On T-Structures and Torsion Theories Induced by Compact Objects
On T-Structures and Torsion Theories Induced by Compact Objects
www.elsevier.com/locate/jpaa
Abstract
First, we show that a compact object C in a triangulated category, which satis5es suitable
conditions, induces a t-structure. Second, in an abelian category we show that a complex P · of
small projective objects of term length two, which satis5es suitable conditions, induces a torsion
theory. In the case of module categories, using a torsion theory, we give equivalent conditions
for P · to be a tilting complex. Finally, in the case of artin algebras, we give a one-to-one
correspondence between tilting complexes of term length two and torsion theories with certain
conditions. c 2002 Elsevier Science B.V. All rights reserved.
0. Introduction
In the representation theory of 5nite dimensional algebras, torsion theories were stud-
ied by several authors in connection with classical tilting modules. For these torsion
theories, there are equivalences between torsion (resp., torsion-free) classes and torsion-
free (resp., torsion) classes, which is known as Theorem of Brenner and Butler ([5]).
One of the authors gave a one-to-one correspondence between classical tilting modules
and torsion theories with certain conditions [7,8]. But in the case of a self-injective
algebra A, tilting modules are essentially isomorphic to A. In [12], Rickard introduced
the notion of tilting complexes as a generalization of tilting modules, and showed that
∗Corresponding author.
E-mail addresses: [email protected] (M. Hoshino), [email protected] (Y. Kato),
[email protected] (J.-I. Miyachi).
0022-4049/02/$ - see front matter c 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 2 2 - 4 0 4 9 ( 0 1 ) 0 0 0 1 2 - 3
16 M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35
In this section, we deal with a triangulated category T and its full subcategory C.
We will call C admissible abelian provided that HomT (C; C[n])=0 for n ¡ 0, and that
all morphisms in C are C-admissible in the sense of [2, 1.2.3]. In this case, according
to [2, Proposition 1:2:4] C is an abelian category. A triangulated category T is said to
contain direct sums if direct sums of objects indexed by any set exist in T. An object
C of T is called compact if HomT (C; −) commutes with direct sums. Furthermore, a
collection S of compact objects of T is called a generating set provided that X = 0
whenever HomT (S; X )=0, and that S is stable under suspension (see [10] for details).
For an object C ∈ T and an integer n, we denote by T¿n (C) (resp., T6n (C)) the
full subcategory of T consisting of X ∈ T with HomT (C; X [i]) = 0 for i ¡ n (resp.,
i ¿ n), and set T0 (C) = T60 (C) ∩ T¿0 (C).
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 17
HomT (X ¿r ; Y ) ∼
= lim HomT (Xn ; Y )
←−
∼
= HomT (X; Y ):
18 M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35
Denition 1.2 (Beilinson et al. [2]). Let T be a triangulated category. For full sub-
categories T60 and T¿0 , (T60 ; T¿0 ) is called a t-structure on T provided that
(i) HomT (T60 ; T¿1 ) = 0,
(ii) T60 ⊂ T61 and T¿0 ⊃ T¿1 ,
(iii) for any X ∈ T, there exists a distinguished triangle
X → X → X →
with X ∈ T60 and X ∈ T¿1 ,
where T6n = T60 [ − n] and T¿n = T¿0 [ − n].
A t-structure (T60 ; T¿0 ) on T is called non-degenerate if n∈Z T
6n
=
¿n
n∈Z T = {0}.
Theorem 1.3. Let T be a triangulated category which contains direct sums; C a com-
pact object satisfying HomT (C; C[n]) = 0 for n ¿ 0; and B = End T (C)op . If {C[i] | i ∈
Z} is a generating set; then the following hold:
(1) (T60 (C); T¿0 (C)) is a non-degenerate t-structure on T.
(2) T0 (C) is admissible abelian.
(3) The functor
Proof. (1) For any object X ∈ T60 (C), we take an object X ¿1 ∈ T¿1 (C) and a
morphism ¿1 : X → X ¿1 satisfying the conditions of Proposition 1.1. Then for any
Y ∈ T¿1 (C), by Proposition 1.1(ii), we have
HomT (X ¿1 ; Y ) ∼
= HomT (X; Y ):
By Proposition 1.1(i), X ∈ T60 (C) implies that HomT (C; X ¿1 [i]) = 0 for all i ∈ Z.
Since {C[i] | i ∈ Z} is a generating set, we have X ¿1 = 0, and hence HomT (X; Y ) = 0.
It is easy to see that T60 (C) ⊂ T61 (C) and T¿0 (C) ⊃ T¿1 (C). For any object
Z ∈ T, we take an object Z ¿1 ∈ T¿1 (C) and a morphism ¿1 : Z → Z ¿1 satisfying
the conditions of Proposition 1.1, and embed ¿1 in a distinguished triangle
Z → Z → Z ¿1 → :
Applying HomT (C; −) to the above triangle, by Proposition 1.1(i), we have Z ∈
T60 (C). Since {C[i] | i ∈ Z} is a generating set, it is easy to see that (T60 (C);
T¿0 (C)) is non-degenerate.
(2) Since T0 (C) is the core of the t-structure (T60 (C); T¿0 (C)), the assertion
follows by [2, Theorem 1:3:6].
(3) Step 1: According to [2, Proposition 1:2:2], the short exact sequences in T0 (C)
are just the distinguished triangles
X →Y →Z →
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 19
with X; Y and Z belonging to T0 (C). It follows that HomT (C; −) : T0 (C) → Mod B
is exact. Let M ∈ Mod B and take a free presentation P1 → P0 → M → 0. We take
C = C ¿0 ∈ T0 (C) and = ¿0 : C → C satisfying the conditions of Proposition 1.1.
Then there exist sets I; J and a collection of morphisms hij : C → C such that
P1 −−−−−−−−−−−−−−→ P0
Hom(C;hij )
HomT (C; C )(J ) −−−−−−−→ HomT (C; C )(I )
ij
is commutative, where the vertical arrows are isomorphisms. We take an exact sequence
in T0 (C)
hij
C (J ) −−→C (I ) → X → 0:
ij
Since C is compact, by the exactness of HomT (C; −), we have HomT (C; X ) ∼
= M.
Step 2: We show that HomT (C; −) reJects isomorphisms. Let
u
X →Y → Z →
be a distinguished triangle in T with X; Y ∈ T0 (C) and with HomT (C; u) an isomor-
phism. Then, by applying HomT (C; −), we get HomT (C; Z[n]) = 0 for all n ∈ Z, and
hence Z = 0. It follows that u is an isomorphism.
Next, we show that HomT (C; −) is faithful. Let v : X → Y be a morphism in
T0 (C) with HomT (C; v) = 0. By the exactness of HomT (C; −), HomT (C; Im v) ∼ =
Im HomT (C; v) = 0. Since Im v ∈ T0 (C), we have HomT (C; Im v[n]) = 0 for all
n ∈ Z, and hence Im v = 0 and v = 0.
Let M be the full subcategory of T0 (C) consisting of objects X such that there
exists an exact sequence C1 → C0 → X → 0 in T0 (C), where C0 ; C1 are direct sums
of copies of C . Since HomT (C; −) is faithful, by the same technique as in Step 1,
it is not hard to see that HomT (C; −)|M is full dense, and hence an equivalence. It
remains to show that M = T0 (C). For an object X ∈ T0 (C), we have a commutative
diagram
(J ) Hom(C;f) Hom(C;g)
HomT (C; (C; C (I ) ) −−−−→ HomT (C; X ) −−→ 0
C ) −−−−→ HomT
HomT (C;1 ) HomT (C;0 )
Hom(C;f )
HomT (C; C (J ) ) −−−−−→ HomT (C; C (I ) )
with the top row being exact and with the vertical arrows being isomorphisms. And
we have a commutative diagram in T
f g
C(J ) −−→ C(I ) −−→ X
1 0
f
C (J ) −−→ C (I )
with gf = 0. By Proposition 1.1(ii), there exists h : C (I ) → X such that g = h0 . Since
HomT (C; hf ) = 0, we have hf = 0. Then there exists w : Cok f → X such that
20 M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35
Remark 1.4. Under the condition of Theorem 1.3, according to [2, Proposition 1:3:3]
there exists a functor (−)¿n : T → T¿n (C) (resp., (−)6n : T → T6n (C)) which is
the right (resp., left) adjoint of the natural embedding functor T¿n (C) → T (resp.,
T6n (C) → T).
Corollary 1.5. Let A be an abelian category satisfying the condition Ab4 (i.e. direct
sums of exact sequences are exact) and T · a bounded complex of small projective
objects of A satisfying
(i) {T ·[i] | i ∈ Z} is a generating set for D(A);
(ii) HomD(A) (T ·; T ·[i]) = 0 for i = 0.
If either of the following conditions (1) or (2) is satis;ed; then we have an equivalence
of triangulated categories
D(A)b (T ·) ∼
= Db (Mod B);
where B = End D(A) (T ·)op .
(1) A has enough projectives.
(2) A has enough injectives and D(A)¿0 (T ·) ⊂ D+ (A).
Moreover; if D(A)0 (T ·) ⊂ Db (A); then we have an equivalence
Db (A) ∼
= Db (Mod B):
Proof. According to [3, Corollary 1:7] D(A) contains direct sums. Since T · is a
bounded complex of small projective objects of A, T · is a compact object in D(A). By
Theorem 1.3 D(A) has a t-structure (D(A)60 (T ·); D(A)¿0 (T ·)), and
HomD(A) (T ·; −) : D(A)0 (T ·) → Mod B is an equivalence.
(1) By the construction of X ¿r in Proposition 1.1, D− (A) also has a t-structure
(D− (A)60 (T ·); D− (A)¿0 (T ·)) and hence by Theorem 1.3(3) we have D− (A)0
(T ·)=D(A)0 (T ·). According to [12, Proposition 10:1] we have a fully faithful @-functor
F : D− (Mod B) → D− (A). Also, since F (B) = T ·, F sends B-modules to objects in
D(A)0 (T ·). Then we have a fully faithful @-functor
F : Db (Mod B) → D(A);
which sends B-modules to objects in D(A)0 (T ·). For any X ∈ D(A)b (T ·), there exist
integers m 6 n such that X ∈ D(A)[m; n] (T ·). Let l = n − m. If l = 0, then there exist,
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 21
X 6n−1 → X → X ¿n →
with X ¿n ∈ D(A)[n] (T ·) and X 6n−1 ∈ D(A)[m; n−1] (T ·). Since F is full, by induction
on l, there exists U ·∈ Db (Mod B) such that X ∼= F(U ·).
(2) By the assumption, D (A) also has a t-structure (D+ (A)60 (T ·); D+ (A)¿0
+
(T ·)). Thus D+ (A)b (T ·) = D(A)b (T ·), and hence D+ (A)0 (T ·) = D(A)0 (T ·). By [2,
Section 3] we have a @-functor real : Db (D(A)0 (T ·)) → D+ (A), and then we have
a @-functor
F : Db (Mod B) → D(A);
which sends B-modules to objects in D(A)0 (T ·). Let f ∈ HomD(A) (X ·; Y ·[n]) with
X ·; Y · ∈ D(A)0 (T ·) and n ¿ 0. Take a distinguished triangle in D+ (A)
X1· → V ·→ X · →
t
0 → X1· → V ·→ X · → 0:
t
Remark 2.1. It is easy to see that P · is a compact object of D(A), and we have
X(P ·) = D(A)60 (P ·) ∩ A and Y(P ·) = D(A)¿1 (P ·) ∩ A.
Lemma 2.5. For any X · ∈ D(A) and n ∈ Z; we have a functorial exact sequence
→ HomD(A) (P ·; Hn (X ·)) → 0:
Moreover; the above short exact sequence commutes with direct sums.
Proof. For X ·[n] ∈ D(A), applying HomD(A) (−; X ·[n]) to a distinguished triangle
d−1
P −1 →
P
P 0 → P · →;
we have a short exact sequence
0 → Cok(HomD(A) (d−1 · · ·
P ; X [n − 1])) → Hom D(A) (P ; X [n])
→ Ker(HomD(A) (d−1 ·
P ; X [n])) → 0:
Ker(HomD(A) (d−1 · ∼ −1 n ·
P ; X [n])) = Ker(Hom A (dP ; H (X )))
= HomD(A) (P ·; Hn (X ·));
∼
Since the P i are small objects, the above short exact sequence commutes with direct
sums.
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 23
Proof. (1) ⇒ (2). For any X ∈ X(P ·) ∩ Y(P ·), by Lemma 2.3(1), HomD(A) (P ·;
X [n]) = 0 for all n ∈ Z and hence X = 0.
(2) ⇒ (1). Let X · ∈ D(A) with HomD(A) (P ·; X ·[n]) = 0 for all n ∈ Z. Then by
Lemma 2.5, Hn (X ·) ∈ X(P ·) ∩ Y(P ·) = {0}.
Remark 2.9. Let A be an abelian category and (X; Y) a torsion theory for A. Then
for any Z ∈ A, the following hold:
(1) Z ∈ X if and only if HomA (Z; Y) = 0.
(2) Z ∈ Y if and only if HomA (X; Z) = 0.
d−1
Theorem 2.10. The following are equivalent for a complex P · : P −1 −→
P
P 0 with the
i
P being small projective objects of A:
(1) {P ·[i] | i ∈ Z} is a generating set for D(A) and HomD(A) (P ·; P ·[i]) = 0 for all
i ¿ 0.
(2) X(P ·) ∩ Y(P ·) = {0} and H0 (P ·) ∈ X(P ·).
(3) X(P ·) ∩ Y(P ·) = {0} and $(X ) ∈ X(P ·); &(X ) ∈ Y(P ·) for all X ∈ A.
(4) (X(P ·); Y(P ·)) is a torsion theory for A.
H−1 (X ·)[1] → X · → H0 (X ·) → :
0 → '6−1 (X ·) → X · → '¿−1 (X ·) → 0;
0 → '¡0 ('¿−1 (X ·)) → '¿−1 (X ·) → '¿0
(X ·) → 0:
Also, '6−1 (X ·) ∼
= H−1 (X ·)[1]; '¡0
('¿−1 (X ·)) ∼
= 0 and '¿0 (X ·) ∼
= H0 (X ·) in
D(A). Thus we get a desired distinguished triangle in D(A).
Lemma 2.13. Assume X(P ·) ∩ Y(P ·) = {0}. Then for any X · ∈ D(A); the following
are equivalent:
(1) X · ∈ C(P ·).
(2) Hn (X ·) = 0 for n ¿ 0 and n ¡ − 1; H0 (X ·) ∈ X(P ·) and H−1 (X ·) ∈ Y(P ·).
is an equivalence.
(3) (Y(P ·)[1]; X(P ·)) is a torsion theory for C(P ·).
Proof. (1) and (2) According to Theorem 2.10, Theorem 1.3 can be applied.
(3) Note 5rst that by Lemma 2.13 we have X(P ·) ⊂ C(P ·) and Y(P ·)[1] ⊂
C(P ·). Also, it is trivial that HomD(A) (Y(P ·)[1]; X(P ·)) = 0. Let X · ∈ C(P ·). Then
by Lemmas 2.12 and 2.13 we have a distinguished triangle in D(A) of the form
H−1 (X ·)[1] → X · → H0 (X ·) → :
0 → H−1 (X ·)[1] → X · → H0 (X ·) → 0
is exact. Thus by Remark 2.14 (Y(P ·)[1]; X(P ·)) is a torsion theory for
C(P ·).
Db (A) ∼
= Db (Mod B):
Proof. Let X · ∈ D(A). According to Lemma 2.5 and Theorem 2.10, it is easy to see
that if X · belongs to D(A)¿0 (P ·) (resp., C(P ·)), then Hn (X ·) = 0 for n ¡ − 1 (resp.,
n ¡ − 1 and n ¿ 0). Thus we have
the A-dual functors HomA (−; A) and set * = D ◦ (−)∗ . Throughtout this section, P · is
d−1
a complex P −1 →P
P 0 with the P i being 5nitely generated projective A-modules.
It is well known that, in a module category, the small projective objects are just the
5nitely generated projective modules. In the following, we deal with the case where
A = Mod A and use the same notation as in Section 2.
HomD(Mod A) (P ·; X [1]) ∼
= H1 ((P ·)∗ ) ⊗A X:
Proof. We have
HomD(Mod A) (P ·; X [1]) ∼
= HomK(Mod A) (P ·; X [1])
= H1 (HomA· (P ·; X ))
∼
= H1 ((P ·)∗ ) ⊗A X:
∼
Proof. This is due essentially to Auslander [1]. We have an exact sequence in Mod A
0 → H−1 (P ·) → P −1 → P 0 → H0 (P ·) → 0
with the P i 5nitely generated projective, and an exact sequence in Mod Aop
+M : M ⊗A X → HomA (M ∗ ; X ); m ⊗ x → (h → h(m)x);
P 0∗
⊗A X −−→ P −1∗⊗A X −−→H1 ((P ·)∗ ) ⊗A X −−→ 0
+P0∗ +P−1∗
HomA (P 0∗∗ ; X ) −−→ HomA (P −1∗∗ ; X ) −−→ HomA (H−1 (P ·); X )
with the top row exact. Since the +Pi∗ are isomorphisms, Ext 1A (H0 (P ·); X ) is embedded
in H1 ((P ·)∗ ) ⊗A X . The assertion follows by Lemma 3.2.
(2) Let X ∈ Mod A. For any Y ∈ Mod A, we have a functorial homomorphism
H0 (P ·
)∗ ⊗A X −−→ P 0∗ ⊗A X −−→ P −1∗⊗A X
-P 0 -P−1
0 ·
0 −−→ HomA (H (P ); X ) −−→ HomA (P ; X ) −−→ HomA (P −1 ; X )
0
with the bottom row exact. Since the -Pi are isomorphisms, Tor A1 (H1 ((P ·)∗ ); X ) is a
homomorphic image of HomA (H0 (P ·); X ). The assertion follows by Lemma 3.2.
d−1
Theorem 3.5. The following are equivalent for a complex P · : P −1 →
P
P 0 with the P i
being ;nitely generated projective A-modules:
(1) X(P ·) ∩ Y(P ·) = {0} and H0 (P ·) ∈ X(P ·).
(2) X(P ·) ∩ Y(P ·) = {0} and $(X ) ∈ X(P ·); &(X ) ∈ Y(P ·) for all X ∈ Mod A.
(3) (X(P ·); Y(P ·)) is a torsion theory for Mod A.
(4) X(P ·) consists of the modules generated by H0 (P ·) and Y(P ·) consists of the
modules cogenerated by H−1 (*(P ·)).
28 M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35
It is obvious that HomA (f; H−1 (*(P ·))) is surjective. Also, by Lemmas 3:3(3) and
3:4(2) we have Ext 1A (Imf; H−1 (*(P ·))) = 0. Applying HomA (−; H−1 (*(P ·))) to the
canonical exact sequence
0 → Ker f → X → Im f → 0;
we get HomA (Ker f; H−1 (*(P ·)))=0. Thus Ker f ∈ X(P ·)∩Y(P ·) and hence Ker f =
0.
(4) ⇒ (1). By Lemma 3.3(2).
Corollary 3.7. For any tilting complexes P1· : P1−1 → P10 ; P2· : P2−1 → P20 for A of
term length two; the following are equivalent:
(1) (X(P1·); Y(P1·)) = (X(P2·); Y(P2·)).
(2) add(P1·) = add(P2·) in Kb (Proj A).
Proof. (1) ⇒ (2). It follows by Corollary 3.6 that Q · = P1· ⊕ P2· is a tilting complex
such that (X(Q ·); Y(Q ·))=(X(Pi·); Y(Pi·)) (i=1; 2). Let B=End D(Mod A) (Q ·)op and for
i=1; 2 denote by ei the composite of canonical homomorphisms Q · → Pi· → Q · . Then,
for i = 1; 2 we have an equivalence D− (Mod B) → D− (Mod ei Bei ) which sends Bei to
ei Bei , so that the Bei are tilting complexes for B, i.e. projective generators for Mod B.
It follows by Morita Theory that add B = add Bei in Mod B. Thus add(P1·) = add(P2·)
in Kb (Proj A).
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 29
(2) ⇒ (1). It is obviously deduced that add(H−1 (*(P1·))) = add(H−1 (*(P2·))) and
add(H0 (P1·)) = add(H0 (P2·)).
−1
dP
Theorem 3.8. Let P · be a complex P −1 → P 0 with the P i being ;nitely generated
projective A-modules. Assume X(P ·) ∩ Y(P ·) = {0} and H0 (P ·) ∈ X(P ·). Then the
following hold:
(1) {P ·[i] | i ∈ Z} is a generating set for D(Mod A).
(2) C(P ·) is admissible abelian.
(3) (Y(P ·)[1]; X(P ·)) is a torsion theory for C(P ·).
(4) The functor
Hom D(Mod A) (P ·; −) : C(P ·) → Mod B
is an equivalence.
Example 3.10 (cf. Hoshino and Kato [9]). Let A be a 5nite dimensional algebra over
a 5eld k given by a quiver
1 −−→
2
3 4
4 −−→ 3
5
with relations 4 = 54 = 35 = 3 = 0. For each vertex i, we denote by S(i); P(i)
the corresponding simple and indecomposable projective left A-modules, respectively.
De5ne a complex P · as the mapping cone of the homomorphism
f 0 0 0
d−1
P = : P(2)2 ⊕ P(4)2 → P(1) ⊕ P(3);
0 0 g 0
where f and g denote the right multiplications of and 5, respectively. Then P · is
not a tilting complex. However, P · satis5es the assumption of Theorem 3.8 and hence
we have an equivalence of abelian categories
Hom (P ·; −) : C(P ·) → Mod B;
D(Mod A)
−1
dP
Throughout this section, P · : P −1 → P 0 is assumed to be a tilting complex. Then
there exists an equivalence of triangulated categories
F : D− (Mod B) → D− (Mod A)
such that F(B) = P ·. Let G : D− (Mod A) → D− (Mod B) be a quasi-inverse of F. For
any n ∈ Z, we have ring homomorphisms
B → End A (Hn (P ·))op and B → End A (Hn ((P ·)∗ )):
In particular, H0 (P ·) is an A-B-bimodule and H1 ((P ·)∗ ) is a B-A-bimodule.
= HomD(Mod A) (P ·; A[i])
∼
= Hi ((P ·)∗ );
∼
we have Q · ∼
= '61 (Q ·) in Kb (proj B). Also, since
Hi (HomB· (Q ·; B)) ∼
= HomD(Mod B) (Q ·; B[i])
= Hi (P ·);
∼
we have HomB· (Q ·; B) ∼
= '60 (HomB· (Q ·; B)) in Kb (proj Bop ) and Q · ∼
= '¿0 (Q ·) in
b i
K (proj B). Thus, we can assume Q = 0 for i ¿ 1 and i ¡ 0.
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 31
= H0 (HomB· (Q ·; B)) ⊗B M
∼
= H0 (P ·) ⊗B M;
∼
= HomD(Mod B) (Q ·; M [ − 1])
H−1 (F(M )) ∼
= H−1 (Hom · (Q ·; M ))
∼
B
= HomB (H1 (Q ·); M )
∼
= HomB (H 1 ((P ·)∗ ); M ):
∼
Proof. (1) According to Lemmas 3.2 and 4.2, we can apply Corollary 3.6 for a tilting
complex Q · to conclude that (U(P ·); V(P ·)) is a torsion theory for Mod B.
(2) For any X ∈ X(P ·), by Lemmas 2:13; 4:1(1) and 4:3(3) we have
Hom (H1 ((P ·)∗ ); Hom (H0 (P ·); X )) ∼
B A = H−1 (F(G(X )))
∼
= H−1 (X )
= 0:
32 M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35
Also, since by Lemma 3.2(1) and Corollary 3.6 H1 ((P ·)∗ )⊗A H0 (P ·)=0, H1 ((P ·)∗ )⊗A
H0 (P ·) ⊗B M = 0 for all M ∈ V(P ·). The last assertion follows by Lemmas 2:13; 4:1
and 4:3.
(3) For any Y ∈ Y(P ·), by Lemmas 2:13; 3:1; 4:1(1) and 4:3(2) we have
H0 (P ·) ⊗B H1 ((P ·)∗ ) ⊗A Y ∼
= H0 (F(G(Y [1])))
∼
= H0 (Y [1])
= 0:
Also, since H1 ((P ·)∗ ) ⊗A H0 (P ·) = 0, for any N ∈ U(P ·) we have
HomA (H0 (P ·); HomB (H1 ((P ·)∗ ); N )) ∼
= HomB (H1 ((P ·)∗ ) ⊗A H0 (P ·); N )
= 0:
The last assertion follows by Lemmas 2:13; 4:1 and 4:3.
Denition 4.5. Let (U; V) be a torsion theory for an abelian category A. Then (U; V)
is called splitting if Ext 1A (V; U) = 0.
For a left A-module M , we denote by proj dimA M (resp., inj dimA M ) the projective
(resp., the injective) dimension of M .
Proposition 4.6. The torsion theory (U(P ·); V(P ·)) for Mod B is splitting if and
only if Ext2A (X(P ·); Y(P ·)) = 0. In particular; (U(P ·); V(P ·)) is splitting if either
proj dim X 6 1 for all X ∈ X(P ·) or inj dim Y 6 1 for all Y ∈ Y(P ·).
In this section, we deal with the case where R is a commutative artin ring, I is an
injective envelope of an R-module R=rad(R) and A is a 5nitely generated R-module. We
denote by mod A the full abelian subcategory of Mod A consisting of 5nitely generated
d−1
modules. Throughout this section, P · is also a complex P −1 → P
P 0 with the P i being
5nitely generated projective A-modules. Note that H (P ); H (*(P ·)) ∈ mod A for all
n · n
n ∈ Z. We set
Xc (P ·) = X(P ·) ∩ mod A and Yc (P ·) = Y(P ·) ∩ mod A:
M. Hoshino et al. / Journal of Pure and Applied Algebra 167 (2002) 15–35 33
*(P) ∼
= DA ⊗A P and P∼
= HomA (DA; *(P)):
Thus,
H0 (*(P ·)) ∼
= DA ⊗A H0 (P ·) and H−1 (P ·) ∼
= HomA (DA; H−1 (*(P ·)))
and hence,
Lemma 5.4. Assume Xc (P ·)∩Yc (P ·)={0} and H0 (P ·) ∈ Xc (P ·). Then the following
are equivalent:
(1) H0 (*(P ·)) ∈ Xc (P ·).
(2) Xc (P ·) is stable under DA ⊗A −.
(3) H−1 (P ·) ∈ Yc (P ·).
(4) Yc (P ·) is stable under HomA (DA; −).
Theorem 5.8. Let (X; Y) be a torsion theory for mod A such that X contains an
Ext-projective module X which generates X; Y contains an Ext-injective module Y
which cogenerates Y; and X is stable under DA⊗A −. Let MX· be a minimal projective
presentation of X and N · a minimal injective presentation of Y . Then
Y
References
[1] M. Auslander, Coherent functors, in: Proceedings of the Conference on Categorical Algebra, La Jolla
1965, Springer, Berlin, 1966, pp. 189 –231.
[2] A.A. Beilinson, J. Bernstein, P. Deligne, Faisceaux pervers, AstNerisque 100 (1982).
[3] M. BOokstedt, A. Neeman, Homotopy limits in triangulated categories, Compositio Math. 86 (1993)
209–234.
[4] S.E. Dickson, A torsion theory for abelian categories, Trans. Amer. Math. Soc. 121 (1966) 233–235.
[5] D. Happel, C.M. Ringel, Tilted algebras, Trans. Amer. Math. Soc. 274 (1982) 399–443.
[6] T. Holm, Derived equivalence classi5cation of algebras of dihedral, semidihedral, and quaternion type,
J. Algebra 211 (1999) 159–205.
[7] M. Hoshino, Tilting modules and torsion theories, Bull. London Math. Soc. 14 (1982) 334–336.
[8] M. Hoshino, On splitting torsion theories induced by tilting modules, Commun. Algebra 11 (4) (1983)
427–439.
[9] M. Hoshino, Y. Kato, Tilting complexes de5ned by idempotents, preprint.
[10] A. Neeman, The Grothendieck duality theorem via Bous5eld’s techniques and Brown representability,
J. Amer. Math. Soc. 9 (1996) 205–236.
[11] R. Hartshorne, Residues and Duality, Lecture Notes in Mathematics, vol. 20, Springer, Berlin, 1966.
[12] J. Rickard, Morita theory for derived categories, J. London Math. Soc. 39 (2) (1989) 436–456.