Hostname: page-component-745bb68f8f-s22k5 Total loading time: 0 Render date: 2025-02-11T11:05:06.286Z Has data issue: false hasContentIssue false

Characterizations of right rejective chains

Published online by Cambridge University Press:  14 July 2021

Mayu Tsukamoto*
Affiliation:
Graduate school of Sciences and Technology for Innovation, Yamaguchi University, 1677-1 Yoshida, Yamaguchi753-8512, Japan
Rights & Permissions [Opens in a new window]

Abstract

In this paper, we give characterizations of the category of finitely generated projective modules having a right rejective chain. By focusing on the characterizations, we give sufficient conditions for right rejective chains to be total right rejective chains. Moreover, we prove that Nakayama algebras with heredity ideals, locally hereditary algebras and algebras of global dimension at most two satisfy the sufficient conditions. As an application, we show that these algebras are right-strongly quasi-hereditary algebras.

Type
Article
Copyright
© Canadian Mathematical Society 2021

1 Introduction

The notion of right rejective subcategories was introduced by Iyama and plays a crucial role in the proof of the finiteness theorem for representation dimensions of artin algebras [Reference Iyama13]. Right rejective subcategories are a special class of coreflective subcategories which appear in the classical theory of localizations of abelian categories. It is known that right rejective subcategories provide effective tools to study representation-finite orders [Reference Iyama12Reference Iyama14, Reference Rump19]. A subcategory $\mathcal {C}'$ of an additive category $\mathcal {C}$ is called a right rejective subcategory if for each $X \in \mathcal {C}$ , there exists a monic right $\mathcal {C}'$ -approximation of X. We call a chain of cosemisimple right rejective (respectively, coreflective) subcategories a right rejective (respectively, coreflective) chain.

Quasi-hereditary algebras were introduced by Cline, Parshall, and Scott to explore highest weight categories arising from the study of semisimple complex Lie algebras and algebraic groups [Reference Coulembier5, Reference Scott20]. As a distinguished class of quasi-hereditary algebras, Ringel [Reference Ringel18] introduced right-strongly quasi-hereditary algebras which frequently appear in the representation theory of algebras [Reference Cline, Parshall and Scott7, Reference Cline, Parshall and Scott8, Reference Geiss, Leclerc and Schröer10, Reference Kalck and Karmazyn16, Reference Tsukamoto21]. It is known that special right rejective chains called total right rejective chains, are characterized by right-strongly quasi-hereditary algebras.

Proposition 1.1 ([Reference Tsukamoto22, Theorem 3.22])

Let A be an artin algebra and $(e_{1},e_{2},\ldots , e_{n})$ a complete set of primitive orthogonal idempotents of A. Let $\varepsilon _{i}:=e_{i} + \cdots + e_{n}$ . Then the following statements are equivalent.

  1. (1) The following chain is a total right rejective chain.

    $$ \begin{align*} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align*} $$
  2. (2) $A> A\varepsilon _{2} A > \cdots > A \varepsilon _{n} A >0$ is a heredity chain (namely A is a quasi-hereditary algebra), and the following chain is a coreflective chain.

    $$ \begin{align*} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset0. \end{align*} $$
  3. (3) $A> A\varepsilon _{2} A > \cdots > A \varepsilon _{n} A >0$ is a right-strongly heredity chain (namely A is a right-strongly quasi-hereditary algebra).

Ágoston et al. [Reference Ágoston, Dlab and Wakamatsu1] introduced neat algebras as a generalization of quasi-hereditary algebras. As an analog of Proposition 1.1, we give characterizations for $\mathsf {proj}\, A$ to admit a right rejective chain using neat algebras. For an idempotent e of A, let $S(e):=eA/eJ$ , where J is the Jacobson radical of A.

Theorem 1.2 (Theorem 3.5)

Let A be an artin algebra and $(e_{1},e_{2},\ldots , e_{n})$ a complete set of primitive orthogonal idempotents of A. Let $\varepsilon _{i}:=e_{i} + \cdots + e_{n}$ . Then the following statements are equivalent.

  1. (1) The following chain is a right rejective chain.

    (1.1) $$ \begin{align} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align} $$
  2. (2) $A> A\varepsilon _{2} A > \cdots > A \varepsilon _{n} A >0$ is a neat chain (namely A is a neat algebra), and the following chain is a coreflective chain.

    $$ \begin{align*} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align*} $$
  3. (3) For each $i \in [1, n]$ , $\operatorname {pd}\nolimits _{\varepsilon _{i} A\varepsilon _{i}} S(e_{i})\varepsilon _{i} \le 1$ holds.

By focusing on the condition (3) in Theorem 1.2, we give sufficient conditions (see Theorem 3.9) for the right rejective chain (1.1) to be a total right rejective chain. Moreover, we prove that the following classes of algebras satisfy the sufficient conditions.

  • Nakayama algebras with heredity ideals (Proposition 3.13).

  • Locally hereditary algebras (Proposition 3.16).

  • Algebras of global dimension at most two (Proposition 3.17).

Therefore, we have the following result.

Theorem 1.3 (Theorem 3.12)

Nakayama algebras with heredity ideals, locally hereditary algebras and algebras of global dimension at most two are right-strongly quasi-hereditary.

Notation

Throughout this paper, A is a basic artin algebra. Let $J=J(A)$ be the Jacobson radical of A. By a module, we mean a finitely generated right A-module. Let $\operatorname {pd}\nolimits _{A} M$ denote the projective dimension of an A-module M. We write $\mathsf {mod}\, A$ for the category of finitely generated right A-modules and $\mathsf {proj}\, A$ for the full subcategory of $\mathsf {mod}\, A$ consisting of projective A-modules. For $M \in \mathsf {mod}\, A$ , let $\mathsf {add}\, M$ denote the full subcategory of $\mathsf {mod}\, A$ whose objects are direct summands of finite direct sums of M.

2 Right rejective chains

In this section, we recall the definition of right rejective chains and collect related results (see [Reference Iyama13, Reference Iyama15 ] for details). By a subcategory, we always mean a full subcategory which is closed under isomorphisms. For a subcategory $\mathcal {C}'$ of an additive category $\mathcal {C}$ , we call a morphism $f: Y \to X$ in $\mathcal {C}$ a right $\mathcal {C}'$ -approximation of X if $Y \in \mathcal {C}'$ and $\mathcal {C}(-, f): \mathcal {C}(-, Y) \rightarrow \mathcal {C}(-, X)$ is an epimorphism on $\mathcal {C}'$ . Dually, we define a left $\mathcal {C}'$ -approximation.

Definition 2.1 Let $\mathcal {C}$ be an additive category and $\mathcal {C}'$ a subcategory of $\mathcal {C}$ .

  1. (1) We call $\mathcal {C}'$ a coreflective subcategory of $\mathcal {C}$ if for any $X\in \mathcal {C}$ , there exists a right $\mathcal {C}'$ -approximation $f \in \mathcal {C}\left (Y,X\right )$ of X such that $\mathcal {C}(-, f): \mathcal {C}(-, Y) \rightarrow \mathcal {C}(-, X)$ is an isomorphism on $\mathcal {C}'$ .

  2. (2) We call $\mathcal {C}'$ a right rejective subcategory of $\mathcal {C}$ if for any $X\in \mathcal {C}$ , there exists a right $\mathcal {C}'$ -approximation of X such that it is a monomorphism in $\mathcal {C}$ .

Dually, reflective subcategories and left rejective subcategories are defined.

Right rejective subcategories are coreflective subcategories, but the converse does not hold in general.

To define a right rejective chain, we need the notion of cosemisimple subcategories. Let $\mathcal {J}_{\mathcal {C}}$ be the Jacobson radical of $\mathcal {C}$ . For a subcategory $\mathcal {C}'$ of $\mathcal {C}$ , let $[\mathcal {C}']$ denote the ideal of $\mathcal {C}$ consisting of morphisms which factor through some object of $\mathcal {C}'$ , and let $\mathcal {C}/[\mathcal {C}']$ denote the factor category.

Definition 2.2 Let $\mathcal {C}$ be an additve category. A subcategory $\mathcal {C}'$ of $\mathcal {C}$ is called a cosemisimple subcategory of $\mathcal {C}$ if $\mathcal {J}_{\mathcal {C}/[\mathcal {C}']}=0$ holds.

We recall characterizations of coreflective subcategories, right rejective subcategories and cosemisimple right rejective subcategories of $\mathsf {proj}\, A$ . We write $\operatorname {gldim}\nolimits A$ for the global dimension of A.

Proposition 2.3 ([Reference Iyama15, Theorem 3.2])

Let A be an artin algebra and $\varepsilon $ an idempotent of A. Then the following statements hold.

  1. (1) $\mathsf {add} \varepsilon A$ is a coreflective subcategory of $\mathsf {proj}A$ if and only if $A\varepsilon \in \mathsf {proj}\, \varepsilon A \varepsilon $ .

  2. (2) $\mathsf {add} \varepsilon A$ is a right rejective subcategory of $\mathsf {proj}A$ if and only if $A \varepsilon A \in \mathsf {add} \varepsilon A$ as a right A-module. In this case, $\operatorname {gldim}\nolimits \varepsilon A \varepsilon \le \operatorname {gldim}\nolimits A$ holds.

  3. (3) $\mathsf {add} \varepsilon A$ is a cosemisimple right rejective subcategory of $\mathsf {proj}A$ if and only if $e J \in \mathsf {add} \varepsilon A$ as a right A-module, where $e:=1- \varepsilon $ . In this case, the inclusion $e J\to e A$ is a right $\mathsf {add}\, \varepsilon A$ -approximation of $eA$ .

Now, we introduce the notion of coreflective chains, right rejective chains and total right rejective chains.

Definition 2.4 Let $\mathcal {C}$ be an additve category. Let

(2.1) $$ \begin{align} \mathcal{C}= \mathcal{C}_{1} \supset \mathcal{C}_{2} \supset \cdots \supset \mathcal{C}_{n} \supset \mathcal{C}_{n+1}=0 \end{align} $$

be a chain of subcategories of $\mathcal {C}$ .

  1. (1) (2.1) is called a coreflective chain (of length n) if $\mathcal {C}_{i+1}$ is a cosemisimple coreflective subcategory of $\mathcal {C}_{i}$ for each $1 \le i \le n$ .

  2. (2) (2.1) is called a right rejective chain (of length n) if $\mathcal {C}_{i+1}$ is a cosemisimple right rejective subcategory of $\mathcal {C}_{i}$ for each $1 \le i \le n$ .

  3. (3) (2.1) is called a total right rejective chain (of length n) if the following conditions hold for each $1 \le i \le n$ :

    1. (a) $\mathcal {C}_{i+1}$ is a right rejective subcategory of $\mathcal {C}$ and

    2. (b) $\mathcal {C}_{i+1}$ is a cosemisimple subcategory of $\mathcal {C}_{i}$ .

One can easily check that total right rejective chains are right rejective chains, and right rejective chains are coreflective chains. However, the converses do not always hold. We refer to Example 3.6 which shows that right rejective chains are not necessarily total right rejective chains.

If $\mathsf {proj}\, A$ has a right rejective chain, then we give an upper bound of the global dimension of A.

Proposition 2.5 ([Reference Iyama15, Theorem 3.3])

Let A be an artin algebra. If $\mathsf {proj} A$ admits a right rejective chain of length n, then the global dimension of A is at most $2n-2$ .

Remark 2.6 The length of a right rejective chain of $\mathsf {proj}\, A$ is bounded by the number of indecomposable direct summands of A, or equivalently the number of isomorphism classes of simple A-modules. Hence, if A has m simple modules and $\mathsf {proj}\, A$ admits a right rejective chain, then the global dimension of A is at most $2m-2$ by Proposition 2.5.

Total right rejective chains are closely related to right-strongly quasi-hereditary algebras which are a special class of quasi-hereditary algebras. We recall the definitions of quasi-hereditary algebras and right-strongly quasi-hereditary algebras (see [Reference Coulembier5, Reference Dlab and Ringel9] and [Reference Ringel18], for details). We call an idempotent ideal $A\varepsilon A$ a heredity ideal of A if $A \varepsilon A \in \mathsf {proj}\, A$ and $\varepsilon A \varepsilon $ is semisimple.

Definition 2.7 Let A be an artin algebra.

  1. (1) ([Reference Coulembier5, Reference Dlab and Ringel9]) We call A a quasi-hereditary algebra if there exists a chain of idempotent ideals

    (2.2) $$ \begin{align} A =A\varepsilon_{1} A> A\varepsilon_{2} A > \cdots > A \varepsilon_{n} A >A \varepsilon_{n+1} A=0 \end{align} $$
    such that $A \varepsilon _{i}A/A\varepsilon _{i+1}A $ is a heredity ideal of $A /A\varepsilon _{i+1}A$ for each $1 \le i \le n$ . In this case, (2.2) is called a heredity chain.
  2. (2) ([Reference Ringel18]) We call A a right-strongly quasi-hereditary algebra if there exists a chain of idempotent ideals

    $$ \begin{align*} A=A\varepsilon_{1} A> A\varepsilon_{2} A > \cdots > A \varepsilon_{n} A >A \varepsilon_{n+1} A=0 \end{align*} $$
    such that $A\varepsilon _{i}A \in \mathsf {proj}\, A$ and $\varepsilon _{i}A \varepsilon _{i}/ \varepsilon _{i} A\varepsilon _{i+1}A \varepsilon _{i}$ is semisimple for each $1 \le i \le n$ .

Since $-\otimes _{A}A/A\varepsilon _{i+1}A$ preserves projective modules, right-strongly quasi-hereditary algebras are quasi-hereditary algebras. We give a characterization of right-strongly quasi-hereditary algebras by total right rejective chains.

Proposition 2.8 ([Reference Tsukamoto22, Theorem 3.22])

Let A be an artin algebra. Then A is a right-strongly quasi-hereditary algebra if and only if $\mathsf {proj}\, A$ admits a total right rejective chain.

3 Main results

In this section, we give characterizations of right rejective chains and constructions of total right rejective chains. Moreover, using the constructions, we provide total right rejective chains for several classes of algebras.

3.1 Characterizations of right rejective chains

In this subsection, using neat algebras, we give characterizations of $\mathsf {proj}\, A$ admitting a right rejective chain.

For an idempotent e of A, $S(e)$ stands for a semisimple A-module $eA/eJ$ . Let $P(e)$ denote a projective cover of $S(e)$ . In the following, we fix a complete set $(e_{1},e_{2},\ldots , e_{n})$ of primitive orthogonal idempotents of A. Then, we have a complete set $\{S(e_{i}) \mid 1 \le i \le n \}$ of representatives of isomorphism classes of simple A-modules. Let $\varepsilon _{i}:=e_{i} + \cdots +e_{n}$ for $1 \le i \le n$ and $\varepsilon _{n+1}:=0$ . We recall the definition of neat algebras (see [Reference Ágoston, Dlab and Wakamatsu1] for details). A typical example of a neat algebra is a quasi-hereditary algebra.

Definition 3.1 [Reference Ágoston, Dlab and Wakamatsu1]

Let A be an artin algebra.

  1. (1) An idempotent e of A is called a neat idempotent if $\operatorname {Ext}\nolimits _{A}^{i}(S(e), S(e))=0$ holds for each $i \ge 1$ . We call $(e_{1},e_{2},\ldots , e_{n})$ a neat sequence if $e_{i}$ is a neat idempotent of $\varepsilon _{i}A\varepsilon _{i}$ for each $1 \le i \le n$ .

  2. (2) An algebra A is called a neat algebra if there exists a neat sequence.

We give an upper bound of the global dimension of a neat algebra.

Proposition 3.2 ([Reference Ágoston, Dlab and Wakamatsu1, Proposition 2])

If A is a neat algebra with a neat sequence $(e_{1},e_{2},\ldots , e_{n})$ , then the global dimension of A is at most $2^{n}-2$ .

We reformulate neat algebras in terms of a chain of strong idempotent ideals. Recall that an idempotent ideal $A \varepsilon A$ is called a strong idempotent ideal (or stratifying ideal) of A if $\operatorname {Ext}\nolimits _{A/A\varepsilon A}^{j}(X, Y) \cong \operatorname {Ext}\nolimits _{A}^{j}(X, Y)$ holds for each $X, Y \in \mathsf {mod}\, (A/ A \varepsilon A)$ and $j \ge 1$ (see [Reference Auslander, Platzeck and Todorov3] and [Reference Conde6] for details).

Proposition 3.3 Let A be an artin algebra. Then A is a neat algebra with a neat sequence $(e_{1},e_{2},\ldots , e_{n})$ if and only if a chain of idempotent ideals

(3.1) $$ \begin{align} A> A\varepsilon_{2} A > \cdots > A \varepsilon_{n} A >A \varepsilon_{n+1} A=0 \end{align} $$

satisfies that $\varepsilon _{i} A \varepsilon _{i+1} A \varepsilon _{i}$ is a strong idempotent ideal of $\varepsilon _{i}A\varepsilon _{i}$ and $\varepsilon _{i} A\varepsilon _{i}/ \varepsilon _{i} A \varepsilon _{i+1} A \varepsilon _{i}$ is semisimple for each $1 \le i \le n$ . In this case, we call (3.1) a neat chain.

Proof. Let e be an idempotent of A and $\varepsilon :=1 -e$ . It suffices to show that e is a neat idempotent of A if and only if $A\varepsilon A$ is a strong idempotent ideal and $A/A\varepsilon A$ is semisimple. Indeed, one can apply this claim to each subalgebra $\varepsilon _{i}A \varepsilon _{i}$ . By [Reference Ágoston, Dlab and Wakamatsu1, Proposition 1.1], e is a neat idempotent of A if and only if it satisfies that (a) the multiplication map: $A \varepsilon \otimes _{\varepsilon A \varepsilon } \varepsilon A \to A \varepsilon A$ is an isomorphism, (b) $\operatorname {Tor}\nolimits _{j}^{\varepsilon A \varepsilon }(A \varepsilon , \varepsilon A)=0$ for each $j \ge 1$ and (c) $e A \varepsilon A e=eJe$ . Moreover, by [Reference Conde6, Remark 2.1.2(a)], the conditions (a) and (b) are equivalent to the condition that $A \varepsilon A$ is a strong idempotent ideal of A. Thus we show that the condition (c) holds if and only if $A/A \varepsilon A$ is semisimple. The “only if” part holds since $A/ A\varepsilon A \cong eAe / eA \varepsilon A e = eAe/ J(eAe)$ . We show the “if” part. By $\varepsilon =1-e$ , we obtain that $eA\varepsilon A e \subseteq eJe$ . Since $A/ A \varepsilon A $ is semisimple, so is $eAe/e A \varepsilon A e$ . Thus $0=J(eAe/eA\varepsilon Ae) =J(eAe)/eA\varepsilon Ae$ , and hence the assertion holds.▪

By Proposition 3.3, we reprove the following result.

Corollary 3.4 ([Reference Ágoston, Dlab and Wakamatsu1, Corollary in $\S$ 0])

Any quasi-hereditary algebra is a neat algebra.

Proof. We assume that A is a quasi-hereditary algebra. Then there exists a heredity chain

$$ \begin{align*} A> A\varepsilon_{2} A > \cdots > A \varepsilon_{n} A > A \varepsilon_{n+1} A=0. \end{align*} $$

We show that this chain is a neat chain. By definition, $\varepsilon _{i} A\varepsilon _{i}/ \varepsilon _{i} A \varepsilon _{i+1} A \varepsilon _{i}$ is semisimple. Thus it suffices from Proposition 3.3 to prove that $\varepsilon _{i}A \varepsilon _{i+1}A \varepsilon _{i}$ is a strong idempotent ideal of $\varepsilon _{i}A \varepsilon _{i}$ for each $1 \le i\le n$ . By $A \varepsilon _{i}A/A\varepsilon _{i+1}A \in \mathsf {proj}\, (A/ A \varepsilon _{i+1}A)$ , it follows from [Reference Auslander, Platzeck and Todorov3, Proposition 5.2] that $A \varepsilon _{i}A/A\varepsilon _{i+1}A$ is a strong idempotent ideal of $A/A\varepsilon _{i+1}A$ . Since

$$ \begin{align*} \frac{A/ A \varepsilon_{i+1}A}{A\varepsilon_{i}A/A\varepsilon_{i+1}A }\cong A/A\varepsilon_{i}A, \end{align*} $$

we obtain that $\operatorname {Ext}\nolimits _{A/A \varepsilon _{i}A}^{j}(X, Y) \cong \operatorname {Ext}\nolimits _{A/A\varepsilon _{i+1}A}^{j}(X, Y)$ for each $X, Y \in \mathsf {mod}\, (A/A \varepsilon _{i}A)$ and $j \ge 1$ . Thus, $A \varepsilon _{i}A$ is inductively a strong idempotent ideal of A for each $2 \le i\le n$ . Since $A \varepsilon _{i+1}A$ is a strong idempotent ideal of A, it follows from [Reference Auslander, Platzeck and Todorov3, Theorem 2.1] that $\varepsilon _{i}A \varepsilon _{i+1}A \varepsilon _{i}$ is also a strong idempotent ideal of $\varepsilon _{i}A \varepsilon _{i}$ . Hence the assertion holds.▪

We give characterizations for $\mathsf {proj}\, A$ to admit a right rejective chain.

Theorem 3.5 Let A be an artin algebra. Then the following statements are equivalent.

  1. (1) The following chain is a right rejective chain.

    $$ \begin{align*} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align*} $$
  2. (2) A is a neat algebra with a neat sequence $(e_{1}, e_{2}, \ldots , e_{n})$ and the following chain is a coreflective chain.

    (3.2) $$ \begin{align} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align} $$
  3. (3) For each $1 \le i \le n$ , $\operatorname {pd}\nolimits _{\varepsilon _{i} A\varepsilon _{i}} S(e_{i})\varepsilon _{i} \le 1$ holds.

Proof. By an equivalence $\operatorname {Hom}\nolimits _{A}(\varepsilon _{i}A, -) : \mathsf {add}\, \varepsilon _{i}A \xrightarrow {\sim } \mathsf {proj}\, \varepsilon _{i}A \varepsilon _{i}$ , we obtain that $\mathsf {add}\, \varepsilon _{i+1}A$ is a cosemisimple right rejective (respectively, coreflective) subcategory of $\mathsf {add}\, \varepsilon _{i}A$ if and only if $\mathsf {add}\, \varepsilon _{i+1}A\varepsilon _{i}$ is a cosemisimple right rejective (respectively, coreflective) subcategory of $\mathsf {proj}\, \varepsilon _{i}A\varepsilon _{i}$ .

(1) $\Leftrightarrow $ (3): We consider an exact sequence in $\mathsf {mod}\, \varepsilon _{i}A\varepsilon _{i}$

$$ \begin{align*} 0 \to e_{i}J\varepsilon_{i} \to e_{i}A\varepsilon_{i} \to S(e_{i})\varepsilon_{i} \to 0. \end{align*} $$

If $\mathsf {add}\, \varepsilon _{i+1} A$ is a cosemisimple right rejective subcategory of $\mathsf {add}\, \varepsilon _{i}A$ , then $e_{i}J\varepsilon _{i} \in \mathsf {add}\, \varepsilon _{i+1} A\varepsilon _{i}$ by Proposition 2.3(3). Thus $\operatorname {pd}\nolimits _{\varepsilon _{i}A\varepsilon _{i}} S(e_{i})\varepsilon _{i} \le 1$ . Conversely, we assume that $\operatorname {pd}\nolimits _{\varepsilon _{i}A\varepsilon _{i}} S(e_{i})\varepsilon _{i} \le 1$ . Then $e_{i}J\varepsilon _{i} \in \mathsf {proj}\, \varepsilon _{i}A\varepsilon _{i}$ . Since $e_{i}A\varepsilon _{i}$ is not a direct summand of $e_{i}J\varepsilon _{i}$ , we have $e_{i}J\varepsilon _{i} \in \mathsf {add}\, \varepsilon _{i+1} A\varepsilon _{i}$ . Hence $\mathsf {add}\, \varepsilon _{i+1} A$ is a cosemisimple right rejective subcategory of $\mathsf {add}\, \varepsilon _{i}A$ by Proposition 2.3(3). Thus the assertion holds.

(1) $\Rightarrow{}$ (2): Since right rejective chains are coreflective chains, it suffices to show that the chain $A>A \varepsilon _{2}A > \cdots > A \varepsilon _{n}A >0$ is a neat chain by Proposition 3.3. Due to our assumption, $\mathsf {add}\, \varepsilon _{i+1}A$ is a right rejective subcategory of $\mathsf {add}\, \varepsilon _{i}A$ . By Proposition 2.3(2), we obtain that $\varepsilon _{i} A\varepsilon _{i+1}A \varepsilon _{i} \in \mathsf {proj}\, \varepsilon _{i}A\varepsilon _{i}$ for each ${1 \le i \le n}$ . Thus $\varepsilon _{i} A\varepsilon _{i+1}A \varepsilon _{i}$ is a strong idempotent ideal of $\varepsilon _{i}A\varepsilon _{i}$ by [Reference Auslander, Platzeck and Todorov3, Proposition 5.2]. Let $\mathcal {C}:=\mathsf {add}\, \varepsilon _{i}A/ [\mathsf {add}\, \varepsilon _{i+1}A]$ . Then $\mathcal {J}_{\mathcal {C}}=0$ holds by cosemisimplicity. Since $J(\varepsilon _{i}A \varepsilon _{i}/ \varepsilon _{i}A \varepsilon _{i+1} A \varepsilon _{i})=J(\operatorname {End}\nolimits _{\mathcal {C}}(\varepsilon _{i}A)) = \mathcal {J}_{\mathcal {C}}(\varepsilon _{i}A, \varepsilon _{i}A)=0$ , we have the assertion.

(2) $\Rightarrow{}$ (1): Since (3.2) is a coreflective chain, $\mathsf {add}\, \varepsilon _{i+1}A$ is a cosemisimple subcategory of $\mathsf {add}\, \varepsilon _{i}A$ . Hence, it is enough to show that $\mathsf {add}\, \varepsilon _{i+1}A$ is a right rejective subcategory of $\mathsf {add}\, \varepsilon _{i}A$ for each $1 \le i \le n-1$ . By Proposition 2.3(2), we prove that $\varepsilon _{i}A \varepsilon _{i+1}A\varepsilon _{i} \in \mathsf {proj}\, \varepsilon _{i}A \varepsilon _{i}$ . Since A is a neat algebra with a neat sequence $(e_{1}, e_{2}, \ldots , e_{n})$ , it follows from Proposition 3.3 that $\varepsilon _{i}A\varepsilon _{i+1}A\varepsilon _{i}$ is a strong idempotent ideal of $\varepsilon _{i}A\varepsilon _{i}$ . Due to our assumption, $\mathsf {add}\, \varepsilon _{i+1}A$ is a coreflective subcategory of $\mathsf {add}\, \varepsilon _{i}A$ . Thus, $\varepsilon _{i}A \varepsilon _{i+1} \in \mathsf {proj}\, \varepsilon _{i+1}A \varepsilon _{i+1}$ by Proposition 2.3(1). Since $\varepsilon _{i}A\varepsilon _{i+1}A\varepsilon _{i}$ is a strong idempotent ideal of $\varepsilon _{i}A\varepsilon _{i}$ and $\varepsilon _{i}A \varepsilon _{i+1} \in \mathsf {proj}\, \varepsilon _{i+1}A\varepsilon _{i+1}$ , it follows from [Reference Auslander, Platzeck and Todorov3, Corollary 3.8(b), Proposition 5.2] that $\varepsilon _{i}A \varepsilon _{i+1}A\varepsilon _{i} \in \mathsf {proj}\, \varepsilon _{i}A \varepsilon _{i}$ . Thus the proof is complete.▪

We give an example of a right rejective chain using Theorem 3.5.

Example 3.6 Let A be the algebra defined by the quiver

with relations $\alpha \beta \gamma \alpha $ and $\gamma \alpha \beta $ . Then we can check that ${\operatorname {pd}\nolimits _{A} S(e_{1})=1}$ , $\operatorname {pd}\nolimits _{\varepsilon _{2}A\varepsilon _{2}} S(e_{2})\varepsilon _{2}=1$ and $\operatorname {pd}\nolimits _{\varepsilon _{3}A\varepsilon _{3}} S(e_{3})\varepsilon _{3}=0$ . Thus $\mathsf {proj}\, A \supset \mathsf {add}\, (e_{2}+e_{3})A \supset \mathsf {add}\, e_{3} A \supset 0$ is a right rejective chain by Theorem 3.5. However, this is not a total right rejective chain since $Ae_{3}A \cong P(e_{3})^{\oplus 2} \oplus P(e_{3})/P(e_{3})J^{2} \not \in \mathsf {proj}\, A$ .

In the rest of this subsection, we give constructions of total right rejective chains using Theorem 3.5. We need the following two lemmas.

Lemma 3.7 Let e and $\varepsilon $ be idempotents of A such that $\mathsf {add}\, eA \subset \mathsf {add}\, \varepsilon A$ . If $\varphi :Y \to X$ is a right $\mathsf {add}\, eA \varepsilon $ -approximation of $X \in \mathsf {proj}\, \varepsilon A \varepsilon $ , then $\varphi \otimes _{\varepsilon A \varepsilon }\varepsilon A: Y\otimes _{\varepsilon A \varepsilon }\varepsilon A \to X\otimes _{\varepsilon A \varepsilon }\varepsilon A$ is a right $\mathsf {add}\, eA$ -approximation of $X \otimes _{\varepsilon A\varepsilon }\varepsilon A$ .

Proof. Let $f \in \operatorname {Hom}\nolimits _{A}(eA, X \otimes _{\varepsilon A \varepsilon } \varepsilon A)$ . Then $f \otimes _{A}A\varepsilon : eA \otimes _{A}A\varepsilon \to X \otimes _{\varepsilon A \varepsilon } \varepsilon A\otimes _{A}A \varepsilon $ . Since $\varphi $ is a right $\mathsf {add}\, eA\varepsilon $ -approximation, there exists $\varphi ' : eA \otimes _{A} A \varepsilon \to Y$ such that $\alpha (f \otimes _{A} A \varepsilon ) = \varphi \varphi '$ , where $\alpha : X \otimes _{\varepsilon A \varepsilon } \varepsilon A \otimes _{A} A\varepsilon \to X$ is an isomorphism. Let $\beta : eA \otimes _{A} A \varepsilon \otimes _{\varepsilon A \varepsilon } \varepsilon A \to eA$ be an isomorphism via $ea \otimes b\varepsilon \otimes \varepsilon c \mapsto eab \varepsilon c$ . Then $(\alpha \otimes _{\varepsilon A \varepsilon } \varepsilon A) (f \otimes _{A} A \varepsilon \otimes _{\varepsilon A \varepsilon } \varepsilon A)= f \beta $ holds. Since $(\varphi \otimes _{\varepsilon A \varepsilon } \varepsilon A) (\varphi ' \otimes _{\varepsilon A \varepsilon } \varepsilon A) \beta ^{-1}=(\varphi \varphi ' \otimes _{\varepsilon A \varepsilon } \varepsilon A) \beta ^{-1}=(\alpha \otimes _{\varepsilon A \varepsilon } \varepsilon A ) (f \otimes _{A} A \varepsilon \otimes _{\varepsilon A \varepsilon } \varepsilon A) \beta ^{-1}=f$ , we have the assertion.▪

Lemma 3.8 Let $\varepsilon $ be an idempotent of A such that $\mathsf {add}\, \varepsilon A$ is a cosemisimple subcategory of $\mathsf {proj}\, A$ . If $\varphi : P \to Q$ is a monomorphism in $\mathsf {add}\, \varepsilon A$ , then $\operatorname {Ker}\nolimits \varphi \in \mathsf {mod}\, (A/A \varepsilon A)$ .

Proof. Let $\varphi : P \to Q$ be a monomorphism in $\mathsf {add}\, \varepsilon A$ . We show that $\operatorname {Hom}\nolimits _{A}(\varepsilon A,\operatorname {Ker}\nolimits \varphi )=0$ . Let $\psi \in \operatorname {Hom}\nolimits _{A}(\varepsilon A, \operatorname {Ker}\nolimits \varphi )$ . Then, we have a composition map

$$ \begin{align*} \varepsilon A \xrightarrow{\psi} \operatorname{Ker}\nolimits \varphi \xrightarrow{i} P \xrightarrow{\varphi} Q \end{align*} $$

such that $\varphi i \psi =0$ holds. Since $\varphi : P \to Q$ is a monomorphism in $\mathsf {add}\, \varepsilon A$ , we have ${i \psi =0}$ . Hence $\psi =0$ holds.▪

Now, we give constructions of total right rejective chains.

Theorem 3.9 Let A be an artin algebra and

(3.3) $$ \begin{align} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0 \end{align} $$

a chain of subcategories. Then the chain (3.3) is a total right rejective chain if one of the following two conditions is satisfied.

  1. (1) For each $1 \le i \le n$ , $e_{i}J\varepsilon _{i} \in \mathsf {add}\, \varepsilon _{n} A \varepsilon _{i}$ and $A \varepsilon _{n}A \in \mathsf {proj}\, A$ hold.

  2. (2) Assume that there exists $1 \le j \le n-1$ such that

    $$ \begin{align*} \operatorname{pd}\nolimits_{\varepsilon_{i}A \varepsilon_{i}} S(e_{i})\varepsilon_{i}= \begin{cases} 1&1 \le i \le j\\ 0&j+1 \le i \le n. \end{cases} \end{align*} $$
    If $\varphi : P \to Q$ is a monomorphism in $\mathsf {add}\, \varepsilon _{i+1}A\varepsilon _{i}$ , then $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, \varepsilon _{i} A \varepsilon _{i}$ for all $1 \le i \le n-1$ .

Proof. First, we assume (1). Then $e_{i}J\varepsilon _{i} \in \mathsf {add}\, \varepsilon _{n} A \varepsilon _{i}$ implies that $\operatorname {pd}\nolimits _{\varepsilon _{i}A \varepsilon _{i}} S(e_{i})\varepsilon _{i} \leq 1$ holds. Thus, by Theorem 3.5, (3.3) is a right rejective chain. Since $\varphi _{j}: e_{j}J \varepsilon _{j} \to e_{j}A \varepsilon _{j}$ is a right $\mathsf {add}\, \varepsilon _{j+1}A \varepsilon _{j}$ -approximation by Proposition 2.3(3), it follows from Lemma 3.7 that a composition map of $\varphi _{j} \otimes _{\varepsilon _{j}A \varepsilon _{j}} \varepsilon _{j}A: e_{j}J \varepsilon _{j}\otimes _{\varepsilon _{j}A \varepsilon _{j}} \varepsilon _{j}A \to e_{j}A \varepsilon _{j}\otimes _{\varepsilon _{j}A \varepsilon _{j}} \varepsilon _{j}A$ and an isomorphism $e_{j}A \varepsilon _{j} \otimes _{\varepsilon _{j}A \varepsilon _{j}}\varepsilon _{j}A \to e_{j}A$ is a right $\mathsf {add}\, \varepsilon _{j+1}A$ -approximation of $e_{j}A$ . By $e_{j}J\varepsilon _{j} \in \mathsf {add}\, \varepsilon _{n} A \varepsilon _{j}$ , we obtain that $e_{j}J \varepsilon _{j}\otimes _{\varepsilon _{j}A \varepsilon _{j}} \varepsilon _{j}A \in \mathsf {add}\, \varepsilon _{n}A$ . Take a minimal right $\mathsf {add}\, \varepsilon _{j+1}A$ -approximation $\varphi ^{\prime }_{j}: P' \to e_{j}A$ of $e_{j}A$ . Then $P' \in \mathsf {add}\, \varepsilon _{n}A$ . On the other hand, we have that $\varepsilon _{n} J \varepsilon _{n} =0$ since $\varepsilon _{n} J \varepsilon _{n} \in \mathsf {add}\, \varepsilon _{n} A \varepsilon _{n}$ and $\varepsilon _{n} A \varepsilon _{n}$ is an indecomposable $\varepsilon _{n} A \varepsilon _{n}$ -module. Thus, we obtain that $A\varepsilon _{n}A$ is a heredity ideal of A since $A \varepsilon _{n}A \in \mathsf {proj}\, A$ . By [Reference Burgess and Fuller4, Lemma 1.7], $\varphi ^{\prime }_{j}$ is a monomorphism, and hence the assertion holds.

Next, we assume (2). We show that (3.3) is a total right rejective chain by induction on n. If $n=1$ , then this is clear. Let $n \ge 2$ . If $j=1$ , then $\varepsilon _{2}A\varepsilon _{2}$ satisfies (1). Thus, we obtain a total right rejective chain of $\mathsf {proj}\, \varepsilon _{2}A\varepsilon _{2}$ . If $j>1$ , then we also have a total right rejective chain of $\mathsf {proj}\, \varepsilon _{2}A\varepsilon _{2}$ by induction hypothesis. Since $\operatorname {pd}\nolimits S(e_{1})=1$ , it follows from Proposition 2.3(3) that $\mathsf {add}\, \varepsilon _{2}A$ is a cosemisimple right rejective subcategory of $\mathsf {proj}\, A$ . Since an equivalence $\mathsf {add}\, \varepsilon _{i}A \varepsilon _{2} \simeq \mathsf {add}\, \varepsilon _{i}A$ holds for each $2 \le i \le n$ , the chain of subcategories

$$ \begin{align*} \mathsf{proj}\, A \supset \mathsf{add}\, \varepsilon_{2} A \supset \cdots \supset \mathsf{add}\, \varepsilon_{n} A \supset 0. \end{align*} $$

satisfies that $\mathsf {add}\, \varepsilon _{2}A$ is a cosemisimple right rejective subcategory of $\mathsf {proj}\, A$ and $\mathsf {add}\, \varepsilon _{2} A \supset \cdots \supset \mathsf {add}\, \varepsilon _{n} A \supset 0$ is a total right rejective chain. Thus, each minimal right $\mathsf {add}\, \varepsilon _{i}A$ -approximation $\varphi : P \to Q$ of $Q \in \mathsf {add}\, \varepsilon _{2}A$ is a monomorphism in $\mathsf {proj}\, \varepsilon _{2}A\varepsilon _{2}$ . Since $\mathsf {add}\, \varepsilon _{2} A$ is a cosemisimple subcategory of $\mathsf {proj}\, A$ , there exists $l \ge 0$ such that $\operatorname {Ker}\nolimits \varphi \cong S(e_{1})^{\oplus l}$ by Lemma 3.8. Suppose to the contrary that $l>0$ . Due to our assumption, we have that $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, A$ . Thus $S(e_{1}) \in \mathsf {proj}\, A$ , a contradiction. Hence, the proof is complete.▪

We give a naive sufficient condition for $\mathsf {proj}\, A$ to satisfy the condition (1) in Theorem 3.9.

Corollary 3.10 Keep the notation in (3.3). If $\operatorname {pd}\nolimits _{\varepsilon _{i}A \varepsilon _{i}} S(e_{i}) \varepsilon _{i} =0$ holds for each $1\le i \le n$ , then A satisfies the condition (1) in Theorem 3.9 . In particular, the chain (3.3) is a total right rejective chain.

Proof. Since $e_{i}J\varepsilon _{i} =0 \in \mathsf {add}\, \varepsilon _{n} A \varepsilon _{i}$ , it is enough to prove $A \varepsilon _{n} A \in \mathsf {proj}\, A$ . By $\operatorname {rad}\nolimits _{A}(\varepsilon _{i}A, e_{i}A) =e_{i}J\varepsilon _{i}=0$ , we obtain that $\operatorname {Hom}\nolimits _{A}(P(e_{n}), e_{i}J)=0$ holds for each $1 \le i \le n$ . Hence $A \varepsilon _{n}A =(e_{1} + \cdots +e_{n})A \varepsilon _{n}A =\varepsilon _{n}A \in \mathsf {add}\, \varepsilon _{n}A$ .▪

The following examples show that the conditions (1) and (2) in Theorem 3.9 are independent.

Example 3.11

  1. (1) Let A be the algebra defined by the quiver

    with relations $\alpha \beta $ and $\gamma \alpha $ . Then we can easily check that the complete set $(e_{1}, e_{2}, e_{3})$ of primitive orthogonal idempotents satisfies the condition (1). However, it dose not satisfy the condition (2). Indeed, $\varphi : P(e_{2})\to P(e_{3})$ is a monomorphism in $\mathsf {add}\, (e_{2}+e_{3})A$ and $\operatorname {Ker}\nolimits \varphi \not \in \mathsf {proj}\, A$ .
  2. (2) Let A be the algebra defined by the quiver

    with relations $\alpha \beta , \alpha \gamma , \delta \beta $ , and $\delta \gamma $ . Then we can check that the complete set $(e_{1}, e_{2}, e_{3})$ of primitive orthogonal idempotents satisfies the condition (2). However, it dose not satisfy the condition (1) since $e_{1}J \cong P(e_{2}) \oplus P(e_{3}) \not \in \mathsf {add}\, \varepsilon _{3}A$ .

3.2 Application

In this subsection, we construct total right rejective chains for three classes of algebras: Nakayama algebras with heredity ideals, locally hereditary algebras and algebras of global dimension at most two. Hence, these algebras are right-strongly quasi-hereditary algebras by Proposition 2.8.

The following theorem is a main result of this subsection.

Theorem 3.12 The following classes of algebras are right-strongly quasi-hereditary algebras.

  1. (1) Nakayama algebras with heredity ideals.

  2. (2) Locally hereditary algebras.

  3. (3) Algebras of global dimension at most two.

It is known that algebras in Theorem 3.12 are quasi-hereditary by [Reference Burgess and Fuller4, Proposition 2.3], [Reference Burgess and Fuller4, Proposition 1.6] and [Reference Dlab and Ringel9, Theorem 2]. Hence, Theorem 3.12 is a refinement of their results. Moreover, Theorem 3.12(3) is proven in [Reference Tsukamoto22, Theorem 4.1].

In the rest of this subsection, we give a proof of Theorem 3.12. First, we construct right rejective chains and total right rejective chains for Nakayama algebras. We say that an algebra A is a Nakayama algebra if every indecomposable projective module and every indecomposable injective module are uniserial.

Proposition 3.13 Let A be a Nakayama algebra. Then the following statements hold.

  1. (1) If A has a simple projective module or a heredity ideal, then $\mathsf {proj}\, A$ admits a total right rejective chain.

  2. (2) If $\operatorname {gldim}\nolimits A<\infty $ , then $\mathsf {proj}\, A$ admits a right rejective chain.

Proof. Let A be a Nakayama algebra and $\varepsilon $ an idempotent of A. Note that $\varepsilon A \varepsilon $ is also a Nakayama algebra since $\operatorname {Hom}\nolimits _{A}(\varepsilon A, -)$ is an exact and dense functor. We may assume that A is connected and fix a complete set $(e_{1},e_{2},\ldots , e_{n})$ of primitive orthogonal idempotents. By [Reference Anderson and Fuller2, Theorem 32.4], we can order the primitive idempotents such that there are projective covers

(3.4) $$ \begin{align} e_{i} A \to e_{i+1}J \to 0 \end{align} $$

for each $1 \le i \le n-1$ and

(3.5) $$ \begin{align} e_{n} A \to e_{1}J \to 0 \end{align} $$

if $e_{1}J \neq 0$ .

(1) If A has a simple projective module $S(e)$ , then $(1-e)A(1-e)$ is also a Nakayama algebra with a simple projective module by [Reference Anderson and Fuller2, Theorem 32.4]. It follows from Corollary 3.10 that $\mathsf {proj}\, A$ admits a total right rejective chain.

We assume that there exists a heredity ideal of A. We prove that A satisfies the condition (1) in Theorem 3.9 by induction on n. If $n=1$ , then this is clear. Let $n \geq 2$ . We assume that A has no simple projective modules. Then $e_{i}J \neq 0$ holds for each $1 \le i \le n$ . By (3.4) and (3.5), we obtain an exact sequence $e_{i} A \to e_{i+1}J \to 0$ for each $1 \le i \le n$ , where $e_{n+1}:=e_{1}$ . Since A has a heredity ideal, we may assume that $Ae_{n}A$ is a heredity ideal of A. Then we have a composition map $e_{n}A \to e_{1} J \to e_{1}A$ . Since $Ae_{n}A$ is a heredity ideal of A, this composition map is a monomorphism by [Reference Burgess and Fuller4, Lemma 1.7]. Thus, we obtain that $e_{n}A \cong e_{1} J \neq 0$ , and hence $\operatorname {pd}\nolimits S(e_{1})=1$ . Let $\varepsilon _{2}:=1-e_{1}$ . Then $\varepsilon _{2} A e_{n} A\varepsilon _{2} \in \mathsf {proj}\, \varepsilon _{2} A \varepsilon _{2}$ and $e_{n} J(\varepsilon _{2} A\varepsilon _{2})e_{n}=e_{n}J(A)e_{n}=0$ hold. Thus, $\varepsilon _{2} A\varepsilon _{2}$ is a Nakayama algebra with a heredity ideal $\varepsilon _{2} A e_{n} A\varepsilon _{2}$ . By induction hypothesis, $\varepsilon _{2} A \varepsilon _{2}$ satisfies the condition (1) in Theorem 3.9. Since $A e_{n} A \in \mathsf {proj}\, A$ and $e_{1} J \in \mathsf {add}\, e_{n} A$ , we have the assertion.

(2) Let A be an algebra of finite global dimension. By (1), we assume that A has no simple projective modules. Take an indecomposable projective module $e_{i}A$ such that its Loewy length $LL(e_{i}A)$ is maximal. By (3.4) and (3.5), we have $LL(e_{i-1}A)\ge LL(e_{i}A)-1$ . If $LL(e_{i-1}A)=LL(e_{i}A)-1$ , then $\operatorname {pd}\nolimits S(e_{i})=1$ . Thus, we assume that $LL(e_{i-1}A)>LL(e_{i}A)-1$ . Then $LL(e_{i-1}A)=LL(e_{i}A)$ by maximality of $LL(e_{i}A)$ . By $\operatorname {gldim}\nolimits A <\infty $ , A is non-self-injective. Therefore, we obtain a simple module with projective dimension exactly one by replacing $e_{i}A$ with $e_{i-1}A$ and repeating this argument. Let $S(e)$ be the simple module with projective dimension one. Since it follows from Proposition 2.3(2) that $\operatorname {gldim}\nolimits (1-e)A(1-e) \le \operatorname {gldim}\nolimits A<\infty $ , we inductively obtain that A satisfies the condition (3) in Theorem 3.5. Hence, $\mathsf {proj}\, A$ admits a right rejective chain.▪

We give applications of Proposition 3.13. In [Reference Gustafson11, Theorem], it is shown that if A is a Nakayama algebra with n simple modules and $\operatorname {gldim}\nolimits A < \infty $ , then $\operatorname {gldim}\nolimits A \le 2n-2$ . We give another proof by Proposition 3.13. Moreover, we give a refinement of [Reference Burgess and Fuller4, Proposition 2.3].

Corollary 3.14 Let A be a Nakayama algebra. Then the following statements hold.

  1. (1) The following statements are equivalent.

    1. (a) $\operatorname {gldim}\nolimits A <\infty $ .

    2. (b) A is a neat algebra.

    3. (c) $\mathsf {proj}\, A$ admits a right rejective chain.

    In this case, if A has n simple modules, then $\operatorname {gldim}\nolimits A \le 2n-2$ .

  2. (2) The following statements are equivalent.

    1. (a) A has a heredity ideal.

    2. (b) A is a quasi-hereditary algebra.

    3. (c) A is a right-strongly quasi-hereditary algebra.

Proof. (1) (c) $\Rightarrow{}$ (b) and (b) $\Rightarrow{}$ (a) follow from Theorem 3.5 and Proposition 3.2, respectively. (a) $\Rightarrow{}$ (c) follows from Proposition 3.13(2). Moreover, if A has n simple modules, then $\operatorname {gldim}\nolimits A \le 2n-2$ by Remark 2.6.

(2) (c) $\Rightarrow{}$ (b) and (b) $\Rightarrow{}$ (a) are clear. By Proposition 3.13(1), if A has a heredity ideal, then $\mathsf {proj}\, A$ has a total right rejective chain. Thus, (a) $\Rightarrow{}$ (c) holds by Proposition 2.8.▪

In Corollary 3.14, (2) $\Rightarrow{}$ (1) clearly holds. However, the converse does not hold in general. Indeed, Example 3.6 satisfies the equivalent conditions in Corollary 3.14(1), but it does not satisfy the equivalence conditions in Corollary 3.14(2).

Next, we prove that locally hereditary algebras satisfy the conditions (1) and (2) in Theorem 3.9. An algebra A is called a locally hereditary algebra if each nonzero morphism between indecomposable projective A-modules is a monomorphism (see [Reference Martinez-Villa17] for details). We need the following lemma.

Lemma 3.15 Let A be a nonsemisimple locally hereditary algebra. Then the following statements hold.

  1. (1) Let e be a primitive idempotent of A and $\varepsilon :=1-e$ . Assume that $\mathsf {add}\, \varepsilon A$ is a cosemisimple subcategory of $\mathsf {proj}\, A$ . If $\varphi : P \to Q$ is a monomorphism in $\mathsf {add}\, \varepsilon A$ , then $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, A$ .

  2. (2) $\operatorname {soc}\nolimits A \in \mathsf {proj}\, A$ holds. In particular, there exists a simple projective module.

  3. (3) There exists a simple module such that its projective dimension is exactly one.

Proof. (1) Let $\varphi : P \to Q$ be a monomorphism in $\mathsf {add}\, \varepsilon A$ . Since $\mathsf {add}\, \varepsilon A$ is a cosemisimple subcategory of $\mathsf {proj}\, A$ , there exists $l \ge 0$ such that $\operatorname {Ker}\nolimits \varphi \cong S(e)^{\oplus l}$ by Lemma 3.8. Thus $\rho ^{\oplus l}: P(e)^{\oplus l} \to \operatorname {Ker}\nolimits \varphi $ is a projective cover of $\operatorname {Ker}\nolimits \varphi $ , where $\rho : P(e) \to S(e)$ is a projective cover of $S(e)$ . On the other hand, there exists an indecomposable direct summand $P'$ of P such that $S(e)$ is a direct summand of $\operatorname {soc}\nolimits P'$ . Let $\iota : S(e) \to P'$ be the inclusion. Then a composition map $\iota \rho : P(e) \to P'$ is a nonzero morphism between indecomposable projective modules. Since $\iota \rho $ is a monomorphism, $\rho $ is an isomorphism. Thus $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, A$ .

(2) Let $\operatorname {soc}\nolimits A =S(e_{i_{1}}) \oplus \cdots \oplus S(e_{i_{t}})$ , where $S(e_{i_{j}})$ is a simple module. For each $1 \le j \le t$ , there exists an indecomposable projective module $P_{j}$ such that $S(e_{i_{j}})$ is a direct summand of $\operatorname {soc}\nolimits P_{j}$ . Since a composition map of a projective cover $p : P(e_{i_{j}}) \to S(e_{i_{j}})$ and the inclusion $S(e_{i_{j}}) \to P_{j}$ is a monomorphism, p is an isomorphism. Thus $\operatorname {soc}\nolimits A \in \mathsf {proj}\, A$ .

(3) We show that there exists an indecomposable projective module P such that its Loewy length $LL (P)$ is two. If $LL(A)=2$ , then this is clear. We assume that $LL (A)\ge 3$ . Let Q be an indecomposable projective module with $LL (Q)=:l \ge 3$ . Then $QJ^{l-2}/QJ^{l-1}\cong S(e_{i_{1}}) \oplus \cdots \oplus S(e_{i_{s}}) \neq 0$ . Note that there exists $1 \le j \le s$ such that $S(e_{i_{j}}) \not \in \mathsf {proj}\, A$ . Thus, $LL(P(e_{i_{j}})) \ge 2$ . Since a composition map of $P(e_{i_{j}}) \to QJ^{l-2}$ and $QJ^{l-2} \to Q$ is a monomorphism, so is $P(e_{i_{j}}) \to QJ^{l-2}$ . Since $LL( P(e_{i_{j}}))\le 2$ , $P(e_{i_{j}})$ is a desired projective module. Thus, we obtain $\operatorname {pd}\nolimits S(e_{i_{j}})=1$ since it follows from (2) that $\operatorname {soc}\nolimits P(e_{i_{j}}) \in \mathsf {proj}\, A$ .▪

Now, we state the following proposition.

Proposition 3.16 Any locally hereditary algebra satisfies the conditions (1) and (2) in Theorem 3.9 . Namely, if A is a locally hereditary algebra, then $\mathsf {proj}\, A$ admits a total right rejective chain.

Proof. Let A be a locally hereditary algebra. If A is semisimple, then it clearly satisfies the conditions (1) and (2) in Theorem 3.9. Thus, we assume that A is not semisimple. First, we show that A satisfies the condition (1). By Lemma 3.15(2), there exists a simple module $S(e)$ such that $\operatorname {pd}\nolimits S(e)=0$ . Let $\varepsilon :=1-e$ . Then $\varepsilon A \varepsilon $ is also a locally hereditary algebra by an equivalence of categories $\mathsf {add}\, \varepsilon A \simeq \mathsf {proj}\, \varepsilon A\varepsilon $ . Due to Corollary 3.10, A inductively satisfies the condition (1) in Theorem 3.9.

Next, we show that A satisfies the condition (2). By Lemma 3.15(3), there exists a simple module $S(e)$ with $\operatorname {pd}\nolimits S(e)=1$ . Let $\varepsilon :=1-e$ . Then $\mathsf {add}\, \varepsilon A$ is a cosemisimple subcategory of $\mathsf {proj}\, A$ by Proposition 2.3(3). For each monomorphism $\varphi : P\to Q$ in $\mathsf {add}\, \varepsilon A$ , we obtain that $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, A$ by Lemma 3.15(1). Since $\varepsilon A \varepsilon $ is also a locally hereditary algebra, A inductively satisfies the condition (2) in Theorem 3.9.▪

Finally, we show the following proposition. This proposition is proven in [Reference Iyama15, Theorem 3.6] and [Reference Tsukamoto22, Theorem 4.1]. In this paper, we give a proof using Theorem 3.9.

Proposition 3.17 If $\operatorname {gldim}\nolimits A \leq 2$ , then $\mathsf {proj}\, A$ admits a total right rejective chain.

Proof. If $\operatorname {gldim}\nolimits A \le 1$ , then A is a locally hereditary algebra. Thus, the assertion follows from Proposition 3.16. We assume that $\operatorname {gldim}\nolimits A =2$ . Let S be an A-module such that its length $\ell (S)$ is minimal among A-modules with the projective dimension one. We show that S is a simple module. Suppose to the contrary that S is not simple. Then there exists an exact sequence $0 \to S' \to S \to S/S' \to 0$ such that $S' \neq 0$ and $S/S' \neq 0$ . Since $\operatorname {pd}\nolimits S'\le \max \{\operatorname {pd}\nolimits S, \operatorname {pd}\nolimits S/S' -1\} = 1$ , it follows from the assumption on S that $\operatorname {pd}\nolimits S'=0$ . Thus, we have $\operatorname {pd}\nolimits S/S' \le \max \{\operatorname {pd}\nolimits S, \operatorname {pd}\nolimits S' +1\} = 1$ . By the assumption on S, we have $\operatorname {pd}\nolimits S/S'=0$ , a contradiction. Thus S is a simple module. We put $S(e):=S$ and $\varepsilon :=1-e$ . Then $\mathsf {add}\, \varepsilon A$ is a cosemisimple right rejective subcategory of $\mathsf {proj}\, A$ by Proposition 2.3(3). Let $\varphi : P \to Q$ be a monomorphism in $\mathsf {add}\, \varepsilon A$ . Then, we have an exact sequence in $\mathsf {mod}\, A$ .

$$ \begin{align*} \operatorname{Ker}\nolimits \varphi \to P \to Q \to \operatorname{Cok}\nolimits \varphi \to 0. \end{align*} $$

By $\operatorname {gldim}\nolimits A = 2$ , we have that $\operatorname {Ker}\nolimits \varphi \in \mathsf {proj}\, A$ . Since $\operatorname {gldim}\nolimits \varepsilon A \varepsilon \le \operatorname {gldim}\nolimits A=2$ by Proposition 2.3(2), we inductively obtain that A satisfies the condition (2) in Theorem 3.9. Hence, we have the assertion.▪

Now we are ready to prove Theorem 3.12.

Proof of Theorem 3.12

Let A be an algebra in Theorem 3.12. By Proposition 2.8, it is enough to show that there exists a total right rejective chain of $\mathsf {proj}\, A$ . Thus, the assertion follows from Propositions 3.13, 3.16 and 3.17.▪

Acknowledgment

The author would like to express her deep gratitude to Takahide Adachi and Ryoichi Kase for stimulating discussions and helpful comments.

References

Ágoston, I., Dlab, V., and Wakamatsu, T., Neat algebras. Comm. Algebra 19(1991), no. 2, 433442.Google Scholar
Anderson, F. W. and Fuller, K. R., Rings and categories of modules . Graduate Texts in Mathematics, 13, Springer-Verlag, Berlin, 1992.Google Scholar
Auslander, M., Platzeck, M. I., and Todorov, G., Homological theory of idempotent ideals. Trans. Amer. Math. Soc. 332(1992), no. 2, 667692.CrossRefGoogle Scholar
Burgess, W. D. and Fuller, K. R., On quasihereditary rings. Proc. Amer. Math. Soc. 106(1989), no. 2, 321328.CrossRefGoogle Scholar
Cline, E., Parshall, B., and Scott, L., Finite-dimensional algebras and highest weight categories, J. Reine Angew. Math. 391(1988), 8599.Google Scholar
Cline, E., Parshall, B., and Scott, L., Stratifying endomorphism algebras. Mem. Amer. Math. Soc. 124(1996), no. 591, viii+119pp.Google Scholar
Conde, T., The quasihereditary structure of the Auslander-Dlab-Ringel algebra. J. Algebra 460(2016), 181202.CrossRefGoogle Scholar
Coulembier, K., Ringel duality and Auslander-Dlab-Ringel algebras, J. Pure Appl. Algebra 222(2018), no. 12, 38313848.CrossRefGoogle Scholar
Dlab, V. and Ringel, C. M., Quasi-hereditary algebras. Illinois J. Math. 33(1989), no. 2, 280291.CrossRefGoogle Scholar
Geiss, C., Leclerc, B., and Schröer, J., Cluster algebra structures and semicanonical bases for unipotentgroups. arXiv:math/0703039v4 Google Scholar
Gustafson, W. H., Global dimension in serial rings. J. Algebra 97(1985), no. 1, 1416.CrossRefGoogle Scholar
Iyama, O., A generalization of rejection lemma of Drozd-Kirichenko. J. Math. Soc. Japan 50(1998), no. 3, 697718.Google Scholar
Iyama, O., Finiteness of representation dimension. Proc. Amer. Math. Soc. 131(2003), no. 4, 10111014.CrossRefGoogle Scholar
Iyama, O., $\tau$ -categories. II. Nakayama pairs and rejective subcategories. Algebr. Represent. Theory 8(2005), no. 4, 449477.10.1007/s10468-005-0969-4CrossRefGoogle Scholar
Iyama, O., Rejective subcategories of artin algebras and orders. arXiv:math/0311281 Google Scholar
Kalck, M. and Karmazyn, J., Ringel duality for certain strongly quasi-hereditary algebras. Eur. J. Math. 4(2018), no. 3, 11001140.CrossRefGoogle Scholar
Martinez-Villa, R., Algebras stably equivalent to ℓ-hereditary. Representation theory, II (Proc. Second Internat. Conf., Carleton Univ., Ottawa, Ont., 1979), Lecture Notes in Math., no. 832, Springer, Berlin, 1980, 396431.CrossRefGoogle Scholar
Ringel, C. M., Iyama’s finiteness theorem via strongly quasi-hereditary algebras. J. Pure Appl. Algebra 214(2010), no. 9, 16871692.10.1016/j.jpaa.2009.12.012CrossRefGoogle Scholar
Rump, W., The category of lattices over a lattice-finite ring. Algebr. Represent. Theory 8(2005), no. 3, 323345.CrossRefGoogle Scholar
Scott, L., Simulating algebraic geometry with algebra. I. The algebraic theory of derived categories. In: The Arcata Conference on Representations of Finite Groups (Arcata, Calif., 1986), Proc. Sympos. Pure Math., 47, Part 2, Amer. Math. Soc., Providence, RI, 1987, pp. 271281.CrossRefGoogle Scholar
Tsukamoto, M., On an upper bound for the global dimension of Auslander–Dlab–Ringel algebras. Arch. Math. (Basel) 112(2019), no. 1, 4151.CrossRefGoogle Scholar
Tsukamoto, M., Strongly quasi-hereditary algebras and rejective subcategories. Nagoya Math. J. 237(2020), 1038.CrossRefGoogle Scholar