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

Endomorphism algebras of silting complexes

Lidia Angeleri Hügel Universitá degli Studi di Verona, Strada Le Grazie 15, 37134 Verona, Italia [email protected] Marcelo Lanzilotta Departamento de Matemática, Facultad de Ingeniería, Universidad de la República, Uruguay [email protected] Jifen Liu Institute of Mathematics, School of Mathematical Sciences, Nanjing Normal University, Nanjing 210023, P. R. China. [email protected]  and  Sonia Trepode Centro Marplatense de Investigaciones Matemáticas. FCEyN, Universidad Nacional de Mar del Plata, CONICET. Dean Funes 3350, Mar del Plata, Argentina. [email protected]
Abstract.

We consider endomorphism algebras of nn-term silting complexes in derived categories of hereditary algebras, and we show that the module category of such an endomorphism algebra has a separated nn-section. For n=3n=3 we obtain a trisection in the sense of [2].

Key words and phrases:
nn-term silting; nn-section; endomorphism algebras; ring epimorphisms.
\ast: Corresponding author.

1. Introduction

In representation theory, algebras are studied in terms of their module category. The category of finite dimensional modules over a finite dimensional hereditary algebra is rather well understood, and it is often taken as a starting point for exploring more complex situations. For example, Happel and Ringel [19] studied tilted algebras, the endomorphism algebras of tilting modules over hereditary finite dimensional algebras. Their module categories can be completely described as a “tilt” of the module category of the underlying hereditary algebra. Later, Happel, Reiten and Smalø [18] extended these results to the class of quasi-tilted algebras, the algebras occurring as endomorphism algebras of tilting objects in hereditary abelian categories with finiteness conditions. They characterized quasi-tilted algebras homologically as the algebras of global dimension at most two such that each indecomposable module has projective or injective dimension at most one. This led Coelho and Lanzilotta [13] to investigate shod algebras, which are defined as the algebras satisfying the latter homological condition on indecomposable modules. Shod algebras always have global dimension at most three, the ones of global dimension three are called strictly shod.

In 2016, Buan and Zhou [12] showed that shod algebras admit a very natural characterization in terms of the notion of a silting complex introduced in [20]. They proved that the strictly shod algebras are precisely the silted algebras, that is, the endomorphism algebras of 2-term silting complexes in the bounded derived category of a hereditary finite dimensional algebra.

The purpose of this work is to investigate nn-silted algebras, i.e. endomorphism algebras of nn-term silting complexes in derived categories of hereditary algebras. We will prove that the module category of an nn-silted algebra has a separated nn-section. To this end, we will employ the fact that the module category of the endomorphism ring of a silting complex is equivalent to the heart of an associated t-structure. We will thus work in the heart of the t-structure induced by our nn-term silting complex. A crucial role will be played by the connection developed in [4] between silting complexes and chains of homological ring epimorphisms over hereditary algebras.

The case n=3n=3 is particularly nice. We start with a 3-term silting complex in the derived category of a hereditary algebra, and we prove that its endomorphism algebra has a separated trisection in its module category given by three functorially finite subcategories. Finally, we also prove that every functorially finite nn-section over a hereditary algebra, under mild conditions, is associated to an nn-term silting complex.

2. Preliminaries

2.1. Notation

Throughout this paper, let AA be a finite dimensional algebra over a field kk. We always assume that all modules are right modules. A composition gfg\circ f of morphisms ff and gg means first ff then gg. But a composition αβ\alpha\beta of arrows α\alpha and β\beta means that first α\alpha then β\beta. The category of all right AA-modules is denoted by ModA\mathrm{Mod}A, and the subcategory of finitely presented AA-modules is denoted by modA\mathrm{mod}A. For a given AA-module XX, we denote by pdX\mathrm{pd}X (resp. idX\mathrm{id}X) the projective (resp. injective) dimension of XX.

Let ModA\mathcal{M}\subset\mathrm{Mod}A be a class of modules. Add\mathrm{Add}\mathcal{M} (resp. add\mathrm{add}\mathcal{M}) denotes the class consisting of all modules isomorphic to direct summands of (finite) direct sums of elements in \mathcal{M}, while Gen\mathrm{Gen}\mathcal{M} (resp. gen\mathrm{gen}\mathcal{M}) is the class of epimorphic images of (finite) direct sums of elements in \mathcal{M}. Dually, we define Cogen\mathrm{Cogen}\mathcal{M} (resp. cogen\mathrm{cogen}\mathcal{M}) as the class of all submodules of (finite) direct sums of elements in \mathcal{M}.

Denote by indA\mathrm{ind}A the subcategory of modA\mathrm{mod}A formed by the indecomposable AA-modules. Given X,YindAX,Y\in\mathrm{ind}A, a path from XX to YY in indA\mathrm{ind}A is a sequence of non-zero morphisms X=X0X1Xt1Xt=Y(t1)X=X_{0}\rightarrow X_{1}\rightarrow\cdots\rightarrow X_{t-1}\rightarrow X_{t}=Y\ (t\geq 1), where XiindAX_{i}\in\mathrm{ind}A for all ii. We say that XX is a predecessor of YY and YY is a successor of XX.

Let 𝒞\mathcal{C} be a subcategory of indA\mathrm{ind}A. Recall that 𝒞\mathcal{C} is closed under predecessors if, whenever there is a path from XX to YY in indA\mathrm{ind}A, with Y𝒞Y\in\mathcal{C}, then X𝒞X\in\mathcal{C}. An example is the left part A\mathcal{L}_{A} of modA\mathrm{mod}A, defined in [18], which is a full subcategory of indA\mathrm{ind}A with object class

A={YindA:pdX1whenever there is a path fromXtoY}.\mathcal{L}_{A}=\left\{Y\in\mathrm{ind}A:\mathrm{pd}X\leq 1\ \text{whenever there is a path from}\ X\ \text{to}\ Y\right\}.

Dually, we define subcategories closed under successors. An example of such categories is the right part A\mathcal{R}_{A} of indA\mathrm{ind}A

A={XindA:idY1whenever there is a path fromXtoY}.\mathcal{R}_{A}=\left\{X\in\mathrm{ind}A:\mathrm{id}Y\leq 1\ \text{whenever there is a path from}\ X\ \text{to}\ Y\right\}.

2.2. Functorially finite subcategories

Let 𝒳\mathcal{X} be a subcategory of an additive category 𝒜\mathcal{A}. For an object MM in 𝒜\mathcal{A}, a right 𝒳\mathcal{X}-approximation of MM is a morphism f:XMf:X\longrightarrow M with X𝒳X\in\mathcal{X} such that any morphism f:XMf^{\prime}:X^{\prime}\longrightarrow M with X𝒳X^{\prime}\in\mathcal{X} factors through ff. If every object in 𝒜\mathcal{A} has a right 𝒳\mathcal{X}-approximation, we call 𝒳\mathcal{X} contravariantly finite in 𝒜\mathcal{A}. The notions of a left 𝒳\mathcal{X}-approximation and a covariantly finite subcategory are defined dually. We say 𝒳\mathcal{X} is functorially finite if it is both contravariantly finite and covariantly finite.

Proposition 2.1.

[9] Let (𝒯,)(\mathcal{T},\mathcal{F}) be a torsion pair in modA\mathrm{mod}A. The following are equivalent:

(1)(1) The torsion class 𝒯\mathcal{T} is functorially finite;

(2)(2) There exists MmodAM\in\mathrm{mod}A such that 𝒯=genM\mathcal{T}=\mathrm{gen}M;

(3)(3) The torsion-free class \mathcal{F} is functorially finite;

(4)(4) There exists NmodAN\in\mathrm{mod}A such that =cogenN\mathcal{F}=\mathrm{cogen}N.

A torsion pair fulfilling the equivalent properties in the above proposition is a functorially finite torsion pair.

2.3. Suspended subcategories and t-structures

Let 𝒯\mathcal{T} be a triangulated category with shift functor [1][1]. For a full subcategory 𝒱\mathcal{V} of 𝒯\mathcal{T} and a subset II\subseteq\mathbb{Z} (which is usually expressed by symbols such as n,n\geq n,\leq n, or just nn), we set

𝒱I={W𝒯|Hom𝒯(V,W[i])=0for anyiIandV𝒱},\mathcal{V}^{\bot_{I}}=\left\{W\in\mathcal{T}\ |\ \mathrm{Hom}_{\mathcal{T}}(V,W[i])=0\ \text{for any}\ i\in I\ \text{and}\ V\in\mathcal{V}\right\},
𝒱I={W𝒯|Hom𝒯(W,V[i])=0for anyiIandV𝒱}.{}^{\bot_{I}}\mathcal{V}=\left\{W\in\mathcal{T}\ |\ \mathrm{Hom}_{\mathcal{T}}(W,V[i])=0\ \text{for any}\ i\in I\ \text{and}\ V\in\mathcal{V}\right\}.

For instance,

𝒱0={W𝒯|Hom𝒯(V,W)=0for anyV𝒱},\mathcal{V}^{\bot_{0}}=\left\{W\in\mathcal{T}\ |\ \mathrm{Hom}_{\mathcal{T}}(V,W)=0\ \text{for any}\ V\in\mathcal{V}\right\},
𝒱0={W𝒯|Hom𝒯(W,V)=0for anyV𝒱}.{}^{\bot_{0}}\mathcal{V}=\left\{W\in\mathcal{T}\ |\ \mathrm{Hom}_{\mathcal{T}}(W,V)=0\ \text{for any}\ V\in\mathcal{V}\right\}.

A full subcategory 𝒱\mathcal{V} of 𝒯\mathcal{T}, closed under direct summands, is said to be suspended if it is closed under positive shifts and extensions, that is, if X𝒱X\in\mathcal{V}, then X[i]𝒱X[i]\in\mathcal{V} for all integers i>0i>0, and if XYZX[1]X\rightarrow Y\rightarrow Z\rightarrow X[1] is a triangle in 𝒯\mathcal{T} with X,Z𝒱X,Z\in\mathcal{V} then Y𝒱Y\in\mathcal{V}. Dually, one can define cosuspended subcategories.

Definition 2.2.

[10] A t-structure on 𝒯\mathcal{T} is a pair of full subcategories (𝒱,𝒲)(\mathcal{V},\mathcal{W}) closed under direct summands such that

(1)(1) Hom𝒯(𝒱,𝒲)=0\mathrm{Hom}_{\mathcal{T}}(\mathcal{V},\mathcal{W})=0, i.e. Hom𝒯(V,W)=0\mathrm{Hom}_{\mathcal{T}}(V,W)=0 for any V𝒱V\in\mathcal{V} and W𝒲W\in\mathcal{W};

(2)(2) 𝒱[1]𝒱\mathcal{V}[1]\subseteq\mathcal{V};

(3)(3) for any XX in 𝒯\mathcal{T}, there exist V𝒱V\in\mathcal{V}, W𝒲W\in\mathcal{W} and a triangle VXWV[1]V\rightarrow X\rightarrow W\rightarrow V[1].

A suspended subcategory 𝒱\mathcal{V} of 𝒯\mathcal{T} is called an aisle if the inclusion functor 𝒱𝒯\mathcal{V}\hookrightarrow\mathcal{T} has a right adjoint u:𝒯𝒱u:\mathcal{T}\rightarrow\mathcal{V}. Similarly, a cosuspended subcategory 𝒲\mathcal{W} of 𝒯\mathcal{T} is called a coaisle if the inclusion functor 𝒲𝒯\mathcal{W}\hookrightarrow\mathcal{T} has a left adjoint v:𝒯𝒲v:\mathcal{T}\rightarrow\mathcal{W}. It is shown in [20] that the following conditions are equivalent for a suspended subcategory 𝒱\mathcal{V} of 𝒯\mathcal{T}:

(i)(i) 𝒱\mathcal{V} is an aisle.

(ii)(ii) (𝒱,𝒱0)(\mathcal{V},\mathcal{V}^{\bot_{0}}) is a t-structure.

(iii)(iii) For any XX in 𝒯\mathcal{T}, there is a triangle VXWV[1]V\rightarrow X\rightarrow W\rightarrow V[1] with V𝒱V\in\mathcal{V} and W𝒱0W\in\mathcal{V}^{\bot_{0}}.

(iv)(iv) 𝒱\mathcal{V} is contravariantly finite in 𝒯\mathcal{T}.

Recall that the heart of the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) is the subcategory =𝒱𝒲[1]\mathcal{H}=\mathcal{V}\cap\mathcal{W}[1]. It is an abelian category by [10]. We denote by H0:𝒯H^{0}:\mathcal{T}\rightarrow\mathcal{H} the associated cohomological functor given by H0(X)=u(v(X)[1])H^{0}(X)=u(v(X)[1]).

The following Lemma will be useful later.

Lemma 2.3.

[17] Let (𝒱,𝒲)(\mathcal{V},\mathcal{W}) be a t-structure in 𝒯\mathcal{T} with heart =𝒱𝒲[1]\mathcal{H}=\mathcal{V}\cap\mathcal{W}[1]. Given a morphism f:XYf:X\rightarrow Y in \mathcal{H}, let ZZ be the cone of ff in 𝒯\mathcal{T}. Consider the canonical triangle

KZLK[1]K\rightarrow Z\rightarrow L\rightarrow K[1]

with K𝒱[1]K\in\mathcal{V}[1] and L𝒲[1]L\in\mathcal{W}[1]. Then Kerf=K[1]\mathrm{Ker}_{\mathcal{H}}f=K[-1] and Cokerf=L\mathrm{Coker}_{\mathcal{H}}f=L.

2.4. Silting complexes

Silting complexes were introduced in [20] to study t-structures in the derived category of a hereditary algebra.

Definition 2.4.

[24] Let D(A)=D(ModA)D(A)=D(\mathrm{Mod}A) be the unbounded derived category of ModA\mathrm{Mod}A. An object TT in D(A)D(A) is silting if the pair (T>0,T0)(T^{\bot_{>0}},T^{\bot_{\leq 0}}) is a t-structure in D(A)D(A), which we call the silting t-structure induced by TT.

Two silting objects T,TT,T^{\prime} in D(A)D(A) are equivalent if they induce the same t-structure.

It is shown in [5, Proposition 4.2] that a bounded complex of finitely generated projective AA-modules TT in Kb(projA)K^{b}(\mathrm{proj}A) is a silting object if and only if it satisfies the following conditions:

(S1)(S1) HomD(A)(T,T[i])=0\mathrm{Hom}_{D(A)}(T,T[i])=0 for all i>0i>0;

(S2)(S2) TT is a generator of D(A)D(A), i.e. T=0T^{\bot_{\mathbb{Z}}}=0.

We call TT a bounded silting complex.

An nn-term silting complex is a bounded silting complex with nn non-zero terms, which we always assume to be concentrated in degrees 0,,n+10,\dots,-n+1. If TT is a 22-term silting complex in Kb(projA)K^{b}(\mathrm{proj}A), then its cohomology in degree zero H0(T)H^{0}(T) is called a silting module [5]. Over a finite dimensional algebra, the finite dimensional silting modules are precisely the support τ\tau-tilting modules from [1]. Every silting module generates a torsion class, called silting class.

2.5. Ring epimorphisms

Definition 2.5.

A ring homomorphism λ:AB\lambda:A\rightarrow B is a ring epimorphism if it is an epimorphism in the category of rings with unit, or equivalently, if the functor given by restriction of scalars λ:ModBModA\lambda_{\ast}:\mathrm{Mod}B\rightarrow\mathrm{Mod}A is a full embedding.

A ring epimorphism λ:AB\lambda:A\rightarrow B is said to be

(i)(i) homological if ToriA(B,B)=0\mathrm{Tor}^{A}_{i}(B,B)=0 for all i>0i>0, or equivalently, the functor given by restriction of scalars λ:D(B)D(A)\lambda_{\ast}:D(B)\rightarrow D(A) is a full embedding.

(ii)(ii) pseudoflat if Tor1A(B,B)=0\mathrm{Tor}^{A}_{1}(B,B)=0.

Two ring epimorphisms λ1:AB1\lambda_{1}:A\rightarrow B_{1} and λ2:AB2\lambda_{2}:A\rightarrow B_{2} are equivalent if there is an isomorphism of rings μ:B1B2\mu:B_{1}\rightarrow B_{2} such that λ2=μλ1\lambda_{2}=\mu\circ\lambda_{1}. We say that λ1\lambda_{1} and λ2\lambda_{2} lie in the same epiclass of AA.

Epiclasses of a ring AA can be classified by suitable subcategories of ModA\mathrm{Mod}A.

Definition 2.6.

A full subcategory 𝒳\mathcal{X} of ModA\mathrm{Mod}A is called bireflective if the inclusion functor 𝒳ModA\mathcal{X}\rightarrow\mathrm{Mod}A admits both a left and a right adjoint, or equivalently, 𝒳\mathcal{X} is closed under products, coproducts, kernels and cokernels.

Theorem 2.7.

[16, 11] The assignment which takes a ring epimorphism λ:AB\lambda:A\rightarrow B to the essential image 𝒳B\mathcal{X}_{B} of λ\lambda_{\ast} defines a bijection between:

(1)(1) epiclasses of ring epimorphisms ABA\rightarrow B,

(2)(2) bireflective subcategories of ModA\mathrm{Mod}A,

which restricts to a bijection between

(1)(1)^{\prime} epiclasses of pseudoflat ring epimorphisms ABA\rightarrow B,

(2)(2)^{\prime} bireflective subcategories closed under extensions in ModA\mathrm{Mod}A.

In particular, if AA is a hereditary ring, then λ:AB\lambda:A\rightarrow B is a homological ring epimorphism if and only if it is pseudoflat, which is equivalent to being a universal localization of AA by [23, Theorem 6.1]. This shows that universal localization provides a powerful tool to construct homological ring epimorphisms for hereditary rings.

Theorem 2.8.

[26, Theorem 4.1] Let AA be a ring and Σ\Sigma be a class of morphisms between finitely generated projective right AA-modules. Then there is a pseudoflat ring epimorphism λ:AAΣ\lambda:A\rightarrow A_{\Sigma} called the universal localization of AA at Σ\Sigma such that

(1)(1) λ\lambda is Σ\Sigma-inverting, i.e. if σ\sigma belongs to Σ\Sigma, then σAAΣ\sigma\otimes_{A}A_{\Sigma} is an isomorphism of right AΣA_{\Sigma}-modules, and

(2)(2) λ\lambda is universal Σ\Sigma-inverting, i.e. for any Σ\Sigma-inverting morphism λ:AB\lambda^{\prime}:A\rightarrow B there exists a unique ring homomorphism g:AΣBg:A_{\Sigma}\rightarrow B such that gλ=λg\circ\lambda=\lambda^{\prime}.

Moreover, every pseudoflat ring epimorphism λ:AB\lambda:A\rightarrow B starting in a hereditary ring AA induces a silting module T=BCokerλT=B\oplus\mathrm{Coker}\lambda, see for example [6].

3. nn-term silting complexes over hereditary algebras

From now on, we assume that AA is a hereditary algebra. We want to exploit a result from [4] stating that bounded silting complexes are closely related to ring epimorphisms.

3.1. The t-structure induced by a silting complex

The partial order on bireflective subcategories given by inclusion corresponds, under the bijection in Theorem 2.7, to a partial order on the epiclasses of AA defined by setting λ1λ2\lambda_{1}\leq\lambda_{2} whenever λ1:AB1\lambda_{1}:A\longrightarrow B_{1} factors through λ2:AB2\lambda_{2}:A\longrightarrow B_{2} via a ring homomorphism μ:B2B1\mu:B_{2}\longrightarrow B_{1}, that is, λ1=μλ2\lambda_{1}=\mu\circ\lambda_{2}.

It is shown in [4, Section 5] that every chain

λn1λnλn+1\cdots\leq\lambda_{n-1}\leq\lambda_{n}\leq\lambda_{n+1}\leq\cdots

of homological ring epimorphisms λn:ABn\lambda_{n}:A\longrightarrow B_{n} induces a t-structure in D(A)D(A). More precisely, consider the corresponding bireflective subcategories 𝒳n\mathcal{X}_{n} of ModAA, which are all extension closed, together with the silting classes

𝒟n=Gen(𝒳n)\mathcal{D}_{n}=\mathrm{Gen}(\mathcal{X}_{n})

induced by the silting AA-modules Tn=BnCoker(λn)T_{n}=B_{n}\bigoplus\mathrm{Coker}(\lambda_{n}). We set

𝒱n=𝒟n𝒳n+1\mathcal{V}_{n}=\mathcal{D}_{n}\cap\mathcal{X}_{n+1}

for all nn\in\mathbb{Z}. Then there is a t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) in D(A)D(A) with aisle

𝒱={XD(A)|Hn(X)𝒱nfor alln}.\mathcal{V}=\left\{X\in D(A)\ |\ H^{-n}(X)\in\mathcal{V}_{n}\ \text{for all}\ n\in\mathbb{Z}\right\}.

The following proposition shows that the t-structure is induced by a silting complex under suitable hypotheses.

Proposition 3.1.

[4, Proposition 5.15] Let AA be a hereditary algebra and let λn1λnλn+1\cdots\leq\lambda_{n-1}\leq\lambda_{n}\leq\lambda_{n+1}\leq\cdots be a chain of homological ring epimorphisms λn:ABn\lambda_{n}:A\longrightarrow B_{n} with induced ring epimorphisms μn:Bn+1Bn\mu_{n}:B_{n+1}\longrightarrow B_{n} given by the commutative diagram

A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λn+1\scriptstyle{\lambda_{n+1}}Bn\textstyle{B_{n}}Bn+1\textstyle{B_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μn\scriptstyle{\mu_{n}}λn\scriptstyle{\lambda_{n}}

Let 𝒳n\mathcal{X}_{n} be the corresponding extension closed bireflective subcategories of ModAA and let 𝒦n=Ker𝐑HomA(Bn,)\mathcal{K}_{n}=\mathrm{Ker}\mathbf{R}Hom_{A}(B_{n},-) be the triangulated subcategories of D(A)D(A) associated with λn\lambda_{n}. Then the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) in D(A)D(A) is induced by a silting object if and only if the conditions

n𝒳n=0andn𝒦n=0\bigcap_{n\in\mathbb{Z}}\mathcal{X}_{n}=0\ and\ \bigcap_{n\in\mathbb{Z}}\mathcal{K}_{n}=0

hold true. In this case, the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) is induced by the silting object

T=nCone(μn)[n].T=\bigoplus_{n\in\mathbb{Z}}\mathrm{Cone}(\mu_{n})[n].

Every bounded silting complex over a hereditary algebra AA has this form, as explained in the following theorem.

Theorem 3.2.

[4, Theorem 6.7] Let AA be a finite dimensional hereditary algebra. Every bounded silting complex TT in Kb(projA)K^{b}(\mathrm{proj}A) arises as in Proposition 3.1 from a finite chain of finite dimensional homological ring epimorphisms 0AλnλmidA0_{A}\leq\lambda_{n}\leq\cdots\leq\lambda_{m}\leq id_{A}.

3.2. The heart of the t-structure

Let now TT be an nn-term silting complex in Kb(projA)K^{b}(\mathrm{proj}A). According to Theorem 3.2, TT arises from a chain of finite dimensional homological ring epimorphisms

0Aλ0λ1λn2idA0_{A}\leq\lambda_{0}\leq\lambda_{1}\leq\cdots\leq\lambda_{n-2}\leq id_{A}

with λi:ABi\lambda_{i}:A\longrightarrow B_{i} for integers 0in20\leq i\leq n-2. Consider the corresponding chains of extension-closed bireflective subcategories

0𝒳0𝒳1𝒳n2ModA0\subseteq\mathcal{X}_{0}\subseteq\mathcal{X}_{1}\subseteq\cdots\subseteq\mathcal{X}_{n-2}\subseteq\mathrm{Mod}A

and of silting classes

0𝒟0𝒟1𝒟n2ModA.0\subseteq\mathcal{D}_{0}\subseteq\mathcal{D}_{1}\subseteq\cdots\subseteq\mathcal{D}_{n-2}\subseteq\mathrm{Mod}A.

We construct a chain

0𝒱0𝒱1𝒱n2ModA0\subseteq\mathcal{V}_{0}\subseteq\mathcal{V}_{1}\subseteq\cdots\subseteq\mathcal{V}_{n-2}\subseteq\mathrm{Mod}A

with

𝒱i=𝒟i𝒳i+1=Gen(𝒳i)𝒳i+1=GenBiModBi+1fori=0,1,,n3\mathcal{V}_{i}=\mathcal{D}_{i}\cap\mathcal{X}_{i+1}=\mathrm{Gen}(\mathcal{X}_{i})\cap\mathcal{X}_{i+1}=\mathrm{Gen}B_{i}\cap\mathrm{Mod}B_{i+1}\ \text{for}\ i=0,1,\dots,n-3

and

𝒱n2=𝒟n2𝒳n1=𝒟n2ModA=Gen(𝒳n2)=GenBn2.\mathcal{V}_{n-2}=\mathcal{D}_{n-2}\cap\mathcal{X}_{n-1}=\mathcal{D}_{n-2}\cap\mathrm{Mod}A=\mathrm{Gen}(\mathcal{X}_{n-2})=\mathrm{Gen}B_{n-2}.

These classes induce a t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) in D(A)D(A), where the aisle

𝒱={XD(A)|Hj(X)𝒱jfor allj}\mathcal{V}=\left\{X\in D(A)\ |\ H^{-j}(X)\in\mathcal{V}_{j}\ \text{for all}\ j\in\mathbb{Z}\right\}

consists of the complexes XX with cohomologies concentrated in degrees 0\leq 0 satisfying Hi(X)GenBiModBi+1fori=0,1,,n3H^{-i}(X)\in\mathrm{Gen}B_{i}\cap\mathrm{Mod}B_{i+1}\ \text{for}\ i=0,1,\dots,n-3 and H(n2)(X)GenBn2H^{-(n-2)}(X)\in\mathrm{Gen}B_{n-2}.

Now we compute the coaisle 𝒲\mathcal{W}. Since AA is hereditary, we know from [4, Section 3.3] that 𝒲\mathcal{W} is determined by its cohomologies, that is,

𝒲={XD(A)|Hj(X)𝒲jfor allj}\mathcal{W}=\left\{X\in D(A)\ |\ H^{j}(X)\in\mathcal{W}_{j}\ \text{for all}\ j\in\mathbb{Z}\right\}

where 𝒲j=Hj(𝒲)\mathcal{W}_{j}=H^{j}(\mathcal{W}) satisfies 𝒲j=𝒱j0𝒱(j+1)1\mathcal{W}_{j}={\mathcal{V}_{-j}}^{\bot_{0}}\cap{\mathcal{V}_{-(j+1)}}^{\bot_{1}}.

Hence we have a chain

0𝒲(n2)𝒲1𝒲0ModA0\subseteq\mathcal{W}_{-(n-2)}\subseteq\cdots\subseteq\mathcal{W}_{-1}\subseteq\mathcal{W}_{0}\subseteq\mathrm{Mod}A

with

𝒲j=𝒱j0𝒱j11forj=1,2,,n2\mathcal{W}_{-j}={\mathcal{V}_{j}}^{\bot_{0}}\cap{\mathcal{V}_{j-1}}^{\bot_{1}}\ \text{for}\ j=1,2,\dots,n-2

and

𝒲0=𝒱00𝒱11=𝒱00ModA=𝒱00.\mathcal{W}_{0}={\mathcal{V}_{0}}^{\bot_{0}}\cap{\mathcal{V}_{-1}}^{\bot_{1}}={\mathcal{V}_{0}}^{\bot_{0}}\cap\mathrm{Mod}A={\mathcal{V}_{0}}^{\bot_{0}}.

We compute the heart of the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}). Since 𝒲[1]\mathcal{W}[1] consists of the complexes XX with Hj(X)𝒲j+1H^{j}(X)\in\mathcal{W}_{j+1} for all jj\in\mathbb{Z}, we have

=𝒱𝒲[1]={XD(A)|Hj(X)𝒱j𝒲j+1for allj}.\mathcal{H}=\mathcal{V}\cap\mathcal{W}[1]=\left\{X\in D(A)\ |\ H^{j}(X)\in\mathcal{V}_{-j}\cap\mathcal{W}_{j+1}\ \text{for all}\ j\in\mathbb{Z}\right\}.

Hence a complex XX lies in \mathcal{H} if and only if

H0(X)𝒱0𝒲1=𝒱0ModA=𝒱0,H^{0}(X)\in\mathcal{V}_{0}\cap\mathcal{W}_{1}=\mathcal{V}_{0}\cap\mathrm{Mod}A=\mathcal{V}_{0},
Hj(X)𝒱j𝒲(j1)forj=1,2,,n2,H^{-j}(X)\in\mathcal{V}_{j}\cap\mathcal{W}_{-(j-1)}\ \text{for}\ j=1,2,\dots,n-2,
H(n1)(X)𝒱n1𝒲(n2)=ModA𝒲(n2)=𝒲(n2),H^{-(n-1)}(X)\in\mathcal{V}_{n-1}\cap\mathcal{W}_{-(n-2)}=\mathrm{Mod}A\cap\mathcal{W}_{-(n-2)}=\mathcal{W}_{-(n-2)},
Hj(X)=0for allj{0,1,2,,(n1)}.H^{j}(X)=0\ \text{for all}\ j\in\mathbb{Z}\setminus\left\{{0,-1,-2,\dots,-(n-1)}\right\}.

The objects in \mathcal{H}, being directs sums of their cohomologies, decompose as direct sums of stalk complexes in 𝒱0[0],(𝒱j𝒲(j1))[j]\mathcal{V}_{0}[0],(\mathcal{V}_{j}\cap\mathcal{W}_{-(j-1)})[j] for j=1,2,,n2j=1,2,\dots,n-2 and 𝒲(n2)[n1]\mathcal{W}_{-(n-2)}[n-1]. We will denote these classes by j\mathcal{H}_{j}, 0j<n0\leq j<n. We have the following result.

Proposition 3.3.

Let AA be a finite dimensional hereditary algebra, and let TT be an nn-term silting complex in Kb(projA)K^{b}(\mathrm{proj}A). Then the heart of the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) in D(A)D(A) induced by TT has a decomposition =01n2n1\mathcal{H}=\mathcal{H}_{0}\vee\mathcal{H}_{1}\vee\dots\vee\mathcal{H}_{n-2}\vee\mathcal{H}_{n-1}.

3.3. The nn-section of the heart

We introduce the definition of an nn-section in indA\mathrm{ind}A.

Definition 3.4.

Let AA be an artin algebra. An nn-section (𝒜1,,𝒜n)(\mathcal{A}_{1},\cdots,\mathcal{A}_{n}) of indA\mathrm{ind}A is an nn-tuple of non-empty disjoint full subcategories of indA\mathrm{ind}A such that

(1)(1) indA=𝒜1𝒜n\mathrm{ind}A=\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n};

(2)(2) HomA(𝒜i,𝒜j)=0\mathrm{Hom}_{A}(\mathcal{A}_{i},\mathcal{A}_{j})=0 for 1j<in1\leq j<i\leq n.

In particular, for n=2n=2 we obtain a split torsion pair, and for n=3n=3 a trisection in the sense of [2, 18].

Motivated by [25, p.120], we say that an nn-section (𝒜1,,𝒜n)(\mathcal{A}_{1},\cdots,\mathcal{A}_{n}) is separated if, for any three adjacent subcategories (𝒜i1,𝒜i,𝒜i+1)(2in1)(\mathcal{A}_{i-1},\mathcal{A}_{i},\mathcal{A}_{i+1})\ (2\leq i\leq n-1) of indA\mathrm{ind}A, any morphism Ai1Ai+1A_{i-1}\longrightarrow A_{i+1} with Ai1𝒜i1A_{i-1}\in\mathcal{A}_{i-1} and Ai+1𝒜i+1A_{i+1}\in\mathcal{A}_{i+1} factors through add𝒜i\mathrm{add}\mathcal{A}_{i}. Moreover, an nn-section (𝒜1,,𝒜n)(\mathcal{A}_{1},\cdots,\mathcal{A}_{n}) is called functorially finite if every add𝒜i\mathrm{add}\mathcal{A}_{i} is a functorially finite subcategory of modA\mathrm{mod}A.

Let AA be a finite dimensional hereditary algebra, and let TT be an nn-term silting complex in Kb(projA)K^{b}(\mathrm{proj}A) with endomorphism ring B=EndD(A)TB=\mathrm{End}_{D(A)}T. Following [21, 24], the heart of the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) induced by TT is equivalent to the module category ModB\mathrm{Mod}B via the functor F=HomD(A)(T,)|:ModBF=\mathrm{Hom}_{D(A)}(T,-)\arrowvert_{\mathcal{H}}:\mathcal{H}\longrightarrow\mathrm{Mod}B. Moreover, the t-structure (𝒱,𝒲)(\mathcal{V},\mathcal{W}) restricts to a bounded t-structure (𝒱,𝒲)(\mathcal{V}^{\prime},\mathcal{W}^{\prime}) in Db(modA)D^{b}(\mathrm{mod}A) and FF restricts to an equivalence between the heart =Db(modA)\mathcal{H}^{\prime}=\mathcal{H}\cap D^{b}(\mathrm{mod}A) and modB\mathrm{mod}B. Put

0=F((𝒱0modA)[0]),1=F((𝒱1𝒲0modA)[1]),,\mathcal{B}_{0}=F((\mathcal{V}_{0}\cap\mathrm{mod}A)[0]),\ \mathcal{B}_{1}=F((\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)[1]),\dots,
n2=F((𝒱n2𝒲(n3)modA)[n2]),n1=F((𝒲(n2)modA)[n1]).\mathcal{B}_{n-2}=F((\mathcal{V}_{n-2}\cap\mathcal{W}_{-(n-3)}\cap\mathrm{mod}A)[n-2]),\mathcal{B}_{n-1}=F((\mathcal{W}_{-(n-2)}\cap\mathrm{mod}A)[n-1]).

Then \mathcal{H}^{\prime} is a Krull-Schmidt kk-category, even a length category, and

indB=ind0ind1indn1\mathrm{ind}B=\mathrm{ind}\mathcal{B}_{0}\cup\mathrm{ind}\mathcal{B}_{1}\cup\cdots\cup\mathrm{ind}\mathcal{B}_{n-1}
Proposition 3.5.

(1)(1) (ind0,ind1,,indn1)(\mathrm{ind}\mathcal{B}_{0},\mathrm{ind}\mathcal{B}_{1},\dots,\mathrm{ind}\mathcal{B}_{n-1}) is a separated nn-section of indB\mathrm{ind}B with ind0\mathrm{ind}\mathcal{B}_{0} closed under predecessors and indn1\mathrm{ind}\mathcal{B}_{n-1} closed under successors.

(2)(2) idXB1\mathrm{id}X_{B}\leq 1 for any object XX in n1\mathcal{B}_{n-1}.

(3)(3) pdXB1\mathrm{pd}X_{B}\leq 1 for any object XX in 0\mathcal{B}_{0}.

Proof..

(1)(1) We have seen that indB=ind0ind1indn1\mathrm{ind}B=\mathrm{ind}\mathcal{B}_{0}\cup\mathrm{ind}\mathcal{B}_{1}\cup\cdots\cup\mathrm{ind}\mathcal{B}_{n-1}, and by construction, there are only morphisms from the left to the right. It is obvious that ind0\mathrm{ind}\mathcal{B}_{0} is closed under predecessors and indn1\mathrm{ind}\mathcal{B}_{n-1} is closed under successors. On the other hand, HomD(A)(X,Y[2])=0\mathrm{Hom}_{D(A)}(X,Y[2])=0 for any X,YindAX,Y\in\mathrm{ind}A, thus there are no morphisms from indk\mathrm{ind}\mathcal{B}_{k} to indk+2\mathrm{ind}\mathcal{B}_{k+2} for k=0,1,,n3k=0,1,\dots,n-3.

(2)(2) Every object XX in n1\mathcal{B}_{n-1} is of the form X=F(W[n1])X=F(W[n-1]) for a module W𝒲(n2)W\in\mathcal{W}_{-(n-2)}. Consider the injective envelope e~:W[n1]E~\widetilde{e}:W[n-1]\rightarrow\widetilde{E} in \mathcal{H}. We have to show that C~=Cokere~\widetilde{C}=\mathrm{Coker}_{\mathcal{H}}\widetilde{e} is an injective object in \mathcal{H}.

Since n1\mathcal{B}_{n-1} is closed under successors, F(E~)F(\widetilde{E}) and F(C~)F(\widetilde{C}) are in n1\mathcal{B}_{n-1}, hence E~=E[n1]\widetilde{E}=E[n-1] and C~=C[n1]\widetilde{C}=C[n-1] with E,C𝒲(n2)E,C\in\mathcal{W}_{-(n-2)} and e~=e[n1]\widetilde{e}=e[n-1] for a morphism e:WEe:W\rightarrow E in ModA\mathrm{Mod}A. Moreover, notice that ExtA1(𝒲(n2),E)HomDb(A)(𝒲(n2),E[1])Ext1(𝒲(n2)[n1],E~)=0\mathrm{Ext}^{1}_{A}(\mathcal{W}_{-(n-2)},E)\cong\mathrm{Hom}_{D^{b}(A)}(\mathcal{W}_{-(n-2)},E[1])\cong\mathrm{Ext}^{1}_{\mathcal{H}}(\mathcal{W}_{-(n-2)}[n-1],\widetilde{E})=0 since E~\widetilde{E} is injective in \mathcal{H}, hence E(𝒲(n2))1E\in(\mathcal{W}_{-(n-2)})^{\bot_{1}}. Now we consider e~=e[n1]:W[n1]E~\widetilde{e}=e[n-1]:W[n-1]\rightarrow\widetilde{E} in D(A)D(A). Its cone is Z:=Cone(e~)=Ker(e)[n]Coker(e)[n1]Z:=\mathrm{Cone}(\widetilde{e})=\mathrm{Ker}(e)[n]\oplus\mathrm{Coker}(e)[n-1]. On the other hand, we know that Ker(e~)=0\mathrm{Ker}_{\mathcal{H}}(\widetilde{e})=0 and Coker(e~)=C~\mathrm{Coker}_{\mathcal{H}}(\widetilde{e})=\widetilde{C} in \mathcal{H}. By Lemma 2.3, we have

C~=Z𝒲[1]={XD(A)|H1(X)𝒲0,,H(n1)(X)𝒲(n2),Hk(X)=0kn},\widetilde{C}=Z\in\mathcal{W}[1]=\left\{X\in D(A)\ |\ H^{-1}(X)\in\mathcal{W}_{0},\dots,H^{-(n-1)}(X)\in\mathcal{W}_{-(n-2)},\right.\\ \left.\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ H^{k}(X)=0\ \forall\ k\leq-n\right\},

hence Ker(e)=0\mathrm{Ker}(e)=0 and Coker(e)=C𝒲(n2)\mathrm{Coker}(e)=C\in\mathcal{W}_{-(n-2)}. It follows that there is an exact sequence in ModA\mathrm{Mod}A

0WeEC0,0\rightarrow W\stackrel{{\scriptstyle e}}{{\rightarrow}}E\rightarrow C\rightarrow 0,

where E,C𝒲(n2)(𝒲(n2))1E,C\in\mathcal{W}_{-(n-2)}\cap(\mathcal{W}_{-(n-2)})^{\bot_{1}} (recall that (𝒲(n2))1(\mathcal{W}_{-(n-2)})^{\bot_{1}} is closed under quotients since 𝒲(n2)ModA\mathcal{W}_{-(n-2)}\subset\mathrm{Mod}A consists of modules of projective dimension at most one). This shows that C~\widetilde{C} is injective in \mathcal{H}. Indeed, if we take its injective envelope and consider the exact sequence

0C~E~C~00\rightarrow\widetilde{C}\rightarrow\widetilde{E^{\prime}}\rightarrow\widetilde{C^{\prime}}\rightarrow 0

in \mathcal{H}, using the same argument as above, we see that it in fact comes from an exact sequence 0CEC00\rightarrow C\rightarrow E^{\prime}\rightarrow C^{\prime}\rightarrow 0 in ModA\mathrm{Mod}A with E,C𝒲(n2)(𝒲(n2))1E^{\prime},C^{\prime}\in\mathcal{W}_{-(n-2)}\cap(\mathcal{W}_{-(n-2)})^{\bot_{1}}, which must split because C𝒲(n2)C^{\prime}\in\mathcal{W}_{-(n-2)} and C(𝒲(n2))1C\in(\mathcal{W}_{-(n-2)})^{\bot_{1}} imply ExtA1(C,C)=0\mathrm{Ext}^{1}_{A}(C^{\prime},C)=0.

(3)(3) is shown by using the dual arguments of the proof of (2)(2). ∎

Theorem 3.6.

Let TT be an nn-term silting complex in Db(modA)D^{b}(\mathrm{mod}A) over a finite dimensional hereditary algebra AA and let B=EndDb(modA)TB=\mathrm{End}_{D^{b}(\mathrm{mod}A)}T be the endomorphism ring of TT. Then indB\mathrm{ind}B has a separated nn-section (ind0,ind1,,indn1)(\mathrm{ind}\mathcal{B}_{0},\mathrm{ind}\mathcal{B}_{1},\dots,\mathrm{ind}\mathcal{B}_{n-1}) with ind0B\mathrm{ind}\mathcal{B}_{0}\subset\mathcal{L}_{B} and indn1B\mathrm{ind}\mathcal{B}_{n-1}\subset\mathcal{R}_{B}.

Proof..

By Proposition 3.5, we know that ind0\mathrm{ind}\mathcal{B}_{0} is closed under predecessors and the projective dimension is at most 1. Hence we have that ind0B\mathrm{ind}\mathcal{B}_{0}\subset\mathcal{L}_{B}. In the same way, since indn1\mathrm{ind}\mathcal{B}_{n-1} is closed under successors and the injective dimension is at most 1, it follows that indn1B\mathrm{ind}\mathcal{B}_{n-1}\subset\mathcal{R}_{B}. ∎

Lemma 3.7.

(Gen𝒱n,Cogen𝒲n)(\mathrm{Gen}\mathcal{V}_{n},\mathrm{Cogen}\mathcal{W}_{-n}) is a silting torsion pair in ModA\mathrm{Mod}A with silting module Tn=BnCokerλnT_{n}=B_{n}\oplus\mathrm{Coker}\lambda_{n}.

Proof..

𝒟n=Gen𝒱n=Cogen0𝒲n=(𝒱n0)0\mathcal{D}_{n}=\mathrm{Gen}\mathcal{V}_{n}={{}^{\bot_{0}}}\mathrm{Cogen}\mathcal{W}_{-n}={{}^{\bot_{0}}}({\mathcal{V}_{n}}^{\bot_{0}}) by [4, Theorem 5.9 and 6.7]. ∎

Proposition 3.8.

In the nn-section (ind0,ind1,,indn1)(\mathrm{ind}\mathcal{B}_{0},\mathrm{ind}\mathcal{B}_{1},\dots,\mathrm{ind}\mathcal{B}_{n-1}), the category add0\mathrm{add}\mathcal{B}_{0} is covariantly finite and addn1\mathrm{add}\mathcal{B}_{n-1} is contravariantly finite in modB\mathrm{mod}B.

Proof..

This follows from the fact that (add(1n1),add0)(\mathrm{add}(\mathcal{B}_{1}\cup\cdots\cup\mathcal{B}_{n-1}),\mathrm{add}\mathcal{B}_{0}) and (addn1,add(0n2))(\mathrm{add}\mathcal{B}_{n-1},\mathrm{add}(\mathcal{B}_{0}\cup\cdots\cup\mathcal{B}_{n-2})) are split torsion pairs in modB\mathrm{mod}B. ∎

Theorem 3.9.

When n=3n=3, the trisection (ind0,ind1,ind2)(\mathrm{ind}\mathcal{B}_{0},\mathrm{ind}\mathcal{B}_{1},\mathrm{ind}\mathcal{B}_{2}) in Theorem 3.6 is functorially finite.

Proof..

For n=3n=3, we know that

0=F((𝒱0modA)[0]),1=F((𝒱1𝒲0modA)[1]),2=F((𝒲1modA)[2]).\mathcal{B}_{0}=F((\mathcal{V}_{0}\cap\mathrm{mod}A)[0]),\ \mathcal{B}_{1}=F((\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)[1]),\ \mathcal{B}_{2}=F((\mathcal{W}_{-1}\cap\mathrm{mod}A)[2]).

Note that 𝒟0=Gen𝒱0𝒟1=𝒱1andCogen𝒲1Cogen𝒲0=𝒲0\mathcal{D}_{0}=\mathrm{Gen}\mathcal{V}_{0}\subset\mathcal{D}_{1}=\mathcal{V}_{1}\ \text{and}\ \mathrm{Cogen}\mathcal{W}_{-1}\subset\mathrm{Cogen}\mathcal{W}_{0}=\mathcal{W}_{0}. By Lemma 3.7, we have two torsion pairs in ModA\mathrm{Mod}A

t0=(Gen𝒱0,𝒲0),t1=(𝒱1,Cogen𝒲1)t_{0}=(\mathrm{Gen}\mathcal{V}_{0},\mathcal{W}_{0}),t_{1}=(\mathcal{V}_{1},\mathrm{Cogen}\mathcal{W}_{-1})

We first prove that add1\mathrm{add}\mathcal{B}_{1} is covariantly finite in modB\mathrm{mod}B.

Step 1: Every MmodAM\in\mathrm{mod}A admits a right (𝒱1𝒲0modA)(\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)-approximation.

For any MmodAM\in\mathrm{mod}A, take f:WMf:W\rightarrow M such that ff is a right (𝒲0modA)(\mathcal{W}_{0}\cap\mathrm{mod}A)-approximation of MM, which is guaranteed by the functorially finite torsion pair (genT0,𝒲0modA)(\mathrm{gen}T_{0},\mathcal{W}_{0}\cap\mathrm{mod}A) in modA\mathrm{mod}A induced by the minimal silting AA-module T0T_{0}. Let tr(W)\mathrm{tr}(W) be the trace of 𝒱1\mathcal{V}_{1} in WW, that is, the sum of the images of all homomorphism from modules in 𝒱1\mathcal{V}_{1} to WW. Because 𝒱1\mathcal{V}_{1} is closed under images and direct (hence arbitrary) sums, tr(W)\mathrm{tr}(W) is the largest submodule of WW that lies in 𝒱1\mathcal{V}_{1}. Denote the inclusion by ι:tr(W)W\iota:\mathrm{tr}(W)\rightarrow W. On the other hand, 𝒲0modA\mathcal{W}_{0}\cap\mathrm{mod}A is a torsion-free class, which is closed under submodules. Hence tr(W)\mathrm{tr}(W) lies in 𝒱1𝒲0modA\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A. It is easy to check that h:=ιfh:=\iota\circ f is a right (𝒱1𝒲0modA)(\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)-approximation of MM.

Step 2: Let {Vi|iI}\left\{V_{i}\ |\ i\in I\right\} be a complete irredundant set of representatives of the isomorphism classes of 𝒱1𝒲0modA\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A, and put V=iIViV=\bigoplus_{i\in I}V_{i}. It is easy to see that V𝒱1𝒲0V\in\mathcal{V}_{1}\cap\mathcal{W}_{0} since the torsion class 𝒱1\mathcal{V}_{1} is closed under direct sums and the torsion-free class 𝒲0\mathcal{W}_{0} is closed under direct products and submodules. By Step 1, any module MmodAM\in\mathrm{mod}A has an addV\mathrm{add}V-precover, which means that HomA(V,M)EndV\mathrm{Hom}_{A}(V,M)_{\mathrm{End}V} is finitely generated as right EndV\mathrm{End}V-module by [3, Lemma 3].

Step 3: For any module XmodAX\in\mathrm{mod}A, there is an EndAV\mathrm{End}_{A}V-linear isomorphism ExtA1EndAV(X,V)ϕDEndAVHomA(V,τX){{}_{\mathrm{End}_{A}V}}\mathrm{Ext}^{1}_{A}(X,V)\stackrel{{\scriptstyle\phi}}{{\cong}}{{}_{\mathrm{End}_{A}V}}D\mathrm{Hom}_{A}(V,\tau X).

Obviously, HomA(V,τX)\mathrm{Hom}_{A}(V,\tau X) is naturally endowed with a structure of right EndAV\mathrm{End}_{A}V-module, by defining (fs)(v):=f(s(v))(fs)(v):=f(s(v)) for any fHomA(V,τX),sEndAVf\in\mathrm{Hom}_{A}(V,\tau X),\ s\in\mathrm{End}_{A}V and vVv\in V. Hence DHomA(V,τX)D\mathrm{Hom}_{A}(V,\tau X) has a left EndAV\mathrm{End}_{A}V-module structure via the formula (sα)f:=α(fs)(s\alpha)f:=\alpha(fs) for any αDHomA(V,τX),sEndAV\alpha\in D\mathrm{Hom}_{A}(V,\tau X),\ s\in\mathrm{End}_{A}V and fHomA(V,τX)f\in\mathrm{Hom}_{A}(V,\tau X).

ExtA1(X,V)\mathrm{Ext}^{1}_{A}(X,V) is also a left EndAV\mathrm{End}_{A}V-module via the map EndAV×ExtA1(X,V)ExtA1(X,V)\mathrm{End}_{A}V\times\mathrm{Ext}^{1}_{A}(X,V)\longrightarrow\mathrm{Ext}^{1}_{A}(X,V), (s,[ε])[sε](s,[\varepsilon])\longmapsto[s\cdot\varepsilon], see, for example, [8, I, §5\S 5].

We have an isomorphism ExtA1(X,V)DHomA(V,τX)\mathrm{Ext}^{1}_{A}(X,V)\cong D\mathrm{Hom}_{A}(V,\tau X), which is functorial in both variables as AA-modules, see [22]. Therefore, for sEndAVs\in\mathrm{End}_{A}V, we have the following commutative diagram:

ExtA1(X,V)\textstyle{\mathrm{Ext}^{1}_{A}(X,V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕX,V\scriptstyle{\phi_{X,V}}ExtA1(X,s)\scriptstyle{\mathrm{Ext}^{1}_{A}(X,s)}DHomA(V,τX)\textstyle{D\mathrm{Hom}_{A}(V,\tau X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DHomA(s,τX)\scriptstyle{D\mathrm{Hom}_{A}(s,\tau X)}ExtA1(X,V)\textstyle{\mathrm{Ext}^{1}_{A}(X,V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕX,V\scriptstyle{\phi_{X,V}}DHomA(V,τX)\textstyle{D\mathrm{Hom}_{A}(V,\tau X)}

For any [ε]ExtA1(X,V)[\varepsilon]\in\mathrm{Ext}^{1}_{A}(X,V), we get ϕX,V(s[ε])=DHomA(s,τX)(ϕX,V([ε]))=ϕX,V([ε])HomA(s,τX)=sϕX,V([ε])\phi_{X,V}(s\cdot[\varepsilon])=D\mathrm{Hom}_{A}(s,\tau X)(\phi_{X,V}([\varepsilon]))=\phi_{X,V}([\varepsilon])\circ\mathrm{Hom}_{A}(s,\tau X)=s\cdot\phi_{X,V}([\varepsilon]). It follows that ϕX,V\phi_{X,V} is EndAV\mathrm{End}_{A}V-linear.

Step 4: Note that there are no morphisms from right to the left in the trisection (ind0,ind1,ind2)(\mathrm{ind}\mathcal{B}_{0},\mathrm{ind}\mathcal{B}_{1},\mathrm{ind}\mathcal{B}_{2}). It suffices to show that there exists a left add1\mathrm{add}\mathcal{B}_{1}-approximation for any module in ind0\mathrm{ind}\mathcal{B}_{0}.

Consider a module B0ind0B_{0}\in\mathrm{ind}\mathcal{B}_{0}. There exists X𝒱0modAX\in\mathcal{V}_{0}\cap\mathrm{mod}A such that B0=FXB_{0}=FX. It follows from Step 2 that HomA(V,τX)EndV\mathrm{Hom}_{A}(V,\tau X)_{\mathrm{End}V} is finitely generated. Thus DEndAVHomA(V,τX){{}_{\mathrm{End}_{A}V}}D\mathrm{Hom}_{A}(V,\tau X) is finitely generated. By Step 3, we have isomorphisms

DEndAVHomA(V,τX)ExtA1EndAV(X,V)HomDb(modA)EndAV(X,V[1]){{}_{\mathrm{End}_{A}V}}D\mathrm{Hom}_{A}(V,\tau X)\cong{{}_{\mathrm{End}_{A}V}}\mathrm{Ext}^{1}_{A}(X,V)\cong{{}_{\mathrm{End}_{A}V}}\mathrm{Hom}_{D^{b}(\mathrm{mod}A)}(X,V[1])
HomEndAV[1](X,V[1])HomBEndBFY(B0,FY)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \cong{{}_{\mathrm{End}_{A}V[1]}}\mathrm{Hom}_{\mathcal{H}^{\prime}}(X,V[1])\cong{{}_{\mathrm{End}_{B}FY}}\mathrm{Hom}_{B}(B_{0},FY)

where Y:=V[1]=iIVi[1]Y:=V[1]=\bigoplus_{i\in I}V_{i}[1] and {F(Vi[1])|iI}\left\{F(V_{i}[1])\ |\ i\in I\right\} is a complete irredundant set of representatives of the isomorphism classes of 1\mathcal{B}_{1}. It follows that the left EndBFY\mathrm{End}_{B}FY-module HomBEndBFY(B0,FY){{}_{\mathrm{End}_{B}FY}}\mathrm{Hom}_{B}(B_{0},FY) is finitely generated. Consequently, B0B_{0} has a left add1\mathrm{add}\mathcal{B}_{1}-approximation by [3, Lemma 3], as desired.

Now we show that add1\mathrm{add}\mathcal{B}_{1} is contravariantly finite in modB\mathrm{mod}B.

Step 11^{\prime}: Every MmodAM\in\mathrm{mod}A admits a left (𝒱1𝒲0modA)(\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)-approximation.

For any MmodAM\in\mathrm{mod}A, since 𝒱1modA\mathcal{V}_{1}\cap\mathrm{mod}A is functorially finite, there exists f:MVf:M\longrightarrow V with V𝒱1V\in\mathcal{V}_{1} such that ff is a left 𝒱1modA\mathcal{V}_{1}\cap\mathrm{mod}A-approximation of MM. Let tr(V)tr(V) be the trace of Gen𝒱0\mathrm{Gen}\mathcal{V}_{0} in VV and π:VV/tr(V)\pi:V\longrightarrow V/tr(V) be the canonical epimorphism. Obviously, V/tr(V)(𝒱1𝒲0modA)V/tr(V)\in(\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A) since 𝒱1\mathcal{V}_{1} is closed under quotients. It is easy to check that πf:MV/tr(V)\pi f:M\longrightarrow V/tr(V) is a left (𝒱1𝒲0modA)(\mathcal{V}_{1}\cap\mathcal{W}_{0}\cap\mathrm{mod}A)-approximation of MM.

Step 22^{\prime}: Let VV be the module as in the Step 2. By Step 11^{\prime}, any module MmodAM\in\mathrm{mod}A has an addV\mathrm{add}V-preenvelope, which means that HomAEndV(M,V){{}_{\mathrm{End}V}}\mathrm{Hom}_{A}(M,V) is finitely generated as left EndV\mathrm{End}V-module by [3, Lemma 3].

Step 33^{\prime}: For any module B2ind2B_{2}\in\mathrm{ind}\mathcal{B}_{2}, there exists X(𝒲1modA)X\in(\mathcal{W}_{-1}\cap\mathrm{mod}A) such that B2=F(X[2])B_{2}=F(X[2]). It follows from Step 22^{\prime} that HomAEndV(τ1X,V){{}_{\mathrm{End}V}}\mathrm{Hom}_{A}(\tau^{-1}X,V) is finitely generated. Thus DHomA(τ1X,V)EndAVD\mathrm{Hom}_{A}(\tau^{-1}X,V)_{\mathrm{End}_{A}V} is finitely generated. Moreover, we have isomorphisms

DHomA(τ1X,V)EndAVExtA1(V,X)EndAVHomDb(modA)(V[1],X[2])EndAV[1]D\mathrm{Hom}_{A}(\tau^{-1}X,V)_{\mathrm{End}_{A}V}\cong\mathrm{Ext}^{1}_{A}(V,X)_{\mathrm{End}_{A}V}\cong\mathrm{Hom}_{D^{b}(\mathrm{mod}A)}(V[1],X[2])_{\mathrm{End}_{A}V[1]}
Hom(V[1],X[2])EndAV[1]HomB(FY,B2)EndBFY\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \cong\mathrm{Hom}_{\mathcal{H}^{\prime}}(V[1],X[2])_{\mathrm{End}_{A}V[1]}\cong\mathrm{Hom}_{B}(FY,B_{2})_{\mathrm{End}_{B}FY}

where Y:=V[1]=iIVi[1]Y:=V[1]=\bigoplus_{i\in I}V_{i}[1] and {F(Vi[1])|iI}\left\{F(V_{i}[1])\ |\ i\in I\right\} is a complete irredundant set of representatives of the isomorphism classes of 1\mathcal{B}_{1}. It follows that the right EndBFY\mathrm{End}_{B}FY-module HomB(FY,B2)EndBFY\mathrm{Hom}_{B}(FY,B_{2})_{\mathrm{End}_{B}FY} is finitely generated. Consequently, B2B_{2} has a right add1\mathrm{add}\mathcal{B}_{1}-approximation by [3, Lemma 3], as desired.

Therefore, add1\mathrm{add}\mathcal{B}_{1} is a functorially finite subcategory of modB\mathrm{mod}B. The assertion follows from Proposition 3.8 and [2, Sec. 2.6 Theorem]. ∎

Finally, we note that every functorially finite nn-section over a hereditary algebra, under mild conditions, is associated to an nn-term silting complex.

Proposition 3.10.

If AA is a hereditary algebra with a functorially finite nn-section (𝒜1,𝒜2,,𝒜n)(\mathcal{A}_{1},\mathcal{A}_{2},\dots,\mathcal{A}_{n}) in modA\mathrm{mod}A and ExtA1(𝒜i,𝒜j)=0\mathrm{Ext}_{A}^{1}(\mathcal{A}_{i},\mathcal{A}_{j})=0 for ij2i-j\geq 2, then there exists a chain of finite dimensional homological ring epimorphisms λi:ABi\lambda_{i}:A\rightarrow B_{i} such that 0Aλ1λn1idA0_{A}\leq\lambda_{1}\leq\cdots\leq\lambda_{n-1}\leq id_{A} and genB1=add𝒜n,genB2=add(𝒜n1𝒜n),,genBn1=add(𝒜2𝒜n)\mathrm{gen}B_{1}=\mathrm{add}\mathcal{A}_{n},\mathrm{gen}B_{2}=\mathrm{add}(\mathcal{A}_{n-1}\cup\mathcal{A}_{n}),\ldots,\mathrm{gen}B_{n-1}=\mathrm{add}(\mathcal{A}_{2}\cup\cdots\cup\mathcal{A}_{n}).

Proof..

Since every add𝒜i\mathrm{add}\mathcal{A}_{i} is functorially finite, there are n1n-1 split torsion pairs with functorially finite torsion classes

(add𝒜n,add(𝒜1𝒜n1)),(add(𝒜n1𝒜n),add(𝒜1𝒜n2)),,(\mathrm{add}\mathcal{A}_{n},\mathrm{add}(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-1})),(\mathrm{add}(\mathcal{A}_{n-1}\cup\mathcal{A}_{n}),\mathrm{add}(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-2})),\ldots,
(add(𝒜3𝒜n),add(𝒜1𝒜2)),(add(𝒜2𝒜n),add𝒜1)(\mathrm{add}(\mathcal{A}_{3}\cup\cdots\cup\mathcal{A}_{n}),\mathrm{add}(\mathcal{A}_{1}\cup\mathcal{A}_{2})),(\mathrm{add}(\mathcal{A}_{2}\cup\cdots\cup\mathcal{A}_{n}),\mathrm{add}\mathcal{A}_{1})

in modA\mathrm{mod}A. By [1, Theorem 2.7], there exist n1n-1 finite dimensional silting modules TiT_{i} such that

genT1=add𝒜n,genT2=add(𝒜n1𝒜n),,genTn2=add(𝒜3𝒜n),\mathrm{gen}T_{1}=\mathrm{add}\mathcal{A}_{n},\mathrm{gen}T_{2}=\mathrm{add}(\mathcal{A}_{n-1}\cup\mathcal{A}_{n}),\ldots,\mathrm{gen}T_{n-2}=\mathrm{add}(\mathcal{A}_{3}\cup\cdots\cup\mathcal{A}_{n}),
genTn1=add(𝒜2𝒜n).\mathrm{gen}T_{n-1}=\mathrm{add}(\mathcal{A}_{2}\cup\cdots\cup\mathcal{A}_{n}).

It follows from [15, (4.4) Lemma] that there are torsion pairs in ModA\mathrm{Mod}A

(GenT1=lim𝒜n,lim(𝒜1𝒜n1)),,(GenTn1=lim(𝒜2𝒜n),lim𝒜1).(\mathrm{Gen}T_{1}=\varinjlim\mathcal{A}_{n},\varinjlim(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-1})),\ldots,(\mathrm{Gen}T_{n-1}=\varinjlim(\mathcal{A}_{2}\cup\cdots\cup\mathcal{A}_{n}),\varinjlim\mathcal{A}_{1}).

By [6, Corollary 5.12], every TiT_{i} corresponds to a homological ring epimorphism λi:ABi\lambda_{i}:A\longrightarrow B_{i} for 1in11\leq i\leq n-1 and then we have that ModBi=α(GenTi):={XGenTi|(g:YX)GenTi,KergGenTi}\mathrm{Mod}B_{i}=\alpha(\mathrm{Gen}T_{i}):=\left\{X\in\mathrm{Gen}T_{i}\ |\ \forall(g:Y\rightarrow X)\in\mathrm{Gen}T_{i},\mathrm{Ker}g\in\mathrm{Gen}T_{i}\right\} by [6, Proposition 5.6].

We need to show that λiλi+1\lambda_{i}\leq\lambda_{i+1}, which is equivalent to α(GenTi)α(GenTi+1)\alpha(\mathrm{Gen}T_{i})\subseteq\alpha(\mathrm{Gen}T_{i+1}). So it suffices to show that GenTiα(GenTi+1)\mathrm{Gen}T_{i}\subseteq\alpha(\mathrm{Gen}T_{i+1}). For any XGenTiX\in\mathrm{Gen}T_{i}, consider g:YXg:Y\rightarrow X in GenTi+1\mathrm{Gen}T_{i+1}. We want to prove K:=KergGenTi+1K:=\mathrm{Ker}g\in\mathrm{Gen}T_{i+1}. In fact, by [27, Proposition 2.14],

GenTi=lim(𝒜ni+1𝒜n)=(𝒜1𝒜ni)0,\mathrm{Gen}T_{i}=\varinjlim(\mathcal{A}_{n-i+1}\cup\cdots\cup\mathcal{A}_{n})={{}^{\bot_{0}}}(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i}),
GenTi+1=lim(𝒜ni𝒜n)=(𝒜1𝒜ni1)0.\mathrm{Gen}T_{i+1}=\varinjlim(\mathcal{A}_{n-i}\cup\cdots\cup\mathcal{A}_{n})={{}^{\bot_{0}}}(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1}).

For any A𝒜1𝒜ni1A^{\prime}\in\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1}, apply HomA(,A)\mathrm{Hom}_{A}(-,A^{\prime}) to the short exact sequences

0KYImg0,0\longrightarrow K\longrightarrow Y\longrightarrow\mathrm{Im}g\longrightarrow 0,
0ImgXCokerg0,0\longrightarrow\mathrm{Im}g\longrightarrow X\longrightarrow\mathrm{Coker}g\longrightarrow 0,

respectively. We obtain two long exact sequences

HomA(Y,A)HomA(K,A)ExtA1(Img,A),\cdots\longrightarrow\mathrm{Hom}_{A}(Y,A^{\prime})\longrightarrow\mathrm{Hom}_{A}(K,A^{\prime})\longrightarrow\mathrm{Ext}^{1}_{A}(\mathrm{Im}g,A^{\prime})\longrightarrow\cdots,
ExtA1(X,A)ExtA1(Img,A)ExtA2(Cokerg,A).\cdots\longrightarrow\mathrm{Ext}^{1}_{A}(X,A^{\prime})\longrightarrow\mathrm{Ext}^{1}_{A}(\mathrm{Im}g,A^{\prime})\longrightarrow\mathrm{Ext}^{2}_{A}(\mathrm{Coker}g,A^{\prime})\longrightarrow\cdots.

Obviously, HomA(Y,A)=0\mathrm{Hom}_{A}(Y,A^{\prime})=0 since YGenTi+1=(𝒜1𝒜ni1)0,A𝒜1𝒜ni1Y\in\mathrm{Gen}T_{i+1}={{}^{\bot_{0}}}(\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1}),\ A^{\prime}\in\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1}, and ExtA2(Cokerg,A)=0\mathrm{Ext}^{2}_{A}(\mathrm{Coker}g,A^{\prime})=0 since AA is hereditary. Moreover, since every finite dimensional module is pure-injective and ExtA1(𝒜i,𝒜j)=0\mathrm{Ext}_{A}^{1}(\mathcal{A}_{i},\mathcal{A}_{j})=0 for ij2i-j\geq 2, we have the isomorphism

ExtA1(lim(𝒜ni+1𝒜n),𝒜1𝒜ni1)limExtA1(𝒜ni+1𝒜n,𝒜1𝒜ni1)=0.\mathrm{Ext}^{1}_{A}(\varinjlim(\mathcal{A}_{n-i+1}\cup\cdots\cup\mathcal{A}_{n}),\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1})\cong\varprojlim\mathrm{Ext}^{1}_{A}(\mathcal{A}_{n-i+1}\cup\cdots\cup\mathcal{A}_{n},\mathcal{A}_{1}\cup\cdots\cup\mathcal{A}_{n-i-1})=0.

Hence ExtA1(X,A)=0\mathrm{Ext}^{1}_{A}(X,A^{\prime})=0 and then ExtA1(Img,A)=0\mathrm{Ext}^{1}_{A}(\mathrm{Im}g,A^{\prime})=0. Consequently HomA(K,A)=0\mathrm{Hom}_{A}(K,A^{\prime})=0, i.e. KGenTi+1K\in\mathrm{Gen}T_{i+1}, as desired. ∎

3.4. Examples

We close this note with some examples. Let us first recall some generalizations of the notion of a tilted algebra appearing in the literature.

Definition 3.11.

An algebra AA is called
\bullet quasi-tilted [18] if it satisfies
(i)(i) gl.dimA2\mathrm{gl.dim}A\leq 2; and
(ii)(ii) indA=AA\mathrm{ind}A=\mathcal{L}_{A}\cup\mathcal{R}_{A}, or equivalently, pdX1\mathrm{pd}X\leq 1 or idX1\mathrm{id}X\leq 1 for each XindAX\in\mathrm{ind}A, cf. [13];
\bullet shod [13] if it satisfies condition (ii)(ii) above;
\bullet weakly shod [14] provided
(iii)(iii) AA\mathcal{L}_{A}\cup\mathcal{R}_{A} is cofinite in indA\mathrm{ind}A, i.e. indA(AA)\mathrm{ind}A\setminus(\mathcal{L}_{A}\cup\mathcal{R}_{A}) has finitely many objects; and
(iv)(iv) no nonsemiregular component of the Auslander-Reiten quiver has oriented cycles;
\bullet laura [7] if it satisfies condition (iii)(iii) above.

We don’t know whether all 3-silted algebras are laura. The following example exhibits a 3-silted algebra which is weakly shod, but not shod.

Example 3.12.

Consider the hereditary algebra A=kQA=kQ with the quiver

1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Q:\textstyle{Q:}3\textstyle{3\ignorespaces\ignorespaces\ignorespaces\ignorespaces}4\textstyle{4\ignorespaces\ignorespaces\ignorespaces\ignorespaces}5\textstyle{5}2\textstyle{2\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Let PiP_{i} (resp., IiI_{i}) denotes the projective (resp., injective) AA-module corresponding to vertex ii. Consider the universal localisation of AA at I3I2I1I_{3}\oplus I_{2}\oplus I_{1}. By [23, Theorem 6.1], it is a homological ring epimorphism, denoted by λ0:AB0:=A{I3I2I1}\lambda_{0}:A\longrightarrow B_{0}:=A_{\left\{I_{3}\oplus I_{2}\oplus I_{1}\right\}}, which corresponds to the bireflective subcategory 𝒳0=(I3I2I1)0,1=Add(P5I5I4)\mathcal{X}_{0}=(I_{3}\oplus I_{2}\oplus I_{1})^{\bot_{0,1}}=\mathrm{Add}(P_{5}\oplus I_{5}\oplus I_{4}). Furthermore, B0=I54P5B_{0}={I_{5}}^{4}\oplus P_{5} and the minimal silting module corresponding to the homological ring epimorphism is equivalent to T0=I5P5I3I2I1T_{0}=I_{5}\oplus P_{5}\oplus I_{3}\oplus I_{2}\oplus I_{1}. Similarly, consider another universal localisation of AA at 13I1\begin{smallmatrix}1\\ 3\end{smallmatrix}\oplus I_{1}. Then the homological ring epimorphism λ1:AB1:=A{13I1}\lambda_{1}:A\longrightarrow B_{1}:=A_{\left\{\begin{smallmatrix}1\\ 3\end{smallmatrix}\oplus I_{1}\right\}} corresponds to the bireflective subcategory 𝒳1=(13I1)0,1=Add(P5P1134I5I4I2)\mathcal{X}_{1}=(\begin{smallmatrix}1\\ 3\end{smallmatrix}\oplus I_{1})^{\bot_{0,1}}=\mathrm{Add}(P_{5}\oplus P_{1}\oplus\begin{smallmatrix}1\\ 3\\ 4\end{smallmatrix}\oplus I_{5}\oplus I_{4}\oplus I_{2}). It is easy to check that B1=P13I5P5B_{1}={P_{1}}^{3}\oplus I_{5}\oplus P_{5} and the minimal silting module is equivalent to T1=P1I5P5I113T_{1}=P_{1}\oplus I_{5}\oplus P_{5}\oplus I_{1}\oplus\begin{smallmatrix}1\\ 3\end{smallmatrix}. Clearly, the partial order 𝒳0𝒳1\mathcal{X}_{0}\subseteq\mathcal{X}_{1} corresponds, under the bijection in Theorem 2.7, to the partial order λ0λ1\lambda_{0}\leq\lambda_{1} and there exists a ring epimorphism μ:B1B0\mu:B_{1}\longrightarrow B_{0} such that λ0=μλ1\lambda_{0}=\mu\circ\lambda_{1}. It is easy to verify that Kerμ=0,Cokerμ=I23,Kerλ1=0,Cokerλ1=I1213\mathrm{Ker}\mu=0,\mathrm{Coker}\mu={I_{2}}^{3},\mathrm{Ker}\lambda_{1}=0,\mathrm{Coker}\lambda_{1}={I_{1}}^{2}\oplus\begin{smallmatrix}1\\ 3\end{smallmatrix}. Then, by Proposition 3.1, we get the 33-term silting complex

T=B0Kerμ[1]CokerμKerλ1[2]Cokerλ1[1]P5I5I213[1]I1[1].T=B_{0}\oplus\mathrm{Ker}\mu[1]\oplus\mathrm{Coker}\mu\oplus\mathrm{Ker}\lambda_{1}[2]\oplus\mathrm{Coker}\lambda_{1}[1]\simeq P_{5}\oplus I_{5}\oplus I_{2}\oplus\begin{smallmatrix}1\\ 3\end{smallmatrix}[1]\oplus I_{1}[1].

It is easy to see that the endomorphism ring B=EndDb(modA)TB=\mathrm{End}_{D^{b}(\mathrm{mod}A)}T of TT is isomorphic to the algebra kQ/IkQ^{\prime}/I^{\prime} with Q:αβγηQ^{\prime}:\bullet\stackrel{{\scriptstyle\alpha}}{{\longrightarrow}}\bullet\stackrel{{\scriptstyle\beta}}{{\longrightarrow}}\bullet\stackrel{{\scriptstyle\gamma}}{{\longrightarrow}}\bullet\stackrel{{\scriptstyle\eta}}{{\longrightarrow}}\bullet and I=αβ,βγ,γηI^{\prime}=\langle\alpha\beta,\beta\gamma,\gamma\eta\rangle. Note that it is weakly shod, but it is not a shod algebra since the global dimension of BB is 44.

We also include an example of a (shod) 3-silted algebra arising from universal localization at regular modules.

Example 3.13.

Let A=kD6~A=k\widetilde{D_{6}} be the path algebra of

1\textstyle{1}6\textstyle{6\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D6~:\textstyle{\widetilde{D_{6}}:}3\textstyle{3\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}4\textstyle{4\ignorespaces\ignorespaces\ignorespaces\ignorespaces}5\textstyle{5\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\textstyle{2}7\textstyle{7\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

It is well known that the Auslander-Reiten quiver Γ(modA)\Gamma(\mathrm{mod}A) of AA is the union of the postprojective component 𝒫(A)\mathcal{P}(A), the preinjective component 𝒬(A)\mathcal{Q}(A) and the union (A)\mathcal{R}(A) of regular components. Let PiP_{i} denotes the projective AA-module corresponding to vertex ii. Consider the stable tube of rank 44 in (A)\mathcal{R}(A) containing the modules F1(1),,F4(1)F_{1}^{(1)},\ldots,F_{4}^{(1)} and, for each s{1,,4}s\in\left\{1,\ldots,4\right\}, there is an isomorphism τAFs+1(1)Fs(1)\tau_{A}F_{s+1}^{(1)}\cong F_{s}^{(1)}, where we set F5(1)=F1(1)F_{5}^{(1)}=F_{1}^{(1)}.

0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F1(1):\textstyle{F_{1}^{(1)}:}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}                    0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F2(1):\textstyle{F_{2}^{(1)}:}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F3(1):\textstyle{F_{3}^{(1)}:}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}              K\textstyle{K}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{1}F4(1):\textstyle{F_{4}^{(1)}:}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{1}1\scriptstyle{1}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{1}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{1}K\textstyle{K}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{1}

Consider the universal localisation of AA at F2(1)F3(1)F_{2}^{(1)}\oplus F_{3}^{(1)} and denote the homological ring epimorphism by λ0:AB0:=AF2(1)F3(1)\lambda_{0}:A\longrightarrow B_{0}:=A_{F_{2}^{(1)}\oplus F_{3}^{(1)}}, which corresponds to the bireflective subcategory 𝒳0=(F2(1)F3(1))0,1\mathcal{X}_{0}=(F_{2}^{(1)}\oplus F_{3}^{(1)})^{\bot_{0,1}}. It is easy to see that B0=P1P2P53P6P7B_{0}=P_{1}\oplus P_{2}\oplus{P_{5}}^{3}\oplus P_{6}\oplus P_{7} and the minimal silting module corresponding to the homological ring epimorphism is equivalent to T0=P1P2P5P6P7F2(1)F3(1)T_{0}=P_{1}\oplus P_{2}\oplus P_{5}\oplus P_{6}\oplus P_{7}\oplus F_{2}^{(1)}\oplus F_{3}^{(1)}. Similarly, the homological ring epimorphism λ1:AB1\lambda_{1}:A\longrightarrow B_{1} corresponds to universal localisation of AA at F2(1)F_{2}^{(1)}, and then the bireflective subcategory is 𝒳0=(F2(1))0,1\mathcal{X}_{0}=(F_{2}^{(1)})^{\bot_{0,1}}. By easy calculation, we have B1=P1P2P42P5P6P7B_{1}=P_{1}\oplus P_{2}\oplus{P_{4}}^{2}\oplus P_{5}\oplus P_{6}\oplus P_{7} and the minimal silting module corresponding to λ1\lambda_{1} is equivalent to T1=P1P2P4P5P6P7F2(1)T_{1}=P_{1}\oplus P_{2}\oplus P_{4}\oplus P_{5}\oplus P_{6}\oplus P_{7}\oplus F_{2}^{(1)}. Clearly, the partial order 𝒳0𝒳1\mathcal{X}_{0}\subseteq\mathcal{X}_{1} corresponds, under the bijection in Theorem 2.7, to the partial order λ0λ1\lambda_{0}\leq\lambda_{1} and there exists a ring epimorphism μ:B1B0\mu:B_{1}\longrightarrow B_{0} such that λ0=μλ1\lambda_{0}=\mu\circ\lambda_{1}. It is easy to verify that Kerμ=0,Cokerμ=(F3(1))2,Kerλ1=0,Cokerλ1=F2(1)\mathrm{Ker}\mu=0,\mathrm{Coker}\mu=(F_{3}^{(1)})^{2},\mathrm{Ker}\lambda_{1}=0,\mathrm{Coker}\lambda_{1}=F_{2}^{(1)}. Therefore, by Proposition 3.1, we get the 33-term silting complex

T=B0Kerμ[1]CokerμKerλ1[2]Cokerλ1[1]P1P2P5P6P7F3(1)F2(1)[1].T=B_{0}\oplus\mathrm{Ker}\mu[1]\oplus\mathrm{Coker}\mu\oplus\mathrm{Ker}\lambda_{1}[2]\oplus\mathrm{Coker}\lambda_{1}[1]\simeq P_{1}\oplus P_{2}\oplus{P_{5}}\oplus P_{6}\oplus P_{7}\oplus F_{3}^{(1)}\oplus F_{2}^{(1)}[1].

It is easy to check that the endomorphism ring B=EndDb(modA)TB=\mathrm{End}_{D^{b}(\mathrm{mod}A)}T of TT is isomorphic to the algebra kQ/IkQ^{\prime}/I^{\prime} with

\textstyle{\bullet\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α1\scriptstyle{\alpha_{1}}\textstyle{\bullet}Q:\textstyle{Q^{\prime}:}\textstyle{\bullet\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β1\scriptstyle{\beta_{1}}β2\scriptstyle{\beta_{2}}β\scriptstyle{\beta}\textstyle{\bullet\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γ\scriptstyle{\gamma}\textstyle{\bullet}\textstyle{\bullet\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α2\scriptstyle{\alpha_{2}}\textstyle{\bullet}

and I=α1β,α2β,βγI^{\prime}=\langle\alpha_{1}\beta,\alpha_{2}\beta,\beta\gamma\rangle. It is a shod algebra.

Acknowledgments

This work was started during a ICTP-INdAM Research in Pairs Programme with the visit of the second and last named authors at the University of Verona in 2019. The first and last named authors would like to thank the Network on Silting Theory funded by the Deutsche Forschungsgemeinschaft and the University of Stuttgart for hospitality during a research visit in 2022 where part of this work was carried out. The first named author also acknowledges support from the project SQUARE: Structures for Quivers, Algebras and Representations, PRIN 2022S97PMY, funded by the Italian Ministry of University and Research. The third named author is supported by China Scholarship Council (Grant No. 202306860057). The fourth named author is supported by the project PICT 2021 01154 from ANPCyT, Argentina.

References

  • [1] T. Adachi, O. Iyama, and I. Reiten. τ\tau-tilting theory. Compos. Math., 150(3):415–452, 2014.
  • [2] E. R. Alvares, I. Assem, F. U. Coelho, M. I. Peña, and S. Trepode. From trisections in module categories to quasi-directed components. J. Algebra Appl., 10(3):409–433, 2011.
  • [3] L. Angeleri Hügel. Endocoherent modules. Pacific Journal of Mathematics 212 (2003), 1-11.
  • [4] L. Angeleri Hügel and M. Hrbek. Parametrizing torsion pairs in derived categories. Represent. Theory, 25:679–731, 2021.
  • [5] L. Angeleri Hügel, F. Marks, and J. Vitória. Silting modules. Int. Math. Res. Not. IMRN, (4):1251–1284, 2016.
  • [6] L. Angeleri Hügel, F. Marks, and J. Vitória. Silting modules and ring epimorphisms. Adv. Math., 303:1044–1076, 2016.
  • [7] I. Assem and F. U. Coelho. Two-sided gluings of tilted algebras. J. Algebra, 269(2):456–479, 2003.
  • [8] M. Auslander, I. Reiten, and S. O. Smalø. Representation theory of Artin algebras, volume 36 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995.
  • [9] M. Auslander and S. O. Smalø. Almost split sequences in subcategories. J. Algebra, 69(2):426–454, 1981.
  • [10] A. A. Beilinson, J. Bernstein, and P. Deligne. Faisceaux pervers. In Analysis and topology on singular spaces, I Astérisque 100. Soc. Math. France, Paris, 5–171, 1982.
  • [11] G. M. Bergman and W. Dicks. Universal derivations and universal ring constructions. Pacific J. Math., 79(2):293–337, 1978.
  • [12] A. B. Buan and Y. Zhou. Silted algebras. Adv. Math., 303:859–887, 2016.
  • [13] F. U. Coelho and M. A. Lanzilotta. Algebras with small homological dimensions. Manuscripta Math., 100(1):1–11, 1999.
  • [14] F. U. Coelho and M. A. Lanzilotta. Weakly shod algebras. J. Algebra, 265(1):379–403, 2003.
  • [15] W. Crawley-Boevey. Locally finitely presented additive categories. Comm. Algebra, 22(5):1641–1674, 1994.
  • [16] P. Gabriel and J. A. de la Peña. Quotients of representation-finite algebras. Comm. Algebra, 15(1-2):279–307, 1987.
  • [17] S. I. Gelfand and Y. I. Manin. Methods of homological algebra. Springer Monographs in Mathematics. Springer-Verlag, Berlin, second edition, 2003.
  • [18] D. Happel, I. Reiten, and S. O. Smalø. Tilting in abelian categories and quasitilted algebras. Mem. Amer. Math. Soc., 120(575):viii+ 88, 1996.
  • [19] D. Happel and C. M. Ringel. Tilted algebras. Trans. Amer. Math. Soc., 274(2):399–443, 1982.
  • [20] B. Keller and D. Vossieck. Aisles in derived categories. Bull. Soc. Math. Belg. Sér. A, 40(2):239–253, 1988.
  • [21] S. Koenig and D. Yang. Silting objects, simple-minded collections, tt-structures and co-tt-structures for finite-dimensional algebras. Doc. Math., 19:403–438, 2014.
  • [22] H. Krause. A short proof for Auslander’s defect formula. volume 365, pages 267–270. 2003. Special issue on linear algebra methods in representation theory.
  • [23] H. Krause and J. Šťovíček. The telescope conjecture for hereditary rings via Ext-orthogonal pairs. Adv. Math., 225(5):2341–2364, 2010.
  • [24] C. Psaroudakis and J. Vitória. Realisation functors in tilting theory. Math. Z., 288(3-4):965–1028, 2018.
  • [25] C. M. Ringel. Tame algebras and integral quadratic forms, volume 1099 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1984.
  • [26] A. H. Schofield. Representation of rings over skew fields. Cambridge University Press, Cambridge, 1985.
  • [27] F. Sentieri. A brick version of a theorem of Auslander. Nagoya Math. J., 249:88–106, 2023.