EQUIVARIANT DERIVED CATEGORIES OF HYPERSURFACES ASSOCIATED TO A SUM OF POTENTIALS by BRONSON LIM A DISSERTATION Presented to the Department of Mathematics and the Graduate School of the University of Oregon in partial fulfillment of the requirements for the degree of Doctor of Philosophy June 2017 DISSERTATION APPROVAL PAGE Student: Bronson Lim Title: Equivariant Derived Categories of Hypersurfaces Associated to a Sum of Potentials This dissertation has been accepted and approved in partial fulfillment of the requirements for the Doctor of Philosophy degree in the Department of Mathematics by: Alexander Polishchuk Chair Dan Dugger Member Victor Ostrik Member Vadim Vologodsky Member Michael Kellman Institutional Representative and Scott L. Pratt Dean of the Graduate School Original approval signatures are on file with the University of Oregon Graduate School. Degree awarded June 2017 ii c© 2017 Bronson Lim iii DISSERTATION ABSTRACT Bronson Lim Doctor of Philosophy Department of Mathematics June 2017 Title: Equivariant Derived Categories of Hypersurfaces Associated to a Sum of Potentials We construct a semi-orthogonal decomposition for the equivariant derived category of a hypersurface associated to the sum of two potentials. More specifically, if f, g are two homogeneous polynomials of degree d defining smooth Calabi-Yau or general type hypersurfaces in Pn, we construct a semi-orthogonal decomposition of D[V (f ⊕ g)/µd]. Moreover, every component of the semi- orthogonal decomposition is explicitly given by Fourier-Mukai functors. iv CURRICULUM VITAE NAME OF AUTHOR: Bronson Lim GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED: University of Oregon, Eugene, Oregon California State University, San Bernardino, California DEGREES AWARDED: Doctor of Philosophy, 2017, University of Oregon Master of Arts, 2011, California State University at San Bernardino Bachelor of Arts, 2009, California State University at San Bernardino AREAS OF SPECIAL INTEREST: Complex Algebraic Geometry Matrix Factorizations Differential Graded Categories Finite Group Theory PROFESSIONAL EXPERIENCE: Graduate Teaching Fellow, University of Oregon, 2011-2017 Teaching Associate, California State University at San Bernardino, 2010-2011 GRANTS, AWARDS AND HONORS: Frank Anderson Teaching Award, University of Oregon, 2016 Rose Hills Scholarship, University of Oregon, 2015 Johnson Fellowship, University of Oregon, 2014 Outstanding Thesis, California State University at San Bernardino, 2011 v ACKNOWLEDGEMENTS Firstly, I thank my advisor Sasha for teaching me algebraic geometry, for introducing me to this problem and other related projects, and for his continuing patience with me. I thank my colleagues at the University of Oregon, especially Nick Howell and Ben Dyer for listening to me ramble at them. I thank Peter Gilkey for helping me find my path. I thank my girlfriend Esther, for believing in me and for nudging me along in the writing process. I thank the mathematics department at the University of Oregon for having me for the past six years. Finally, I thank my mother for making all of this possible and encouraging me to pursue what I care about. vi TABLE OF CONTENTS Chapter Page I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1. Semi-orthogonal decompositions in algebraic geometry . . . . . . 1 1.2. Orlov’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3. Adding two potentials. . . . . . . . . . . . . . . . . . . . . . . . . 3 1.4. Main result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.5. Further Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 II. PRELIMINARIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.1. Triangulated Categories . . . . . . . . . . . . . . . . . . . . . . . 14 2.2. Semi-orthogonal decompositions . . . . . . . . . . . . . . . . . . 15 2.3. Admissible and saturated triangulated subcategories . . . . . . . 18 2.4. Equivariant triangulated categories . . . . . . . . . . . . . . . . . 21 2.5. Fourier-Mukai functors . . . . . . . . . . . . . . . . . . . . . . . 23 2.6. Spanning classes . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.7. Derived Category of [Pm+n−1/µd] . . . . . . . . . . . . . . . . . . 28 III. SETUP AND EMBEDDINGS . . . . . . . . . . . . . . . . . . . . . . . . 33 3.1. Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 vii Chapter Page 3.2. Equivariant geometry of X . . . . . . . . . . . . . . . . . . . . . 34 3.3. Subcategory of exceptional line bundles. . . . . . . . . . . . . . . 36 3.4. Embedding D(Xf ) . . . . . . . . . . . . . . . . . . . . . . . . . . 38 3.5. Embedding D(Xg) . . . . . . . . . . . . . . . . . . . . . . . . . . 40 3.6. Embedding D(Xf ×Xg) . . . . . . . . . . . . . . . . . . . . . . . 42 IV. PROOF OF MAIN THEOREM . . . . . . . . . . . . . . . . . . . . . . . 50 4.1. Semi-orthogonality between Dfg and A . . . . . . . . . . . . . . 50 4.2. Semi-orthogonality between Dg,Df ,A. . . . . . . . . . . . . . . . 51 4.3. Koszul complexes and Joins . . . . . . . . . . . . . . . . . . . . . 53 4.4. Other kernels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 4.5. Proof of Fullness . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 4.6. The Case m = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 V. COMPARISON WITH ORLOV’S FUNCTORS ON POINTS . . . . . . . 70 5.1. Graded Matrix Factorizations . . . . . . . . . . . . . . . . . . . . 70 5.2. Singularity Categories . . . . . . . . . . . . . . . . . . . . . . . . 75 5.3. Orlov’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 5.4. Computations in the Calabi-Yau case . . . . . . . . . . . . . . . 81 5.5. Comparison of Functors . . . . . . . . . . . . . . . . . . . . . . . 85 REFERENCES CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 viii CHAPTER I INTRODUCTION 1.1 Semi-orthogonal decompositions in algebraic geometry To a space X, i.e. a smooth and projective variety or more generally a smooth and proper Deligne-Mumford stack, we can associate the bounded derived category of coherent sheaves on the space, denoted D(X). The category D(X) lives in the intersection between homological algebra and algebraic geometry and has proved to be a useful tool when applied to algebro-geometric problems. The derived category is a strong invariant of X. For example, if X has ample or anti-ample canonical bundle, a derived equivalence D(X) ' D(Y ) gives rise to an isomorphism X ' Y . Additionally, additive invariants such as Hochschild homology and K-theory factor through D(X). Of particular interest is when D(X) admits a semi-orthogonal decomposition (see Section 2.2 for the definition). Roughly, a semi-orthogonal decomposition is the analogue of a group extension for triangulated categories. If D(X) admits a semi- orthogonal decomposition, one can hope to further understand D(X), or sometimes X, using the components of the decomposition. For example, projective spaces admit the Beilinson exceptional collection: D(Pn) = 〈O(−n), . . . ,O〉. This can be regarded as a categorical analogue of the isomorphism H∗(Pn;Z) ∼= Z[x]/(xn+1). 1 An interesting example relating the derived category to birational geometry is given by cubic fourfolds. Let X ⊂ P5 be a cubic fourfold, then there is a decomposition D(X) = 〈O(−2),O(−1),O,AX〉, where AX is characterized as the left-orthogonal to the exceptional collection of line bundlees 〈O(−2),O(−1),O〉 and is sometimes called the Kuznetsov subcategory (or the Kuznetsov component). It is conjectured, [17, Conjecture 1.1], that X is rational if and only if AX is equivalent to D(S) for S a K3-surface. We will come back to this example shortly. There are many more examples of using derived categories and semi- orthogonal decompositions to study varieties. We refer the reader to the surveys [4, Sections 4 and 5] and [7, Section 2]. 1.2 Orlov’s Theorem This project was discovered when trying to understand Orlov’s theorem relating the derived category of a projective hypersurface to the category of graded matrix factorizations of the defining function. We recall this theorem now to motivate the main result. Let k be an algebraically closed field of characteristic zero and V a vector space over k of dimension n. Assume f ∈ Symd(V ∨) defines a smooth projective hypersurface, say Xf = V (f) ⊂ P(V ). We call f a potential. Let HMFgr(f) denote the homotopy category of graded matrix factorizations of the potential f (see Section 5.1). Objects of HMFgr(f) are Z/2-graded, curved complexes of Gm- equivariant vector bundles on V with curvature f . There is a natural differential on 2 the space of morphisms between two matrix factorizations. The category HMFgr(f) is the corresponding homotopy category. A relationship between HMFgr(f) and D(Xf ) was discovered by Orlov in [20]. Orlov constructs two Z-indexed families of exact functors Ψi : D(Xf ) → HMFgr(f) and Φi : HMF gr(f) → D(Xf ). If Xf is Fano or Calabi-Yau, then Φi is a full embedding. If Xf is general type or Calabi-Yau, then Ψi is a full embedding. Moreover, the semi-orthogonal complement is explicitly determined. Orlov’s Theorem. [20, Theorem 3.11] Let f be as above. For each i ∈ Z, we have the following semi-orthogonal decompositions: Fano : D(Xf ) = 〈OXf (−i− n+ d+ 1), . . . ,OXf (−i),ΦiHMFgr(f)〉; General Type : HMFgr(f) = 〈kstab(−i), . . . , kstab(−i+ n− d+ 1),ΨiD(Xf )〉; Calabi-Yau : Φi,Ψi induce mutual inverse equivalences D(Xf ) ∼= HMFgr(f). Here kstab is a certain Koszul matrix factorization (see Example 5.1.2) associated to the residue field of Sym(V ∨) at the origin. 1.3 Adding two potentials. Let f, g be homogeneous polynomials of degree d defining smooth hypersurfaces Xf ⊂ Pm−1 and Xg ⊂ Pn−1. Let X = V (f ⊕ g) ⊂ Pm+n−1. Then X is smooth since Xf and Xg are smooth (see Proposition 3.1.1). Suppose d ≥ max{m,n}. Then there is a Z-indexed family of embeddings Ψi : D(Xf ) → HMFgr(f) and similarly Ψj : D(Xg) → HMFgr(g). By tensoring, we 3 can consider the family of embeddings: Ψi,j : D(Xf ×Xg) ∼= D(Xf )⊗D(Xg)→ HMFgr(f)⊗ HMFgr(g). where Ψi,j = Ψi ⊗ Ψj and the tensor product is understood to be taken in suitable dg-enhancements1. We further have an identification, see [1, Corollary 5.18] HMFgr(f)⊗ HMFgr(g) ∼= HMFgr,µd(f ⊕ g). where µd acts on the g variables. If in addition d ≤ n + m, it was noticed in [2, Example 3.10] that we can then embed HMFgr,µd(f ⊕ g) into D[X/µd] using Orlov’s Theorem a second time. Fix one such embedding to get a doubly indexed family of fully-faithful functors Ξi,j : D(Xf ×Xg) → D[X/µd]. The complement consists of mn exceptional objects, d−m copies of D(Xg), and d− n copies of D(Xf ). Specifically, we have D[X/µd] = 〈A,K,Df ,Dg,Dfg〉 where A consists of (m+n−d)d line bundles, K ∼= 〈kstab(−i), . . . , kstab(−i+m−d+ 1)〉⊗〈kstab(−j), . . . , kstab(−j+n−d+1)〉, Df = ΦiD(Xf )⊗〈kstab(−j), . . . , kstab(−j+ n − d + 1)〉, Dg = 〈kstab(−i), . . . , kstab(−i + m − d + 1)〉 ⊗ ΦjD(Xg), and Dfg = Ξi,jD(Xf ×Xg). 1It is known that Orlov’s functors lift to the dg level [10] 4 These functors are not easy to compute and, with the exception of A, explicitly understanding the left and right semi-orthogonal complements to the image of Ξi,j as µd-equivariant complexes of sheaves on X is not easy. 2 1.4 Main result In this dissertation, we give a more geometric definition of the functors Ξi,j and show that they, miraculously, remain embeddings even if d > n + m. Moreover, we explicitly determine the other components in the associated semi-orthogonal decomposition. Main Theorem. If m,n ≥ 2 and d ≥ max{m,n}, then there is a semi-orthogonal decomposition D[X/µd] = 〈Dg1,Dfg,Dg2,Df ,A〉. Here Dg1 and Dg2 collectively consist of d −m twists of D(Xg) (Section 3.5), Df consists of d − n twists of D(Xf ) (Section 3.4), Dfg is the image of Ξ−m,−n (Section 3.6), and A consists of an exceptional collection of line bundles (Section 3.3). To align with the picture given by Orlov’s theorem we can mutate the decomposition; however, it gets complicated quickly as we will see in Chapter V. As stated, each of the components has a simple description given by explicit Fourier- Mukai functors. Using Orlov’s theorem we can verify the main theorem in the cases that it holds. We do that, numerically, for the special case of surfaces in P3. In particular, our decomposition furnishes a full exceptional collection. We then show our 2In low dimensional cases, one can compute the image of K and, for an appropriate choice of functor, get twists of powers of ΩPN |X . 5 theorem gives a full exceptional collection in the cases Orlov’s theorem doesn’t hold. We finish by discusssing an interesting class of cubic fourfolds. In the case of surfaces in P3, Orlov’s theorem holds for d ≤ 4. Example 1.4.1 (Surfaces with d ≤ 4). Suppose d ≤ 4 and m = n = 2. Let f, g be homogeneous polynomials defining d points in P1. Set S = V (f ⊕ g) ⊂ P3 so that S is a degree d surface in P3. If d = 2, then Orlov’s decomposition is D[S/µ2] = 〈O(−1)(χ0,1),O(χ0,1),HMFgr,µ2(f ⊕ g)〉 and HMFgr,µ2(f ⊕ g) ∼= D(Xf ×Xg) which contributes 4 exceptional objects for a total of 8. Alternatively, S ∼= P1 × P1 and Beilinson’s decomposition gives D[S/µ2] = 〈O(−1,−1)(χ0,1),O(−1, 0)(χ0,1),O(0,−1)(χ0,1),O(χ0,1)〉 which is 8 exceptional objects. Our decomposition gives D[S/µ2] = 〈Dfg,A〉 where A consists of 4 exceptional line bundles and Dfg consists of 4 exceptional objects for a total of 8. 6 If d = 3, we have Orlov’s decomposition D[S/µ3] = 〈O(−1)(χ0,1,2),HMFgr,µ3(f ⊕ g)(χ0,1,2)〉. The category HMFgr,µ3(f ⊕ g) has a semi-orthogonal decomposition HMFgr,µ3(f ⊕ g) = 〈kstab(−i),D(Xf ),D(Xg),D(Xf ×Xg)〉 which is 16 exceptional objects for a grand total of 19 exceptional objects. Our decomposition is D[S/µ3] = 〈Dg,Dfg,Df ,A〉 where A again has 4 exceptional objects, Df has 3, Dfg has 9, and Dg has 3 for a grand total of 19. If d = 4, we have a quartic K3 surface. Orlov’s theorem gives an equivalence of categories: D[S/µ4] ∼= HMFgr,µ4(f ⊕ g) and we have the equivalence HMFgr,µ4(f ⊕ g) = HMFgr(f)⊗ HMFgr(g) so that HMFgr,µ4 = 〈K,G,F ,D(Xf ×Xg)〉 where G consists of two copies of D(Xg) and similarly for F . The category K consists of stabilizations of the residue field. There are 4 exceptional objects here. 7 The categories G,F both have 8 exceptional objects and the category D(Xf × Xg) has 16. The grand total is 36. Our decomposition yields D[S/µ4] ∼= 〈Dg,Dfg,Df ,A〉. In this case A consists of 4 exceptional objects, Dg consists of 8 exceptional objects, Dfg consists of 16 exceptional objects, and Df consist of 8 exceptional objects. The grand total is 36 exceptional objects. Example 1.4.2 (Surfaces with d > 4). If d > 4, Orlov’s theorem no longer holds. In this case, we get an exceptional collection. Indeed, the decomposition in the main theorem gives D[S/µd] = 〈Dg,Dfg,Df ,A〉, which is an exceptional collection as in Example 1.4.1. The category A still has 4 exceptional objects. The categories Dg and Df have d(d − 2) exceptional objects. The category Dfg has d2 exceptional objects for a total of 3d2 − 4d + 4 exceptional objects. Example 1.4.3 (Cubic fourfolds from genus 1 curves). Here is an interesting example which recovers Orlov’s theorem in the case m = n = d = 3. In this case Xf and Xg are genus 1 curves and X is a cubic fourfold. The µ3-equivariant Orlov decomposition is D[X/µ3] = 〈O(−2)(χ0,1,2),O(−1)(χ0,1,2),O(χ0,1,2),Aµ3X 〉. 8 The µ3-equivariant Kuznetsov component is equivalent to HMF gr,µ3(f ⊕ g) which is equivalent to D(Xf ×Xg). The decomposition we give is D[X/µ3] = 〈D(Xf ×Xg),O(−4)(χ),O(−3)(χ1,2),O(−2)(χ0,1,2),O(−1)(χ0,1),O〉. It’s easy to see that the decompositions differ by a sequence of mutations and possibly tensoring by a line bundle.3 This example was considered in [2, Example 4.7]. We conjecture that these cubic fourfolds are rational. If so, then by Kuznetsov’s conjecture, AX ∼= D(S) for some K3-surface S and there is an action of µ3 on D(S) such that the µ3- equivariant category, D(S)µ3 is equivalent to D(Xf × Xg). It would be very interesting to verify this conjecture: explicitly determine the K3 surface S and understand the µ3-action on D(S). 1.5 Further Work We discuss four conjectures for future work related to the Main Theorem and discuss the organization of Chapters 2-5. The first natural extension is to the case of k ≥ 2 hypersurfaces. In particular, suppose f1, . . . , fk defined k smooth degree d-hypersurfaces Xi ⊂ P(Vi) which are Calabi-Yau or general type. Let X = V (f1 ⊕ · · · ⊕ fk) ⊂ P(⊕iVi). Then we can use the technique discussed in Section 1.3 to see that the µk−1d -equivariant derived category of X (with the natural action of µk−1d ) has semi-orthogonal components of 3Theorem 5.5.1 shows they agree on points. This implies they differ by an autoequivalence that fixes points. Since Xf ×Xg is an Abelian variety, the result follows. 9 the form D(X1 × · · · × Xk), copies of subproducts of the Xi, and an exceptional collection of line bundles. Conjecture 1.5.1. Suppose d ≥ max{dim(Vi)}, then there is a semi-orthogonal decomposition of D[X/µk−1d ] with components: (i) a copy of D(X1 × · · · ×Xk); (ii) copies of subproducts D(Xi1 × · · · ×Xit), for a subset {i1, . . . , it} ⊂ {1, . . . , k}, depending on d and dim(Vi); (iii) an exceptional collection of line bundles. The decomposition in the Main Theorem should theoretically also hold in weighted projective spaces and for finitely many stacky hypersurfaces. Indeed, Orlov’s theorem is still valid here and the functors involved are still well defined. We state the conjecture for two stacky hypersurfaces. Conjecture 1.5.2. Let Xf = V (f) ⊂ P(a1, · · · , am) and Xg = V (g) ⊂ P(b1, . . . , bn) be smooth, degree d hypersurfaces (regarded as stacks) in the weighted projective stacks with weights a1, . . . , am and b1, . . . , bn. Set X = V (f ⊕ g) ⊂ P(a1, . . . , am, b1, . . . , bn). Suppose d ≥ max{ ∑ ai, ∑ bi}, then there is a semi- orthogonal decomposition of D[X/µd], where the components consist of – an exceptional collection of line bundles of length; – a copy of D(Xf ×Xg); – copies of D(Xg) and D(Xf ) depending on d and the weights ai, bi. The Main Theorem should extend to families of hypersurfaces. More specifically, let V1, V2 be finite dimensional vector spaces. Consider the hypersurface 10 Y ⊂ P(Sd(V1)∨ ⊕ Sd(V2)∨) × P(V1 ⊕ V2) given by ([f1 : f2], [x1 : x2]) such that f1(x1) + f2(x2) = 0. The multiplicative group Gm acts on Y by t · ([f1 : f2], [x1 : x2]) = ([f1 : t −df2], [x1 : tx2]). Set X = [Y/Gm] to be the corresponding quotient stack. This quotient stack is the relative analogue of [V (f ⊕ g)/µd] over the base P(Sd(V1)∨)× P(Sd(V2)∨). Let Hi ⊂ P(Sd(Vi)∨) × P(Vi) be the universal hypersurfaces. The analogue of Xf × Xg is unfortunately not H1 × H2 and is slightly more complicated. Consider the affine version of the universal hypersurfaces: H1 ×H2 ⊂ Sd(V1)∨)× Sd(V2)∨ × P(V1)× P(V2). Define an action of G2m on H1 ×H2 by (t1, t2) · (f1, f2, [v1, v2]) = (t1f1, t1t−d2 f2, [v1, v2]). Let B = [H1 × H2/G2m] be the corresponding quotient stack. This is the analogue of Xf1 ×Xf2 in the sense that the coarse moduli of B is H1 ×H2 and the fiber over ([f1], [f2]) under the projection H1 ×H2 → P(Sd(V1)∨)× P(Sd(V2)∨) is Xf1 ×Xf2 . The action of G2m is not effective with kernel µd and so B is a µd-gerbe over H1 × H2. Over B there is a P1-bundle given by descending the bundle PH1×H2(OH1(−1)  OX2(−1)) to B, i.e. the G2m action on H1 × H2 lifts to this bundle via (t1, t2) · (f1, f2, [v1 : v2]) = (t1f1, t1t−d2 f2, [v1 : t2v2]). Let P be the corresponding µd-gerbe. We can now use OP to define a Fourier- Mukai functor Φ: D(B)→ D(X ). 11 Conjecture 1.5.3. The functor Φ is fully-faithful under appropriate conditions on d and dim(Vi) with explicit semi-orthogonal complement consisting of copies of D(B), D(H1), D(H2), and an exceptional collection of line bundles. By specializing to a specific fiber, we could hope to extend the Main Theorem to possibly singular hypersurfaces, which is very interesting. The Main Theorem may also generalize to the case where Xf and Xg are replaced with smooth complete intersections. However, it appears that the number of hypersurfaces we intersect to get Xf and Xg must be equal. Otherwise it is unclear how to add them and get something smooth. Here is a toy example indicating how it might be setup. Example 1.5.1. Suppose f1, f2 are degree d polynomials defining a smooth complete intersection Xf in P(V1) and g1, g2 are degree d defining a smooth complete intersection Xg in P(V2). Then X = V (f1 ⊕ g1, f2 ⊕ g2) define a smooth complete intersection in P(V1 ⊕ V2). There is again an action of µd which scales the gi variables. If d >> 0, then it’s easy to see the natural inclusions Xf , Xg ↪→ X induce fully-faithful functors D(Xf ),D(Xg) ↪→ D[X/µd]. The functors Ξ−m,−n still make sense and the proof of fully-faithfulness in the Main Theorem should carry over to this case. We conjecture that it is still fully-faithful and that there is a semi- orthogonal decomposition of D[X/µd] analogous to the Main Theorem. Example 1.5.2. If say Xf = V (f1, f2) ⊂ P(V1) and Xg = V (g) ⊂ P(V2), then the only possible choice for X is X = V (f1 ⊕ g, f2 ⊕ g) ⊂ P(V1 ⊕ V2). This is the same as V (f1 − f2, g) which is not always smooth even if Xf and Xg are smooth. In Chapter II, we recall preliminary facts about triangulated and equivariant triangulated categories.In Chapter III we define all of the terms in the above 12 decomposition and show they are embeddings. In Chapter IV we prove the Main Theorem and also discuss the special case m = 1, which is not covered by the Main Theorem. In Chapter V we compare our functors to the ones described in Section 1.3 on the structure sheaf of points. 13 CHAPTER II PRELIMINARIES 2.1 Triangulated Categories Throughout k is an algebraically closed field of characteristic zero. For an overview of triangulated categories in algebraic geometry see [14]. Definition 2.1.1. A triangulated category T is a k-linear category together with an autoequivalence [1] : T → T and a class of exact triangles t→ u→ v → t[1] certain axioms, see [13, Section 1.1]. The autoequivalence [1] is sometimes called a shift functor. Example 2.1.1. Most (but not all) examples come from Abelian categories. If A is an Abelian category, then the derived category of A is given by D(A) := Com(A)[S−1]. That is, we take the category of chain complexes1 of objects in A. Then we localize at the class of quasi-isomorphisms. The shift functor is given by shifting the complex to the left : K·[1] = K·+1. 1We only consider bounded complexes in this dissertation. However, the use of unbounded complexes still has many uses. 14 Example 2.1.2. If (X,OX) is a locally ringed space (for us X will be either a scheme or a stack), then the category of OX-modules is an Abelian category. By Example 2.1.1, we can consider the derived category of OX-modules, denoted D(OX −Mod). If X is a locally Noetherian scheme (or stack), then we can consider the subcategory, Coh(X), of coherent OX-modules. This is still Abelian and we define the derived category of X to be the derived category of the Abelian category Coh(X): D(X) := D(Coh(X)). As mentioned in the introduction, the derived category is a strong invariant of X and is our primary object of study. Another subcategory of OX-modules which is interesting is the subcategory of perfect complexes, denote Perf(X). A complex F · is perfect if it is locally quasi-isomorphic to a bounded complex of locally free sheaves. In the case X is Noetherian, this is the same as being quasi-isomorphic to a complex of vector bundles. If X is quasi-projective and smooth, then Perf(X) = D(X). Thus the subcategory of perfect complexes is a categorical measure of smoothness (see Definition 5.2.1). 2.2 Semi-orthogonal decompositions Semi-orthogonal decompositions allow us to decompose our triangulated category into simpler pieces. The first example was found by Beilinson in [3]. Roughly, the idea is the following: Suppose we have a triangulated subcategory K of a triangulated category T . We can form the Drinfeld-Verdier localization of T 15 by K. Recall, the localization C is C = T [Σ−1] where Σ is the class of morphisms, σ, such that Cone(σ) ∈ K. Consider the sequence of triangulated categories K ι−→ T pi−→ C. If K is nice enough (see 2.3), then the embedding functor (ι : K → T ) will have left and right adjoints, say ιL,R. Take an object t ∈ T , then there is an exact triangle in T of the form: ι(ιL(t))→ t→ Cone→ where Cone is the cone of the map ι(ιL(t)) → t. If C is also nice enough, then Cone is in the image of the right adjoint to pi. It’s not hard to see that ι(ιL(t)) and Cone are unique up to isomorphism and HomT (ι(ιL(t),Cone)) = 0. This is made precise with the next definition. Definition 2.2.1. Let T be a triangulated category. A semi-orthogonal decomposition of T , written T = 〈A1, . . . ,An〉 is a sequence of full triangulated subcategories A1, . . . ,An of T such that: (i) HomT (ai, aj) = 0 for ai ∈ Ai, aj ∈ Aj, and i > j; 16 (ii) For any t ∈ T , there is a sequence of morphisms 0 = an → an−1 → · · · → a1 → a0 = t where Cone(ai → ai−1) ∈ Ai. Condition (i) is called the semi-orthogonality condition and condition (ii) is called the fullness or generation condition. Example 2.2.1. The most common examples of semi-orthogonal decompositions occur when T = D(X) for some smooth projective scheme over k. In this case, there is often a vector bundle E such that Ext∗X(E , E) ∼= k[0]. Such an object in D(X) is called exceptional. The subcategory generated by E is also abusively denoted by E and there are two semi-orthogonal decompositions (that it exists follows from Example 2.3.1): D(X) = 〈E⊥, E〉 = 〈E , ⊥E〉 where E⊥ = {F · ∈ D(X) | Ext∗X(E ,F ·) = 0} and ⊥E is defined similarly. Example 2.2.2. Beilinson’s example, [3], is given by D(Pn) = 〈O(−n), . . . ,O〉. Semi-orthogonality is Serre’s theorem and fullness is given by the Beilinson spectral sequence (or equivalently Beilinson’s resolution of the diagonal). In other words, D(Pn) is made of iterated extensions of D(Spec(k)). As mentioned before, this is, in 17 a very precise way, a categorical analogue of the fact H∗(Pn;Z) ∼= Z[x]/(xn+1). Projective spaces are an important example of a scheme with the (rare) Diagonal Resolution Property2: If X is any smooth projective variety such that the structure sheaf of the diagonal O∆ in X × X has a resolution by sheaves of the form pi∗1Ei ⊗ pi∗2Fi, then it has the Diagonal Resolution Property. An important consequence is that the full triangulated subcategory of D(X) generated by either the Ei or the Fi is D(X) itself. If in addition we have semi-orthogonality between them, then this furnishes a semi-orthogonal decomposition of D(X). Example 2.2.3. It is not always the case that the derived category of projective scheme admits a semi-orthogonal decomposition. Indeed, let C be a smooth and proper curve of genus g > 1. Then Okawa proves, see [18], that D(C) does not admit a semi-orthogonal decomposition. For g = 1, there is no semi-orthogonal decomposition because the Serre functor is the shift functor. For g = 0, we have Beilinson’s exceptional collection of Example 2.2.2. In a sense, this result can be viewed as saying that curves provide one dimensional building blocks for triangulated categories. 2.3 Admissible and saturated triangulated subcategories As mentioned in Section 2.2, triangulated subcategories (or quotients) possessing adjoints lead the way to semi-orthogonal decompositions. In this section, 2There is a weaker diagonal resolution property, where we require that the diagonal is cut out by the zero locus of a section of a vector bundle on X ×X. See [22] for results in this direction. 18 we introduce admissibility and saturatedness that characterize those subcategories which possess adjoints. Definition 2.3.1. Let A ⊂ T be a full triangulated subcategory of a triangulated category. We say A is admissible if the embedding functor ι : A → T has a left and right adjoint3. If A is admissible, then it follows formally that T admits two semi-orthogonal decompositions T = 〈A⊥,A〉 = 〈A, ⊥A〉. where A⊥ := {t ∈ T | HomT (a[i], t) = 0 for all a ∈ A, i ∈ Z}; ⊥A := {t ∈ T | HomT (t, a[i]) = 0 for all a ∈ A, i ∈ Z}. We refer to A⊥ as the right orthogonal and ⊥A as the left orthogonal to A in T . Indeed, semi-orthogonality is by definition. Then suppose ι : A → T is the embedding functor and ιL,R are the left and right adjoints. It follows that ι(ιL(t))→ t→ Cone→, where Cone is the cone of the obvious mapping, gives T = 〈A, ⊥A〉. To see Cone ∈ ⊥A we apply HomT (−, ι(a)) for some a ∈ A: HomT (Cone, ι(a))→ HomT (t, ι(a)) −→ HomT (ι(ιL(t)), ι(a))→ . 3Sometimes, we define left (or right) admissible to mean the existence of a left (or right) adjoint. We won’t need these more general notions in this dissertation. 19 The last term is HomT (ι(ιL(t)), ι(a)) ∼= HomA(ιL(t), a) ∼= HomT (t, ι(a)) and  is the obvious isomorphism. Definition 2.3.2. A triangulated category T is called saturated if every cohomological functor (contravariant or covariant) H : T → Vectk of finite type is representable. We have the following important proposition regarding saturated subcategories, see [5, Proposition 2.6]. Proposition 2.3.1. Let A be a saturated triangulated category and ι : A → T is a full embedding. Then A is an admissible subcategory of T . Example 2.3.1. The derived category of coherent sheaves on a smooth projective variety, X, is saturated, [5, Theorem 2.14]. If E is an exceptional object of D(X), then there is a full embedding ιE : D(Spec(k)) → D(X) given by ιE(V ) = E ⊗ V . This justifies Example 2.2.1. Example 2.3.2. Let X be smooth projective and let A be the full triangulated subcategory of D(X) generated by the structure sheaves of closed points, Op for p ∈ X. Then A is not saturated. If it were, then the functor of cohomology HomD(X)(OX ,−) would be representable. This is not possible as the only objects in A are torsion sheaves (with zero dimensional support) and there are always nonzero extensions between a torsion sheaf and itself, whereas cohomology of a sheaf with zero dimensional support is just global sections. 20 We will use the following proposition in conjuction with Proposition 2.6.1 in Section 4.5. Proposition 2.3.2. [5, Theorem 2.10] If A,B ⊂ T are full, saturated, triangulated subcategories such that 〈A,B〉 is semi-orthogonal, then 〈A,B〉 is saturated. 2.4 Equivariant triangulated categories The main triangulated categories we will concern ourselves with are derived categories of quotient stacks. The notion of a G-equivariant objects is central here. Definition 2.4.1. Suppose G is a finite group. An action of G on a triangulated category T is the following data, [16, §3.1]: – For every g ∈ G, an exact autoequivalence g∗ : T → T ; – For every g, h ∈ G, an isomorphism of functors εg,h : (gh)∗ ∼−→ h∗ ◦ g∗ satisfying the usual associativity conditions. Definition 2.4.2. A G-equivariant object of T is a pair (t, θ), where t ∈ T and θ is a collection of isomorphisms θg : t ∼−→ g∗t for all g ∈ G satifying the usual associativity diagram. An object of t ∈ T together with an equivariant structure θ is called a linearization of t. For any action of G on a triangulated category T , we can form the category of equivariant objects of T denoted T G.4 Example 2.4.1. Suppose a finite group G acts on a scheme X, then there is an exact equivalence D[X/G] ∼= D(X)G, see [25, Section 3.8]. So in this case there is a 4It is not true that T G is always triangulated, i.e. that the triangulated structure on T descends to T G, see [11, Example 8.4]. In fact, this example shows that even if the G-action lifts to the dg level, it may not be true that the G-equivariant objects of a pretriangulated dg category is pretriangulated. 21 natural triangulated structure on D(X)G. A good reference for equivariant derived categories of coherent sheaves is [8, Section 4]. Example 2.4.2. Let (F , θ) be an equivariant object in D(X). Further, let χ : G→ Gm be a multiplicative character of G. Define a new equivariant object (F(χ), θ · χ) where F(χ) = F as an object of D(X) but the maps θg · χ : F → g∗F are twisted by χ. If F is a vector bundle, then an equivariant structure is equivalent to a compatible fiberwise G-representation. Tensoring with a character corresponds to tensoring the fiberwise G-respresentation by the same character. Thus if F admits one linearization it can admit several distinct linearizations. We denote these adjusted linearizations by F(χ) and say F(χ) is the twist of F by χ. More generally, if V is a representation of G, and (F , θ) is a G-equivariant sheaf, we can tensor with V to get a new G-equivariant sheaf (F ⊗V, θ⊗ ρV ), where ρV : G → GL(V ) is the representation and F ⊗ V is F⊕ dim(V ). We will denote this new G-equivariant sheaf by F ⊗ V . This can be made very precise as follows. Let pi : X → ? be projection to a point. Giving ? the trivial G-action makes pi equivariant and so we have a G- equivariant pullback functor pi∗ : D[?/G]→ D[X/G]. The category D[?/G] is the derived category of the category of G-representations. So by F ⊗ V , we mean the object F ⊗ pi∗V . 22 2.5 Fourier-Mukai functors Let X and Y be smooth projective varieties and denote the two projections piX : X × Y → X and piY : X × Y → Y. Definition 2.5.1. Let K ∈ D(X ×Y ). The induced Fourier-Mukai transform or Fourier-Mukai functor is the functor ΦK : D(X)→ D(Y ) given by E · 7→ piY ∗(pi∗X(E ·)⊗K). The object K is called a Fourier-Mukai kernel. It can be used to define a functor D(Y ) → D(X) in an analogous way. In this dissertation, we will explicitly state which way the functor goes to avoid awful notation like ΦX→YK . Remark 2.5.1. If G is a finite group acting on Y . We can define equivariant Fourier-Mukai functors. See [8, Section 4.1] for a short discussion and [1, Section 2] for a long discussion. In this case, where X, Y are smooth projective, all triangulated functors D(X) → D(Y ) are of Fourier-Mukai type. In fact, this is still true after in the quasi-compact separated world, [24, Theorem 8.9] One important aspect of having a Fourier-Mukai functor is the existence of adjoints. Indeed, suppose K is a Fourier-Mukai kernel defining an exact functor 23 ΦK : D(X)→ D(Y ). Define KL = K∨ ⊗ pi∗Y ωY [dim(Y )] and KR = K∨ ⊗ pi∗XωX [dim(Y )]. Proposition 2.5.1 ([14, Proposition 5.9]). Let ΦK : D(X) → D(Y ) be the Fourier- Mukai transform with Fourier-Mukai kernel K. Then ΦKL is the left adjoint and ΦKR is the right adjoint to ΦK. We shall also need the well known fully-faithfulness criterion of Bondal and Orlov. A consequence of the proof, which doesn’t seem to be widely used, is that the target category need not be D(Y ) for some smooth stack Y . The functor just needs to have a right adjoint. We will use Op to denote the structure sheaf of a closed point p ∈ X. Theorem 2.5.1 ([14, Proposition 7.1]). Let X be a smooth projective variety over k and T be a triangulated category. Suppose F : D(X) → T is an exact functor with a right adjoint G. Then F is fully-faithful if and only if for any two closed points x, y ∈ X we have HomT (F (Ox), F (Oy)[i]) =  k if x = y and i = 0 0 if x 6= y and i /∈ [0, dim(X)]. Proof. The proof in [14, Proposition 7.1] only requires that the functor F has a right adjoint. 24 Remark 2.5.2. We will use Theorem 2.5.1 when T is the derived category of a smooth quotient stack over k. In this case, the existence of a (left and) right adjoint is given by Proposition 2.5.1 and the existence of dualizing sheaves in this case. Remark 2.5.3. It would be interesting to know if X can be replaced with a DM stack or, more specifically, a global quotient stack [X/G] with G a finite group. That is, suppose Ψ: D[X/G]→ T is an exact functor, then Ψ is fully-faithful if and only if for all stacky points x, y ∈ [X/G] we have HomT (F (Ox), F (Oy)[i]) =  k if x = y and i = 0 0 if x 6= y and i /∈ [0, dim(X)]. . Note, since G is finite dim[X/G] = dim(X). The current proof relies on using the Hilbert scheme; however, the G-Hilbert scheme is not as well understood, if it exists. 2.6 Spanning classes The easy part to prove in a semi-orthogonal decomposition is vanishing of extension groups. The harder part is proving fullness. The notion of a spanning class, introduced by Bridgeland in [6], gives a natural way of proving fullness. Definition 2.6.1. Let T be a triangulated category. A subclass of objects Ω ⊂ T is called a spanning class if for every t ∈ T the following two conditions hold: – HomT (t, ω[i]) = 0 for all ω ∈ Ω and all i ∈ Z implies t = 0; – HomT (ω[i], t) = 0 for all ω ∈ Ω and all i ∈ Z implies t = 0. 25 In the presence of a Serre functor, it suffices to require only one of the two conditions. Example 2.6.1. If X is a smooth projective variety over k, then a spanning class is furnished by the structure sheaves of closed points: Ω = {Ox | x ∈ X is a closed point}. This is a derived analogue of the fact that a sheaf on a variety is zero if and only if the stalks at each closed point are zero. More generally, if X is a smooth and proper Deligne-Mumford stack over k, then we can consider the coarse moduli space pi : X → X and take the following collection as a spanning class: Ω = {OZ | Z is a closed substack of X and pi(Z) is a closed point of X}. In the cases we will consider, X is a global quotient stack X = [X/G] with G a finite group. The class Ω is provided by (twists of) orbits of points under the G-action. On one extreme, the orbit could be free: if p ∈ X and |G · p| = |G|, then set Z = G · p to be the corresponding G-cluster. Then OZ ∈ Ω. On the other extreme, the orbit could be a single point: if G · p = p, then for each irreducible representation of V , we have Op ⊗ V ∈ Ω. Our actions will either be one of these extremes so we do not discuss the intermediate cases. The next proposition shows how useful spanning classes can be. Proposition 2.6.1. Suppose Ω is a spanning class for T and A is a full, admissible, triangulated subcategory containing Ω, then A = T . 26 Proof. Since A is admissible, there is a semi-orthogonal decomposition of T of the form: T = 〈A⊥,A〉 The condition that A contains a spanning class implies that A⊥ must be trivial. Example 2.6.2. The condition that A is admissible cannot be removed. Indeed, let X be a smooth projective scheme. Let A be the full subcategory of D(X) from Example 2.3.2. Then, by definition, A has a spanning class; however, the only objects in A are torsion (with zero dimensional support) and so they cannot generate. It will be convenient to use more than one embedding in Section 4.4. The following theorem tells us which other functors we can use. Theorem 2.6.1. Suppose F : D(X) → T is a full embedding where X is smooth and projective over k. Further suppose there exists a saturated subcategory A containing F (Ω), where Ω is a spanning class for D(X). Then F (D(X)) ⊂ A. Proof. It is sufficient to prove that the right adjoint G : T → D(X) is zero on A⊥. Suppose b ∈ A⊥. For every ω ∈ Ω and i ∈ Z we have ExtiX(ω,G(b)) ∼= HomA(F (ω), b[i]) = 0. Since Ω is a spanning class, we conclude G(b) = 0 for all b ∈ A⊥. 27 2.7 Derived Category of [Pm+n−1/µd] This section serves to illustrate how to work with equivariant triangulated categories of quotient stacks as well as provide a useful semi-orthogonal decomposition for D[P1/µd] and more generally a semi-orthogonal decomposition for cyclic quotients with fixed locus a smooth divisor (see Theorem 2.7.2). For the rest of this dissertation, we set χ : µd → Gm denote the standard primitive character χ(λ) = λ. There is an action of µd on Pm+n−1, where Pm+n−1 has coordinates [x1 : . . . : xm : y1 : . . . : yn] and µd acts by scaling the variables y1, . . . , yn: λ · [x1 : · · · : xm : y0 : · · · : yn] = [x1 : · · · : xm : λy1 : · · · : λyn]. In terms of the homogeneous coordinate algebra k[x1, . . . , xm, y1, . . . , yn], the variables yi have weight χ −1 and the variables xi have trivial weight. Let Hy = V (x1, . . . , xm) and Hx = V (y1, . . . , yn). The fixed locus of the µd action is (Pm+n−1)µd = Hx unionsqHy. Therefore the sheaves OHx and OHy have a natural equivariant structure given by the identity morphism. As in Example 2.4.2, we can form the equivariant sheaves OHx(χi) and OHy(χi) for i = 0, . . . , d− 1. We equip O(−1) with the µd-linearization θλ : O(−1) → λ∗O(−1) given by fiberwise multiplication by λ and consider O(i) with the induced µd-linearizations. We can also twist these sheaves by characters to get the equivariant line bundles OPm+n−1(i)(χj) for i ∈ Z and j = 0, . . . , d− 1. The canonical bundle on Pm+n−1 is O(−m − n). It is locally trivial as a µd- equivariant bundle; however, the identification ωPm+n−1 ∼= O(−m − n) may involve twisting by a character. To determine the twist, we recall the Euler exact sequence 28 on Pm+n−1 0→ Ω1 → O(−1)⊕m+n α−→ O → 0 where α = (x1, . . . , xm, y1, . . . , yn). Since the sections yi have weight −1, the above Euler exact sequence admits the following µd-linearization: 0→ Ω1 → (⊕mi=1O(−1))⊕ (⊕nj=1O(−1)(χ−1)) α−→ O → 0 Now taking determinants yields ω[Pm+n−1/µd] ∼= OPm+n−1(−m − n)(χ−n) as µd- equivariant sheaves. Serre duality therefore takes the following form: Proposition 2.7.1 (Serre Duality). For any F ,G ∈ D[Pm+n−1/µd] there is a natural isomorphism Ext∗[Pm+n−1/µd](F ,G) ∼= Extm+n−1−∗[Pm+n−1/µd](G,F(−m− n)(χ−n)). Let us consider the case m = n = 1. In this case, we can describe a useful semi-orthogonal decomposition of [P1/µd]. The projective line is a coarse moduli space for [P1/µd] and the mapping pi : [P1/µd] → P1 is defined by the µd-equivariant morphism p˜i : P1 → P1 given by [x : y] 7→ [xd : yd]. Since p˜i can be described as as the d-uple embedding P1 → P 12d(d+1) followed by the linear projection onto the xd, yd variables, we have pi∗OP1(−1) ∼= O[P1/µd](−d). The fixed orbit consists of two points {p = [1 : 0], q = [0 : 1]}. In the notation before, we have Hx = {p} and Hy = {q}. The following semi-orthogonal decomposition is used in Section 4.5. 29 Theorem 2.7.1. For each i ∈ Z there is a semi-orthogonal decomposition D([P1/µd]) = 〈Op(χd−1), . . . ,Op(χ),Oq(χ−(d−1)), . . . ,Oq(χ−1), pi∗D(P1)〉 = 〈Op(χd−1), . . . ,Op(χ),Oq(χ−(d−1)), . . . ,Oq(χ−1),O(−di),O(−d(i− 1))〉. We prove something slightly more general than Theorem 2.7.1 regarding µd- actions. See also [15] for more on the derived categories of cyclic quotients. Theorem 2.7.2. Let µd act on a smooth projective variety X of dimension n. Suppose the geometric quotient pi : X → X/µd is smooth and the fixed locus Xµd = Z is a smooth divisor such that µd acts freely on X\Z such that NZ/X ∼= L(χ−1) for some fixed line bundle L on Z, where NZ/X is the normal bundle. Let ι : Z ↪→ X denote the inclusion. Then there is a semi-orthogonal decomposition of D[X/µd]: D[X/µd] = 〈ι∗(D(Z))(χ), . . . , ι∗(D(Z))(χd−1), pi∗D(X/µd)〉. Proof. We first show ι∗ : D(Z) → D[X/µd] is fully-faithful using Theorem 2.5.1. Pick z ∈ Z. Since NZ/X,z ∼= χ−1, we have an isomorphism of µd-representations: Ext∗X(Oz,Oz) ∼= Λ∗(TzX) ∼= Λ∗(1⊕n−1 ⊕ χ−1). It follows that ι∗ is fully-faithful. Semi-orthogonality follows from this identification as well. 30 To see fullness, take an object F ∈ D[X/µd] such that F is left orthogonal to 〈ι∗(D(Z))(χ), . . . , ι∗(D(Z))(χd−1)〉. As the action of µd on X \ Z is free, we can apply [23, Theorem 2.4] to see F ∈ pi∗D(X/µd). Remark 2.7.1. Of course the theorem can be adapted to the case where NZ/X has different weights. However, the components ι∗D(Z)(χi) may need to be reordered to ensure semi-orthogonality. The correct reordering for [P1/µd] is provided in the statement and so Theorem 2.7.2 proves Theorem 2.7.1. Remark 2.7.2. The sheaves O(n) have a natural µd-equivariant structure and so we could equally as well have considered an equivariantized Beilinson’s exceptional collection: D[P1/µd] = 〈O(−1),O(−1)(χ) . . . ,O(−1)(χd−1),O,O(χ), . . . ,O(χd−1)〉. We could then tediously argue that the decomposition of Theorem 2.7.1 is a mutation of Beilinson’s collection. The above argument is more pleasant and Theorem 2.7.2 will be needed in Section 4.6. The classical Grothendieck splitting theorem decomposes any vector bundle on P1 as a sum of line bundles. We have the following equivariant version of this result which is used in Section 3.6 in the case G = µd. Theorem 2.7.3 (Equivariant Grothendieck Splitting). Let E be a rank r vector bundle on [P1/G], where G is a finite Abelian group. Then there exists ni ∈ Z, 31 χi ∈ Gˆ for i = 1, . . . , r such that E ∼= ⊕ri=1O(ni)(χi) Proof. The proof is almost identical to the classical proof, see [19, Section 2.1]. The Abelian condition ensures the irreducible representations are one dimensional and the twist by characters χi show up when looking for an equivariant global section of E(ni) with ni >> 0. Remark 2.7.3. In the case G = µd, the character group is generated by χ. If E is a µd-equivariant vector bundle of rank r on P1, then we have E ∼= r⊕ i=1 O(ni)(χsi) for 0 ≤ si < d− 1. 32 CHAPTER III SETUP AND EMBEDDINGS 3.1 Setup Let Xf ⊂ Pm−1 and Xg ⊂ Pn−1 be smooth degree d hypersurfaces. Let X = V (f ⊕ g) ⊂ Pm+n−1 be the hypersurface associated to the Thom- Sebastiani sum of potentials. We impose the conditions d ≥ n ≥ m ≥ 2, i.e. the hypersurfaces involved are Calabi-Yau or general type and are non-empty and dim(Xg) ≥ dim(Xf ). The latter condition is for purely computational purposes. Proposition 3.1.1. The hypersurface X is smooth. Proof. The gradient of f ⊕ g is ∇(f ⊕ g) = [∇f ∇g]. Suppose ∇(f ⊕ g)([p : q]) = 0. Then for each i and j, we have ∂xi(f)(p) = 0 and ∂yj(g)(q) = 0. By the Euler formula, we have f(p) = 1 d ∑ i ∂xif(p) = 0 and g(q) = 1 d ∑ j ∂yjg(q) = 0. Since f, g both define smooth hypersurfaces, it must be that p = [0 : · · · : 0] and q = [0 : · · · : 0] which is impossible. We conclude X is smooth. 33 The action of µd on Pm+n−1 from Section 2.7 descends to X and we consider the quotient stack [X/µd]. The fixed loci are given by the intersections with (Pm+n−1)µd = Hx unionsqHy: Xµd = X ∩ (Hx unionsqHy) ∼= Xf unionsqXg. 3.2 Equivariant geometry of X Line bundles associated to hyperplane sections OX(iH) have d distinct equivariant structures. These equivariant line bundles are of the form OX(iH)(χj). Proposition 3.2.1 (Serre Duality). The triangulated category D[X/µd] has the Serre functor (−)⊗OX(d−m− n)(χ−n)[m+ n− 2]. Proof. Since [X/µd] is a smooth substack of [Pm+n−1/µd] with normal bundle isomorphic to OX(d), we can use the adjunction formula ω[X/µd] ∼= ω[Pm+n−1/µd] ⊗O[X/µd](d) ∼= OX(d−m− n)(χ−n). For Fano hypersurfaces it is easy to see that line bundles are exceptional. With this extra µd action, all line bundles on [X/µd] are exceptional. Proposition 3.2.2. Line bundles are exceptional objects of D[X/µd]. Proof. It is sufficient to prove H∗(OX)µd ∼= k. We have an equivariant exact sequence on Pm+n−1: 0→ OPm+n−1(−d) f⊕g−−→ OPm+n−1 → OX → 0. 34 We therefore only need to show the vanishing of Hm+n−2(OX)µd . We have an isomorphism Hm+n−2(OX)µd ∼= Hm+n−1(OPm+n−1(−d))µd . If d < m+ n, then the latter group is zero and we are finished. Suppose d ≥ m+ n. Then Hm+n−1(OPm+n−1(−d))µd ∼= H0(OPm+n−1(d−m− n)(χ−n))µd . The latter has a basis of monomials of the form xIyJ where I = (i1, . . . , im), J = (j1, . . . , jn) and x I = xi11 · · ·ximm , yJ = yj11 . . . yjnn such that |I| + |J | = d−m− n and |J | + n is a positive multiple of d. It follows that |J | = d − n is the only possible option. Hence, d−m−n = |I|+ |J | = |I|+d−n from which we conclude |I| = −m, impossible. Proposition 3.2.3. For 0 < i− j < m, H∗(OX(χj−i)) = 0. Proof. Clearly equivariant global sections are zero. By Proposition 3.2.2, there is an isomorphism Hm+n−2(OX(χj−i))µd ∼= Hm+n−1(Om+n−1P (d−m− n)(χi−j−n))µd . As in the preceeding proof, we must have a monomial xIyJ where |I| + |J | = d − m− n and |J |+ n+ j − i is a positive multiple of d. Thus |J | = d− n− j + i is the only possible obtion. Hence, d −m − n = |I| + |J | = |I| + d − n − j + i. Therefore |I| = i− j −m < m−m = 0, which is impossible. 35 3.3 Subcategory of exceptional line bundles. Define subcategories A1,A2,A3 of D[X/µd] as follows. A1 = 〈OX(−(n− 1)− (m− 1))(χ−(n−1)), OX(−(n− 1)− (m− 1) + 1)(χ−(n−2),−(n−1)), . . . ,OX(−(n− 1)− 1)(χ−(n−m)−1,...,−(n−1))〉; A2 = 〈OX(−(n− 1))(χ−(n−m),...,−(n−1)), OX(−(n− 1) + 1)(χ−(n−m)+1,...,−(n−1)+1), . . . ,OX(−(m− 1)− 1)(χ−1,...,−(m−1))〉; A3 = 〈OX(−(m− 1))(χ0,...,−(m−1)), OX(−(m− 2))(χ0,...,−(m−2)), . . . ,OX〉. It is understood that if m = n, then A2 is zero. Further, the notation OX(i)(χj1,...,jk) means the subcategory generated by the exceptional objects OX(i)(χj1), . . . ,OX(i)(χjk). By Proposition 3.2.3, there is a semi-orthogonal decomposition: OX(i)(χj1,...,jk) = 〈OX(i)(χjk), . . . ,OX(i)(χj1)〉, where j1 > j2 > · · · > jk. Proposition 3.3.1. The decomposition of the subcategories Ai for i = 1, 2, 3 is semi-orthogonal. Moreover, A = 〈A1,A2,A3〉 is semi-orthogonal. Proof. We only show the semi-orthogonality of the decomposition for A3. The semi-orthogonality of the decomposition for A1 and A2 is similar. 36 The sheaves in A3 are of the form O(−i1)(χ−j1) for 0 ≤ i ≤ m − 1 and 0 ≤ i1 ≤ i2. Pick two such sheaves with i1 ≤ i2 and j1 ≤ j2. By Serre Duality and the closed substack exact sequence: Hm+n−2(OX(i1 − i2)(χj1−j2)) ∼= H0(OPm+n−1(d+ i2 − i1 − n−m)(χj2−j1−n)). We have the following inequalities: m− 1 ≥ i2 − i1 ≥ 0, m− 1 ≥ j2 − j1 ≥ 0. An equivariant global section is of the form xIyJ where |I|+ |J | = d+ i2− i1− n−m such that |J | = j2 − j1 − n+ d. But d+ i2 − i1 − n−m = |I|+ |J | = |I|+ d+ j2 − j1 − n. Hence, |I| = i2−i1−(j2−j1)−m ≤ i2−i1−m ≤ m−1−m = −1, which is impossible. Thus H∗(OX(i1 − i2)(χj1−j2)) = 0 and the semi-orthogonal decomposition for A3 is verified. We now check 〈A2,A3〉, the semi-orthogonality computations for 〈A1,A2〉 and 〈A1,A3〉 are similar. Recall, for A2 to exist we require n > m. The relevant group is Hm+n−2(OX(i1 − (m− 1)− i2)(χj1−j2))µd ∼= H0(OPm+n−1(d+ i2 − i1 − n− 1)(χj2−j1−n))µd 37 for i1 = 0, . . . ,m− 1, j1 = 0, . . . , i1, i2 = 1, . . . , n−m, j2 = i2, . . . , (m− 1) + i2. If d+i2−i1−n−1 < 0, there is nothing to prove. Assume d+i2−i1−n−1 ≥ 0. Let xIyJ be an equivariant global section. Since j2 − j1 − n < 0 we require |J | = d+ j2 − j1 − n. Then d+ i2 − i1 − n− 1 = |I|+ |J | = |I|+ j2 − j1 − n forces |I| = i2 − j2 + j1 − i1 − 1. However, i2 − j2 ≤ 0 and j1 − i1 ≤ 0 so |I| ≤ −1 which is impossible. This finishes the proof. 3.4 Embedding D(Xf ) Let ιf : Xf → X be given by ιf ([x1 : . . . : xm]) = [x1 : . . . : xm : 0 : . . . : 0]. Clearly ιf is µd-equivariant as it coincides with a component of the fixed locus. Let ιf∗ denote the corresponding equivariant pushforward functor ιf∗ : D(Xf ) → D[X/µd]. This means first include D(Xf ) into the trivial component of D[Xf/µd] = d−1⊕ i=0 D(Xf )(χi), then use the equivariant pushforward. Proposition 3.4.1. If d > n, then ιf∗ is fully-faithful. Proof. We use Theorem 2.5.1. Let p ∈ Xf be a closed point and identify p with ιf (p). It is sufficient to show vanishing of Ext ∗ [X/µd] (Op,Op) ∼= (Λ∗TpX)µd for ∗ > m− 2. From the normal bundle exact sequence 0→ TX → TPm+n−1|X → OX(d)→ 0 38 and the identification TpPm+n−1 ∼= 1⊕m−1⊕χ⊕n coming from the µd-linearized Euler exact sequence, we see TpX ∼= 1⊕m−2 ⊕ χ⊕n If d > n, then (Λ∗TpX)µd = 0 for ∗ > m− 2. Define the following subcategories of D(X)µd : Dif = ιf∗(D(Xf ))(χi). Orlov’s theorem predicts that for some i, these subcategories appear in the semi- orthogonal decomposition. The next proposition tells us which ones are needed. Proposition 3.4.2. For 0 < i1 − i2 < d− n we have the semi-orthogonality 〈Di1f ,Di2f 〉. Proof. Pick a closed point p ∈ Xf , then we have the isomorphism: Ext∗X(Op(χi2),Op(χi1))µd ∼= Λ∗(1⊕m−2 ⊕ χ⊕n)(χi1−i2). Provided 0 < i1− i2 < d−n, we will have a nontrivial weight on the extension group for every ∗. Definition 3.4.1. Let Df be the strictly full subcategory of D[X/µd] generated by D1f , . . . ,Dd−nf . The follwing corollary is immediate. 39 Corollary 3.4.1. For d > n, we have a semi-orthogonal decomposition Df = 〈Dd−nf ,Dd−n−1f , . . . ,D1f〉. 3.5 Embedding D(Xg) Similarly to Df , we have a closed embedding ιg : Xg → X given by ιg([y1 : . . . : yn]) = [0 : . . . : 0 : y1 : . . . : yn], which is the inclusion of the other component of the fixed locus and so is µd-equivariant. Let ιg∗ : D(Xg) → D[X/µd] be the associated equivariant pushforward. The following results are analogous to Propositions 3.4.1, 3.4.2 and Corollary 3.4.1. We include the proofs for completeness Proposition 3.5.1. If d > m, then ιg∗ is fully-faithful. Proof. As in Proposition 3.4.1, for each closed point q ∈ Xg, we have an identification TpX ∼= 1⊕n−2 ⊕ (χ−1)⊕m. If d > m, then (Λ∗TqX)µd = 0 for ∗ > n− 2. Define the subcategories Dig = ιg∗(D(Xg))(χi) of D([X/µd]). 40 Proposition 3.5.2. For m− d < i1 − i2 < 0 we have 〈Di1g ,Di2g 〉. Proof. Pick a closed point q ∈ Xg, then have the isomorphism Ext∗X(Oq(χi2 ,Oq(χi1))) ∼= Λ∗(1⊕n−2 ⊕ (χ−1)⊕m)⊗ χi1−i2 . Provided 0 > i1 − i2 > m − d, the extension group will always have a nontrivial weight and so the it must vanish. Corollary 3.5.1. For d > m we have a semi-orthogonal decomposition Dg = 〈Dm−dg ,Dm−d+1g , . . . ,D−1g 〉. In the case n > m, it will be necessary to split Dg into two subcategories. Definition 3.5.1. Define subcategories of D[X/µd]: Dg1 = 〈Dm−dg ,Dm−d+1g , . . . ,Dm−n−1g 〉. and Dg2 = 〈Dm−ng ,Dm−n+1g , . . . ,D−1g 〉. 41 We have Dg = 〈Dg1,Dg2〉, where it is understood that if m = n, then Dg = Dg1. 3.6 Embedding D(Xf ×Xg) Let Y = P(OXf (−1)OXg(−1)), i.e. the projectivization of the rank 2 vector bundle O(−1, 0)⊕O(0,−1) over Xf ×Xg. Let pi : Y → Xf ×Xg be the projection. Consider the commutative diagram: Y Xf ×Xg ×X Xf ×Xg X ι pi σ piX The cyclic group µd acts on Y by scaling the second coordinate of the fiber. We endow Xf ×Xg with the trivial action rendering the diagram µd-equivariant. Define a family of (equivariant) Fourier-Mukai functors Ξi,j : D(Xf ×Xg)→ D[X/µd] using the kernel ι∗OY ⊗ pi∗XOX(iH)(χj), i.e. Ξi,j(F ·) = RpiX∗(pi∗Xf×Xg(F ·)⊗ ι∗OY ⊗OX(iH)(χj)), where it is understood that before applying Ξi,j we precompose with the embedding D(Xf ×Xg) ↪→ D[Xf ×Xg/µd] = d−1⊕ i=0 D(Xf ×Xg)(χi) 42 into the trivial component. Then the derived push and pull functors are taken equivariantly. Proposition 3.6.1. Let (p, q) ∈ Xf ×Xg be a closed point. Then Ξi,j(Op,q) = Ol(p,q)(i)(χj) where lp,q is the line joining ιf (p) to ιg(q) inside X. Proof. Since the kernel is flat over Xf × Xg, we just have to take the restriction of the kernel to {(p, q)} × X. This is precisely the structure sheaf of the line l(p, q) with a twist. We show Ξi,j is an embedding using Theorem 2.5.1. Lemma 3.6.1. Let N denote the normal bundle to l(p, q) inside X. Then N ∼= ( m−2⊕ i=1 Ol(p,q)(1) ) ⊕ ( n−2⊕ j=1 Ol(p,q)(1)(χ) ) ⊕Ol(p,q)(2− d)(χ). Proof. By the µd-equivariant Grothendieck splitting theorem (Theorem 2.7.3), we have an isomorphism N ∼= m+n−3⊕ i=1 Ol(p,q)(ni)(χji) for some ni ∈ Z and weights ji. As X is a degree d hypersurface in Pm+n−1 and l(p, q) is a linear subvariety of Pm+n−1, the normal bundle N fits into the following equivariant exact sequence: 43 0→ N → ( m−1⊕ i=1 Ol(p,q)(1) ) ⊕ ( n−1⊕ j=1 Ol(p,q)(1)(χ) ) → Ol(p,q)(d)→ 0 on l(p, q). The weights come from the description of the morphism Ol(p,q)(1)⊕m+n−2 → Ol(p,q)(d). It is given by multiplication by (∂u1f |l(p,q), . . . , ∂um−1f |l(p,q), ∂v1g|l(p,q), . . . , ∂vn−1g|l(p,q)), where u1, . . . , um−1 are linear sections cutting out p ∈ Pm−1 and v1, . . . , vn−1 are linear sections cutting out q ∈ Pm−1. Up to a linear change of coordinates, we can assume this mapping is (ud−11 , 0, . . . , 0, v d−1 1 , 0, . . . , 0). Hence, N ∼= ( m−2⊕ i=1 Ol(p,q)(1) ) ⊕ ( n−2⊕ j=1 Ol(p,q)(1)(χ) ) ⊕O(i)(χj) Since deg(N ) = m+ n− 2− d we must have i = 2− d. By checking the fibers of the normal bundle exact sequence at p = [1 : 0], we have 0→ Np → 1⊕m−1 ⊕ χ⊕n−1 → 1→ 0 as µd-representations. Therefore Np ∼= 1⊕m−2 ⊕ χ⊕n−1 and it follows that j = 1. 44 Lemma 3.6.2. For (p, q), (p′, q′) ∈ Xf ×Xg. If p 6= p′ or q 6= q′, then Ext∗[X/µd](Ol(p,q),Ol(p′,q′)) = 0. Proof. If p 6= p′ and q 6= q′, then the subvarieties l(p, q) and l(p′, q′) are disjoint. The vanishing follows. Without loss of generality, suppose p = p′. We must compute Ext∗[X/µd](Ol(p,q),Ol(p,q′)) ∼= Ext∗OX,p(Ol(p,q),p,Ol(p,q′),p)µd Let R = ÔX,p. Then by the Cohen structure theorem, we have R ∼= k[[x1, . . . , xm−2, y1, . . . , yn]]. The action of µd on Spec(R) endows x1, . . . , xm−2 with trivial weight and y1, . . . , yn with weight -1. The completions of Ol(p,q),p and Ol(p,q′),p are isomorphic to the modules Mq = R/(x1, . . . , xm−2, y2, . . . , yn) ∼= k[[y1]] and Mq′ = R/(x1, . . . , xm−2, y1, y3, . . . , yn) ∼= k[[y2]], respectively. Since Mq is cut out by the regular sequence x1, . . . , xm−2, y2, . . . , yn, we have the following equivariant Koszul resolution (⊗m−2i=1 Rexi xi−→ R)⊗ (⊗nj=2Reyj(χ−1) yj−→ R) 45 of Mq. We apply HomR(−,Mq′): (⊗m−2i=1 Mq′e∨xi 0−→Mq′)⊗ (⊗nj=3Mq′e∨yj 0−→Mq′(χ))⊗ (Mq′ y2−→Mq′(χ)) ∼= (⊗m−2i=1 Mq′e∨xi 0−→Mq′)⊗ (⊗nj=3Mq′e∨yj 0−→Mq′χ)⊗ k(χ) Since d ≥ n, the terms appearing in Ext∗R(Mq,Mq′) will all have nontrivial weight. Indeed, the weights will be between 1 and n−1. Hence, Ext∗R(Mq,Mq′)µd = 0. Since completion is faithful, we have the desired vanishing. We can now prove Ξi,j is fully-faithful. Theorem 3.6.1. The functors Ξi,j are fully-faithful for all i, j. Proof. Using Theorem 2.5.1 and Lemma 3.6.2 we only need to show Ext∗[X/µd](Ol(p,q),Ol(p,q)) =  k ∗ = 0 0 ∗ /∈ [0,m+ n− 4] . That Hom[X/µd](Ol(p,q),Ol(p,q)) ∼= k is clear. For vanishing, we use the local-to-global spectral sequence. Since l(p, q) and X are smooth, this reduces to: Hr(ΛsN )⇒ Extr+s[X/µd](Ol(p,q),Ol(p,q)). here N is the normal bundle from Lemma 3.6.1. So we must compute Hr(ΛsN ) for (r, s) = (0,m+ n− 3), (1,m+ n− 4). We will compute separately. For the case (r, s) = (0,m+ n− 3), we have Λm+n−3N ∼= Ol(p,q)(m+ n− 4 + 2− d)(χn−1) ∼= O(m+ n− d− 2)(χn−1). 46 Suppose m + n − d − 2 ≥ 0 (otherwise there is nothing to check), then we would require a monomial of the form xayn−1 with a ≥ 0 and a + n − 1 = m + n − d − 2. Solving for a, we have a = m− d− 1 ≤ −1, but d ≥ m, which is impossible. Now for the case (r, s) = (1,m + n − 4). Since l(p, q) ∼= P1, the only way H1(Λm+n−4N ) can be nonvanishing is if O(2 − d)(χ) is involved in the product. In which case, the isotypical summands of Λm+n−4N involving O(2− d)(χ) are: O(m+ n− 3− d)(χn−2),O(m+ n− 3− d)(χn−1). If m + n − 3 − d ≥ −1, then the first cohomology group is zero without equivariance. Assume m+n−3−d ≤ −2. By Serre duality, we have an isomorphism H1(Ol(p,q)(m+ n− 3− d))µd ∼= H0(Ol(p,q)(d+ 1−m− n)(χ−n,1−n))µd We remark that d > n here; otherwise, if d = n, then m−3 ≤ −2 forces m = 1 and we assume m ≥ 2. In particular, the weights χ−n,1−n are nontrivial above. We must find a monomial of the form xayd−n where a ≥ 0 and a+d−n = d+1−m−n. This forces a = 1−m, which is absurd. Similarly, we would need a monomial of the form xayd−n+1 with a ≥ 0 and a + d − n + 1 = d + 1 −m − n and hence a = −m, which is still absurd. Thus there are no equivariant global sections and the group vanishes. We conclude Ξ0,0 is fully-faithful and so Ξi,j is fully-faithful for all i, j ∈ Z as it differs from Ξ0,0 by an autoequivalence. Definition 3.6.1. Let Dfg = Ξ−m,−nD(Xf ×Xg). By Lemma 3.6.1 we have Ξ−m,−n is a full embedding. The main result can now be stated. 47 Main Theorem. In the above notation, we have a semi-orthogonal decomposition D[X/µd] = 〈Dg1,Dfg,Dg2,Df ,A〉. The proof of this theorem will occupy Chapter IV. In Sections 4.1 and 4.2 we finish proving that the decomposition is semi-orthogonal. In Sections 4.3 and 4.4, we analyze other sheaves that we can construct from the components present. In Section 4.5 we complete the proof of fullness. It is worth noting that in the cases (m,n) = (2, 2), (2, 3), (3, 3), there is an easier proof of this result. The idea of the proof is what is used in the subsequent sections and so we believe it does no harm in proving these special cases now. Pf. in the special cases. We will only do the case (m,n) = (2, 2) with the understanding that the other two are similar. The subcategories are as follows: Dg1 = Dg = 〈D2−dg , . . . ,D−1g 〉, Dfg = Ξ−2,−2(D(Xf ×Xg)), Df = 〈Dd−2f , . . . ,D1f〉, A = 〈OX(−2)(χ−1),OX(−1)χ0,−1,OX〉, Define T = 〈Dg,Dfg,Df ,A〉. We will prove orthogonality in §IV, the difficult part is fullness. To do this we show T has a spanning class. This will be sufficient to conclude D[X/µd] = T . 48 Using Example 2.6.1, we see that the collection of objects consisting of free orbits, say OZ where Z = {λ · z}λ∈µd and λ · z 6= z}, as well as the sheaves Oιf (p)(χi) and Oιg(q)(χi) for i = 1, . . . , d form a spanning class. Let J = J(Xf , Xg) inside of X denote the join of Xf and Xg. A free orbit Z ⊂ X \ J is a complete intersection with respect to two sections sx ∈ Γ(OX(1)) and sy ∈ Γ(OX(1)(χ)). It follows that the corresponding resolution of OZ given by 0→ OX(−2)(χ−1)→ OX(−1)⊕OX(−1)(χ−1)→ OX is in the subcategory A, hence the free orbits OZ ⊂ T provided we are away from J . To see we have the remaining objects of the spanning class, we will use Theorem 2.7.1. Let l(p, q) denote the line joining ιf (p) to ιg(q). We will see in §4.3 that the objects Ol(p,q)(−d) and Ol(p,q) are in T for all p, q. It remains to see that the twists of the fixed orbits are in T . Using Dg and Df we only need one additional twist, say Op, Oq (or in the case (m,n) = (2,3),(3,3) we will also need Op(χ−1),Oq(χ)). To do that, we notice that Cone(OX(−1) → OX) ∼= OJ(p,Xg) ∈ A by cutting out p with a section of OX(1). We then have the exact sequence 0→ OJ(p,Xg) → ⊕ q∈Xg Ol(p,q) → O⊕d−1p → 0. Since T is saturated, it follows that Op ∈ T . In the case (m,n) = (2, 3), (3, 3) we can look at a similar sequence using OJ(p,Xg)(−1)(χ−1) ∈ A. A similar argument shows Oq ∈ T and Theorem 2.7.1 finishes the proof. 49 CHAPTER IV PROOF OF MAIN THEOREM In this chapter we finish proving semi-orthogonality (Section 4.1) and prove fullness (Sections 4.3) We finish the chapter by studying the case m = 1. This is the case of a cyclic cover. 4.1 Semi-orthogonality between Dfg and A As before, we let Dfg be the image of the fully-faithful functor Ξ−m,−n. Let us compute the semi-orthogonality 〈Dfg,A〉. We have the formula Ext∗[X/µd](OX(−i)(χ−j),Ol(p,q)(−m)(χ−n)) ∼= H∗(P1;O(i−m)(χj−n))µd . Lemma 4.1.1. There is a semi-orthogonal decomposition 〈Dfg,A〉. Proof. We only check the semi-orthogonality 〈Dfg,A3〉 as the other computations are similar. The objects in A3 are of the form O(−i)(χ−j), where 0 ≤ i ≤ m − 1 and 0 ≤ j ≤ i. In this case we have O(i−m)(χj−n) is a negative line bundle so H1(O(i−m)(χj−n)) ∼= H0(O(m− i− 2)(χn−j−1))∨ Since i ≥ j ≥ 0 we have n− 1 ≥ n− j − 1 ≥ n− i− 1 50 and so we need a monomial of the form xayn−j−1 where a ≥ 0 and a + n − j − 1 = m− i− 2. This is impossible because a+ n− j − 1 ≥ n− i− 1 ≥ m− i− 1 > m− i− 2. 4.2 Semi-orthogonality between Dg,Df ,A. For Df to be present in the semi-orthogonal decomposition, we need d > n and for Dg to be present, we require d > m. Lemma 4.2.1. We have the semi-orthogonality 〈Dg,Df ,A〉. Proof. That Dg and Df are semi-orthogonal is clear. We only show Df is right orthogonal to A. The claim that Dg is also right orthogonal is analgous. Let p ∈ Xf and consider the sheaves Op(χ−j). We compute Ext∗[X/µd](O(−i1)(χ−i2),Op(χ−j)) ∼= Γ(Op(χi2−j))µd which is nonzero if and only if i2 − j = 0. Since 0 ≤ i2 ≤ n − 1, if we choose n ≤ j ≤ d− 1, then 1− d ≤ i2 − j ≤ −1. These are precisely the weights in Df . This shows the semi-orthogonality 〈Df ,A〉. Lemma 4.2.2. We have the semi-orthogonality 〈Dg1,Dfg,Dg2,Df〉. Proof. Again, we only prove the semi-orthogonality 〈Dfg,Df〉 the other claims are analogous. The only possible nonzero extension group in the standard spanning 51 class for Dfg is Ext∗[X/µd](Op(χ−j),Ol(−m)(χ−n)) ∼= (Ext∗X(Op,Ol)(χj−n))µd , where l is the line between p ∈ Xf and and any point q ∈ Xg. Set R = ÔX,p, then R ∼= k[[x1, x2, . . . , xm−1, y1, . . . , yn−1]]. Here the variables xi have weight 0 and the variables yj have weight -1. The sheaf Op corresponds to the graded module kp = R/(x1, . . . , xm−1, y1, . . . , yn−1) The sheaf Ol corresponds to the graded module Mq = R/(x1, . . . , xm−1, y2, . . . , yn−1). We can take the Koszul resolution of kp: ( ⊗m−1i=1 Rexi xi−→ R ) ⊗ ( ⊗n−1j=1Reyj(χ−1) yj−→ R ) → kp and apply Hom(−,Mq). This will kill all of the maps except y1. The resulting complex has general term Mqe ∨ xI ∧ e∨yJ , where 0 ≤ |I| ≤ m− 1 and 0 ≤ |J | ≤ n− 1. It’s easy to see that the cohomology of this complex has general term ke∨xI ∧ e∨yJ where 0 ≤ |I| ≤ m − 1 and 1 ≤ |J | ≤ n − 1. The weights therefore vary between 1 and n− 1. Thus the summands of the extension group are of the form: k(χ), . . . , k(χn−1). 52 Thus the general term of Ext∗[X/µd](Op(χ−j),Ol(−m)(χ−n)) is of the form k(χ1+j−n), . . . , k(χn−1+j−n) = k(χj−1). Since n ≤ j ≤ d − 1, these terms all have nonzero weight and so the equivariant extension group vanishes. This completes the semi-orthogonal claim in the Main Theorem. It remains to see fullness. Definition 4.2.1. Define a full-subcategory T of D[X/µd] by T = 〈Dg1,D,Dg2,Df ,A〉. 4.3 Koszul complexes and Joins By Proposition 2.3.2, the subcategory T is saturated and hence admissible. Using Proposition 2.6.1, it suffices to check that T has a spanning class. This is done in Section 4.5. To do so, we need to construct more sheaves in T using the subcategories present. Essential to these constructions is the Koszul complex of a regular section of a vector bundle. Let E be a µd-equivariant locally free sheaf of rank r over X and s ∈ Γ(E)µd be an equivariant global section. Then we have the corresponding Koszul complex: 0→ ΛrE∨ → Λr−1E∨ → · · · → E∨ s∨−→ OX . Denote this complex by K(E, s). 53 If the zero locus of s is codimension r, then the Koszul complex is exact and is a locally free resolution of OZ(s), where Z(s) is the vanishing locus of s. Even if the Koszul complex is not exact we can still learn information from its cohomology sheaves, see Lemma 4.3.2. Let Z ⊂ X be a free orbit of the µd action away from the join of Xf and Xg inside X, i.e. pick p /∈ X \ J(Xf , Xg) and let Z = {λp | λ ∈ µd}. These free orbits are also called µd-clusters. Lemma 4.3.1. For each µd-cluster Z ⊂ X, we have OZ ∈ T . Proof. In this case, we notice that Z is the intersection of X with m + n − 2 hyperplanes. Moreover, we can pick sections si ∈ Γ(OX(1))µd and section tj ∈ Γ(OX(1)(χ))µd where i = 1, . . . ,m − 1 and j = 1, . . . , n − 1 such that Z is the vanishing locus of s = (s1, . . . , sm−1, t1, . . . , tn−1) ∈ Γ(E), where E = (⊕m−1 i=1 OX(1) )⊕ (⊕n−1j=1 OX(1)(χ)). The summands of K(E, s) are precisely the sheaves that occur in A. We conclude for any free orbit Z in X \ J(Xf , Xg), we know OZ ∈ A. Let p ∈ Xf and q ∈ Xg. As before, denote by l(p, q) ∼= P1 the line joining ιf (p) to ιg(q). Contrary to the case of free orbits outside of J(Xf , Xg), the structure sheaves of both fixed orbits and free orbits in the join are not complete intersections. Moreover, the structure sheaf Ol(p,q) is not a complete intersection subvariety. We can still take the corresponding Koszul complex cutting it out. Indeed, there exists a section s = (s1, . . . , sm−1, t1, . . . , tn−1) ∈ Γ(E), where E is as before, such that V (s) = l(p, q). 54 Lemma 4.3.2. As above, let K(E, s) be the Koszul complex cutting out l(p, q). Then H∗(K(E, s)) =  Ol(p,q) ∗ = 0 Ol(p,q)(−d) ∗ = −1 0 ∗ 6= 0,−1 Proof. By Bezout’s theorem, we can assume the intersection Xf ∩ V (s1, . . . , sm−2) consists of d points, say p1, . . . , pd. The intersection Xg ∩ V (t1, . . . , tn−1) is {q}. Let J denote the join in X of {p1, . . . , pd} with {q}. The Koszul complex associated to the sections (s1, . . . , sm−2, t1, . . . , tn−1) is quasi-isomorphic to the following complex 0→ OJ(−1)→ OJ → 0. To each point {p1, . . . , pd} ∈ Xf ⊂ P1 there exists a linear section si such that V (si) = pi. We have the exact sequence 0→ Ol(p,q)(−d) s2···sd−−−→ OJ(−1) s1−→ OJ(−1)→ Ol(p1,q) → 0. The claim follows. Lemma 4.3.3. The subvarieties Xg and Xf are complete intersection subvarieties. Proof. We show how to get Xg, Xf is analogous. The zero locus of the section sXg = (x1, . . . , xm) of the vector bundle EXg = OX(1)⊕m is Xg. The summands of the Koszul resolution, K(EXg , sXg) are of the form OX(−m+ i) for i = 0, . . . ,m. 55 Lemma 4.3.4. For i = 1, . . . , d −m. The components of K(EXg , sXg)(−(n − 1) + i+ t)(χ−(n−1)+t) are in T for t = 0, . . . , n− 1. Remark 4.3.1. The restriction of the equivariant structure on a hyperplane divisor of X to Xg is not the trivial structure. In particular, we have isomorphisms: OX(iH)|Xg ∼= OXg(ih)(χ−i). Proof. We check the base case i = 1. In this case, we have explicitly: K(EXg , sXg)(1)→ OXg(1)(χ−1) K(EXg , sXg)(χ−1)→ OXg(χ−1) ...→ ... K(EXg , sXg)(−(n− 1) + 1)(χ−(n−1))→ OXg(−(n− 1) + 1)(χ−1) All line bundles appearing in the resolution are already in T except the line bundle appearing in degree zero. Since OXg(j)(χ−1) ∈ T for all j, we know that the line bundles in degree zero are also in T . Suppose true for 1, . . . , i we show true for i + 1 ≤ d − m. In which case we have the following twists of the above diagram: K(EXg , sXg)(i+ 1)→ OXg(i+ 1)(χ−i−1) K(EXg , sXg)(i)(χ−1)→ OXg(i)(χ−i−1) ...→ ... K(EXg , sXg)(−(n− 1) + i+ 1)(χ−(n−1))→ OXg(−(n− 1) + i+ 1)(χ−i−1). 56 Again, all of the sheaves except those in the extreme right of the resolution are already in T by induction. That the line bundles in degree zero are in T follows since OXg(j)(χ−i−1) ∈ T as i+ 1 ≤ d−m. Lemma 4.3.5. Let J(Xf , q) denote the join of Xf and q inside X. Then OJ(Xf ,q)(i) ∈ T for i = 0, . . . , d−m. Proof. The subvariety J(Xf , q) is a complete interesection: (⊗n−1i=1OX(−1)(χ−1)→ OX)→ OJ(Xf ,q). Twisting by OX(i) for i = 0, . . . , d − m gives a general component of the Koszul resolution as OX(−(n− 1) + i+ t)(χ−(n−1)+t). The statement now follows from Lemma 4.3.4. As Xf is also a complete intersection subvariety, we have the following similar statement for OJ(p,Xg) with p ∈ Xf . Lemma 4.3.6. Let OJ(p,Xg) denote the join of p ∈ Xf with Xg. Then OJ(p,Xg)(i)(χi) ∈ T for i = 0, . . . , d− n. Proof. The presence of twists comes from the fact that Xf is cut out be the section s = (y1, . . . , yn) ∈ Γ(OX(1)⊕n(χ)). 57 Our goal is to show T has a spanning class. Recall, Example 2.6.1, if X is a smooth DM stack with coarse moduli space pi : X → X, the sheaves Ω = {Z ⊂ X | Z is a closed substack of X and pi(Z) is a closed point of X} form a spanning class. For X = [X/µd], these sheaves are the structure sheaves of the free orbits and twists of the structure sheaves of fixed orbits by all characters. In Lemma 4.3.1 we saw that the structure sheaves of the free orbits away from J(Xf , Xg) have Koszul resolutions using the sheaves in A. We will get the remaining sheaves by showing for all p ∈ Xf and q ∈ Xg we have D[l(p, q)/µd] ⊂ D[X/µd]. For that we use Theorem 2.7.1 and is carried out in the next two sections. 4.4 Other kernels. It will be convenient to use the images of other Fourier-Mukai kernels from Ξ−m,−n to Ξd−m,0 and Ξd−n,d−n. We justify their use in this subsection. Using Theorem 2.6.1 we must verify that the image of the spanning class {O(p,q)} under Ξd−m,0 and Ξd−n,d−n factors through T . We will need to start by verifying the line bundles in Theorem 2.7.1 are in T . Lemma 4.4.1. For all p ∈ Xf and q ∈ Xg we have Ol(p,q)(−d) and Ol(p,q) in T . Proof. By Lemma 4.3.2, it suffices to show Ol(p,q)(−d) ∈ T . We have the exact sequences 0→ Ol(p,q)(−m− i− 1)(χ−n)→ Ol(p,q)(−m− i)(χ−n)→ Oq(χ−(n−m)+i)→ 0. 58 Since Ol(p,q)(−m)(χ−n),Oq(χ−(n−m)), . . . ,Oq(χ−1) ∈ T , we have Ol(p,q)(−n)(χ−n) ∈ T by induction. Now consider the sequences 0→ Ol(p,q)(−n− i− 1)(χ−n−i−1)→ Ol(p,q)(−n− i)(χ−n−i)→ Op(χ−n−i)→ 0. Since Ol(p,q)(−n)(χ−n),Op(χ1), . . . ,Op(χd−n) ∈ T , we have Ol(p,q)(−d) ∈ T by induction and this completes the proof. Lemma 4.4.2. For all p ∈ Xf and q ∈ Xg we have Ol(p,q)(d − m),Ol(p,q)(d − n)(χd−n) ∈ T . Proof. Use the exact sequences in the proof of Lemma 4.4.1. Proposition 4.4.1. The functors Ξd−m,0 and Ξd−n,d−n factor through T . Proof. By Proposition 2.3.2, it follows from Lemma 4.4.2 and Theorem 2.6.1. 4.5 Proof of Fullness We now compute Ξd−m,0(OXf×{q}(−i)) and Ξd−n,d−n(O{p}×Xg(−j)) for various i = 0, . . . ,m− 1 and j = 0, . . . , n− 1 to show that this gives us the missing sheaves: Oq(χ0,1,...,m−1),Op(χ0,−1,...,−(n−1)). In particular, we show there exists triangles OJ(Xf ,q)(d−m− i)→ Ξd−m,0(OXf×{q}(−i))→ Tq → and OJ(p,Xg)(d− n− j)(χd−n)→ Ξd−n,d−n(O{p}×Xg)→ Tp → for every p ∈ Xf and q ∈ Xg, where Tq and Tp are certain torsion sheaves supported at q and p, respectively. Since the first two objects are in T we have Tq, Tp ∈ T . 59 We then build a filtration of Tq, Tp and argue then that T has the remaining elements of the spanning class. To perform these computations, we need to add an auxiliary blowup. Let q ∈ Xg and let Pm be the linear subspace spanned by x1, . . . , xm, q. Consider the following commutative diagram P(OXf (−1)OXg(−1)) P(OXf (−1)⊕OXf ) = Z Bl = Blq Pm Xf ×Xg Xf Pm ιq pi σ β j where j includes Xf via j(x) = (x, q), Z is the fibered product, and σ also denotes the restriction of σ : Y → Pm+n−1 which factors through Pm. The mapping ιq includes Y as the divisor dH − dE in Bl. Here Pm has coordinates [x1 : · · · : xm : y] and µd acts by scaling the y coordinate. The µd-action lifts to Bl fixing the exceptional divisor pointwise and thus rendering the enter diagram µd-equivariant. Recall, the canonical bundle of [Pm/µd] is isomorphic to OPm(−m − 1)χ−1. The usual formula for the canonical bundle of a blowup yields ωBl ∼= β∗OPm ⊗ O((m − 1)E) which admits a µd-linearization since the divisors involved are invariant under the µd-action. It remains to determine if there is a twist by a character. Restricting to Bl \E gives the isomorphism ωBl|Bl \E ∼= β∗ωPm|Bl \E and so it follows ωBl ∼= β∗OPm(−m− 1)⊗OBl((m− 1)E)(χ−1). Let H1 be a hyperplane section of Xf and H be a hyperplane section of Pm which restricts to H1 under the inclusion Xf → Pm where [x] 7→ [x : 0]. Then H − E|Y ∼= pi∗H1. 60 Theorem 4.5.1 (Equivariant Grothendieck Duality). There is a natural isomorphism: Rβ∗O[Bl /µd](D) ∼= RHom[Pm/µd](Rβ∗(ω[Bl /µd](−D)), ω[Pm/µd]) for any µd-invariant divisor D on Bl. Proof. This follows since β is µd-equivariant and the usual Grothendieck duality, [14, Theorem 3.34], is natural, hence commutes with automorphisms. The divisors on Bl are, up to equivalence, well known to be of the form aH + bE for a, b ∈ Z. Using the projection formula, we have Rβ∗ωBl(−(aH + bE) ∼= (Rβ∗O[Bl /µd]((m− 1− b)E))⊗ ω[Pm/µd](−aH) and by Grothendieck duality Rβ∗(O[Bl /µd](aH + bE)) ∼= RHom[Pm/µd](Rβ∗(O[Bl /µd]((m− 1− b)E),OPm)(−aH). Remark 4.5.1. Since {q} is codimension m, there is a canonical isomorphism Rβ∗O[Bl /µd](kE) ∼= O[X/µd] for k = 0, . . . ,m− 1. If k > 0, then Rβ∗O[Bl /µd](−kE) ∼= Ikq , where Iq is the ideal sheaf for the closed subscheme {q} in Pm. Moreover, Rβi∗O[Bl /µd](kE) = 0 unless i = 0,m− 1. 61 For i = 0, . . . ,m − 1, we consider Ξd−m,0(OXf (−iH1)). On Bl we have the divisor exact sequence 0→ O[Bl /µd](−dH + dE)→ O[Bl /µd] → ιq∗OY → 0. Since ιq∗pi∗OXf (−iH1) = ιq∗OY (−iH1) ∼= ιq∗OY ⊗ O[Bl /µd](−i(H − E)), we can consider the twist of the divisor exact sequence 0→ O[Bl /µd](−dH + dE − iH + iE)→ O[Bl /µd](−iH + iE)→ OY (−iH1)→ 0. Using the long exact sequence of cohomology sheaves for Rβ∗, we see there is an isomorphism Rσk∗OY (−iH1) ∼= Rβk+1∗ O[Bl /µd](−(d + i)H + (d + i)E). It follows from Remark 4.5.1, that the only possible nonzero higher direct image is Rσm−2∗ OY (−iH1). Lemma 4.5.1. If m > 2, then for i = 0, . . . ,m− 1, we have a distinguished triangle OJ(Xf ,q)(d−m− i)→ Ξd−m,0(OXf (−iH1))→ Hm−1((Id−m+1+iq )∨)(−m− i)[2−m]→ where Iq is the ideal sheaf of {q} in Pm, and (Id−m+1+iq )∨ is the derived dual. In particular, H∗(Ξd−m,0(OXf (−iH1))) ∼=  OJ(Xf ,q)(d−m− i) ∗ = 0 Hm−1((Id−m+1+iq )∨)(−m− i) ∗ = m− 2 0 ∗ 6= 0,m− 2 . 62 If m = 2, then for i = 0, 1 there is an exact sequence 0→ OJ(Xf ,q)(d− 2− i)→ Ξd−2,0(OXf (−iH1))→ O⊕d−1q (χ2+i−d)→ 0. Proof. The case m = 2 is easy to see directly and the vanishing statements for m > 2 follow from the preceeding discussion. It remains to show the isomorphisms. Since d + i > 0, the only possible higher direct image is in degree m − 1. We have an exact sequence 0→ O[Pm/µd](−d− i)→ O[Pm/µd](−i)→ OJ(Xf ,q)(−i)→ 0. Now H0(Ξd−m,0(OXf (−iH1))) ∼= H0(Ξ0,0(OXf (−iH1)))(d−m) ∼= OJ(Xf ,q)(d−m− i). This gives us the first arrow for m > 2. If m = 2, the first arrow is defined similarly but it is not surjective onto H0(Ξd−m,0(OXf (−iH1))). For the second we need to compute Rβm−1O[Bl /µd](−(d + i)H + (d + i)E). By Grothendieck duality and the derived functor spectral sequence, we have an isomorphism Rβm−1∗ O[Bl /µd](−(d+ i)H + (d+ i)E) ∼= RHom[Pn/µd](Rβ∗(O[Bl /µd](m− 1− d− i)E),O[Pn/µd])(−d− i) ∼= RHom[Pn/µd](Id−m+i+1q ,O[Pn/µd])(−d− i). and the second isomorphism follows by twisting by (d−m). Corollary 4.5.1. For i = 0, . . . , d−m, we have Hm−1((Id−m+i+1q )∨)(−m− i) ∈ T . 63 Proof. This now follows by Lemmas 4.3.5 and 4.5.1. It remains to compute Hm−1((Id−m+i+1q )∨)(−m− i) for i = 0, . . . ,m− 1. We seek to understand the sheaves in Corollary 4.5.1. To that end, there is an exact sequence of sheaves on Pm, where we have identified the conormal bundle of {q} with ΩPm,q: 0→ Ir+1q → Irq → Sr(ΩPm,q) ∼= O⊕N(r)q (χr)→ 0, where N(r) = ( m+r−1 r ) . Since Oq is a smooth closed subscheme of codimension m, we know Hm((Oq(χr))∨) ∼= Oq ⊗ ω∨Pm(χ−r). Since ω[Pm/µd] ∼= OPm(−m− 1)(χ−1), we see Hm((Oq(χr))∨) ∼= Oq(χ−r−m). Taking the derived dual of the above sequence now yields the short exact sequence 0→ Hm−1((Irq )∨)→ Hm−1((Ir+1q )∨)→ O⊕N(r)q (χ−r−m)→ 0. Lemma 4.5.2. For r > 0, the sheaves Hm−1((Irq )∨) have a filtration by sheaves 0 = F0 ⊂ F1 ⊂ · · · ⊂ Fr+1 = Hm−1((Irq )∨) such that Fi/Fi+1 ∼= O⊕N(r)q (χ−i−m). In particular, Hm−1((Ir+1q )∨) ∈ 〈Oq(χ−m),Oq(χ−1−m), . . . ,Oq(χ−r−m)〉. 64 Proof. Immediate from the previous discussion and the observation Hm−1(I∨q ) ∼= Oq(χ−m). Using Lemma 4.5.2, we have the following filtration of Hm−1((Id−m+1+iq )∨): Hm−1((Id−m+1+iq )∨)(d−m− i) O⊕N(d−m+i)q (χ−(d−m)) Hm−1((Id−m+iq )∨)(d−m− i) O⊕N(d−m−1+i)q (χ−(d−m)+1) ... ... Hm−1((I2q )∨)(d−m− i) O⊕N(1)q (χ−1+i) Hm−1((Iq)∨)(d−m− i) Oq(χi) Lemma 4.5.3. For all i = 0, . . . , d− 1 we have Oq(χi) ∈ T . Proof. We proceed by induction on i. When i = 0 we use the filtration given by Lemma 4.5.2 and the fact that Oq(χ−j) ∈ T for j = 1, . . . , d − m to see Oq ∈ T . Suppose we have Oq, . . . ,Oq(χi), then we use the filtration again with i + 1 to see Oq(χi+1) ∈ T . Lemma 4.5.4. For all i = 0, . . . , d− 1 we have Op(χi) ∈ T . Proof. This is analogous to the results proved in this section using the images Ξd−n,d−n(OXg(−jH2)). One requires the extra twist because the hyperplane section that restricts to H2 has nontrivial equivariant structure. Proof of Main Result. By Lemmas 4.4.1, 4.5.3, and 4.5.4 we have shown for all p ∈ Xf and q ∈ Xg we have D[l(p, q)/µd] ⊂ T . We have also shown in Lemma 65 4.3.1 that the structure sheaves of free orbits not in the join are in T . Thus T has a spanning class. Since T is admissible, by Proposition 2.6.1 we conclude T = D[X/µd]. 4.6 The Case m = 1 For completeness, we devote this section to understanding [X/µd] when m = 1. We will independently study n = 1 and n > 1. In the case n > 1, the hypersurface X is called a cyclic hypersurface. We also compare the decompositions to the work in [16] Let m,n = 1, then D(Xf ) and D(Xg) will not appear. The associated quotient stack is, up to a change of coordinates, X = V (xd+yd) ⊂ P1. In particular, |X| = d and the µd action permutes the linear factors. It follows that [X/µd] is a scheme and is represented by Spec(k). There is a single exceptional object given by OX and so D[X/µd] = 〈OX〉. Let n > 1. Then f(x) = xd and g(y1, . . . , yn) be a degree d polynomial defining a smooth hypersurface in Pn−1. Let pi : X → Pn−1 be the linear projection onto the y variables. This is well defined since [1 : 0 : . . . : 0] /∈ X. The map pi is a degree d mapping ramified along the divisor ιg : Xg ↪→ X. In particular, we have the following commutative diagram. X Xg Pn−1 pi ιg 66 Endowing Pn−1 with the trivial µd action renders the diagram commutative. Moreover, it is not hard to see that pi exhibits Pn−1 as a coarse moduli space. Theorem 4.6.1. There is a semi-orthogonal decomposition D[X/µd] = 〈D1g , . . . ,Dd−1g , pi∗D(Pn−1)〉, where Dig = ιg∗D(Xg)(χi). Proof. The action of µd on X \ Xg is easily seen to be free. Hence Theorem 2.7.2 applies. The case of a cyclic cover of a variety was investigated in [16, §8.3]. In particular, they discuss the equivariant derived category of cyclic hypersurfaces where d ≤ n, here X ⊂ Pn and Xg ⊂ Pn−1. For completeness, we recall their result. Since d ≤ n, we have the standard semi-orthogonal decomposition of a hypersurface D(X) = 〈AX ,OX , . . . ,OX(d− n)〉, where AX is characterized as the right orthogonal to 〈OX , . . . ,OX(d − n)〉. The category AX is also quasi-equivalent to the homotopy category of graded matrix factorizations of the potential f , where f is the defining equation for X. Theorem 4.6.2. In the above notation, if d ≤ n, then there is a decomposition D[X/µd] = 〈AµdX ,OX(χ0,...,d−1), . . .OX(d− n)(χ0,...,d−1)〉 where AµdX = 〈AXg ,AXg(χ), . . . ,AXg(χn−2)〉. 67 where AXg is viewed as a subcategory of D[X/µd] via ιg∗. Remark 4.6.1. Note, their results do not apply when d > n because pi : X → Pn−1 is not a cyclic cover in the sense of [16]. Indeed, suppose X is the relative spectrum associted to the the line bundle O(i) with a section s ∈ Γ(O(di)) over Pn−1. That is, we consider the sheaf of algebras A = O(−di)⊕ · · · ⊕ O on Pn−1 and X ∼= Spec(A). Then Rpi∗OX ∼= A as sheaves on Pn−1. Therefore we would have an isomorphism Hn−1(OX) ∼= d⊕ i=1 Hn−1(OPn−1(−di)). Using the divisor exact sequence for X we see Hn−1(OX) ∼= Hn(OPn(−d)) ∼= H0(OPn(d − n − 1)) which is of dimension ( d−1 d−n−1 ) . For the right hand side we have d⊕ i=0 Hn−1(OPn−1(−di)) ∼= d⊕ i=0 H0(OPn−1(di− n)), which is of dimension d∑ i=1 ( di di− n ) > ( d d− n ) = ( d n ) > ( d− 1 n ) = ( d− 1 d− 1− n ) . It follows that pi cannot be a cyclic cover. When d = n the subcategory AXg is all of D(Xg). Using the notation Dig = ιg∗(D(Xg))(χi), the decomposition of Theorem 4.6.2 is D[X/µd] = 〈D0g ,D1g , . . . ,Dn−2g ,OX(χ0,...,n−1)〉. 68 The decomposition of Theorem 4.6.1 is D[X/µd] = 〈D1g , . . . ,Dn−1g , pi∗(D(Pn−1))〉. It follows that our decomposition agrees with theirs up to a twist by a character. 69 CHAPTER V COMPARISON WITH ORLOV’S FUNCTORS ON POINTS In this chapter, we show in the case of two Calabi-Yau hypersurfaces, i.e. m = n = d, that the functor Ξ agrees with Orlov’s on points. To do so we remind the reader of graded matrix factorizations and singularity categories in Section 5.1. In Section 5.3, we discuss Orlov’s theorem and how to compute Orlov’s functors. In Section 5.4, we do some computations on special objects. In Section 5.5 we show that our functor agree, up to a mutation with Orlov’s functor on points. 5.1 Graded Matrix Factorizations Let R = Spec(Sym(V ∨)) be functions on a finite dimensional vector space V . Then R has a natural Z-grading so that Rd = Sym d(V ∨), which corresponds to the diagonal action of Gm on V with weight 1. By a graded R-module, we mean an R-module M together with a Z-grading such that Rm ·Mn ⊂Mm+n. We will refer to the Z-grading on M as the internal grading. Morphisms between graded R-modules are R-module morphisms that preserve the grading. Denote the category of Z-graded R-modules by Gr −R. This is an Abelian category. If M ∈ Gr − R, then there is a internal grading shift functor, denoted by (1), where we define the graded R-Module M(1) by (M(1))m = Mm+1. Denote by (q) 70 the qth power of (1). If M,N ∈ Gr − R, then there is a natural grading on the (ungraded) R-module HomR(M,N), where HomR(M,N)m = ⊕ m HomGr−R(M,N(m)). Call this, now graded, R-module HomR(M,N). As usual, it is the internal hom in Gr −R. Let f ∈ Rd be a degree d polynomial. Then f defines a hypersurface X = V (f) ⊂ P(V ). We assume X is smooth or equivalently the affine cone C(X) ⊂ V has an isolated singularity at the origin. Definition 5.1.1. A graded matrix factorization is a pair of morphism δ0 : P−1 → P0, δ−1 : P0 → P−1(d) between graded projective R-modules such that δ−1δ0 = f = δ0δ−1. When no confusion arises, we denote a matrix factorization by P . Equivalently, a graded matrix factorization is a curved complex of Gm- equivariant vector bundles on V with curvature f . Example 5.1.1. Consider the case dim(V ) = 2. Then R ∼= k[x, y] with its usual Z- grading. Consider f = xy. Then there are two (inequivalent) matrix factorizations lx : R x−→ R(1) y−→ R(2) 71 and ly : R y−→ R(1) x−→ R(2). It will follow from Orlov’s Theorem that these matrix factorizations are inequivalent. Example 5.1.2 (Koszul matrix factorizations). Let s1, . . . , sr be a homogeneous regular sequence in R such that f ∈ (s1, . . . , sr). Pick homogeneous elements t1, . . . , tr such that f = t1s1 + · · ·+ trsr. Consider the Koszul complex associated to si: K · : ⊗ri=1 (R(| − si|)e∨i si−→ R), where the e∨i is just a place-holder. Objects of this complex are of the form R(−|si1 | − · · · − |sik |)e∨i1 ∧ · · · ∧ e∨ik for a subset {i1, . . . , ik} ⊂ {1, . . . , r}. The differential is given by contraction with s = s1e1 + · · ·+ srer. Call this morphism ιs. Define the morphism ds = (t1e ∨ 1 + · · ·+ tre∨r ) ∧ (−). Define the associated Koszul matrix factorization as follows. Set P−1 = ⊕ i odd Ki(d(i− 1)), P0 = ⊕ i even Ki(di) 72 with differential given by ιs and ds. Clearly dsιs = ιsds = f . An important example of a Koszul matrix factorization is furnished by R = k[x1, . . . , xn] and si = xi for i = 1, . . . , n. Then we can take, up to scale, ti = ∂xif . This matrix factorization is often called the stabilization of the residue field (see Proposition 5.2.1). Morphisms between two graded matrix factorizations are morphisms of underlying R-modules making the relevant diagram commute. Suppose we have two matrix factorizations P ,P ′. Then the set of morphisms between them can be enriched to a Z-graded vector space. For n = 2l, define HomMF gr(f)(P ,P ′)n = HomGr−R(P−1, P ′−1(dl))⊕ HomGr−R(P0, P ′0(dl)) and for n = 2l + 1, define HomMF gr(f)(P ,P ′)n = HomGr−R(P0, P ′−1(d(l + 1)))⊕ HomGr−R(P−1, P ′0(dl)). This gives a category enriched in graded vector spaces. We can further enhance it to a dg category, see [1, Definition 3.1] for a more precise statement, by defining dϕ = δP ′ ◦ ϕ± ϕ ◦ δP . There is also a natural triangulated structure on the category MFgr(f) making it a pretriangulated dg category. We set HMFgr(f) to be the homotopy category. Matrix factorizations show up, in different guises, in several places in algebraic geometry, commutative algebra, and physics. One interesting example 73 connections matrix factorization to commutative algebra through maximal Cohen- Macaulay modules over the hypersurface algebra A = R/(f). Recall, a graded maximal Cohen-Macaulay A-module is a graded A-module such that depthA(M) = depth(A) = n− 1. If we have a graded maximal Cohen-Macaulay A-module, then by the Auslander- Buchsbaum formula, M has a projective resolution as an R-module of length 2: F · : F−1 δ−1−−→ F0 →M → 0. Multiplication by f induces a map of chain complexes ·f : F · → F ·(d). Since f acts as zero on M , this map is null-homotopic. Define δ0 : F0 → F−1(d) to be this nullhomotopy. Then by definition δ0δ−1 = δ−1δ0 = f . Conversely, given a matrix factorization (F·, δ·), we can construct a maximal Cohen-Macaulay module cok(δ−1) by taking the cokernel of δ−1: F−1 δ−1−−→ F0 → cok(δ−1)→ 0. Since δ0δ−1 = f , which is injective, we know δ−1 is injective. So pdR(M) = 1. Moreover, M is an A-module. Indeed, pick m ∈ M and m¯ ∈ F0 a preimage under the projection. Then fm = pi(fm¯) = pi(δ−1δ0(m¯)) = 0 since pi ◦ δ−1 = 0. Finally, M is maximal Cohen-Macaulay since depthA(M) = n− 1. There is also a natural triangulated structure on MCMgr(A) and we have proved part of the following theorem well known triangulated equivalence. 74 Theorem 5.1.1. The functor cok: HMFgr(f) ∼−→ MCMgr(A) is a triangulated equivalence between the category of graded Maximal Cohen- Macualay modules over A and the cagtegory of matrix factorizations of f . 5.2 Singularity Categories Let’s define singularity categories in full generality first. We then specialize to the affine cone over X = V (f). Let X be a scheme, not necessarily smooth, and let Perf(X) denote the full subcategory of D(X) consisting of perfect complexes from Example 2.1.2. Definition 5.2.1. The singularity category of X is the Drinfeld-Verdier localization DSg(X) = D(X)/Perf(X). It is immediate from these definitions that if X is quasi-projective and smooth, then the singularity category vanishes. If in addition X has an action of an algebraic group G, we can take G- equivariant objects in Definition 5.2.1 to define the G-equivariant (or G-graded) singularity category DSg,G(X) = D(X)G/Perf(X)G. We will be primarily interested in the following example. 75 Example 5.2.1. Let X = V (f) ⊂ P(V ) be a smooth degree d hypersurface. Then C(X) = V (f) ⊂ V , the affine cone over X, has an isolated singularity at the origin. There is a natural Gm-action by scaling the variables. We consider the graded singularity category denoted DSg,gr(C(X)) = DSg,Gm(C(X)). In the case where X = V (f ⊕ g), where f, g are degree d polynomials in different variables, then there is an additional µd-action on the g-variables (or the f -variables). We can then also consider the Gm × µd-graded singularity category DSg,gr,µd = DSg,Gm×µd(C(X)). Example 5.2.2. Objects in the singularity category DSg,gr(C(X)) can be thought of as objects with support at the origin. Indeed, the primary example of such an object is k(i), for i ∈ Z. Here k is the graded A-module given by A/(x1, . . . , xn). Since A is not smooth at the origin, k does not possess a finite free resolution and so these objects are nontrivial. Theorem 5.2.1. Let X = V (f) ⊂ P(V ). If C(X) has an isolated singularity at the origin, then there is an equivalence of categories HMFgr(f) ' DSg,gr(C(X)) 76 Proof. We only describe the functor. Let cok: HMFgr(f) → D(gr − A) denote the cokernel functor, which takes a graded matrix factorization (F·, δ·) to the cokernel of δ−1. Remark 5.2.1. By Theorem 5.1.1, we also have a triangulated equivalence MCMgr(A) ∼= DSg,gr(C(X)). Remark 5.2.2. There are natural extensions of Theorem 5.2.1 and Proposition 5.2.1 to the case of a Gm × µd-action (as in Example 5.2.1). Definition 5.2.2. In view of Theorem 5.2.1, we define the stabilization functor, stab: DSg,gr(C(X))→ HMFgr(f) to be the quasi-inverse to cok. We end with a proposition which tells us how to compute stabilizations for complete intersection subschemes. Proposition 5.2.1 ([21, Lemma 1.6.2]). Let K(s, t) be a Koszul matrix factorization cutting out a subscheme of Z of C(X), then stab(OZ) ∼= K(s, t). 5.3 Orlov’s Theorem For the rest of this chapter, we work with the hypersurface algebra A from Example 5.2.1. To get an understanding of Orlov’s functors we need to understand various semi-orthogonal decompositions that define them. Let gr − A denote the full subcategory of Gr − A generated by coherent modules. For each i ∈ Z, we have truncation functors tr≥i on Gr− A, defined by tr≥i(M)n =  Mn n ≥ i 0 n < i . 77 Define the full triangulated subcategory gr − A≥i to be the image of tr≥i on the subcategory gr − A. Define S −i. In other words, these are finite dimensional A- modules concentrated in degrees less than i. The kernel of tr≥i is clearly S −i. Similarly define P≥i. Let tors − A denote the full subcategory of gr − A generated by finite dimensional A-modules and grproj − A the full subcategory of graded projective A-modules. Lemma 5.3.1 ([20, Lemma 2.3]). The subcategories S 0, Φ is fully-faithful and there is a semi-orthogonal decomposition D(X) = 〈OX(−i− a− 1), . . . ,OX(−i),ΦiDSg,gr(C(X))〉; (ii) if a < 0, Ψ is fully-faithful and there is a semi-orthogonal decomposition DSg,gr(C(X)) = 〈kstab(−i), . . . , kstab(−i+ a+ 1),ΨiD(X)〉; (iii) if a = 0, the functors Ψi and Φi induce mutually inverse equivalences of categories D(X) ' DSg,gr(C(X)). Remark 5.3.1. In view of Theorem 5.2.1, we will can replace all instances of DSg,gr(C(X)) with cok(HMFgr(f)). Remark 5.3.2. Theorem 5.3.1 holds equally as well in the case of the Gm×µd action of Example 5.2.1. We will need an understanding of both the functors Φi and Ψi. Let’s start with Φi. This functor is defined in the proof of the theorem as the composite DSg,gr(A) ∼−→ Ti ↪→ D(gr − A) sh−→ D(X). The most difficult part of this computation is the first step, where we need to compute the image of an object under the quasi-inverse to pii. Using Lemmas 5.3.1 and 5.3.2, we have a semi-orthogonal decomposition for each i ∈ Z: D(gr − A) = 〈S 0 will be iterated extensions of shifts of k. 5.5 Comparison of Functors In this section, we compare our functors to Orlov’s on points. Let Xf and Xg define Calabi-Yau hypersurface in P(V ), where dim(V ) = n + 1. Let Af and Ag denote the corresponding graded hypersurface algebras. Let R = SpecSym(V ∨ ⊕ V ∨) and A = R/f ⊕ g. Then A is Z × µd-graded. If we set x0, . . . , xn to be the f -variables and y0, . . . , yn to be the y-variables, then the µn+1-weight of the coordinate functions xi is 0, while for yi it is -1. Denote Orlov’s functors for Xf by Ψ f i and for Xg by Ψ g j . Define Ψi,j = Ψ f i ⊗Ψgj : D(Xf ×Xg)→ HMfgr,µd(f ⊕ g) to be the tensor product of Ψfi and Ψ g j (in a dg enhancement) using the equivalences: D(Xf )⊗D(Xg) ∼= D(Xf ⊗Xg) via (F ·,G ·) 7→ pi∗XfF · ⊗ pi∗XgG · and HMFgr(f)⊗ HMFgr(g) ∼= HMFgr,µd(f ⊕ g) described in [2, Proposition 2.39]. 85 Proposition 5.5.1. Let p ∈ Xf and q ∈ Xg. Let Sp,q denote the Z × µd-graded A-module corresponding to the structure sheaf of the plane containing lp and lq. Then Ψi,j(Op,q) ∼= Sstabp,q (−i− j)(χj). Proof. Using Lemma 5.4.1, we know Ψfi (Op) ∼= lp(−i) and Ψgj (Op) ∼= lq(−j). Both of these have Koszul resolutions cutting them out and so the stabilization is given by Koszul matrix factorizations. Thus lp(−i)stab  lq(−j)stab is also given by a Koszul resolution, where we take all of the sections cutting out lp and lq. The zero locus of this Koszul resolution is precisely the plane containing lp and lq in V × V . The last step is to check the grading shift and the weights involved. The Z- grading is evidently −i − j. However, the µd-grading is not. Indeed, twisting by (−j) in the y-variables corresponds to a µn+1-twist by j. This completes the proof. Let us recall our decomposition for this case D[X/µn+1] = 〈D(Xf ×Xg),A〉 where A is the subcategory A = 〈OX(−2n)(χ−n),OX(−2n− 1)(χ−n,−n+1), . . . ,OX(−1)(χ0,−1),OX〉. Using the inverse Serre functor (−)⊗OX(n+ 1), we have D[X/µd] = 〈A,D(Xf ×Xg)(n+ 1)〉. 86 Let A≥−n be the subcategory of A given by B = 〈O(−n)(χ0,...,−n), . . . ,OX(−1)(χ0,−1),OX〉 and A<−n be the subcategory generated by the remaining exceptional objects so that A = 〈A<−n,A≥−n〉. We now match Orlov’s decomposition (at least in spirit): D[X/µn+1] = 〈A<−n,A≥−n,D〉 = 〈A≥−n,D,A<−n(n+ 1)〉 use Serre functor = 〈A≥−n,A<−n(n+ 1), RA<−n(n+1)D〉 mutate to the right By definition of A, we have 〈A≥−n,A<−n(n+ 1)〉 = 〈OX(−n)(χall), . . . ,OX(χall)〉, where χall means take all weights. We are left to compare Orlov’s composite functor Φ0 ◦ cok ◦Ψ0,0 : D(Xf ×Xg)→ D[X/µd] to the composite given by the Main Theorem followed by a right mutation: RA<−n(n+1) ◦ Ξ0,0 : D(Xf ×Xg)→ D[X/µd]. 87 To being comparing, we will need the following lemma. Lemma 5.5.1 ([9, Lemma 3.16]). Let M be a finitely generated graded A-module, then there is an isomorphism of graded A-modules Ext∗gr−A(M,A)) ∼= Ext∗+1gr−R(M,R(−d))). Proof. The proof given in [9, Lemma 3.16] extends to the category of graded A- modules. Lemma 5.5.2. We have vanishing Ext∗A(Sp,q, A(e)(χ i)) = 0 for −n ≥ e ≥ 0 and e ≤ i ≤ 0. Proof. Using Lemma 5.5.1, we have Ext∗A(Sp,q, A(e)(χ i)) = Sp,q(n− 1)(χi)[1− 2n]. The statement follows by taking degree zero pieces. Lemma 5.5.3. There is an isomorphism Ext∗A(Sp,q, A(e)(χ i)) ∼= Ext∗X(Ol(p,q),OX(e)(χi)) for −n ≤ e ≤ 0 and −n− 1 ≤ i < e. 88 Proof. In the proof of Lemma 5.5.2, we saw Ext∗A(Sp,q, A(e)(χ i)) = (Sp,q(n− 1)(χi))µn+10 [1− 2n] To finish the claim, we compute the other extension group. We have Ext∗X(Ol(p,q),OX(e)(χi)) ∼= Ext2n−∗X (OX(e)(χi),Ol(p,q)(−n− 1)) ∼= H2n−∗(Ol(p,q)(−e− n− 1)(χ−i)) ∼= H∗+1−2n(Ol(p,q)(n+ e− 1)(χi−1)) ∼= (Sp,q(n+ e− 1)(χi−1))µn+10 [1− 2n] Theorem 5.5.1. For a point (p, q) ∈ Xf ×Xg, we have (Φ0 ◦ cok ◦Ψ)(Op,q) = (RA<−n(n+1) ◦ Ξ0,0)(Op,q). Proof. Follows from Lemmas 5.5.2 and 5.5.3 and the discussion on how to compute Φ0 in Section 5.3. It follows from Theorem 5.5.1, that Orlov’s functors and ours (after a mutation) are the same up to a twist by a line bundle. That is, there exists a line bundle L on Xf ×Xg such that (Φ0 ◦ cok ◦Ψ) ◦ (L ⊗ (?)) ∼= (RA<−n(n+1) ◦ Ξ0,0) 89 Conjecture 5.5.1. Orlov’s functors and our functors agree, after an appropriate choice, and up to a mutation. That is, L ∼= O. It would also be interesting to do these computations in the case d,m, n arbitrary (but d ≤ m + n). An argument similar to the one carried out should still hold for points. For now we just leave it as a conjecture. Conjecture 5.5.2. There exists appropriate choices of i, j such that Conjecture 5.5.1 extends to arbitrary Calabi-Yau or general type hypersurfaces provided d ≤ n+m. 90 REFERENCES CITED [1] Matthew Ballard, David Favero, and Ludmil Katzarkov. A category of kernels for equivariant factorizations and its implications for Hodge theory. Publ. Math. Inst. Hautes E´tudes Sci., 120:1–111, 2014. [2] Matthew Ballard, David Favero, and Ludmil Katzarkov. A category of kernels for equivariant factorizations, II: further implications. J. Math. Pures Appl. (9), 102(4):702–757, 2014. [3] A. A. Be˘ılinson. Coherent sheaves on Pn and problems in linear algebra. Funktsional. Anal. i Prilozhen., 12(3):68–69, 1978. [4] A. Bondal and D. Orlov. Derived categories of coherent sheaves. In Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002), pages 47–56. Higher Ed. Press, Beijing, 2002. [5] A. I. Bondal and M. M. Kapranov. Representable functors, Serre functors, and reconstructions. Izv. Akad. Nauk SSSR Ser. Mat., 53(6):1183–1205, 1337, 1989. [6] Tom Bridgeland. Equivalences of triangulated categories and Fourier-Mukai transforms. Bull. London Math. Soc., 31(1):25–34, 1999. [7] Tom Bridgeland. Derived categories of coherent sheaves. In International Congress of Mathematicians. Vol. II, pages 563–582. Eur. Math. Soc., Zu¨rich, 2006. [8] Tom Bridgeland, Alastair King, and Miles Reid. The McKay correspondence as an equivalence of derived categories. J. Amer. Math. Soc., 14(3):535–554 (electronic), 2001. [9] W. Bruns and H.J. Herzog. Cohen-Macaulay Rings. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1998. [10] Andrei Ca˘lda˘raru and Junwu Tu. Curved A∞ algebras and Landau-Ginzburg models. New York J. Math., 19:305–342, 2013. [11] A. Elagin. On equivariant triangulated categories. http://adsabs.harvard.edu/abs/2014arXiv1403.7027E, March 2014. [12] Lennart Galinat. Orlov’s equivalence and maximal Cohen-Macaulay modules over the cone of an elliptic curve. Math. Nachr., 287(13):1438–1455, 2014. 91 [13] Sergei I. Gelfand and Yuri I. Manin. Methods of homological algebra. Springer Monographs in Mathematics. Springer-Verlag, Berlin, second edition, 2003. [14] D. Huybrechts. Fourier-Mukai transforms in algebraic geometry. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, Oxford, 2006. [15] A. Krug, D. Ploog, and P. Sosna. Derived categories of resolutions of cyclic quotient singularities. ArXiv e-prints, January 2017. [16] A. Kuznetsov and A. Perry. Derived categories of cyclic covers and their branch divisors. http://adsabs.harvard.edu/abs/2014arXiv1411.1799K, November 2014. [17] Alexander Kuznetsov. Derived categories of cubic fourfolds. In Cohomological and geometric approaches to rationality problems, volume 282 of Progr. Math., pages 219–243. Birkha¨user Boston, Inc., Boston, MA, 2010. [18] Shinnosuke Okawa. Semi-orthogonal decomposability of the derived category of a curve. Adv. Math., 228(5):2869–2873, 2011. [19] Christian Okonek, Michael Schneider, and Heinz Spindler. Vector bundles on complex projective spaces. Modern Birkha¨user Classics. Birkha¨user/Springer Basel AG, Basel, 2011. Corrected reprint of the 1988 edition, With an appendix by S. I. Gelfand. [20] Dmitri Orlov. Derived categories of coherent sheaves and triangulated categories of singularities. In Algebra, arithmetic, and geometry: in honor of Yu. I. Manin. Vol. II, volume 270 of Progr. Math., pages 503–531. Birkha¨user Boston, Inc., Boston, MA, 2009. [21] Alexander Polishchuk and Arkady Vaintrob. Matrix factorizations and cohomological field theories. J. Reine Angew. Math., 714:1–122, 2016. [22] Piotr Pragacz, Vasudevan Srinivas, and Vishwambhar Pati. Diagonal subschemes and vector bundles. Pure Appl. Math. Q., 4(4, Special Issue: In honor of Jean-Pierre Serre. Part 1):1233–1278, 2008. [23] Sophie Te´rouanne. Sur la cate´gorie Db,G(X) pour G re´ductif fini. C. R. Math. Acad. Sci. Paris, 336(6):483–486, 2003. [24] Bertrand Toe¨n. The homotopy theory of dg-categories and derived Morita theory. Invent. Math., 167(3):615–667, 2007. [25] Angelo Vistoli. Grothendieck topologies, fibered categories and descent theory. In Fundamental algebraic geometry, volume 123 of Math. Surveys Monogr., pages 1–104. Amer. Math. Soc., Providence, RI, 2005. 92