\mathcal{T} has all small coproducts;

  • that 𝒯\mathcal{T} is compactly generated by a generating set 𝒢\mathcal{G}; and

  • that the subcategory 𝒯c\mathcal{T}^{c} 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 𝒯\mathcal{T} is introduced.

Definition 3.4.

In what follows let 𝒮\mathcal{S} denote a fixed set of objects in 𝒯c\mathcal{T}^{c} given by choosing exactly one representative of each isomorphism class. Such a set 𝒮\mathcal{S} exists because we are assuming that 𝒯c\mathcal{T}^{c} is skeletally small.

[8, §3] The canonical language 𝔏𝒯\mathfrak{L}^{\mathcal{T}} of 𝒯\mathcal{T} is given by an 𝒮\mathcal{S}-sorted predicate language pred𝒮,func𝒮,ar𝒮,sort𝒮\langle\mathrm{pred}_{\mathcal{S}},\,\mathrm{func}_{\mathcal{S}},\,\mathrm{ar}_{\mathcal{S}},\,\mathrm{sort}_{\mathcal{S}}\rangle, defined as follows. The set pred𝒮\mathrm{pred}_{\mathcal{S}} consists of a symbol 𝟢G\mathsf{0}_{G} with sort𝒮(𝟢G)=G\mathrm{sort}_{\mathcal{S}}(\mathsf{0}_{G})=G for each G𝒮G\in\mathcal{S}. The set func𝒮\mathrm{func}_{\mathcal{S}} consists of: a ternary symbol +G+_{G} with sort𝒮(+G)=(G,G,G)\mathrm{sort}_{\mathcal{S}}(+_{G})=(G,G,G) for each G𝒮G\in\mathcal{S}; and a unary operation a-\circ a with sort𝒮(a)=(H,G)\mathrm{sort}_{\mathcal{S}}(-\circ a)=(H,G) for each morphism a:GHa:G\to H with G,H𝒮G,H\in\mathcal{S}. Variables of sort G𝒮G\in\mathcal{S} will be denoted vGv_{G}.

Notation 3.5.

Suppose 𝒜\mathcal{A} is any additive category. We write 𝒜-𝐌𝐨𝐝\mathcal{A}\text{-}\mathbf{Mod} (respectively 𝐌𝐨𝐝-𝒜\mathbf{Mod}\text{-}\mathcal{A}) for the category of additive covariant (respectively contravariant) functors 𝒜𝐀𝐛\mathcal{A}\to\mathbf{Ab} where 𝐀𝐛\mathbf{Ab} is the category of abelian groups.

For any object MM of 𝒯\mathcal{T} we let 𝒯(,M)|\mathcal{T}(-,M)| denote the object of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} defined by restriction of 𝒯(,M)\mathcal{T}(-,M) to compact objects. We write

𝐘:𝐌𝐨𝐝-𝒯c𝐀𝐛\mathbf{Y}\colon\mathbf{Mod}\text{-}\mathcal{T}^{c}\to\mathbf{Ab}

to denote the restricted Yoneda functor. That is, 𝐘\mathbf{Y} takes an object MM to 𝐘(M)=𝒯(,M)|\mathbf{Y}(M)=\mathcal{T}(-,M)|, and takes a morphism h:LMh:L\to M to the natural transformation 𝐘(h):𝒯(,L)|𝒯(,M)|\mathbf{Y}(h)\colon\mathcal{T}(-,L)|\to\mathcal{T}(-,M)| given by defining, for each compact object XX, the map 𝐘(h)X:𝒯(X,L)𝒯(X,M)\mathbf{Y}(h)_{X}\colon\mathcal{T}(X,L)\to\mathcal{T}(X,M) by ghgg\mapsto hg.

Remark 3.6.

[8, §3] Consider the theory given from the set of axioms expressing the positive atomic diagram of the objects in 𝒯c\mathcal{T}^{c} including the specification that all functions are additive. In this way, the category of models for the above theory is equivalent to the category 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} where objects MM of 𝒯\mathcal{T} are regarded as structures 𝖬\mathsf{M} for this language.

That is, in the notation of Definition 2.3, we let G(𝖬)=𝒯(G,M)G(\mathsf{M})=\mathcal{T}(G,M), we interpret the predicate symbol 0G0_{G} as the identitly element of G(𝖬)G(\mathsf{M}), we interpret +G+_{G} as the additive group operation on G(𝖬)G(\mathsf{M}), and we interpret a-\circ a as the map G(𝖬)H(𝖬)G(\mathsf{M})\to H(\mathsf{M}) given by ffaf\mapsto fa.

Lemma 3.7.

Let LL and MM be objects in 𝒯\mathcal{T} with corresponding 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-structures 𝖫\mathsf{L} and 𝖬\mathsf{M}. Then the choice of 𝒮\mathcal{S} made in Definition 3.4 defines a bijection between morphisms 𝐘(L)𝐘(M)\mathbf{Y}(L)\to\mathbf{Y}(M) (that is, natural transformations) in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} and 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-homomorphisms 𝖫𝖬\mathsf{L}\to\mathsf{M}.

Proof.

Any object XX of 𝒯c\mathcal{T}^{c} lies in the same isoclass as some unique c(X)𝒮c(X)\in\mathcal{S}, in which case we choose an isomorphism ϕX:c(X)X\phi_{X}\colon c(X)\to X. In this way, any object NN defines an isomorphism ϕX,N:𝒯(X,N)𝒯(c(X),N)-\circ\phi_{X,N}\colon\mathcal{T}(X,N)\to\mathcal{T}(c(X),N) by precomposition with ϕX\phi_{X}. Recall that here the 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-structure 𝖭\mathsf{N} is defined by setting G(𝖭)=𝒯(G,N)G(\mathsf{N})=\mathcal{T}(G,N) for each sort G𝒮G\in\mathcal{S}. In case X𝒮X\in\mathcal{S} we assume, without loss of generality, that ϕX=1X\phi_{X}=1_{X}. Define the required bijection as follows. Fix an 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-homomorphism 𝗁:𝖫𝖬\mathsf{h}\colon\mathsf{L}\to\mathsf{M}. For any object XX of 𝒯c\mathcal{T}^{c} define the function (𝗁)X:𝒯(X,L)𝒯(X,M)\mathscr{H}(\mathsf{h})_{X}\colon\mathcal{T}(X,L)\to\mathcal{T}(X,M) by l(𝗁c(X)(lϕX))ϕX1l\mapsto(\mathsf{h}_{c(X)}(l\phi_{X}))\phi^{-1}_{X}. Converlsey, fixing a natural transformation :𝐘(L)𝐘(M)\mathscr{H}\colon\mathbf{Y}(L)\to\mathbf{Y}(M), let 𝗁()G=G\mathsf{h}(\mathscr{H})_{G}=\mathscr{H}_{G} for each G𝒮G\in\mathcal{S}.

It suffices to explain why these assignments swap between morphisms in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} and 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-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 0G0_{G} and the function symbols +G+_{G}) is equivalent to saying that each function X\mathscr{H}_{X} is a homomorphism of abelian groups. Letting b:XYb\colon X\to Y be a morphism in 𝒯c\mathcal{T}^{c} and a=ϕY1bϕXa=\phi_{Y}^{-1}b\phi_{X}, for any object NN of 𝒯\mathcal{T} the sorted function symbol a-\circ a is interpreted in 𝖭\mathsf{N} by the equation (a)(𝖭)=(ϕN,Y1)bϕN,X(-\circ a)(\mathsf{N})=-\circ(\phi_{N,Y}^{-1})b\phi_{N,X}. Thus, by construction, saying that the function symbols a-\circ a are preserved is equivalent to saying that the collection of X\mathscr{H}_{X} (for XX 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 h:LMh\colon L\to M in 𝒯\mathcal{T} is a pure monomorphism if 𝐘(h)X:𝒯(X,L)𝒯(X,M)\mathbf{Y}(h)_{X}\colon\mathcal{T}(X,L)\to\mathcal{T}(X,M) is injective for each compact object XX.

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 h:LMh\colon L\to M is a pure monomorphism if and only if the image 𝗁:𝖫𝖬\mathsf{h}\colon\mathsf{L}\to\mathsf{M} under the bijection in Lemma 3.7 is an 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-pure embedding.

Proof.

By [8, Proposition 3.1] any pp-formula φ(vG)\varphi(v_{G}) is equivalent to a divisibility formula uH:vG=uHa\exists u_{H}\colon v_{G}=u_{H}a where a:GHa\colon G\to H is morphism and G,H𝒮G,H\in\mathcal{S}. By Definition 2.6, 𝗁\mathsf{h} is an 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-pure embedding if and only if, for any morphism a:GHa\colon G\to H with G,H𝒮G,H\in\mathcal{S} and any pair (f,g)G(𝖫)×H(𝖫)(f,g)\in G(\mathsf{L})\times H(\mathsf{L}), if hg=hfahg=hfa then g=fag=fa. Since any compact object is isomorphic to an object in 𝒮\mathcal{S}, this is equivalent to the condition which says that, for each compact object XX, the morphism 𝒯(X,L)𝒯(X,M)\mathcal{T}(X,L)\to\mathcal{T}(X,M) given by ghgg\mapsto hg is injective. ∎

4. Products, coproducts, coherent functors and pp-formulas.

Recall, from Definition 2.1, that pp-formulas in 𝔏𝒯\mathfrak{L}^{\mathcal{T}} are those lying in the closure of the set of equations under conjunction and existential quantification.

Definition 4.1.

[8, §2] Given G𝒮G\in\mathcal{S} and an object MM of 𝒯\mathcal{T} with 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-structure 𝖬\mathsf{M}, a pp-definable subgroup of MM of sort GG is the set φ(M)={fG(𝖬)𝖬φ(f)}\varphi(M)=\{f\in G(\mathsf{M})\mid\mathsf{M}\models\varphi(f)\} of solutions (in 𝖬\mathsf{M}) to some pp-formula φ(vG)\varphi(v_{G}) in one free variable of sort G𝒮G\in\mathcal{S}.

For any morphism b:XYb\colon X\to Y in 𝒯c\mathcal{T}^{c} and any object MM in 𝒯\mathcal{T} recall the map 𝐘(b):𝒯(Y,M)𝒯(X,M)\mathbf{Y}(b)\colon\mathcal{T}(Y,M)\to\mathcal{T}(X,M) is defined by precomposition. In this case let

Mb=im(𝐘(b))={fb𝒯(X,M)f𝒯(Y,M)}.Mb=\mathrm{im}(\mathbf{Y}(b))=\{fb\in\mathcal{T}(X,M)\mid f\in\mathcal{T}(Y,M)\}.

If G,H𝒮G,H\in\mathcal{S} and ϕX:GX\phi_{X}\colon G\to X and ϕY:HY\phi_{Y}\colon H\to Y are isomorphisms in 𝒯\mathcal{T} (as in the proof of Lemma 3.7), then fbfϕY1bϕXfb\mapsto f\phi^{-1}_{Y}b\phi_{X} defines a isomorphism Mbφ(M)Mb\to\varphi(M) in 𝐀𝐛\mathbf{Ab} where φ(vG)\varphi(v_{G}) is the pp-formula (uH:vG=uHa)(\exists u_{H}\colon v_{G}=u_{H}a) where a=ϕY1bϕXa=\phi^{-1}_{Y}b\phi_{X}.

Lemma 4.2.

Let G𝒮G\in\mathcal{S} and let φ(vG)\varphi(v_{G}) be a pp-formula in one free variable of sort GG. If h:LMh\colon L\to M is a pure monomorphism then φ(L)={g𝒯(G,L)hgφ(M)}\varphi(L)=\{g\in\mathcal{T}(G,L)\mid hg\in\varphi(M)\}.

Proof.

The claim follows from Lemma 3.9, together with Definition 2.6, which we recall defines 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-pure embeddings as those which preserve solutions to the negations of all pp-formulas. ∎

We continue, slightly abusing terminology, by reffering to any set of the form MbMb (for some b𝒯(X,Y)b\in\mathcal{T}(X,Y)) as a pp-definable subgroup of MM of sort XX. 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 MM in 𝒯\mathcal{T} and any morphism b𝒯(X,Y)b\in\mathcal{T}(X,Y) the pp-definable subgroup MbMb is a left End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodule of 𝒯(X,M)\mathcal{T}(X,M).

Proof.

By the associativity of the composition of morphisms in 𝒯\mathcal{T} the set MbMb is closed under postcomposition with endomorphisms of MM. ∎

Recall that a covariant functor :𝒯𝐀𝐛\mathscr{F}:\mathcal{T}\to\mathbf{Ab} is coherent if there is an exact sequence in 𝒯-𝐌𝐨𝐝\mathcal{T}\text{-}\mathbf{Mod} of the form 𝒯(A,)𝒯(B,)0\mathcal{T}(A,-)\to\mathcal{T}(B,-)\to\mathscr{F}\to 0. If t:MNt\colon M\to N and a:GHa:G\to H are morphisms in 𝒯\mathcal{T} with G,H𝒮G,H\in\mathcal{S}, then tvNatv\in Na for any vMav\in Ma. So, for the pp-formula φ(vG)=(uH:vG=uHa)\varphi(v_{G})=(\exists u_{H}\colon v_{G}=u_{H}a) in 𝔏𝒯\mathfrak{L}^{\mathcal{T}} the assignment of objects Mφ(M)M\mapsto\varphi(M) 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.

As a result of Assumption 3.3, by the Brown representability theorem we have that 𝒯\mathcal{T} has all small products. See [17, Lemma 1.5] for details.

Lemma 4.5 is analogous to [14, Proposition 6.7(i,ii)]. Recall Notation 3.1.

Lemma 4.5.

Let G𝒮G\in\mathcal{S} and let φ(vG)\varphi(v_{G}) be a pp-formula in one free variable of sort GG. For any set II and any collection M={Mii𝙸}\mathrm{M}=\{M_{i}\mid i\in\mathtt{I}\} of objects in 𝒯\mathcal{T} the restrictions of γG,M\gamma_{G,\mathrm{M}} and λG,M\lambda_{G,\mathrm{M}} define isomorphisms of abelian groups

iφ(Mi)φ(iMi),φ(iMi)iφ(Mi).\begin{array}[]{cc}\coprod_{i}\varphi(M_{i})\to\varphi(\coprod_{i}M_{i}),&\varphi(\prod_{i}M_{i})\to\prod_{i}\varphi(M_{i}).\end{array}
Proof.

By the existence of small products and coproducts in 𝒯\mathcal{T} and the functorality of φ\varphi, the universal properties give morphisms δ:iφ(Mi)φ(iMi)\delta\colon\coprod_{i}\varphi(M_{i})\to\varphi(\coprod_{i}M_{i}) and μ:φ(iMi)iφ(Mi)\mu\colon\varphi(\prod_{i}M_{i})\to\prod_{i}\varphi(M_{i}). By [8, Lemma 4.3] the functor φ\varphi is coherent, so by the equivalence of statements (1) and (3) from [18, Theorem A] the morphisms δ\delta and μ\mu are isomorphisms. It is straightforward to check that δ\delta and μ\mu are the respectively restrictions of γG,M\gamma_{G,\mathrm{M}} and λG,M\lambda_{G,\mathrm{M}} from Notation 3.1. ∎

We now adapt some technical results from work of Huisgen-Zimmerman [11], in which a (now well-known) charaterisation of Σ\Sigma-pure-injective modules was given. Our adaptations, namely Lemmas 4.8 and 7.7, are used in the sequel.

Notation 4.6.

Fix collections M={Mii}\mathrm{M}=\{M_{i}\mid i\in\mathbb{N}\} and L={Ljj𝙹}\mathrm{L}=\{L_{j}\mid j\in\mathtt{J}\} of objects in 𝒯\mathcal{T} and let M=LM=L where M=iMiM=\prod_{i}M_{i} and L=jLjL=\coprod_{j}L_{j} exist. By Notation 3.1, pi,M:MMip_{i,\mathrm{M}}\colon M\to M_{i} and uj,L:LjLu_{j,\mathrm{L}}\colon L_{j}\to L denote the morphisms equipping the product and coproduct, and vi,M:MiMv_{i,\mathrm{M}}\colon M_{i}\to M and qj,L:LLjq_{j,\mathrm{L}}\colon L\to L_{j} denote the morphisms given by universal properties such that pi,Mvi,M=1Mip_{i,\mathrm{M}}v_{i,\mathrm{M}}=1_{M_{i}} and qj,Luj,L=1Ljq_{j,\mathrm{L}}u_{j,\mathrm{L}}=1_{L_{j}} for each ii and jj.

Corollary 4.7.

Consider Notation 4.6, let a:GHa\colon G\to H be a morphism with G,H𝒮G,H\in\mathcal{S}, and let φ(vG)=(uH:vG=uHa)\varphi(v_{G})=(\exists u_{H}\colon v_{G}=u_{H}a). Then there is an isomorphism

κφ:iφ(Mi)j𝙹φ(Lj),(fii)(qj,Lfj𝙹)\begin{array}[]{c}\kappa\langle\varphi\rangle\colon\prod_{i\in\mathbb{N}}\varphi(M_{i})\to\coprod_{j\in\mathtt{J}}\varphi(L_{j}),\,(f_{i}\mid i\in\mathbb{N})\mapsto(q_{j,\mathrm{L}}f\mid j\in\mathtt{J})\end{array}

where f:GiMif\colon G\to\prod_{i}M_{i} is given by the universal property, and whose inverse is

κ1φ:j𝙹φ(Lj)iφ(Mi),(gjj𝙹)(j𝙹pi,Muj,Lgji).\begin{array}[]{c}\kappa^{-1}\langle\varphi\rangle\colon\coprod_{j\in\mathtt{J}}\varphi(L_{j})\to\prod_{i\in\mathbb{N}}\varphi(M_{i}),\,(g_{j}\mid j\in\mathtt{J})\mapsto(\sum_{j\in\mathtt{J}}p_{i,\mathrm{M}}u_{j,\mathrm{L}}g_{j}\mid i\in\mathbb{N}).\end{array}
Proof.

By Lemma 4.5 the restriction of γG,L\gamma_{G,\mathrm{L}} and λG,M\lambda_{G,\mathrm{M}} define isomorphisms

δ:jφ(Lj)φ(jLj),μ:φ(iMi)iφ(Mi).\begin{array}[]{cc}\delta\colon\coprod_{j}\varphi(L_{j})\to\varphi(\coprod_{j}L_{j}),\mu\colon\varphi(\prod_{i}M_{i})\to\prod_{i}\varphi(M_{i}).\end{array}

Letting κφ=δ1μ1\kappa\langle\varphi\rangle=\delta^{-1}\mu^{-1} and κφ1=μδ\kappa\langle\varphi\rangle^{-1}=\mu\delta, the proof is straightforward. ∎

The proof of Lemma 4.8 follows the proof of the cited result of Huisgen-Zimmerman.

Lemma 4.8.

[11, Lemma 4] Consider Notation 4.6 and let φ1(M)φ2(M)\varphi_{1}(M)\supseteq\varphi_{2}(M)\supseteq\dots be a descending chain of pp-definable subgroups of MM of some sort G𝒮G\in\mathcal{S}. Let

Ψ(n)={φn(Lj)j𝙹},Π(n)={i<nφn(Mi),inφn(Mi)},\begin{array}[]{cc}\Psi(n)=\{\varphi_{n}(L_{j})\mid j\in\mathtt{J}\},&\Pi(n)=\{\prod_{i<n}\varphi_{n}(M_{i}),\prod_{i\geq n}\varphi_{n}(M_{i})\},\end{array}

for any nn\in\mathbb{N}, and for any jj consider the map ρn,j=qj,Ψ(n)κφnu,Π(n)\rho_{n,j}=q_{j,\Psi(n)}\kappa\langle\varphi_{n}\rangle u_{\geq,\Pi(n)} given by

inφn(Mi)\textstyle{\prod_{i\geq n}\varphi_{n}(M_{i})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u,Π(n)\scriptstyle{u_{\geq,\Pi(n)}}iφn(Mi)\textstyle{\prod_{i\in\mathbb{N}}\varphi_{n}(M_{i})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κφn\scriptstyle{\kappa\langle\varphi_{n}\rangle}jJφn(Lj)\textstyle{\coprod_{j\in J}\varphi_{n}(L_{j})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qj,Ψ(n)\scriptstyle{q_{j,\Psi(n)}}φn(Lj).\textstyle{\varphi_{n}(L_{j}).}

For some rr\in\mathbb{N} and some 𝙹𝙹\mathtt{J}^{\prime}\subseteq\mathtt{J} finite, im(ρr,j)φn(Lj)\mathrm{im}(\rho_{r,j})\subseteq\varphi_{n}(L_{j}) for all nrn\geq r and j𝙹j\notin\mathtt{J}^{\prime}.

Proof.

Considering Notation 3.1, for each nn\in\mathbb{N}, we have that u,Π(n)u_{\geq,\Pi(n)} is (the inclusion) given by sending (xn+ii)=(xn,xn+1,)(x_{n+i}\mid i\in\mathbb{N})=(x_{n},x_{n+1},\dots) (where xiφn(Mi)x_{i}\in\varphi_{n}(M_{i})) to the sequence (0,,0,xn,xn+1,)(0,\dots,0,x_{n},x_{n+1},\dots), the initial nn terms of which are 0. Similarly, we have that qj,Ψ(n)q_{j,\Psi(n)} is the restriction of qj,𝒯(G,L)q_{j,\mathcal{T}(G,\mathrm{L})} for each jj and each nn.

We may assume 𝙹\mathtt{J} is non-empty since otherwise the statement is automatic. Assume for a contradiction that for any rr\in\mathbb{N} and any finite subset 𝙹\mathtt{J}^{\prime} of 𝙹\mathtt{J}, there exists nn\in\mathbb{N} with nrn\geq r and there exists j𝙹𝙹j\in\mathtt{J}\setminus\mathtt{J}^{\prime} such that im(ρr,j)φn(Lj)\mathrm{im}(\rho_{r,j})\not\subseteq\varphi_{n}(L_{j}). Following [11, Lemma 4], we firstly claim that there exists: a strictly increasing sequence of integers r(0)<r(1)<r(2)<r(3)<r(0)<r(1)<r(2)<r(3)<\dots; a sequence j(0),j(1),j(2),j(0),j(1),j(2),\dots of pairwise distinct elements of 𝙹\mathtt{J} (so, where j(n){j(0),,j(n1)}j(n)\notin\{j(0),\dots,j(n-1)\} for all n>0n>0); and a sequence of \mathbb{N}-tuples m¯0,m¯1,m¯2,\underline{m}_{0},\underline{m}_{1},\underline{m}_{2},\dots where

m¯dsr(d)φr(d)(Ms),ρr(d),j(d)(m¯d)φr(d+1)(Lj(d)),ρr(d),j(d)(m¯t)=0,\begin{array}[]{ccc}\underline{m}_{d}\in\prod_{s\geq r(d)}\varphi_{r(d)}(M_{s}),&\rho_{r(d),j(d)}(\underline{m}_{d})\notin\varphi_{r(d+1)}(L_{j(d)}),&\rho_{r(d),j(d)}(\underline{m}_{t})=0,\end{array}

for all dd\in\mathbb{N} and all tt\in\mathbb{N} with t<dt<d. We proceed inductively. Choose an element j(1)𝙹j(-1)\in\mathtt{J}. Let 𝙹0={j(1)}\mathtt{J}^{\prime}_{0}=\{j(-1)\} and r(0)=1r(0)=1. By our assumption that the conclusion is false, there exists an integer r(1)>0r(1)>0 and an element j(0)𝙹{j(1)}j(0)\in\mathtt{J}\setminus\{j(-1)\} where im(ρr(0),j(0))φr(1)(Lj(0))\mathrm{im}(\rho_{r(0),j(0)})\not\subseteq\varphi_{r(1)}(L_{j(0)}). Hence there exists n¯0ir(0)φr(0)(Mi)\underline{n}_{0}\in\prod_{i\geq r(0)}\varphi_{r(0)}(M_{i}) where ρr(0),j(0)(n¯0)φr(1)(Lj(0))\rho_{r(0),j(0)}(\underline{n}_{0})\notin\varphi_{r(1)}(L_{j(0)}), and so r(1)>r(0)r(1)>r(0).

We now iterate this process, yielding sequences (r(d)d)(r(d)\mid d\in\mathbb{N}), (j(d)d)(j(d)\mid d\in\mathbb{N}) and (n¯dd)(\underline{n}_{d}\mid d\in\mathbb{N}) where r(d)r(d) is a positive integer, j(d)𝙹j(d)\in\mathtt{J} and n¯dir(d)φr(d)(Mi)\underline{n}_{d}\in\prod_{i\geq r(d)}\varphi_{r(d)}(M_{i}) and such that r(d+1)>r(d)r(d+1)>r(d), j(d+1)J{j(1),,j(d)}j(d+1)\in J\setminus\{j(-1),\dots,j(d)\}, and ρr(d),j(d)(n¯d)φr(d+1)(Lj(d))\rho_{r(d),j(d)}(\underline{n}_{d})\notin\varphi_{r(d+1)}(L_{j(d)}). Now fix tt\in\mathbb{N}. Since we are considering coproducts of abelian groups, note that we have ρr(t),j(n¯t)=0\rho_{r(t),j}(\underline{n}_{t})=0 for all but finitely many j𝙹j\in\mathtt{J}. For each jj define the map ljt:GLjl_{j}^{t}\colon G\to L_{j} in φr(t)(Lj)\varphi_{r(t)}(L_{j}) by ljt=ρr(t),j(n¯t)l_{j}^{t}=\rho_{r(t),j}(\underline{n}_{t}) if j=j(d)j=j(d) for some d>td>t, and ljt=0l_{j}^{t}=0 otherwise. Now let

m¯~d=u,Π(r(d))(n¯d)κ1φr(d)(ljdj𝙹)iφr(d)(Mi).\begin{array}[]{c}\tilde{\underline{m}}_{d}=u_{\geq,\Pi(r(d))}(\underline{n}_{d})-\kappa^{-1}\langle\varphi_{r(d)}\rangle(l^{d}_{j}\mid j\in\mathtt{J})\in\prod_{i\in\mathbb{N}}\varphi_{r(d)}(M_{i}).\end{array}

By construction we have pi,M(m¯~d)=0p_{i,\mathrm{M}}(\tilde{\underline{m}}_{d})=0 for all i<r(d)i<r(d), and if t<dt<d then

qj(d),𝒯(G,L)(κφr(d)(m¯~t))=qj(d),𝒯(G,L)(κφr(d)(u,Π(r(d))(n¯t)))lj(d)t=0.\begin{array}[]{c}q_{j(d),\mathcal{T}(G,\mathrm{L})}(\kappa\langle\varphi_{r(d)}\rangle(\tilde{\underline{m}}_{t}))=q_{j(d),\mathcal{T}(G,\mathrm{L})}(\kappa\langle\varphi_{r(d)}\rangle(u_{\geq,\Pi(r(d))}(\underline{n}_{t})))-l^{t}_{j(d)}=0.\end{array}

Now, writing m¯~d=(md,0,md,1,)\tilde{\underline{m}}_{d}=(m_{d,0},m_{d,1},\dots) where md,iφr(d)(Mi)m_{d,i}\in\varphi_{r(d)}(M_{i}) for all ii\in\mathbb{N}, the above gives md,i=0m_{d,i}=0 for all i<r(d)i<r(d), so we may define

m¯d=(md,r(d)+ii)=(md,r(d),md,r(d)+1,md,r(d)+2)srdφr(d)(Ms).\begin{array}[]{c}\underline{m}_{d}=(m_{d,r(d)+i}\mid i\in\mathbb{N})=(m_{d,r(d)},m_{d,r(d)+1},m_{d,r(d)+2}\dots)\in\prod_{s\geq r_{d}}\varphi_{r(d)}(M_{s}).\end{array}

So we have m¯~d=u,Π(r(d))(m¯d)\tilde{\underline{m}}_{d}=u_{\geq,\Pi(r(d))}(\underline{m}_{d}), and therefore

ρr(d),j(d)(m¯d)=qj(d),𝒯(G,L)(κφr(d)(m¯~d))=ρr(d),j(d)(n¯d)lj(d)dφr(d+1)(Lj(d)),\begin{array}[]{c}\rho_{r(d),j(d)}(\underline{m}_{d})=q_{j(d),\mathcal{T}(G,\mathrm{L})}(\kappa\langle\varphi_{r(d)}\rangle(\tilde{\underline{m}}_{d}))=\rho_{r(d),j(d)}(\underline{n}_{d})-l^{d}_{j(d)}\notin\varphi_{r(d+1)}(L_{j(d)}),\end{array}

since by definition lj(d)d=0l^{d}_{j(d)}=0. Our calculations above likewise show that if t<dt<d then we have ρr(d),j(d)(m¯t)=0\rho_{r(d),j(d)}(\underline{m}_{t})=0. This verifies our initial claim. Now let dd vary. Since r(d)<r(d+1)r(d)<r(d+1) for all dd, for each ii\in\mathbb{N} observe that there are finitely many dd with r(d)ir(d)\leq i. Hence the sum

m¯~=im¯~i=(d,r(d)imd,ii)i𝒯(G,Mi)\begin{array}[]{c}\tilde{\underline{m}}=\sum_{i\in\mathbb{N}}\tilde{\underline{m}}_{i}=(\sum_{d\in\mathbb{N},r(d)\leq i}m_{d,i}\mid i\in\mathbb{N})\in\prod_{i\in\mathbb{N}}\mathcal{T}(G,M_{i})\end{array}

is well-defined. Now, for each ll\in\mathbb{N}, by combining everything so far with Corollary 4.7, we have

qj(l),𝒯(G,L)(γG,L1(λG,M1(m¯~)))=iqj(l),𝒯(G,L)(γG,L1(λG,M1(m¯~i)))=iqj(l),𝒯(G,L)(κφr(i)(m¯~i))=ilqj(l),𝒯(G,L)(κφr(i)(m¯~i))=ilρr(i),j(l)(m¯i)=ρr(l),j(l)(m¯l)+i>lρr(i),j(l)(m¯i).\begin{array}[]{c}q_{j(l),\mathcal{T}(G,\mathrm{L})}(\gamma^{-1}_{G,\mathrm{L}}(\lambda^{-1}_{G,\mathrm{M}}(\tilde{\underline{m}})))=\sum_{i\in\mathbb{N}}q_{j(l),\mathcal{T}(G,\mathrm{L})}(\gamma^{-1}_{G,\mathrm{L}}(\lambda^{-1}_{G,\mathrm{M}}(\tilde{\underline{m}}_{i})))\\ =\sum_{i\in\mathbb{N}}q_{j(l),\mathcal{T}(G,\mathrm{L})}(\kappa\langle\varphi_{r(i)}\rangle(\tilde{\underline{m}}_{i}))=\sum_{i\geq l}q_{j(l),\mathcal{T}(G,\mathrm{L})}(\kappa\langle\varphi_{r(i)}\rangle(\tilde{\underline{m}}_{i}))\\ =\sum_{i\geq l}\rho_{r(i),j(l)}(\underline{m}_{i})=\rho_{r(l),j(l)}(\underline{m}_{l})+\sum_{i>l}\rho_{r(i),j(l)}(\underline{m}_{i}).\end{array}

Now recall that ρr(l),j(l)(m¯l)φr(l+1)(Lj(l))\rho_{r(l),j(l)}(\underline{m}_{l})\notin\varphi_{r(l+1)}(L_{j(l)}). Since φ1(M)φ2(M)\varphi_{1}(M)\supseteq\varphi_{2}(M)\supseteq\dots is descending and r(i+1)>r(i)r(i+1)>r(i) for all ii, we have ρr(i),j(l)(m¯i)φr(i)(Lj(l))φr(l+1)(Lj(l))\rho_{r(i),j(l)}(\underline{m}_{i})\in\varphi_{r(i)}(L_{j(l)})\subseteq\varphi_{r(l+1)}(L_{j(l)}) whenever i>li>l. Together with the above, this shows

qj(l),𝒯(G,L)(γG,L1(λG,M1(m¯~)))0q_{j(l),\mathcal{T}(G,\mathrm{L})}(\gamma^{-1}_{G,\mathrm{L}}(\lambda^{-1}_{G,\mathrm{M}}(\tilde{\underline{m}})))\neq 0

for all ll. Since γG,L1(λG,M1(m¯~))\gamma^{-1}_{G,\mathrm{L}}(\lambda^{-1}_{G,\mathrm{M}}(\tilde{\underline{m}})) lies in the coproduct j𝒯(G,Lj)\coprod_{j}\mathcal{T}(G,L_{j}), and therefore must have had finite support over j𝙹j\in\mathtt{J}, we have a contradiction, since the set {j(l)l}\{j(l)\mid l\in\mathbb{N}\} is in bijection with \mathbb{N}. ∎

5. Annihilator subobjects and pp-definable subgroups.

Recall that, in a category with all small coproducts, a set {𝒢αα𝚇}\{\mathscr{G}_{\alpha}\mid\alpha\in\mathtt{X}\} of objects is called a set of generators provided, for each object 𝒬\mathscr{Q}, there is an epimorphism α𝒢α𝒬\coprod_{\alpha}\mathscr{G}_{\alpha}\to\mathscr{Q}. In case 𝚇\mathtt{X} is a singleton we say the category has a generator. Recall an additive category 𝒜\mathcal{A} is Grothendieck provided: 𝒜\mathcal{A} is abelian; 𝒜\mathcal{A} has all small coproducts; 𝒜\mathcal{A} has a generator; and the direct limit of any short exact sequence in 𝒜\mathcal{A} is again exact.

Remark 5.1.

Let 𝒜\mathcal{A} be a Grothendieck category. Recall that an object 𝒬\mathscr{Q} of 𝒜\mathcal{A} is finitely presented provided the functor 𝒜(𝒬,):𝒜𝐀𝐛\mathcal{A}(\mathscr{Q},-)\colon\mathcal{A}\to\mathbf{Ab} commutes with direct limits. Recall that an object 𝒮\mathscr{S} of 𝒜\mathcal{A} is finitely generated provided there is an exact sequence 𝒬0\mathscr{R}\to\mathscr{Q}\to 0 in 𝒜\mathcal{A}. 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.

Following Krause [15], a category 𝒜\mathcal{A} is said to be locally coherent provided: 𝒜\mathcal{A} is a Grothendieck category; 𝒜\mathcal{A} has a set {𝒢αα𝚇}\{\mathscr{G}_{\alpha}\mid\alpha\in\mathtt{X}\} of generators such that each 𝒢α\mathscr{G}_{\alpha} is finitely presented; and the full subcategory of 𝒜\mathcal{A} consisting of finitely presented objects is abelian. As noted at the top of [8, p.3], 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is locally coherent.

Thus, in Definition 5.2 and Lemmas 5.3 and 5.4, we specify various definitions and results from [7] and [9] to 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} (which can be done by Remark 5.1). We now recall a notion introduced by Harada.

Definition 5.2.

[9, §1] Let 𝒜\mathcal{A} be a Grothendieck category with a set {𝒢αα𝚇}\{\mathscr{G}_{\alpha}\mid\alpha\in\mathtt{X}\} of finitely generated generators. Let 𝒬\mathscr{Q} and \mathscr{R} be objects in 𝒜\mathcal{A}. A subobject 𝒫\mathscr{P} of 𝒬\mathscr{Q} is said to be an \mathscr{R}-annihilator subobject of 𝒬\mathscr{Q} provided 𝒫=𝔣𝙺ker(𝔣)\mathscr{P}=\bigcap_{\mathfrak{f}\in\mathtt{K}}\mathrm{ker}(\mathfrak{f}) where the intersection is taken over morphisms 𝔣:𝒬\mathfrak{f}\colon\mathscr{Q}\to\mathscr{R} running through some 𝙺𝒜(𝒬,)\mathtt{K}\subseteq\mathcal{A}(\mathscr{Q},\mathscr{R}).

Lemma 5.3 focuses on a particular context of Definition 5.2. That is, we specifiy to the locally coherent category 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} and consider the image of objects under 𝐘\mathbf{Y}. Note Lemma 5.3 was written only to simplify the proof of Lemma 5.4.

Lemma 5.3.

Let 𝔮:𝐘(X)𝒬\mathfrak{q}\colon\mathbf{Y}(X)\to\mathscr{Q} be an epimorphism in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} where the object XX of 𝒯\mathcal{T} is compact. If MM and ZZ are objects of 𝒯\mathcal{T} and 𝒯c\mathcal{T}^{c} respectively, then for any 𝐘(M)\mathbf{Y}(M)-annihilator subobject 𝒫=𝔣𝙺ker(𝔣)\mathscr{P}=\bigcap_{\mathfrak{f}\in\mathtt{K}}\mathrm{ker}(\mathfrak{f}) of 𝒬\mathscr{Q} (where 𝙺𝒜(𝒬,)\mathtt{K}\subseteq\mathcal{A}(\mathscr{Q},\mathscr{R})) we have

𝒫(Z)={𝔮Z(g)g𝒯(Z,X) and 𝔣X(𝔮X(1X))g=0 for all 𝔣𝙺}.\mathscr{P}(Z)=\{\mathfrak{q}_{Z}(g)\mid g\in\mathcal{T}(Z,X)\text{ and }\mathfrak{f}_{X}(\mathfrak{q}_{X}(1_{X}))g=0\text{ for all }\mathfrak{f}\in\mathtt{K}\}.
Proof.

It suffices to assume 𝙺\mathtt{K}\neq\emptyset. Let p𝒬(Z)p\in\mathscr{Q}(Z). Since 𝔮Z\mathfrak{q}_{Z} is onto, p=𝔮Z(g)p=\mathfrak{q}_{Z}(g) for some g𝒯(Z,X)g\in\mathcal{T}(Z,X). Since 𝔮\mathfrak{q} and each 𝔣\mathfrak{f} are morphisms in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} the diagram of abelian groups given by

𝒯(X,X)\textstyle{\mathcal{T}(X,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯(g,X)\scriptstyle{\mathcal{T}(g,X)}𝔮X\scriptstyle{\mathfrak{q}_{X}}𝒬(X)\textstyle{\mathscr{Q}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒬(g)\scriptstyle{\mathscr{Q}(g)}𝔣X\scriptstyle{\mathfrak{f}_{X}}𝒯(X,M)\textstyle{\mathcal{T}(X,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯(g,M)\scriptstyle{\mathcal{T}(g,M)}𝒯(Z,X)\textstyle{\mathcal{T}(Z,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔮Z\scriptstyle{\mathfrak{q}_{Z}}𝒬(Z)\textstyle{\mathscr{Q}(Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔣Z\scriptstyle{\mathfrak{f}_{Z}}𝒯(Z,M)\textstyle{\mathcal{T}(Z,M)}

commutes. By the commutativity of the diagram we have 𝔮Z(g)=𝒬(g)(𝔮X(1X))\mathfrak{q}_{Z}(g)=\mathscr{Q}(g)(\mathfrak{q}_{X}(1_{X})) and (hence) 𝔣Z(p)=𝔣X(𝔮X(1X))g\mathfrak{f}_{Z}(p)=\mathfrak{f}_{X}(\mathfrak{q}_{X}(1_{X}))g. Now suppose p𝔣𝙺ker(𝔣Z)p\in\bigcap_{\mathfrak{f}\in\mathtt{K}}\mathrm{ker}(\mathfrak{f}_{Z}). By the above this means 𝔣X(𝔮X(1X))g=0\mathfrak{f}_{X}(\mathfrak{q}_{X}(1_{X}))g=0 for all 𝔣\mathfrak{f}. Hence 𝒫(Z)\mathscr{P}(Z) 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 Σ\Sigma-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 MM and XX be objects in 𝒯\mathcal{T} and 𝒯c\mathcal{T}^{c} respectively, and let 𝒯(,X)𝒬0\mathcal{T}(-,X)\to\mathscr{Q}\to 0 be a sequence in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} which is exact. Any strictly ascending chain of (𝐘(M)\mathbf{Y}(M)-annihilator subobjects of 𝒬\mathscr{Q}) gives a strictly descending chain of (pp-definable subgroups of MM of sort XX).

Proof.

Suppose 𝒫1𝒫2\mathscr{P}_{1}\subsetneq\mathscr{P}_{2}\subsetneq\dots is a strictly ascending chain of 𝐘(M)\mathbf{Y}(M)-annihilator subobjects of 𝒬\mathscr{Q}, say where, for each integer n>0n>0, we have 𝒫n=𝔣𝙺[n]ker(𝔣)\mathscr{P}_{n}=\bigcap_{\mathfrak{f}\in\mathtt{K}[n]}\mathrm{ker}(\mathfrak{f}) for some subset 𝙺[n]\mathtt{K}[n] of morphisms in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} of the form 𝔣:𝒬𝐘(M)\mathfrak{f}\colon\mathscr{Q}\to\mathbf{Y}(M).

We assume 𝙺[1]𝙺[2]\mathtt{K}[1]\supsetneq\mathtt{K}[2]\supsetneq\dots without loss of generality. For each nn there is an object ZnZ_{n} of 𝒯c\mathcal{T}^{c} for which 𝒫n(Zn)𝒫n+1(Zn)\mathscr{P}_{n}(Z_{n})\subsetneq\mathscr{P}_{n+1}(Z_{n}), and we choose hn𝒫n+1(Zn)𝒫n(Zn)h_{n}\in\mathscr{P}_{n+1}(Z_{n})\setminus\mathscr{P}_{n}(Z_{n}). Let 𝔮:𝒯(X,)𝒬\mathfrak{q}\colon\mathcal{T}(X,-)\to\mathscr{Q} be the epimorphism in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} giving the exact sequence 𝒯(,X)𝒬0\mathcal{T}(-,X)\to\mathscr{Q}\to 0. Since 𝔮Zn\mathfrak{q}_{Z_{n}} is onto and hn𝒬(Zn)h_{n}\in\mathscr{Q}(Z_{n}) we have hn=𝔮Zn(gn)h_{n}=\mathfrak{q}_{Z_{n}}(g_{n}) for some morphism gn:ZnXg_{n}\colon Z_{n}\to X. By Lemma 5.3, since hn𝒫n+1(Zn)h_{n}\in\mathscr{P}_{n+1}(Z_{n}) we have that 𝔣X(𝔮X(1X))gn=0\mathfrak{f}_{X}(\mathfrak{q}_{X}(1_{X}))g_{n}=0 for all 𝔣𝙺[n+1]\mathfrak{f}\in\mathtt{K}[n+1]. Similarly, since hn𝒫n(Zn)h_{n}\notin\mathscr{P}_{n}(Z_{n}) there exists 𝔰(n)𝙺[n]\mathfrak{s}(n)\in\mathtt{K}[n] such that 𝔰(n)X(𝔮X(1X))gn0\mathfrak{s}(n)_{X}(\mathfrak{q}_{X}(1_{X}))g_{n}\neq 0.

We now follow the proof of [8, Proposition 3.1]. Since 𝒯c\mathcal{T}^{c} is triangulated any morphism g:ZXg:Z\to X in 𝒯c\mathcal{T}^{c} yields a triangle in 𝒯c\mathcal{T}^{c} and an exact sequence in 𝐀𝐛\mathbf{Ab}

ZgXbYΣZ,𝒯(Y,W)b𝒯(X,W)g𝒯(Z,W),\begin{array}[]{cc}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 6.77083pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\crcr}}}\ignorespaces{\hbox{\kern-6.77083pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 14.35782pt\raise 5.1875pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{g}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 30.77083pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 30.77083pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 53.07776pt\raise 5.43056pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.43056pt\hbox{$\scriptstyle{b}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 69.84026pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 69.84026pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 107.86803pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 107.86803pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\Sigma Z}$}}}}}}}\ignorespaces}}}}\ignorespaces,&\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 21.31944pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\crcr}}}\ignorespaces{\hbox{\kern-21.31944pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{T}(Y,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 28.18889pt\raise 5.43056pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.43056pt\hbox{$\scriptstyle{-b}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 45.31944pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 45.31944pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{T}(X,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 96.46198pt\raise 5.1875pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{-g}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 114.1111pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 114.1111pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{T}(Z,W),}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{array}

given by applying the contravaraint functor 𝒯(,W):𝒯𝐀𝐛\mathcal{T}(-,W)\colon\mathcal{T}\to\mathbf{Ab}. In other words, bb is a pseudocokernel of gg, and so any morphism t:XWt\colon X\to W in 𝒯c\mathcal{T}^{c} with tg=0tg=0 must satisfy t=sbt=sb for some s:YZs\colon Y\to Z.

Let tn=𝔰(n)X(𝔮X(1X))t_{n}=\mathfrak{s}(n)_{X}(\mathfrak{q}_{X}(1_{X})) for each nn. Combining what we have so far, for each nn we have 𝔰(n+1)𝙺[n+1]\mathfrak{s}(n+1)\in\mathtt{K}[n+1], so tn+1gn=0t_{n+1}g_{n}=0, and so tn+1=snbnt_{n+1}=s_{n}b_{n} for some morphism sn:YnMs_{n}\colon Y_{n}\to M, and so tn+1Mbnt_{n+1}\in Mb_{n}. On the other hand, if tn+1Mbn+1t_{n+1}\in Mb_{n+1} then tn+1gn+1=0t_{n+1}g_{n+1}=0 which contradicts that 𝔰(n+1)X(𝔮X(1X))gn+10\mathfrak{s}(n+1)_{X}(\mathfrak{q}_{X}(1_{X}))g_{n+1}\neq 0, and so tn+1Mbn+1t_{n+1}\notin Mb_{n+1}. This gives a strict descending chain

Mb1Mb1Mb2Mb1Mb2Mb3i=1dMbi\begin{array}[]{c}Mb_{1}\supsetneq Mb_{1}\cap Mb_{2}\supsetneq Mb_{1}\cap Mb_{2}\cap Mb_{3}\supsetneq\dots\supsetneq\bigcap_{i=1}^{d}Mb_{i}\supsetneq\dots\end{array}

A direct application of [8, Proposition 3.1] shows that each finite intersection i=1dMbi\bigcap_{i=1}^{d}Mb_{i} has the form MadMa_{d} for some morphism in 𝒯c\mathcal{T}^{c} of the form ad:XWda_{d}\colon X\to W_{d}. So the chain above is, as required, a strictly descending chain of pp-definable subgroups of MM of sort XX. ∎

Definition 5.5.

[9, §1] Let 𝒜\mathcal{A} be a Grothendieck category with a set of finitely generated generators. Fix an object \mathscr{M} of 𝒜\mathcal{A}. We say that \mathscr{M} is Σ\Sigma-injective if, for any set 𝙸\mathtt{I}, the coproduct (𝙸)=i𝙸\mathscr{M}^{(\mathtt{I})}=\coprod_{i\in\mathtt{I}}\mathscr{M} is injective. We say that \mathscr{M} is fp-injective if, whenever 0𝒫𝒬00\to\mathscr{P}\to\mathscr{R}\to\mathscr{Q}\to 0 is an exact sequence in 𝒜\mathcal{A} where 𝒬\mathscr{Q} is finitely presented, any morphism 𝒫\mathscr{P}\to\mathscr{M} extends to a morphism \mathscr{R}\to\mathscr{M}; see [7, §1].

For the proof of Corollary 5.8 we recall two results: Proposition 5.6, due to Garcia and Dung, characterises Σ\Sigma-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 \mathscr{M} be an fp-injecitve in a Grothendieck category 𝒜\mathcal{A} which has a set 𝙶\mathtt{G} of finitely presented generators 𝒢\mathscr{G}. Then \mathscr{M} is Σ\Sigma-injective if and only if, for each 𝒢𝙶\mathscr{G}\in\mathtt{G}, every ascending chain of \mathscr{M}-annihilator subobjects of 𝒢\mathscr{G} must stabilise.

Lemma 5.7.

[17, Lemma 1.6] For any MM in 𝒯\mathcal{T} the image 𝐘(M)\mathbf{Y}(M) is fp-injective.

Corollary 5.8.

Let MM be an object in 𝒯\mathcal{T} such that for any compact object XX of 𝒯\mathcal{T} we have that each descending chain Ma1Ma2Ma_{1}\supseteq Ma_{2}\supseteq\dots of pp-definable subgroups of MM of sort XX must stabilise. Then the image 𝐘(M)\mathbf{Y}(M) of MM in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is Σ\Sigma-injective.

Proof.

We prove the contrapositive, so we assume 𝐘(M)\mathbf{Y}(M) is not Σ\Sigma-injective. Recall, from Remark 5.1, that 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is locally coherent, and so it is a Grothendieck category with a set 𝙶\mathtt{G} of finitely presented generators.

Note =𝐘(M)\mathscr{M}=\mathbf{Y}(M) is fp-injective by Lemma 5.7, and combining our initial assumption with Proposition 5.6 shows that, for some 𝒢𝙶\mathscr{G}\in\mathtt{G}, there exists a strictly ascending chain of \mathscr{M}-annihilator subobjects of 𝒢\mathscr{G}. Since 𝒢\mathscr{G} is finitely presented, there is an exact sequence of the form 𝒯(,Y)𝒯(,X)𝒢0\mathcal{T}(-,Y)\to\mathcal{T}(-,X)\to\mathscr{G}\to 0 in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} where XX and YY lie in 𝒯c\mathcal{T}^{c}.

By Lemma 5.4 the aforementioned ascending chain strict ascending chain gives rise to a strictly descending chain of pp-definable subgroups of MM of sort XX. ∎

6. Σ\Sigma-pure-injective objects and canonical morphisms.

Definition 6.1.

Recall Notation 3.1. Let 𝙸\mathtt{I} be a set and let MM be an object of 𝒯\mathcal{T}. By the universal properties of the product and coproduct of the collection M={Mi𝙸}\mathrm{M}=\{M\mid i\in\mathtt{I}\}, there exists a unique summation morphism σ𝙸,M:iMM\sigma_{\mathtt{I},\mathrm{M}}:\coprod_{i}M\to M and a unique canonical morphism ι𝙸,M:iMiM\iota_{\mathtt{I},\mathrm{M}}:\coprod_{i}M\to\prod_{i}M for which σ𝙸,Mui,M=1M\sigma_{\mathtt{I},\mathrm{M}}u_{i,\mathrm{M}}=1_{M} and ι𝙸,Mui,M=vi,M\iota_{\mathtt{I},\mathrm{M}}u_{i,\mathrm{M}}=v_{i,\mathrm{M}} for each ii.

Proposition 6.2.

Let MM be an object of 𝒯\mathcal{T} and let 𝙸\mathtt{I} be a set. Then the canonical morphism ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is a pure monomorphism.

Proof.

Let XX be a object in 𝒯\mathcal{T} which is compact. In general: the morphism λX,M\lambda_{X,\mathrm{M}} is an isomorphism; the canonical morphism ι𝙸,𝒯(X,M)\iota_{\mathtt{I},\mathcal{T}(X,\mathrm{M})} is injective; and ι𝙸,𝒯(X,M)\iota_{\mathtt{I},\mathcal{T}(X,\mathrm{M})} is the composition λX,M𝒯(X,ι𝙸,M)γX,M\lambda_{X,\mathrm{M}}\mathcal{T}(X,\iota_{\mathtt{I},\mathrm{M}})\gamma_{X,\mathrm{M}}.

Since XX is compact the morphism γX,M\gamma_{X,\mathrm{M}} is an isomorphism. This shows 𝒯(X,ι𝙸,M)\mathcal{T}(X,\iota_{\mathtt{I},\mathrm{M}}) is injective if XX is compact, and so ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is a pure monomorphism. ∎

Definition 6.3.

[17, Definition 1.1] An object MM of 𝒯\mathcal{T} is called pure-injective if each pure monomorphism MNM\to N is a section, and MM is called Σ\Sigma-pure-injective if, for any set 𝙸\mathtt{I}, the coproduct i𝙸M=M(𝙸)\coprod_{i\in\mathtt{I}}M=M^{(\mathtt{I})} 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 MM of 𝒯\mathcal{T} the following statements are equivalent.

  1. (1)

    The object MM of 𝒯\mathcal{T} is pure-injective.

  2. (2)

    The object 𝐘(M)\mathbf{Y}(M) of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is injective.

  3. (3)

    For any set 𝙸\mathtt{I} the morphism σ𝙸,M\sigma_{\mathtt{I},\mathrm{M}} factors through the morphism ι𝙸,M\iota_{\mathtt{I},\mathrm{M}}.

Proposition 6.5 is analogous to parts (i) and (ii) in [14, Theorem 8.1].

Proposition 6.5.

An object MM of 𝒯\mathcal{T} is Σ\Sigma-pure-injective if and only if, for each set 𝙸\mathtt{I}, the canonical morphism ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is a section.

Proof.

Assume that MM is Σ\Sigma-pure-injective and that 𝙸\mathtt{I} is a set. By assumption the domain of ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is pure-injective. Since ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is a pure monomorphism, this means it is a section. Supposing conversley that ι𝙸,M\iota_{\mathtt{I},\mathrm{M}} is a section for each set 𝙸\mathtt{I}, it remains to show that MM is Σ\Sigma-pure-injective. Choose a set 𝚃\mathtt{T} and let N=t𝚃MN=\coprod_{t\in\mathtt{T}}M. It suffices to prove NN is pure-injective. Let 𝚂\mathtt{S} be any set and consider the collection N={Ns𝚂}\mathrm{N}=\{N\mid s\in\mathtt{S}\}. By Theorem 6.4 it suffices to find a map θ𝚂,N:sNN\theta_{\mathtt{S},\mathrm{N}}\colon\prod_{s}N\to N such that σ𝚂,N=θ𝚂,Nι𝚂,N\sigma_{\mathtt{S},\mathrm{N}}=\theta_{\mathtt{S},\mathrm{N}}\iota_{\mathtt{S},\mathrm{N}}. Let M={Mt𝚃}\mathrm{M}=\{M\mid t\in\mathtt{T}\}.

For each (s,t)𝚂×𝚃(s,t)\in\mathtt{S}\times\mathtt{T} the morphisms us,Nut,Mu_{s,\mathrm{N}}u_{t,\mathrm{M}} satisfy the universal property of the coproduct s,tM\coprod_{s,t}M, and so we assume us,t,M=us,Nut,Mu_{s,t,\mathrm{M}}=u_{s,\mathrm{N}}u_{t,\mathrm{M}} without loss of generality. Consider the morphisms φs,t,M=qt,Mps,N\varphi_{s,t,\mathrm{M}}=q_{t,\mathrm{M}}p_{s,\mathrm{N}} for each (s,t)𝚂×𝚃(s,t)\in\mathtt{S}\times\mathtt{T}. Since us,t,M=us,Nut,Mu_{s,t,\mathrm{M}}=u_{s,\mathrm{N}}u_{t,\mathrm{M}} we have qs,t,M=qt,Mqs,Nq_{s,t,\mathrm{M}}=q_{t,\mathrm{M}}q_{s,\mathrm{N}} by uniqueness. Consequently φs,t,Mvs,Nqs,Nus,t,M\varphi_{s,t,\mathrm{M}}v_{s,\mathrm{N}}q_{s,\mathrm{N}}u_{s,t,\mathrm{M}} is the identity on MM. By the universal property of the product, there is a morphism Ξ:sNs,tM\Xi\colon\prod_{s}N\to\prod_{s,t}M such that ps,t,MΞ=φs,t,Mp_{s,t,\mathrm{M}}\Xi=\varphi_{s,t,\mathrm{M}} for each (s,t)(s,t). It suffices to let θ𝚂,N=σ𝚂,Nπ𝚂×𝚃,MΞ\theta_{\mathtt{S},\mathrm{N}}=\sigma_{\mathtt{S},\mathrm{N}}\pi_{\mathtt{S}\times\mathtt{T},\mathrm{M}}\Xi. By the uniqueness of the involved morphisms, it is straightforward to see that σ𝚂,N=θ𝚂,Nι𝚂,N\sigma_{\mathtt{S},\mathrm{N}}=\theta_{\mathtt{S},\mathrm{N}}\iota_{\mathtt{S},\mathrm{N}}.

Lemma 6.6 is analogous to [14, Theorem 8.1(ii,iii)].

Lemma 6.6.

If G𝒮G\in\mathcal{S} and MM is a Σ\Sigma-pure-injective object in 𝒯\mathcal{T} then every descending chain of pp-definable subgroups of MM of sort GG stabilises.

Proof.

For a contradiction we assume the existence of a strictly descending chain Ma0Ma1Ma2Ma_{0}\supsetneq Ma_{1}\supsetneq Ma_{2}\supsetneq\dots of (pp-definable subgroups of MM of sort GG) for some G𝒮G\in\mathcal{S}. Hence there is a collection of compact objects Hn𝒮H_{n}\in\mathcal{S} such that an𝒯(G,Hn)a_{n}\in\mathcal{T}(G,H_{n}) for each nn\in\mathbb{N}. By our assumption we may choose elements bn𝒯(Hn,M)b_{n}\in\mathcal{T}(H_{n},M) such that bnanMan+1b_{n}a_{n}\notin Ma_{n+1}. By Proposition 6.5 the canonical morphism ι,M:MM\iota_{\mathbb{N},\mathrm{M}}:\coprod_{\mathbb{N}}M\to\prod_{\mathbb{N}}M is a section, and so there is some morphism π,M:MM\pi_{\mathbb{N},\mathrm{M}}\colon\prod_{\mathbb{N}}M\to\coprod_{\mathbb{N}}M such that π,Mι,M\pi_{\mathbb{N},\mathrm{M}}\iota_{\mathbb{N},\mathrm{M}} is the identity on M\coprod_{\mathbb{N}}M. After applying the functor 𝒯(G,)\mathcal{T}(G,-) (from 𝒯\mathcal{T} to the category of abelian groups) this means 𝒯(G,π,M)𝒯(G,ι,M)\mathcal{T}(G,\pi_{\mathbb{N},\mathrm{M}})\mathcal{T}(G,\iota_{\mathbb{N},\mathrm{M}}) is the identity on 𝒯(G,M)\mathcal{T}(G,\coprod_{\mathbb{N}}M). Let ba¯=(bnann)n𝒯(G,M)\underline{ba}=(b_{n}a_{n}\mid n\in\mathbb{N})\in\prod_{n}\mathcal{T}(G,M). Fix nn\in\mathbb{N} and let φn(vG)\varphi_{n}(v_{G}) be the formula (uHn:vG=uHnan)(\exists u_{H_{n}}\colon v_{G}=u_{H_{n}}a_{n}). Let M={Mn}\mathrm{M}=\{M\mid n\in\mathbb{N}\}. Recall λG,M\lambda_{G,\mathrm{M}} is always an isomorphism, and since GG is compact, γG,M\gamma_{G,\mathrm{M}} is an isomorphism. Define μ\mu by

μ=(γG,M)1𝒯(G,π,M)(λG,M)1:n𝒯(G,M)n𝒯(G,M).\begin{array}[]{c}\mu=(\gamma_{G,\mathrm{M}})^{-1}\mathcal{T}(G,\pi_{\mathbb{N},\mathrm{M}})(\lambda_{G,\mathrm{M}})^{-1}\colon\prod_{n\in\mathbb{N}}\mathcal{T}(G,M)\to\coprod_{n\in\mathbb{N}}\mathcal{T}(G,M).\end{array}

Let μ(ba¯)=(cnn)\mu(\underline{ba})=(c_{n}\mid n\in\mathbb{N}). The contradiction we will find is that cl0c_{l}\neq 0 for all ll\in\mathbb{N}, which contradicts that μ\mu has codomain n𝒯(G,M)\coprod_{n\in\mathbb{N}}\mathcal{T}(G,M). Fix ll\in\mathbb{N}. Now let

ba¯l=(b0a0,,blal,0,0,) and ba¯>l=(0,,0,bl+1al+1,bl+2al+2,),\underline{ba}_{\leq l}=(b_{0}a_{0},\dots,b_{l}a_{l},0,0,\dots)\text{ and }\underline{ba}_{>l}=(0,\dots,0,b_{l+1}a_{l+1},b_{l+2}a_{l+2},\dots),

where the first ll entries of ba¯>l\underline{ba}_{>l} are 0.

Note that ba¯ln𝒯(G,M)\underline{ba}_{\leq l}\in\coprod_{n}\mathcal{T}(G,M). Furthermore, since the chain Ma0Ma1Ma_{0}\supseteq Ma_{1}\supseteq\cdots is descending, we have bnanφl(M)b_{n}a_{n}\in\varphi_{l}(M) for all n>ln>l and so ba¯>lnφl+1(M)\underline{ba}_{>l}\in\prod_{n}\varphi_{l+1}(M). By Lemma 4.2(ii), the restrictions of (λG,M)1(\lambda_{G,\mathrm{M}})^{-1} and (γG,M)1(\gamma_{G,\mathrm{M}})^{-1} respectively define isomorphisms nφl+1(M)φl+1(nM)\prod_{n}\varphi_{l+1}(M)\to\varphi_{l+1}(\prod_{n}M) and φl+1(nM)nφl+1(M)\varphi_{l+1}(\coprod_{n}M)\to\coprod_{n}\varphi_{l+1}(M). Similarly 𝒯(G,π,M)\mathcal{T}(G,\pi_{\mathbb{N},\mathrm{M}}) restricts to define a morphism φl+1(nM)φl+1(nM)\varphi_{l+1}(\prod_{n}M)\to\varphi_{l+1}(\coprod_{n}M). Altogether we have that μ\mu restricts to a morphism nφl+1(M)nφl+1(M)\prod_{n}\varphi_{l+1}(M)\to\coprod_{n}\varphi_{l+1}(M). Let μ(ba¯>l)=(dnn)\mu(\underline{ba}_{>l})=(d_{n}\mid n\in\mathbb{N}), and so dnφl+1(M)d_{n}\in\varphi_{l+1}(M) for all nn.

Recall it suffices to show cl0c_{l}\neq 0 where μ(ba¯)=(cnn)\mu(\underline{ba})=(c_{n}\mid n\in\mathbb{N}). From the above,

(c0,,cl,cl+1,)=μ(ba¯)=μ(ba¯l+ba¯>l)=ba¯l+μ(ba¯>l)=(b0a0+d0,,blal+dl,bl+1al+1+dl+1,),\begin{array}[]{c}(c_{0},\dots,c_{l},c_{l+1},\dots)=\mu(\underline{ba})=\mu(\underline{ba}_{\leq l}+\underline{ba}_{>l})=\underline{ba}_{\leq l}+\mu(\underline{ba}_{>l})\\ =(b_{0}a_{0}+d_{0},\dots,b_{l}a_{l}+d_{l},b_{l+1}a_{l+1}+d_{l+1},\dots),\end{array}

and so cl0c_{l}\neq 0 as otherwise φl+1(M)dl=blalφl+1(M)\varphi_{l+1}(M)\ni-d_{l}=b_{l}a_{l}\notin\varphi_{l+1}(M). ∎

We now recall a result of Krause which is used heavily in the sequel.

Corollary 6.7.

[17, Corollary 1.10] The functor 𝐘\mathbf{Y} gives an equivalence between the full subcategory of 𝒯\mathcal{T} consisting of pure-injective objects and the full subcategory of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} consitsting of injective objects.

Note that, since 𝐘\mathbf{Y} is additive, by Corollary 6.7 a pure-injective object MM is indecomposable if and only if 𝐘(M)\mathbf{Y}(M) is an indecomposable (injective) object.

Remark 6.8.

Fix a collection M={Mii𝙸}\mathrm{M}=\{M_{i}\mid i\in\mathtt{I}\} of objects in 𝒯\mathcal{T}. Since small coproducts exist in 𝒯\mathcal{T} (by Assumption 3.3) and 𝐀𝐛\mathbf{Ab}, the morphisms γX,M\gamma_{X,\mathrm{M}} combine to define a natural transformation i𝒯(,Mi)𝒯(,iMi)\coprod_{i}\mathcal{T}(-,M_{i})\to\mathcal{T}(-,\coprod_{i}M_{i}). By Definition 3.2 this transformation is in fact an isomorphism i𝐘(Mi)𝐘(iMi)\coprod_{i}\mathbf{Y}(M_{i})\simeq\mathbf{Y}(\coprod_{i}M_{i}) in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} which shows that 𝐘\mathbf{Y} preserves small coproducts.

Corollary 6.9.

The functor 𝐘\mathbf{Y} gives an equivalence between the full subcategory of 𝒯\mathcal{T} consisting of Σ\Sigma-pure-injective objects and the full subcategory of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} consitsting of Σ\Sigma-injective objects.

Proof.

By definition, an object MM of 𝒯\mathcal{T} is Σ\Sigma-pure-injective if and only if, for every set 𝙸\mathtt{I}, the coproduct M(𝙸)M^{(\mathtt{I})} is pure-injective. By Corollary 6.7 and Remark 6.8, given any such 𝙸\mathtt{I}, M(𝙸)M^{(\mathtt{I})} is pure-injective if and only if 𝐘(M(𝙸))(𝐘(M))(𝙸)\mathbf{Y}(M^{(\mathtt{I})})\simeq(\mathbf{Y}(M))^{(\mathtt{I})} is injective.

This shows that 𝐘\mathbf{Y} induces a functor from the full subcategory of Σ\Sigma-pure-injectives in 𝒯\mathcal{T} and the full subcategory of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} consisting of Σ\Sigma-injectives. That this functor is full, faithful and dense, follows by Corollary 6.7 together with the fact that any Σ\Sigma-pure-injective object of 𝒯\mathcal{T} is pure-injective. ∎

Corollary 6.10 is analogous to [14, Corollary 8.2(i,ii)].

Corollary 6.10.

Let MM be a Σ\Sigma-pure-injective object of 𝒯\mathcal{T} and let G𝒮G\in\mathcal{S}.

  1. (1)

    For any set 𝙸\mathtt{I}, the objects M(𝙸)M^{(\mathtt{I})} and M𝙸M^{\mathtt{I}} are Σ\Sigma-pure-injective.

  2. (2)

    If h:LMh\colon L\to M is a pure monomorphism in 𝒯\mathcal{T} then LL is Σ\Sigma-pure-injective and hh is a section.

Proof.

It is worth noting that we now have the equivalence of (1) and (5) of Theorem 1.1. That is, by Lemma 6.6 and Corollaries 5.8 and 6.9, an object MM is Σ\Sigma-pure-injective if and only if every descending chain of pp-definable subgroups of MM of each sort stabilises.

The proof of part (1) is now straightforward, recalling that, by Lemma 4.5, we have that φ(M)(𝙸)φ(M(𝙸))\varphi(M)^{(\mathtt{I})}\simeq\varphi(M^{(\mathtt{I})}) and φ(M𝙸)φ(M)𝙸\varphi(M^{\mathtt{I}})\simeq\varphi(M)^{\mathtt{I}} for any pp-formula φ\varphi of sort GG. For (2), as above it suffices to recall that φ(L)={g𝒯(G,L)hgφ(M)}\varphi(L)=\{g\in\mathcal{T}(G,L)\mid hg\in\varphi(M)\} for any pp-formula φ\varphi of sort GG, by Lemma 4.2. ∎

To prove Lemma 6.12 we use Corollary 6.11, a result of Garcia and Dung.

Corollary 6.11.

[7, Corollary 1.6] Let 𝒜\mathcal{A} be a Grothendieck category with a set of finitely generated generators. Any Σ\Sigma-injective object of 𝒜\mathcal{A} 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 𝙸={0}\mathtt{I}=\{0\} shows that Σ\Sigma-pure-injectives decompose into indecomposables.

Lemma 6.12.

If MM is a Σ\Sigma-pure-injective object of 𝒯\mathcal{T} then for any set 𝙸\mathtt{I} the product M𝙸M^{\mathtt{I}} is a coproduct of indecomposable Σ\Sigma-pure-injectives.

Proof.

Recall 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is a Grothendieck category with a set of finitely generated generators. Let K=M𝙸K=M^{\mathtt{I}} so that, by Corollary 6.10(1), KK is Σ\Sigma-pure-injective, and so 𝐘(K)\mathbf{Y}(K) is Σ\Sigma-injective by Corollary 6.9. By Corollary 6.11 this means 𝐘(K)j𝙹j\mathbf{Y}(K)\simeq\coprod_{j\in\mathtt{J}}\mathscr{L}_{j} where j\mathscr{L}_{j} is an indecomposable object of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} for each jj.

For each jj, there is a section uj:j𝐘(K)u_{j}\colon\mathscr{L}_{j}\to\mathbf{Y}(K), which means j\mathscr{L}_{j} is injective, and so by Corollary 6.7 we have j𝐘(Lj)\mathscr{L}_{j}\simeq\mathbf{Y}(L_{j}) for some pure-injective object LjL_{j}. Since j\mathscr{L}_{j} is indecomposable, we have that LjL_{j} is indecomposable. Again, applying Corollary 6.7 gives a section hj:LjKh_{j}\colon L_{j}\to K in 𝒯\mathcal{T} with 𝐘(hj)=uj\mathbf{Y}(h_{j})=u_{j}. Altogether this means hjh_{j} is a pure monomorphism into a Σ\Sigma-pure-injective object, and so LjL_{j} is Σ\Sigma-pure-injective by Corollary 6.10(2). By Remark 6.8 we have that 𝐘(jLj)𝐘(K)\mathbf{Y}(\coprod_{j}L_{j})\simeq\mathbf{Y}(K) is injective, and so jLjK\coprod_{j}L_{j}\simeq K 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 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-structure 𝖬\mathsf{M} underlying any object MM of 𝒯\mathcal{T} is defined and denoted |𝖬|=|G𝒮𝒯(G,M)||\mathsf{M}|=|\bigsqcup_{G\in\mathcal{S}}\mathcal{T}(G,M)| where 𝒮\mathcal{S} 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 κ\kappa such that the 𝔏𝒯\mathfrak{L}^{\mathcal{T}}-structure underlying any indecomposable pure-injective object of 𝒯\mathcal{T} has cardinality at most κ\kappa.

We delay the proof of Corollary 7.1 until after Corollary 7.3.

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 𝔏A\mathfrak{L}_{A} of modules over a fixed ring AA from Example 2.2, any 𝔏A\mathfrak{L}_{A}-structure is an object in the category of AA-modules, and any elementary embedding of 𝔏A\mathfrak{L}_{A}-structures is a pure embedding of AA-modules. The same correspondence between structures need not be true here. Although objects of 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} correspond to structures over 𝔏𝒯\mathfrak{L}^{\mathcal{T}} by Remark 3.6, the former need not be given by objects of 𝒯\mathcal{T} as 𝐘\mathbf{Y} 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 𝚂𝚙\mathtt{Sp} of isomorphism classes of objects in 𝒯\mathcal{T} which are pure-injective indecomposable.

Proof of Corollary 7.1.

It suffices to let κ=|𝒯(G,M)|\kappa=|\bigsqcup\mathcal{T}(G,M)| where GG runs through 𝒮\mathcal{S} and MM runs through the set 𝚂𝚙\mathtt{Sp} 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 𝒜\mathcal{A} be an abelian category, and let L={jj𝙹}\mathrm{L}=\{\mathscr{L}_{j}\mid j\in\mathtt{J}\} and N={𝒩kk𝙺}\mathrm{N}=\{\mathscr{N}_{k}\mid k\in\mathtt{K}\} be collections of objects in 𝒜\mathcal{A}. The following statements hold.

  1. (1)

    [20, p.82, Theorem 4.A7] Suppose the collections L\mathrm{L} and N\mathrm{N} consist of indecomposable objects with local endomorphism rings. If we have j𝙹jk𝙺𝒩k\coprod_{j\in\mathtt{J}}\mathscr{L}_{j}\simeq\coprod_{k\in\mathtt{K}}\mathscr{N}_{k} then there is a bijection σ:𝙹𝙺\sigma\colon\mathtt{J}\to\mathtt{K} with j𝒩σ(j)\mathscr{L}_{j}\simeq\mathscr{N}_{\sigma(j)} for all jj.

  2. (2)

    [20, p.82, Theorem 4.A11] Suppose that 𝙹=𝙺\mathtt{J}=\mathtt{K}, and that for all jj, j\mathscr{L}_{j} is indecomposable and j\mathscr{L}_{j} is the injective hull of 𝒩j\mathscr{N}_{j}. Suppose there is a section 𝒫j𝙹j\mathscr{P}\to\coprod_{j\in\mathtt{J}}\mathscr{L}_{j}. Then there is a subset 𝙷𝙹\mathtt{H}\subseteq\mathtt{J} with j𝙹j𝒫j𝙷j\coprod_{j\in\mathtt{J}}\mathscr{L}_{j}\simeq\mathscr{P}\amalg\coprod_{j\in\mathtt{H}}\mathscr{L}_{j}.

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 MM of 𝒯\mathcal{T} is indecomposable if and only if the endomorphism ring End𝒯(M)\mathrm{End}_{\mathcal{T}}(M) is local.

Corollary 7.6.

Let L={Ljj𝙹}\mathrm{L}=\{L_{j}\mid j\in\mathtt{J}\} be a collection of indecomposable objects in 𝒯\mathcal{T} such that the coproduct j𝙹Lj\coprod_{j\in\mathtt{J}}L_{j} is pure injective. If PP is a summand of j𝙹Lj\coprod_{j\in\mathtt{J}}L_{j} then there exists 𝙷𝙹\mathtt{H}\subseteq\mathtt{J} such that j𝙹LjPj𝙷Lj\coprod_{j\in\mathtt{J}}L_{j}\simeq P\amalg\coprod_{j\in\mathtt{H}}L_{j}.

If additionally Pk𝙺NkP\simeq\coprod_{k\in\mathtt{K}}N_{k} where each NkN_{k} is indecomposable, then there is a bijection σ:𝙹𝙺𝙷\sigma\colon\mathtt{J}\to\mathtt{K}\sqcup\mathtt{H} with LjNσ(j)L_{j}\simeq N_{\sigma(j)} for all j𝙹j\in\mathtt{J}.

Proof.

Let j=𝐘(Lj)\mathscr{L}_{j}=\mathbf{Y}(L_{j}) for each jj. Let L𝚃=j𝚃LjL_{\mathtt{T}}=\coprod_{j\in\mathtt{T}}L_{j} and 𝚃=j𝚃j\mathscr{L}_{\mathtt{T}}=\coprod_{j\in\mathtt{T}}\mathscr{L}_{j} for any subset 𝚃𝙹\mathtt{T}\subseteq\mathtt{J}. Since 𝐘\mathbf{Y} preserves small coproducts by Remark 6.8, and since each LjL_{j} is a summand of L𝙹L_{\mathtt{J}}, each j\mathscr{L}_{j} is a summand of 𝙹\mathscr{L}_{\mathtt{J}}. By Theorem 6.4 and Remark 6.8 𝙹\mathscr{L}_{\mathtt{J}} is an injective object. Thus each j\mathscr{L}_{j} is injective. Since each LjL_{j} is indecomposable, each j\mathscr{L}_{j} is indecomposable. Similarly, each NkN_{k} is pure-injective.

Let 𝒫=𝐘(P)\mathscr{P}=\mathbf{Y}(P). Since PP is a direct summand of L𝙹L_{\mathtt{J}} by assumption, 𝒫\mathscr{P} is a direct summand of 𝙹\mathscr{L}_{\mathtt{J}}. Hence by Theorem 7.4(2) there exists a subset 𝙷𝙹\mathtt{H}\subseteq\mathtt{J} such that 𝙹𝒫𝙷\mathscr{L}_{\mathtt{J}}\simeq\mathscr{P}\amalg\mathscr{L}_{\mathtt{H}}. By Corollary 6.7, since we assume L𝙹L_{\mathtt{J}} is pure injective, 𝙹\mathscr{L}_{\mathtt{J}} is injective, and hence so too are 𝒫\mathscr{P} and 𝙷\mathscr{L}_{\mathtt{H}}. Again, by Corollary 6.7 this gives L𝙹PL𝙷L_{\mathtt{J}}\simeq P\amalg L_{\mathtt{H}}.

Now suppose also Pk𝙺NkP\simeq\coprod_{k\in\mathtt{K}}N_{k} where each NkN_{k} is indecomposable and, as above, neccesarilly pure-injective. By Lemma 7.5 each of the objects LjL_{j} and NkN_{k} has a local endomorphism ring. By Corollary 6.7, for any pure-injective object ZZ of 𝒯\mathcal{T} there is a ring isomorphism

End𝒯(Z)End𝐌𝐨𝐝-𝒯c(𝐘(Z)).\mathrm{End}_{\mathcal{T}}(Z)\simeq\mathrm{End}_{\mathbf{Mod}\text{-}\mathcal{T}^{c}}(\mathbf{Y}(Z)).

Since 𝙹𝒫𝙷\mathscr{L}_{\mathtt{J}}\simeq\mathscr{P}\amalg\mathscr{L}_{\mathtt{H}}, the second claim follows, as above, by Theorem 7.4(1), using again that 𝐘\mathbf{Y} 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 M=iMi=j𝙹LjM=\prod_{i\in\mathbb{N}}M_{i}=\coprod_{j\in\mathtt{J}}L_{j}. Let XX be compact and let φ1(M)φ2(M)\varphi_{1}(M)\supseteq\varphi_{2}(M)\supseteq\dots be a descending chain of pp-definable subgroups of sort XX. Suppose additionally that each object LjL_{j} is both pure-injective and indecomposable.

Then there exists rr\in\mathbb{N} such that, for each collection N={Nii}\mathrm{N}=\{N_{i}\mid i\in\mathbb{N}\} where NiN_{i} is an indecomposable summand of MiM_{i}, we have φr(Ni)=φn(Ni)\varphi_{r}(N_{i})=\varphi_{n}(N_{i}) for all nn\in\mathbb{N} with nrn\geq r and all but finitely many ii\in\mathbb{N}.

Proof.

Recall the morphisms ρn,j=qj,Ψ(n)κφnu,Π(n)\rho_{n,j}=q_{j,\Psi(n)}\kappa\langle\varphi_{n}\rangle u_{\geq,\Pi(n)} defined by the composition

inφn(Mi)\textstyle{\prod_{i\geq n}\varphi_{n}(M_{i})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u,Π(n)\scriptstyle{u_{\geq,\Pi(n)}}iφn(Mi)\textstyle{\prod_{i\in\mathbb{N}}\varphi_{n}(M_{i})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κφn\scriptstyle{\kappa\langle\varphi_{n}\rangle}j𝙹φn(Lj)\textstyle{\coprod_{j\in\mathtt{J}}\varphi_{n}(L_{j})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qj,Ψ(n)\scriptstyle{q_{j,\Psi(n)}}φn(Lj).\textstyle{\varphi_{n}(L_{j}).}

Here the isomorphism κφn\kappa\langle\varphi_{n}\rangle is defined in Corollary 4.7, and the maps u,Π(n)u_{\geq,\Pi(n)} and qj,Ψ(n)q_{j,\Psi(n)} are the canonical inclusion and projection respectively. By Lemma 4.8 there exists rr\in\mathbb{N} and a finite subset 𝙹\mathtt{J}^{\prime} of 𝙹\mathtt{J} such that we have the containment im(ρr,j)φn(Lj)\mathrm{im}(\rho_{r,j})\subseteq\varphi_{n}(L_{j}) for all nn\in\mathbb{N} with nrn\geq r and all j𝙹𝙹j\in\mathtt{J}\setminus\mathtt{J}^{\prime}.

Choose arbitrary mm\in\mathbb{N} with m>|𝙹|m>|\mathtt{J}^{\prime}|, and choose arbitrary i1,,imi_{1},\dots,i_{m}\in\mathbb{N} with ipri_{p}\geq r for each pp. Let P=Ni1NimP=N_{i_{1}}\amalg\dots\amalg N_{i_{m}} so that PP is a summand of M=jLjM=\coprod_{j}L_{j}. By Corollary 7.6, since PP is a coproduct of mm indecomposable pure-injective objects, we have MPj𝙺LjM\simeq P\amalg\coprod_{j\in\mathtt{K}}L_{j} and a bijection σ:𝙹𝙺{i1,,im}\sigma\colon\mathtt{J}\to\mathtt{K}\sqcup\{i_{1},\dots,i_{m}\} with LjNσ(j)L_{j}\simeq N_{\sigma(j)} for all j𝙹j\in\mathtt{J}.

Since m>|𝙹|m>|\mathtt{J}^{\prime}| there exists p=1,,mp=1,\dots,m such that σ1(ip)𝙹\sigma^{-1}(i_{p})\notin\mathtt{J}^{\prime}. Let j=σ1(ip)j=\sigma^{-1}(i_{p}). It is straightforward to check that post composition with the isomorphism LjNσ(j)L_{j}\simeq N_{\sigma(j)} defines an isomorphism φn(Lj)φn(Nσ(j))\varphi_{n}(L_{j})\simeq\varphi_{n}(N_{\sigma(j)}) for each nn\in\mathbb{N}. Since Nσ(j)N_{\sigma(j)} is a summand of Mσ(j)M_{\sigma(j)} and ipri_{p}\geq r the group φn(Nσ(j))\varphi_{n}(N_{\sigma(j)}) embedds into irφr(Mi)\prod_{i\geq r}\varphi_{r}(M_{i}). Considering the image under ρr,j\rho_{r,j} the result follows by Lemma 4.8 as in the proof of [11, Lemma 5]. ∎

Corollary 7.8.

Let MM be a pure-injective object of 𝒯\mathcal{T} such that MIM^{I} is a coproduct of indecomposable pure-injectives for any set II. Then MM is Σ\Sigma-pure-injective.

Proof.

Let K=MK=M^{\mathbb{N}} and Mi=MM_{i}=M for each ii\in\mathbb{N}, so that K=iMiK=\prod_{i\in\mathbb{N}}M_{i}. By hypothesis we have K=j𝙹LjK=\coprod_{j\in\mathtt{J}}L_{j} where each LjL_{j} is an indecomposable pure-injective object of 𝒯\mathcal{T}. By Lemma 7.7 there exists rr\in\mathbb{N} such that, for each collection N={Nii}\mathrm{N}=\{N_{i}\mid i\in\mathbb{N}\} where NiN_{i} is an indecomposable summand of MiM_{i}, we have φr(Ni)=φn(Ni)\varphi_{r}(N_{i})=\varphi_{n}(N_{i}) for all nn\in\mathbb{N} with nrn\geq r and all but finitely many ii\in\mathbb{N}. Fixing j𝙹j\in\mathtt{J}, for the collection given by Ni=LjN_{i}=L_{j} for all ii, we have φr(Lj)=φn(Lj)\varphi_{r}(L_{j})=\varphi_{n}(L_{j}) for all nn\in\mathbb{N}, and φn(K)jφn(Lj)\varphi_{n}(K)\simeq\coprod_{j}\varphi_{n}(L_{j}) for all nn by Lemma 4.5. ∎

Theorem 7.9 goes back to a characterisation due to Faith [6, Proposition 3].

Theorem 7.9.

[9, Theorem 1] (see also [7, Lemma 1.1]). Let \mathscr{M} be an injective object in a Grothendieck category 𝒜\mathcal{A} which has a set 𝙶\mathtt{G} of finitely generated generators. Then the following statements are equivalent.

  1. (1)

    \mathscr{M} is Σ\Sigma-injective.

  2. (2)

    The countable coproduct ()\mathscr{M}^{(\mathbb{N})} is injective.

  3. (3)

    For each 𝒢𝙶\mathscr{G}\in\mathtt{G}, any ascending chain of \mathscr{M}-annihilator subobjects of 𝒢\mathscr{G} eventually stabilises.

Finally, we may now complete the proof of our main result.

of Theorem 1.1.

Taking I=I=\mathbb{N} 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 MM of a fixed sort XX define End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodules of 𝒯(X,M)\mathcal{T}(X,M). 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 MM of 𝒯\mathcal{T} is said to be endoperfect if for all compact objects XX of 𝒯\mathcal{T} the left End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-module 𝒯(X,M)\mathcal{T}(X,M) has the descending chain condition on cyclic submodules. That is, MM is endoperfect provided 𝒯(X,M)\mathcal{T}(X,M) is a perfect End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-module in the sense of [3, §1] for each compact object XX.

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 Σ\Sigma-pure-injective modules.

Definition 8.2.

An object MM of 𝒯\mathcal{T} is called endonoetherian (respectively, endoartinian) if 𝒯(X,M)\mathcal{T}(X,M) is noetherian (respectively, artinian) over End𝒯(M)\mathrm{End}_{\mathcal{T}}(M) for each compact object XX.

By using a result of Björk [3] we now see a direct application of Theorem 1.1.

Corollary 8.3.

Any endoperfect endonoetherian object of 𝒯\mathcal{T} is Σ\Sigma-pure-injective, and hence decomposes into a coproduct of indecomposable objects with local endomorphism rings.

Proof.

Let XX be a compact object of 𝒯\mathcal{T} and φ1(M)φ2(M)\varphi_{1}(M)\supseteq\varphi_{2}(M)\supseteq\dots be a descending chain of pp-definable subgroups of MM of sort XX. By Lemma 4.3 this defines a descending chain of End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodules of 𝒯(X,M)\mathcal{T}(X,M). Since MM is endonoetherian each submodule (of the form φn(M)\varphi_{n}(M)) is finitely generated. Since MM is endoperfect the module 𝒯(X,M)\mathcal{T}(X,M) has the descending chain condition on cyclic submodules. This means 𝒯(X,M)\mathcal{T}(X,M) is perfect (over End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)) in the sense of Björk [3, §1]. By [3, Theorem 2] this means that the descending chain of finitely generated submodules φ1(M)φ2(M)\varphi_{1}(M)\supseteq\varphi_{2}(M)\supseteq\dots of 𝒯(X,M)\mathcal{T}(X,M) must terminate.

By the equivalence of (1) and (5) in Theorem 1.1 this means MM is Σ\Sigma-pure-injective. That MM decomposes into indecomposables with local endomorphism rings follows by the equivalence of (1) and (6) in Theorem 1.1 (setting 𝙸={0}\mathtt{I}=\{0\}). ∎

Lemma 8.4 was motivated by a result due to Crawley-Boevey [4, §4, Lemma].

Lemma 8.4.

If MM is a pure-injective object of 𝒯\mathcal{T} and f:XMf\colon X\to M is a morphism in 𝒯\mathcal{T} with XX compact, then the cyclic End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodule End𝒯(M)f\mathrm{End}_{\mathcal{T}}(M)f of 𝒯(X,M)\mathcal{T}(X,M) is an intersection of pp-definable subgroups of MM of sort XX.

Proof.

We follow [4, §1.7]. Suppose 𝒦\mathscr{K} is a finitely generated subobject of ker(𝐘(f))\mathrm{ker}(\mathbf{Y}(f)), so that we have an exact sequence in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} of the form 𝐘(Z)𝒦0\mathbf{Y}(Z)\to\mathscr{K}\to 0 with ZZ compact. By construction the quotient 𝐘(X)/𝒦\mathbf{Y}(X)/\mathscr{K} is finitely presented, so there is a morphism k:ZXk\colon Z\to X such that 𝐘(k)\mathbf{Y}(k) is the composition 𝐘(Z)𝒦𝐘(X)\mathbf{Y}(Z)\to\mathscr{K}\subseteq\mathbf{Y}(X). Similarly it is straightforward to check that any morphism 𝐘(X)/𝒦𝐘(M)\mathbf{Y}(X)/\mathscr{K}\to\mathbf{Y}(M) is given by a morphism h:XMh\colon X\to M in 𝒯\mathcal{T} such that hk=0hk=0.

By [8, Proposition 3.1] this shows that the set of morphisms 𝐘(X)/𝒦𝐘(M)\mathbf{Y}(X)/\mathscr{K}\to\mathbf{Y}(M) is the set of 𝐘(h)\mathbf{Y}(h) where hh lies in a pp-definable subgroup of MM of sort XX. We now follow [4, §4, Lemma]. Since 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is a locally finitely generated Grothendieck category we have that ker(𝐘(f))\mathrm{ker}(\mathbf{Y}(f)) is the sum 𝒦\sum\mathscr{K} over the finitely generated subobjects 𝒦\mathscr{K} of ker(𝐘(f))\mathrm{ker}(\mathbf{Y}(f)). Hence the set of morphisms 𝐘(X)/ker(𝐘(f))𝐘(M)\mathbf{Y}(X)/\mathrm{ker}(\mathbf{Y}(f))\to\mathbf{Y}(M) is the intersection over each 𝒦\mathscr{K} of the set of morphisms 𝐘(X)/𝒦𝐘(M)\mathbf{Y}(X)/\mathscr{K}\to\mathbf{Y}(M). Altogether this shows the set of 𝐘(X)/ker(𝐘(f))𝐘(M)\mathbf{Y}(X)/\mathrm{ker}(\mathbf{Y}(f))\to\mathbf{Y}(M) in 𝐌𝐨𝐝-𝒯c\mathbf{Mod}\text{-}\mathcal{T}^{c} is given by the set of 𝐘(h)\mathbf{Y}(h) where hh lies in an intersection of pp-definable subgroup of MM of sort XX.

The pure-injectivity of MM in 𝒯\mathcal{T} is equivalent, by Corollary 6.7, to the injectivity of its image 𝐘(M)\mathbf{Y}(M). Thus, considering the canonical embedding 𝐘(X)/ker(𝐘(f))𝐘(M)\mathbf{Y}(X)/\mathrm{ker}(\mathbf{Y}(f))\to\mathbf{Y}(M) given by 𝐘(f)\mathbf{Y}(f), any morphism 𝐘(X)/ker(𝐘(f))𝐘(M)\mathbf{Y}(X)/\mathrm{ker}(\mathbf{Y}(f))\to\mathbf{Y}(M) factors through an endomorphism of 𝐘(M)\mathbf{Y}(M). Altogether this shows that the cyclic module End𝒯(M)f\mathrm{End}_{\mathcal{T}}(M)f is the intersection 𝒦φ𝒦(M)\bigcap_{\mathscr{K}}\varphi_{\mathscr{K}}(M) where each pp-formula φ𝒦\varphi_{\mathscr{K}} is given by φ𝒦(vG)=uZ:vG=uZk\varphi_{\mathscr{K}}(v_{G})=\exists u_{Z}\colon v_{G}=u_{Z}k in the above notation. ∎

Hence we may now provide partial a converse to Corollary 8.3.

Corollary 8.5.

Any Σ\Sigma-pure-injective object MM of 𝒯\mathcal{T} is endoperfect.

Proof.

By the equivalence of (1) and (5) in Theorem 1.1 we have that for each compact object XX any descending chain of pp-definable subgroups of MM of sort XX must stabilise. By Lemma 8.4 any cyclic End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodule of 𝒯(X,M)\mathcal{T}(X,M) is an interesction, and hence a finite intersection, of such subgroups. Thus any cyclic End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodule of 𝒯(X,M)\mathcal{T}(X,M) is a pp-definable subgroup of MM of sort XX. So, by definition, and by the equivalence of (1) and (5) in Theorem 1.1, any descending chain of cyclic End𝒯(M)\mathrm{End}_{\mathcal{T}}(M)-submodules stabilises. ∎

The author would be interested in finding, if it exists, an endoperfect object in a compactly generated triangulated category which is not Σ\Sigma-pure-injective. Of course, by Corollary 8.3 such an example would not be endonoetherian.

Definition 8.6.

[16, Definition 1.1] An object MM of 𝒯\mathcal{T} is called endofinite if 𝒯(X,M)\mathcal{T}(X,M) has finite length over End𝒯(M)\mathrm{End}_{\mathcal{T}}(M) for all compact objects XX.

To conclude, we recover a result of Krause. Note endofinite objects are endorartinian, endonoetherian, hence endoperfect and Σ\Sigma-pure-injective by Corollary 8.3.

Theorem 8.7.

[16, Theorem 1.2(1)] Every endofinite object of 𝒯\mathcal{T} 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 KK-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