Semiorthogonal decomposition

In mathematics, a semiorthogonal decomposition is a way to divide a triangulated category into simpler pieces. One way to produce a semiorthogonal decomposition is from an exceptional collection, a special sequence of objects in a triangulated category. For an algebraic variety X, it has been fruitful to study semiorthogonal decompositions of the bounded derived category of coherent sheaves, D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} .

Semiorthogonal decomposition

Alexei Bondal and Mikhail Kapranov (1989) defined a semiorthogonal decomposition of a triangulated category T {\displaystyle {\mathcal {T}}} to be a sequence A 1 , , A n {\displaystyle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}} of strictly full triangulated subcategories such that:[1]

  • for all 1 i < j n {\displaystyle 1\leq i<j\leq n} and all objects A i A i {\displaystyle A_{i}\in {\mathcal {A}}_{i}} and A j A j {\displaystyle A_{j}\in {\mathcal {A}}_{j}} , every morphism from A j {\displaystyle A_{j}} to A i {\displaystyle A_{i}} is zero. That is, there are "no morphisms from right to left".
  • T {\displaystyle {\mathcal {T}}} is generated by A 1 , , A n {\displaystyle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}} . That is, the smallest strictly full triangulated subcategory of T {\displaystyle {\mathcal {T}}} containing A 1 , , A n {\displaystyle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}} is equal to T {\displaystyle {\mathcal {T}}} .

The notation T = A 1 , , A n {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}\rangle } is used for a semiorthogonal decomposition.

Having a semiorthogonal decomposition implies that every object of T {\displaystyle {\mathcal {T}}} has a canonical "filtration" whose graded pieces are (successively) in the subcategories A 1 , , A n {\displaystyle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}} . That is, for each object T of T {\displaystyle {\mathcal {T}}} , there is a sequence

0 = T n T n 1 T 0 = T {\displaystyle 0=T_{n}\to T_{n-1}\to \cdots \to T_{0}=T}

of morphisms in T {\displaystyle {\mathcal {T}}} such that the cone of T i T i 1 {\displaystyle T_{i}\to T_{i-1}} is in A i {\displaystyle {\mathcal {A}}_{i}} , for each i. Moreover, this sequence is unique up to a unique isomorphism.[2]

One can also consider "orthogonal" decompositions of a triangulated category, by requiring that there are no morphisms from A i {\displaystyle {\mathcal {A}}_{i}} to A j {\displaystyle {\mathcal {A}}_{j}} for any i j {\displaystyle i\neq j} . However, that property is too strong for most purposes. For example, for an (irreducible) smooth projective variety X over a field, the bounded derived category D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} of coherent sheaves never has a nontrivial orthogonal decomposition, whereas it may have a semiorthogonal decomposition, by the examples below.

A semiorthogonal decomposition of a triangulated category may be considered as analogous to a finite filtration of an abelian group. Alternatively, one may consider a semiorthogonal decomposition T = A , B {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}},{\mathcal {B}}\rangle } as closer to a split exact sequence, because the exact sequence 0 A T T / A 0 {\displaystyle 0\to {\mathcal {A}}\to {\mathcal {T}}\to {\mathcal {T}}/{\mathcal {A}}\to 0} of triangulated categories is split by the subcategory B T {\displaystyle {\mathcal {B}}\subset {\mathcal {T}}} , mapping isomorphically to T / A {\displaystyle {\mathcal {T}}/{\mathcal {A}}} .

Using that observation, a semiorthogonal decomposition T = A 1 , , A n {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}}_{1},\ldots ,{\mathcal {A}}_{n}\rangle } implies a direct sum splitting of Grothendieck groups:

K 0 ( T ) K 0 ( A 1 ) K 0 ( A n ) . {\displaystyle K_{0}({\mathcal {T}})\cong K_{0}({\mathcal {A}}_{1})\oplus \cdots \oplus K_{0}({\mathcal {A_{n}}}).}

For example, when T = D b ( X ) {\displaystyle {\mathcal {T}}={\text{D}}^{\text{b}}(X)} is the bounded derived category of coherent sheaves on a smooth projective variety X, K 0 ( T ) {\displaystyle K_{0}({\mathcal {T}})} can be identified with the Grothendieck group K 0 ( X ) {\displaystyle K_{0}(X)} of algebraic vector bundles on X. In this geometric situation, using that D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} comes from a dg-category, a semiorthogonal decomposition actually gives a splitting of all the algebraic K-groups of X:

K i ( X ) K i ( A 1 ) K i ( A n ) {\displaystyle K_{i}(X)\cong K_{i}({\mathcal {A}}_{1})\oplus \cdots \oplus K_{i}({\mathcal {A_{n}}})}

for all i.[3]

Admissible subcategory

One way to produce a semiorthogonal decomposition is from an admissible subcategory. By definition, a full triangulated subcategory A T {\displaystyle {\mathcal {A}}\subset {\mathcal {T}}} is left admissible if the inclusion functor i : A T {\displaystyle i\colon {\mathcal {A}}\to {\mathcal {T}}} has a left adjoint functor, written i {\displaystyle i^{*}} . Likewise, A T {\displaystyle {\mathcal {A}}\subset {\mathcal {T}}} is right admissible if the inclusion has a right adjoint, written i ! {\displaystyle i^{!}} , and it is admissible if it is both left and right admissible.

A right admissible subcategory B T {\displaystyle {\mathcal {B}}\subset {\mathcal {T}}} determines a semiorthogonal decomposition

T = B , B {\displaystyle {\mathcal {T}}=\langle {\mathcal {B}}^{\perp },{\mathcal {B}}\rangle } ,

where

B := { T T : Hom ( B , T ) = 0 } {\displaystyle {\mathcal {B}}^{\perp }:=\{T\in {\mathcal {T}}:\operatorname {Hom} ({\mathcal {B}},T)=0\}}

is the right orthogonal of B {\displaystyle {\mathcal {B}}} in T {\displaystyle {\mathcal {T}}} .[2] Conversely, every semiorthogonal decomposition T = A , B {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}},{\mathcal {B}}\rangle } arises in this way, in the sense that B {\displaystyle {\mathcal {B}}} is right admissible and A = B {\displaystyle {\mathcal {A}}={\mathcal {B}}^{\perp }} . Likewise, for any semiorthogonal decomposition T = A , B {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}},{\mathcal {B}}\rangle } , the subcategory A {\displaystyle {\mathcal {A}}} is left admissible, and B = A {\displaystyle {\mathcal {B}}={}^{\perp }{\mathcal {A}}} , where

A := { T T : Hom ( T , A ) = 0 } {\displaystyle {}^{\perp }{\mathcal {A}}:=\{T\in {\mathcal {T}}:\operatorname {Hom} (T,{\mathcal {A}})=0\}}

is the left orthogonal of A {\displaystyle {\mathcal {A}}} .

If T {\displaystyle {\mathcal {T}}} is the bounded derived category of a smooth projective variety over a field k, then every left or right admissible subcategory of T {\displaystyle {\mathcal {T}}} is in fact admissible.[4] By results of Bondal and Michel Van den Bergh, this holds more generally for T {\displaystyle {\mathcal {T}}} any regular proper triangulated category that is idempotent-complete.[5]

Moreover, for a regular proper idempotent-complete triangulated category T {\displaystyle {\mathcal {T}}} , a full triangulated subcategory is admissible if and only if it is regular and idempotent-complete. These properties are intrinsic to the subcategory.[6] For example, for X a smooth projective variety and Y a subvariety not equal to X, the subcategory of D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} of objects supported on Y is not admissible.

Exceptional collection

Let k be a field, and let T {\displaystyle {\mathcal {T}}} be a k-linear triangulated category. An object E of T {\displaystyle {\mathcal {T}}} is called exceptional if Hom(E,E) = k and Hom(E,E[t]) = 0 for all nonzero integers t, where [t] is the shift functor in T {\displaystyle {\mathcal {T}}} . (In the derived category of a smooth complex projective variety X, the first-order deformation space of an object E is Ext X 1 ( E , E ) Hom ( E , E [ 1 ] ) {\displaystyle \operatorname {Ext} _{X}^{1}(E,E)\cong \operatorname {Hom} (E,E[1])} , and so an exceptional object is in particular rigid. It follows, for example, that there are at most countably many exceptional objects in D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} , up to isomorphism. That helps to explain the name.)

The triangulated subcategory generated by an exceptional object E is equivalent to the derived category D b ( k ) {\displaystyle {\text{D}}^{\text{b}}(k)} of finite-dimensional k-vector spaces, the simplest triangulated category in this context. (For example, every object of that subcategory is isomorphic to a finite direct sum of shifts of E.)

Alexei Gorodentsev and Alexei Rudakov (1987) defined an exceptional collection to be a sequence of exceptional objects E 1 , , E m {\displaystyle E_{1},\ldots ,E_{m}} such that Hom ( E j , E i [ t ] ) = 0 {\displaystyle \operatorname {Hom} (E_{j},E_{i}[t])=0} for all i < j and all integers t. (That is, there are "no morphisms from right to left".) In a proper triangulated category T {\displaystyle {\mathcal {T}}} over k, such as the bounded derived category of coherent sheaves on a smooth projective variety, every exceptional collection generates an admissible subcategory, and so it determines a semiorthogonal decomposition:

T = A , E 1 , , E m , {\displaystyle {\mathcal {T}}=\langle {\mathcal {A}},E_{1},\ldots ,E_{m}\rangle ,}

where A = E 1 , , E m {\displaystyle {\mathcal {A}}=\langle E_{1},\ldots ,E_{m}\rangle ^{\perp }} , and E i {\displaystyle E_{i}} denotes the full triangulated subcategory generated by the object E i {\displaystyle E_{i}} .[7] An exceptional collection is called full if the subcategory A {\displaystyle {\mathcal {A}}} is zero. (Thus a full exceptional collection breaks the whole triangulated category up into finitely many copies of D b ( k ) {\displaystyle {\text{D}}^{\text{b}}(k)} .)

In particular, if X is a smooth projective variety such that D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} has a full exceptional collection E 1 , , E m {\displaystyle E_{1},\ldots ,E_{m}} , then the Grothendieck group of algebraic vector bundles on X is the free abelian group on the classes of these objects:

K 0 ( X ) Z { E 1 , , E m } . {\displaystyle K_{0}(X)\cong \mathbb {Z} \{E_{1},\ldots ,E_{m}\}.}

A smooth complex projective variety X with a full exceptional collection must have trivial Hodge theory, in the sense that h p , q ( X ) = 0 {\displaystyle h^{p,q}(X)=0} for all p q {\displaystyle p\neq q} ; moreover, the cycle class map C H ( X ) Q H ( X , Q ) {\displaystyle CH^{*}(X)\otimes \mathbb {Q} \to H^{*}(X,\mathbb {Q} )} must be an isomorphism.[8]

Examples

The original example of a full exceptional collection was discovered by Alexander Beilinson (1978): the derived category of projective space over a field has the full exceptional collection

D b ( P n ) = O , O ( 1 ) , , O ( n ) {\displaystyle {\text{D}}^{\text{b}}(\mathbf {P} ^{n})=\langle O,O(1),\ldots ,O(n)\rangle } ,

where O(j) for integers j are the line bundles on projective space.[9] Full exceptional collections have also been constructed on all smooth projective toric varieties, del Pezzo surfaces, many projective homogeneous varieties, and some other Fano varieties.[10]

More generally, if X is a smooth projective variety of positive dimension such that the coherent sheaf cohomology groups H i ( X , O X ) {\displaystyle H^{i}(X,O_{X})} are zero for i > 0, then the object O X {\displaystyle O_{X}} in D b ( X ) {\displaystyle {\text{D}}^{\text{b}}(X)} is exceptional, and so it induces a nontrivial semiorthogonal decomposition D b ( X ) = ( O X ) , O X {\displaystyle {\text{D}}^{\text{b}}(X)=\langle (O_{X})^{\perp },O_{X}\rangle } . This applies to every Fano variety over a field of characteristic zero, for example. It also applies to some other varieties, such as Enriques surfaces and some surfaces of general type.

A source of examples is Orlov's blowup formula concerning the blowup X = Bl Z ( Y ) {\displaystyle X=\operatorname {Bl} _{Z}(Y)} of a scheme Y {\displaystyle Y} at a codimension k {\displaystyle k} locally complete intersection subscheme Z {\displaystyle Z} with exceptional locus ι : E P Z ( N Z / Y ) X {\displaystyle \iota :E\simeq \mathbb {P} _{Z}(N_{Z/Y})\to X} . There is a semiorthogonal decomposition D b ( X ) = Φ 1 k ( D b ( Z ) ) , , Φ 1 ( D b ( Z ) ) , π ( D b ( Y ) ) {\displaystyle D^{b}(X)=\langle \Phi _{1-k}(D^{b}(Z)),\ldots ,\Phi _{-1}(D^{b}(Z)),\pi ^{*}(D^{b}(Y))\rangle } where Φ i : D b ( Z ) D b ( X ) {\displaystyle \Phi _{i}:D^{b}(Z)\to D^{b}(X)} is the functor Φ i ( ) = ι ( O E ( k ) ) p ( ) ) {\displaystyle \Phi _{i}(-)=\iota _{*}({\mathcal {O}}_{E}(k))\otimes p^{*}(-))} with p : X Y {\displaystyle p:X\to Y} is the natural map.[11]

While these examples encompass a large number of well-studied derived categories, many naturally occurring triangulated categories are "indecomposable". In particular, for a smooth projective variety X whose canonical bundle K X {\displaystyle K_{X}} is basepoint-free, every semiorthogonal decomposition D b ( X ) = A , B {\displaystyle {\text{D}}^{\text{b}}(X)=\langle {\mathcal {A}},{\mathcal {B}}\rangle } is trivial in the sense that A {\displaystyle {\mathcal {A}}} or B {\displaystyle {\mathcal {B}}} must be zero.[12] For example, this applies to every variety which is Calabi–Yau in the sense that its canonical bundle is trivial.

See also

Notes

  1. ^ Huybrechts 2006, Definition 1.59.
  2. ^ a b Bondal & Kapranov 1990, Proposition 1.5.
  3. ^ Orlov 2016, Section 1.2.
  4. ^ Kuznetsov 2007, Lemmas 2.10, 2.11, and 2.12.
  5. ^ Orlov 2016, Theorem 3.16.
  6. ^ Orlov 2016, Propositions 3.17 and 3.20.
  7. ^ Huybrechts 2006, Lemma 1.58.
  8. ^ Marcolli & Tabuada 2015, Proposition 1.9.
  9. ^ Huybrechts 2006, Corollary 8.29.
  10. ^ Kuznetsov 2014, Section 2.2.
  11. ^ Orlov, D O (1993-02-28). "PROJECTIVE BUNDLES, MONOIDAL TRANSFORMATIONS, AND DERIVED CATEGORIES OF COHERENT SHEAVES". Russian Academy of Sciences. Izvestiya Mathematics. 41 (1): 133–141. doi:10.1070/im1993v041n01abeh002182. ISSN 1064-5632.
  12. ^ Kuznetsov 2014, Section 2.5.

References

  • Bondal, Alexei; Kapranov, Mikhail (1990), "Representable functors, Serre functors, and reconstructions", Mathematics of the USSR-Izvestiya, 35: 519–541, doi:10.1070/IM1990v035n03ABEH000716, MR 1039961
  • Huybrechts, Daniel (2006), Fourier–Mukai transforms in algebraic geometry, Oxford University Press, ISBN 978-0199296866, MR 2244106
  • Kuznetsov, Alexander (2007), "Homological projective duality", Publications Mathématiques de l'IHÉS, 105: 157–220, arXiv:math/0507292, doi:10.1007/s10240-007-0006-8, MR 2354207
  • Kuznetsov, Alexander (2014), "Semiorthogonal decompositions in algebraic geometry", Proceedings of the International Congress of Mathematicians (Seoul, 2014), vol. 2, Seoul: Kyung Moon Sa, pp. 635–660, arXiv:1404.3143, MR 3728631
  • Marcolli, Matilde; Tabuada, Gonçalo (2015), "From exceptional collections to motivic decompositions via noncommutative motives", Journal für die reine und angewandte Mathematik, 701: 153–167, arXiv:1202.6297, doi:10.1515/crelle-2013-0027, MR 3331729
  • Orlov, Dmitri (2016), "Smooth and proper noncommutative schemes and gluing of DG categories", Advances in Mathematics, 302: 59–105, arXiv:1402.7364, doi:10.1016/j.aim.2016.07.014, MR 3545926