Homological Algebra
Homological Algebra
Yuri Berest1
Fall 2013 Spring 2014
ii
Contents
1 Standard complexes in Geometry
1
Complexes and cohomology . . . . . . . . . . . . . . . .
2
Simplicial sets and simplicial homology . . . . . . . . . .
2.1
Motivation . . . . . . . . . . . . . . . . . . . . .
2.2
Definitions . . . . . . . . . . . . . . . . . . . . .
2.3
Geometric realization . . . . . . . . . . . . . . .
2.4
Homology and cohomology of simplicial sets . . .
2.5
Applications . . . . . . . . . . . . . . . . . . . .
2.6
Homology and cohomology with local coefficients
3
Sheaves and their cohomology . . . . . . . . . . . . . . .
3.1
Presheaves . . . . . . . . . . . . . . . . . . . . .
3.2
Definitions . . . . . . . . . . . . . . . . . . . . .
3.3
Kernels, images and cokernels . . . . . . . . . . .
3.4
Germs, stalks, and fibers . . . . . . . . . . . . . .
3.5
Coherent and quasi-coherent sheaves . . . . . . .
3.6
Motivation for sheaf cohomology . . . . . . . . .
3.7
Sheaf cohomology . . . . . . . . . . . . . . . . .
3.8
Applications . . . . . . . . . . . . . . . . . . . .
3.9
Cech
cohomology . . . . . . . . . . . . . . . . . .
2 Standard complexes in algebra
1
Group cohomology . . . . . . . . . . . . . . .
1.1
Definitions and topological origin . . .
1.2
Interpretation of H1 (G, A) . . . . . . .
1.3
Interpretation of H2 (G, A) . . . . . . .
2
Hochschild (co)homology . . . . . . . . . . .
2.1
The Bar complex . . . . . . . . . . . .
2.2
Differential graded algebras . . . . . .
2.3
Why DG algebras? . . . . . . . . . . .
2.4
Interpretation of bar complex in terms
2.5
Hochschild (co)homology: definitions .
2.6
Centers and Derivations . . . . . . . .
iii
1
. . . . . . . . . . . . 1
. . . . . . . . . . .
3
. . . . . . . . . . .
3
. . . . . . . . . . .
5
. . . . . . . . . . .
6
. . . . . . . . . . .
8
. . . . . . . . . . .
9
. . . . . . . . . . .
10
. . . . . . . . . . .
12
. . . . . . . . . . .
12
. . . . . . . . . . .
12
. . . . . . . . . . .
13
. . . . . . . . . . .
14
. . . . . . . . . . .
16
. . . . . . . . . . .
17
. . . . . . . . . . .
19
. . . . . . . . . . . . 21
. . . . . . . . . . .
23
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
of DG algebras
. . . . . . . . .
. . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
25
25
25
26
27
29
29
30
33
34
35
37
2.7
Extensions of algebras . . . . . . . . . .
2.8
Crossed bimodules . . . . . . . . . . . .
2.9
The characteristic class of a DG algebra
Deformation theory . . . . . . . . . . . . . . . .
3.1
Motivation . . . . . . . . . . . . . . . .
3.2
Formal deformations . . . . . . . . . . .
3.3
Deformation theory in general . . . . . .
3.4
The Gerstenhaber bracket . . . . . . . .
3.5
Stasheff construction . . . . . . . . . . .
3.6
Kontsevich Formality Theorem . . . . .
3.7
Deformation theory in algebraic number
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
theory
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
37
.
38
.
40
. . 41
. . 41
.
42
.
44
.
49
.
50
.
52
.
53
3 Category theory
1
Basic category theory . . . . . . . . . . . . . . . . . .
1.1
Definition of categories . . . . . . . . . . . . . .
1.2
Functors and natural transformations . . . . .
1.3
Equivalences of categories . . . . . . . . . . . .
1.4
Representable functors and the Yoneda lemma
1.5
Adjoint functors . . . . . . . . . . . . . . . . .
1.6
Limits and colimits . . . . . . . . . . . . . . . .
2
Special topics in category theory . . . . . . . . . . . .
2.1
Brief introduction to additive categories . . . .
2.2
Center of a category and Bernstein trace . . . .
2.3
Morita theory . . . . . . . . . . . . . . . . . . .
2.4
Recollement (gluing) of abelian sheaves . . . .
2.5
Kan extensions . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
57
. . . . . . . . .
57
. . . . . . . . .
57
. . . . . . . . .
59
. . . . . . . . . . 61
. . . . . . . . .
62
. . . . . . . . .
66
. . . . . . . . .
68
. . . . . . . . . . 71
. . . . . . . . . . 71
. . . . . . . . .
72
. . . . . . . . .
76
. . . . . . . . .
78
. . . . . . . . .
79
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
iv
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
83
.
83
.
83
.
85
.
87
. . 91
.
95
.
96
.
99
.
99
. . 101
. 102
. 104
. 106
. 106
. 108
3.3
3.4
3.5
3.6
3.7
5 Derived categories
1
Localization of categories . . . . . .
1.1
Motivation . . . . . . . . . .
1.2
Definition of localization . . .
1.3
Calculus of fractions . . . . .
1.4
Ore localization . . . . . . . .
2
Localization of abelian categories . .
2.1
Serre quotients . . . . . . . .
2.2
Injective envelopes . . . . . .
2.3
Localizing subcategories . . .
3
Derived categories . . . . . . . . . .
3.1
Definition and basic examples
3.2
The homotopy category . . .
3.3
Verdier theorem . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
. . 111
. 113
. 115
. 117
. . 121
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
125
. 125
. 125
. 126
. 128
. 130
. 134
. 134
. 137
. 138
. 139
. 139
. . 141
. 143
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
145
145
145
146
148
152
152
153
153
154
155
156
160
160
. 161
162
163
164
165
165
167
168
6 Triangulated categories
1
The basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1
Definitions . . . . . . . . . . . . . . . . . . . . . . . .
1.2
Examples of triangulated categories . . . . . . . . . .
1.3
Basic properties of triangulated categories . . . . . . .
2
Further properties . . . . . . . . . . . . . . . . . . . . . . . .
2.1
Abstract cone and octahedron axiom . . . . . . . . . .
2.2
The homotopy category is triangulated (need proofs!!)
2.3
Localization of triangulated categories . . . . . . . . .
2.4
Exact functors . . . . . . . . . . . . . . . . . . . . . .
2.5
Verdier quotients . . . . . . . . . . . . . . . . . . . . .
2.6
Exact categories . . . . . . . . . . . . . . . . . . . . .
3
t-structures and the recollement . . . . . . . . . . . . . . . . .
3.1
Motivation and definition . . . . . . . . . . . . . . . .
3.2
The core of a t-structure and truncation functors . . .
3.3
Cohomological functors . . . . . . . . . . . . . . . . .
3.4
t-exact functors . . . . . . . . . . . . . . . . . . . . . .
3.5
Derived category of t-structure . . . . . . . . . . . . .
3.6
Gluing t-structures . . . . . . . . . . . . . . . . . . . .
3.7
Examples of gluing . . . . . . . . . . . . . . . . . . . .
3.8
Recollement for triangulated categories . . . . . . . .
3.9
Example: topological recollement . . . . . . . . . . . .
v
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3.10
168
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
171
. 171
. 171
172
172
174
175
176
177
177
177
183
184
186
189
. 191
195
195
197
198
199
199
199
199
. 201
203
204
204
205
206
208
210
212
212
214
A Miscellaneous topics
1
Characteristic classes of representations (after
2
Generalized manifolds . . . . . . . . . . . . .
3
Quivers and path algebras . . . . . . . . . . .
3.1
Basic definitions . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
217
217
220
222
222
vi
Quillen)
. . . . .
. . . . .
. . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3.2
3.3
3.4
B Exercises
1
Standard complexes in Algebra and Geometry . . . . . . .
2
Classical homological algebra . . . . . . . . . . . . . . . .
3
Residues and Lie cohomology . . . . . . . . . . . . . . . .
3.1
Commensurable subspaces . . . . . . . . . . . . . .
3.2
Traces . . . . . . . . . . . . . . . . . . . . . . . . .
3.3
Residues . . . . . . . . . . . . . . . . . . . . . . . .
3.4
Interpretation in terms of Lie algebra cohomology
3.5
Adeles and residues on algebraic curves . . . . . .
vii
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
223
224
225
233
233
233
235
235
236
237
238
238
viii
Chapter 1
Definition 1.0.1. A chain complex C is a sequence of abelian groups together with group
homomorphisms
C :
dn+1
n
Cn+1 Cn
Cn1
dn
dn+1
C n C n+1 C n+2
dn
C
/ C n+1
fn
Dn
dn
D
f n+1
/ D n+1
/ C n1
g n1
f n1
}
/ D n1
n1
dC
hn
n1
dD
/ Cn
gn
fn
}
/ Dn
dn
C
hn+1
/ C n+1
g n+1
dn
D
f n+1
/ D n+1
2
2.1
There are a number of complexes that appear quite algebraic, but whose construction
involves topology.
Definition 2.1.1. The geometric n-dimensional simplex is the topological space
(
)
n
X
n+1
n = (x0 , . . . , xn ) R
:
xi = 1, xi > 0
i=0
dn : Cn Cn1 , i 7
n
X
k=0
(1)k k ik
In the above definition of dn each simplex i is equipped with an orientation (i.e. choice
of an ordering of its vertices). Then ik denotes the simplex {e0 , . . . , ek , . . . , en }, and k = +1
or 1 depending on the sign of the permutation that maps the sequence {e0 , . . . , ek , . . . , en }
to the sequence of vertices of ik determined by its orientation.
As an exercise, show that dn dn1 = 0.
L
Theorem 2.1.6. The homology groups H (X) =
Hn (X) of the complex C = (Cn , dn )
are independent of the choice of triangulation and orientation of simplices.
Proof. See any book on algebraic topology.
It follows that the Hn (X) are invariants of X as a topological space.
Geometric intuition A homology cycle c Hn (X) can be viewed as n-dimensional chains
(n-cycles) modulo the equivalence relation c c0 if there exists an (n + 1)-cycle of which c
and c0 are the boundary.
Definition 2.1.7.
1. A simplicial set is a family of sets X = {Xn }n>0 and a family of
maps {X(f ) : Xn Xm }, one for each non-decreasing function f : [m] [n], where
[n] = {0, . . . , n}, satisfying
X(id) = id
X(f g) = X(g) X(f )
2. A map of simplicial sets : X Y is a family of maps {n : Xn Yn }n>0 such that
for all f : [m] [n]:
Xn
X(f )
Xm
/ Yn
Y (f )
/ Ym
commutes.
Remark 2.1.8. A simplicial set is just a contravariant functor from the simplicial category
to the category of sets Set, and a map of simplicial sets is just a natural transformation of
functors. The simplicial category has finite sets [n] as objects, and non-decreasing functions
as morphisms. We denote the category ofsimplicial sets by Set.
Definition 2.1.9. Let X be a simplicial set. The geometric realization of X is
,
a
|X| =
(n Xn )
n=0
where the equivalence relation is defined by (s, x) (t, y) if, for (s, x) n Xn and
(t, y) m Xm , there exists f : [m] [n] non-decreasing
such that y = X(f )x and
`
t = f s. We give |X| the weakest topology such that (n Xn ) |X| is continuous.
4
2.2
Definitions
Recall that we defined simplicial set as a family X = {Xn }n>0 of sets and a family of maps
X(f ) : Xn Xm , one for each non-decreasing map f : [m] [n], such that X(id) = id and
X(f g) = X(g)X(f ) when the compositions are defined. This can be rephrased more
conceptually using the simplicial category.
Definition 2.2.1. The simplicial category has as objects all finite well ordered sets.
That is, Ob = {[n] = {0 < 1 < < n}}. Morphisms are order-preserving maps (i.e.
i 6 j f (i) 6 f (j)).
A simplicial set is just a contravariant functor X : Set. Thus the category of
simplicial sets is just the category Set = Set . There are two distinguished classes of
maps in :
in : [n] , [n + 1]
06i6n
ji :
06j 6n+1
[n + 1] [n]
called the face maps in and degeneracy maps jn . They are defined by
(
k
if k < i
in (k) =
k+1 if k > i
(
k
if k 6 j
jn (k) =
k 1 if k > j
Theorem 2.2.2. Any morphism f Hom ([n], [m]) can be decomposed in a unique way as
f = i1 i2 ir j1 js
such that m = n s + r and i1 6 6 ir and j1 6 6 js .
The proof of this theorem is a little technical, but a few examples make it clear what is
going on.
Example 2.2.3. Let f : [3] [1] be {0, 1 7 0; 2, 3 7 1}. One can easily check that
f = 11 22 .
Corollary 2.2.4. For any f Hom ([n], [m]), there is a unique factorization
[n]
/ [m]
O
?
[k]
Corollary 2.2.5. The category can be presented by {i } and {j } as generators with the
following relations:
j i = i j
i<j
j i = i j+1
i j1
j i = id
i1 j
i6j
(1.1)
if i < j
if i = j or i = j + 1
if i > j + 1
Corollary 2.2.6. Giving a simplicial set X = {Xn }n>0 is equivalent to giving a family of
sets {Xn } equipped with morphisms in : Xn Xn1 and sni : Xn Xn+1 satisfying
i j = j i
i<j
si sj = sj+1 si
sj1 i
j si = id
sj i1
i6j
(1.2)
if i < j
if i = j or i = j + 1
if i > j + 1
The relation between (1.1) and (1.2) is given by in = X(in1 ) and sni = X(in ).
P
Consider the n-dimensional geometric simplex n = {(x0 , . . . , xn ) Rn+1
xi = 1}.
>0 :
For aPnon-empty subset I [n], define the I-th face of n by eI = {(x0 , . . . , xn )
n :
iI xi = 1}. In particular, if I = {i}, then the I-th face of n is just the i-th vertex
ei = (0, . . . , 1, . . . , 0).
It is more convenient to parametrize faces by maps f : [m] [n] for m 6 n with
Im(f ) = I.
Example 2.2.7. Let I = {0, 1, 3} [3]. The corresponding map f : [2] [3] is just
{0 7 0, 1 7 1, 2 7 3} = 23 .
In general, given f : [m] [n], the corresponding f : m n is defined to be the
restriction of the linear map Rm+1 Rn+1 sending ei to ef (i) .
2.3
Geometric realization
Recall that given any simplicial set X = {Xn }, we defined the geometric realization of X as
|X| =
.
(n Xn )
n=0
n Xn
n=0
m X(m)
m=0
by (s, x
) 7 (g (s), x) m X(m) . Clearly if (s, x
) (s0 , x
0 ) then (s, x
) = (s0 , x
0 ).
Hence induces a continuous map : |X| X.The homotopically inverse map is induced
by the map
a
a
:
m X(m)
n Xn
m=0
n=0
2.4
Let X = {Xn } be a simplicial set. Recall that for each n Z and for each fixed abelian
group A, we defined
(
0
if n < 0
Cn (X, A) =
.
AXn = A ZXn otherwise
The differential dn : Cn Cn1 is defined by
X
a(x) x 7
xXn
a(x)
xXn
a(x)
n
X
(1)i X(in1 )x
i=0
n
X
(1)i in x
i=0
xXn
n
P
(1)i i .
i=0
(dn f )(x) =
i=0
n+1
X
(1)i f (in+1 x)
i=0
Theorem 2.4.1. The objects Cn (X, A) and C n (X, A) are actually complexes, i.e. dn1 dn =
0 and dn+1 dn = 0.
"
#
n
P
j
(1) j =
Proof. Lets check that dn1 dn = 0. We have dn1 dn = dn1
j=0
n n1
P
P
(1)i+j i j . Then we can split this sum into two parts and use the relations (1.2)
j=0 i=0
(1)i+j i j +
i<j
i+j
(1)
j1 i +
ij1
(1)i+j i j
ij
X
X
X
(1)i+j i j
ij
i0 +j 0 +1
(1)
i0 j 0
i0 j 0 +
X
ij
=0
8
(1)i+j i j
Remark 2.4.2. For any category C we can define simplicial objects in C as functors op C.
If category C is abelian (for example abelian groups Ab, or vector spaces Vect, or modules
Mod(R) over some algebra), then we can define homology of a simplicial object X op C as
the homology P
of complex X, where the differential dn : Xn Xn1 is defined by the same
formula dn = (1)i X(i ) as before.
Then the above definition of homology of a simplicial set coincides with homology of
free
the simplicial abelian group S : op Ab defined by the composition op Sets A Ab,
where freeA sends a set X to the abelian group A ZX.
Dually we can define cosimplicial objects in C as functors C. Again, if C is abelian,
then we can define cohomology of a cosimplicial object Y in CPas the cohomology of a
complex Y where the differential is defined by the formula dn = (1)i Y (i ).
2.5
Applications
Example 2.5.1 (Singular (co)homology). Let X be a topological space. A (singular) nsimplex of X is a continuous map : n X. Put, for n > 0, Xn = HomTop (n , X) =
{singular n-simplices in X}. For f Hom ([m], [n]), define X(f ) : Xm Xn by X(f )() =
f .
If A is an abelian group, then we can define singular homology and cohomology of X by
Hsing
(X, A) = H (X, A) and Hsing (X, A) = H (X, A) respectively.
Note that if X has some extra structure (e.g. is a C -manifold or a complex manifold)
then it is often convenient to take simplices compatible with that structure.
Example 2.5.2 (Nerve of a covering). Let X be a topological space, U = {U }I a
covering of X. Define
Xn = {(0 , . . . , n ) I n+1 : U0 Un 6= }.
For f Hom ([m], [n]), the morphism X(f ) : Xn Xm is given by (0 , . . . , n ) 7
gj if f (i 1) 6= f (i)
hi = f (i1)<j6f (i)
e
otherwise
G
We call |BG| the classifying space of G. Group (co)homology is defined as H (G, A) =
Hsing
(BG, A), H (G, A) = Hsing (BG, A).
9
Example 2.5.4 (Hochschild homology). Let A be an algebra over a field k and M be Abimodule. Then we can form a simplicial module C (A, M ) by setting Cn (A, M ) = M An
and defining face maps and degeneracy maps as follows:
d0 (m, a1 , . . . , an ) = (ma1 , a2 , . . . , an )
di (m, a1 , . . . , an ) = (m, a1 , . . . , ai ai+1 , . . . , an ), i = 1, . . . , n 1
dn (m, a1 , . . . , an ) = (an m, a1 , . . . , an1 )
sj (m, . . . , an ) = (m, a0 , . . . , aj , 1, aj+1 , . . . , an )
Homology HH (A, M ) := H(C (A, M )) of this simplicial module is called Hochschild
homology of A with coefficients in bimodule M . We will consider this cohomology theory in
more details in section 2.
What is a (co)homology theory? A (co)homology theory should be a function of
two arguments: H(X, A ), where X is a nonabelian argument, and A is an object in
some abelian category. For example, X could be a topological space, algebra, group etc.
and usually A will be a sheaf, (bi)module, representation etc. The modern perspective is
that we should fix X and think of H(X, ) as a functor from some abelian category to
abelian groups. More formally, we have some non-abelian (that is, arbitrary) category C,
and an additive category A over C fibred in abelian categories. For example, we can consider
C = Top, and the fiber of A over Top being Sh(X).
2.6
L
Recall that given a simplicial P
set X we defined Cn (X, A) = xXn Ax. That is, elements
a Cn (X, A) are of the form xX a(x) x with a(x) A. What if we allowed the a(x) to
live in different abelian groups? That is exactly what we will try to do!
Definition 2.6.1. A homological system of coefficients for X consists of
1. a family of abelian groups {Ax }xXn one for each simplex x Xn
2. a family of group homomorphisms {A (f, x) : Ax AX(f )x }xXn ,f : [m][n]
satisfying
1. A (id, x) = idAx for all x Xn
2. the following diagram commutes:
Ax
A (f g,x)
A (f,x)
/ AX(f )x
AX(f g)x
A (g,X(f )x)
AX(g)X(f )x
that is, A (f g, x) = A (g, X(f )x)A (f, x). (this is a cocycle condition).
10
Cn (X, A ) =
Ax x
xXn
!
X
dn
a(x) x
xXn i=0
xXn
functions f : Xn
Bx
xXn
3.1
Presheaves
if U 6 V
HomOpen(X) (U, V ) =
U , V if U V
Here and elsewhere, U V means that U is not necessarily a proper subset of V . We will
use notation U ( V if U is proper subset of V .
Definition 3.1.1. Let X be a topological space. A presheaf on X with values in a category
C is a contravariant functor F : Open(X) C.
Common categories are C = Set, Grp, Ring, . . . . Elements s F (U ) are called sections
of F over U , and F (X) is the set of global sections. One often writes (U, F ) instead of
F (U ), and thinks of (U, ) as a functor on F . From the definition, we see that for U V
we have maps VU : F (V ) F (U ); these are called the restriction maps from V to U . Since
F is a functor, these satisfy:
1. U
U = id for all open U .
V
W
2. if U V W , then W
U = U V .
F (U )
3.2
(V )
(U )
/ G (V )
V
U
/ G (U )
Definitions
Definition 3.2.1. A presheaf F is a sheaf if given any open U X, any open cover
S
U
/ F (U ) F (V )
// F (U V )
(1.3)
We will mostly deal with abelian sheaves, that is sheaves of abelian groups. The following
are all examples of sheaves.
c (U ) = Hom
Example 3.2.3. Let X be a topological space. Set OX
Top (U, C).
diff
c
If X is a differentiable manifold, we can define OX OX , the sheaf of differentiable
diff (U ) be the ring of C -functions U C.
functions by letting OX
an O diff to
If X is a complex analytic manifold, e.g. X = P1 (C), then we can define OX
X
be the sheaf of holomorphic functions.
If we go even further and stipulate that X is an algebraic variety over C, then we can
alg
define OX
to be the sheaf of regular functions.
All the above sheaves are often called structure sheaves. Indeed, smooth manifolds,
analytic manifolds. . . can be defined to be topological spaces along with a sheaf of rings
satisfying certain properties.
For abelian sheaves, (1.3) is exact in the usual sense:
0
/ F (U V )
/ F (U ) F (V )
/ F (U V )
3.3
The hope would be that K and I 0 are the category-theoretic kernel / image of .
Lemma 3.3.2. With the above notation, K is a sheaf, but I 0 is not a sheaf in general.
S
Proof. Given open U X, U = U an open cover, and s K (U ) such that
U
U
U U (s ) = U U (s ), since F is a sheaf, there exists a unique s F (U ) such that
U
U (s) = s for all . We need to show that s is actually in K . Since is a morphism, for
any we have
U
U
U ((U )s) = (U )(U (s)) = (U )(s ) = 0
since each s K (U ). Since G is also a sheaf, this force (U )(s) = 0, hence s K (U ).
To show that I 0 is not in general a sheaf, we give a counterexample. Let X = C = C\{0}
an O an given by f 7 f 0 = df . One can check
with the analytic topology. Consider : OX
X
dz
that for all x X, there exists a neighborhood Ux X with x Ux such that (Ux ) :
+
P
an (U ) O an (U ) is surjective. But, the equation df = g for g =
an ),
OX
an z n (X, OX
x
x
X
dz
n=
has a global solution if and only if a1 = 0, which implies that the function z1 is not in the
image Im ((X)) although z1 Im ((Ux )) for any x X. This violates the sheaf axiom.
The moral of this is that I 0 needs to be redefined in order to be a sheaf.
Definition 3.3.3. Given a morphism of sheaves : F G , define Im() by
Im()(U ) = {s G (U ) : x U, Ux U : U
Ux (s) Im (Ux )}
It is a good exercise to show that Im() actually is a sheaf, and is moreover the categorytheoretic image of . If one is more ambitious, it is not especially difficult to show that
AbSh(X) is an abelian category.
3.4
U 3x
that is, HomSh(X) (F + , G ) ' HomPSh(X) (F , G ). The map F is the one induced by idF + ,
when we set G = F + .
Definition 3.4.7. For sheaves F and G with an embedding G , F define the quotient
F /G as sheafification of the presheaf U 7 F (U )/G (U ).
3.5
In this section we will introduce some basic types of sheaves (finitely generated, coherent,etc.)
that we will widely use later.
Let (X, O) be a topological space with the structure sheaf O of continuous functions
on it.
Definition 3.5.1. A sheaf F on X of O-modules is called finitely generated (or of finite
type) if every point x X has an open neighbourhood U such that there is a surjective
morphism of restricted sheaves
O n |U F |U , n N
In other words, locally such sheaf is generated by finite number of sections. That is, for any
x X and small enough open U 3 x, for any V U the abelian group F (V ) is finitely
generated as a module over O(V ).
Example 3.5.2. The structure sheaf O itself is of finite type, as well as O n .
Example 3.5.3. If F is finitely generated, then any quotient F /G and any inverse image
1 (F ) will be finitely generated.
Proposition 3.5.4. Suppose F is finitely generated. Suppose for some point x X and
open U 3 x the images of sections s1 , . . . , sn F (U ) in Fx generate the stalk Fx . Then
there exists an open subset V U s.t. the images of s1 , . . . , sn F (U ) in Fy generate Fy
for all y V .
Proof. Since F is finitely generated, there is some V 0 U and t1 , . . . , tm F (V 0 ) such
that t1 , . . . , tm generate Fy for anyP
y V 0 . Since Fx is also generated by s1 , . . . , sn , we
can express ti in terms of sj : ti =
aij sj , where aij Ox . There are finitely many aij ,
and since they are germs, there is a small open neighbourhood V 00 s.t. aij are actually
restrictions of some a
ij F (V 00 ). If we now take V = V 0 V 00 , sections si generate Fy for
any y V .
Corollary 3.5.5. If F is of finite type and Fx = 0 for some x X, then F |V = 0 for
some small open neighbourhood V of x.
Definition 3.5.6. A sheaf F is called quasi-coherent if it is locally presentable, i.e. for
every x X there is an open U X containing x s.t. there exist an exact sequence
O I |U O J |U F |U 0,
where I and J may be infinite, i.e. if F is locally the cokernel of free modules. If both I
and J can be chosen to be finite then F is called finitely presented.
16
Definition 3.5.7. A sheaf F is called coherent if it is finitely generated and for every open
U X and every finite n N, every morphism O n |U F |U of O|U -modules has a finitely
generated kernel.
Example 3.5.8. If X is Noetherian topological space, i.e. such that any chain V1 V2 . . .
of closed subspaces stabilizes, then the structure sheaf OX is coherent.
Example 3.5.9. The sheaf of complex analytic functions on a complex manifold is coherent.
This is a hard theorem due to Oka, see [Oka50].
Example 3.5.10. The sheaf of sections of a vector bundle on a scheme or a complex analytic
space is coherent.
Example 3.5.11. If Z is a closed subscheme of a scheme X, the sheaf IZ of all regular
functions vanishing on Z is coherent.
Lemma 3.5.12. If X is Noetherian then F is of finite type if and only if F is finitely
presented, if and only if F is coherent.
Lemma 3.5.13. For coherent and quasi-coherent sheaves the two out of three property
holds. Namely, if there is a short exact sequence
0
/F
/G
/H
/0
and two out of three sheaves F , G , H are coherent (resp. quasi-coherent), then the third
one is also coherent (resp. quasi-coherent).
Theorem 3.5.14. On affine variety (or better affine scheme) X with affine algebra of
functions A = OX (X) the global section functor gives equivalence of categories Qcoh(X)
Mod(A). Moreover, restriction of to coh(X) Qcoh(X) gives equivalence of categories
coh(X) fgMod(A), where fgMod(A) Mod(A) is a full subcategory of finitely generated
A-modules.
Proof. The inverse functor is given by tilde-construction. For details see, for example,
[EH00].
3.6
/ F (U )
F (Ui )
F (Ui Uj )
i,j
Uj
Ui
where the first map is s 7 (U
.
Ui (s))i and the second is (si )i 7 Ui Uj (s) Ui Uj (t)
i,j
17
Warning. Some textbooks only require that this sequence be exact for finite open covers.
This does not yield the same notion of a sheaf. For example, let X = Cn and F be the
sheaf of bounded continuous C-valued functions. Then F satisfies the sheaf axiom for all
finite covers, but it is easily seen that F is not a sheaf.
In the previous section, we defined kernels and images of sheaves. This enables us to
define an exact sequence of sheaves. In particular, we can consider short exact sequences
/K
/F
/G
/0
(1.4)
The global section functor (X, ) : Sh(X) Ab is left exact. That is, if we apply
(X, ) to the sequence (1.4), then
0
/ K (X)
/ F (X)
/ G (X)
P
where [xi ] is a formal basis element. Define : F G by (U )f = xU f (xi ) [xi ]. Set
K = Ker(). Taking global sections, we obtain (G ) = Ck , and, by Louivilles theorem,
an ) = C. If k > 1, then it is certainly not possible for (F ) (G ) to be
(F ) = (X, OX
surjective, even though it is trivial to check that is surjective (at the level of sheaves).
Given an exact sequence 0 K F G 0, we can apply the global sections
functor to obtain an exact sequence 0 (X, K ) (X, F ) (X, G ). The main idea
of classical homological algebra is to canonically construct groups Hi (X, F ) that extend the
exact sequence on the right:
0 (K ) (F ) (G ) H1 (K ) H1 (F ) H1 (G ) H2 (K )
Here, and elsewhere, we will write (F ) and Hi (F ) for (X, F ) and Hi (X, F ) when X is
clear from the context.
18
3.7
Sheaf cohomology
In this section we will define sheaf cohomology using the classical Godement resolution, and
compare it to Cech
cohomology.
Definition 3.7.1. A sheaf F is called flabby if for all U X, the restriction map
X
U : F (X) F (U ) is surjective.
Note that if F is flabby, then the sequence
0
/ F (U U 0 )
/ F (U ) F (U 0 )
/ F (U U 0 )
/0
is exact on the right. Indeed, given any r F (U U 0 ), we can choose r F (X) with
U U 0 (
r) = r. Then (X
r), 0) maps to r.
U (
Lemma 3.7.2. Let 0 K F G 0 be an exact sequence of sheaves, and assume
K is flabby. Then 0 K (X) F (X) G (X) 0 is also exact.
Proof. Recall that the surjectivity of : F G implies that (X) : F (X) G (X)
is locally surjective. That is, for all t G (X) and for all x X, there exists an open
neighborhood U of x and s F (U ) such that X
U (t) = (U )(s). Assume now that a
given t G (X) lifts to local sections s F (U ) and s0 F (U 0 ). Put (U )s = X
and
U (t)
0 agree over U U 0 , i.e. U
U0
0 ),
(U 0 )s0 = X
(t).
If
it
happens
that
s
and
s
(s)
=
(s
U0
U U 0
U U 0
then we can glue s and s0 along U U 0 . Unfortunately s and s0 do not always agree over
U0
0
U U 0 . However, if we let r = U
U U 0 (s) U U 0 (s ), then
0
U
0
(U U 0 )(r) = U
U U 0 ((U )(s)) U U 0 ((U )(s))
0
X
U
X
= U
U U 0 U (t) U U 0 U 0 (t)
=0
In other words, r K (U U 0 ). Since K is flabby, there exists r K (X) such that
X
r) = r. Using r, we can correct s0 by replacing it with s00 = s0 + X
r) F (U 0 ).
U (
U U 0 (
Now
0
00
U
0
U
X
U
r)
U U 0 (s ) = U U 0 (s ) + U U 0 U 0 (
0
0
X
= U
r)
U U 0 (s ) + U U 0 (
= U
U U 0 (s)
0
U = s and U U (
thus there exists s F (U U 0 ) such that U
s) = s00 . By (transfinite)
U
U0
induction, the result follows.
Lemma 3.7.3. There are enough flabby sheaves. More precisely, for any abelian sheaf
F , there is a (functorial) embedding : F C 0 (F ), where C 0 (F ) is flabby.
Q
Proof. Define C 0 (F )(U ) = xU Fx and (U )(s) = (sx )xU , where sx = U
x (s) is the germ
0
of s at x. It is not difficult to check that C (F ) is actually a flabby sheaf.
19
d1
C (F ) : 0 0 C 0 (F ) C 1 (F )
together with the morphism of complexes : F C (F ), where we regard F as the complex
0 0 F 0 , is called the Godement resolution of F .
Note that by construction, Hn (C (F )) = 0 for all n > 0, and H0 (C (F )) ' F .
Equivalently, we can say that is a quasi-isomorphism. Applying (X, ) termwise to
C (F ), we get a new complex (X, C (F )) which may not be acyclic.
Definition 3.7.5. The cohomology of X with coefficients in F is Hn (X, F ) = Hn (C (F )(X)).
It follows immediately from the definition that Hn (X, F ) = 0 if n < 0, and that
H0 (X, F ) ' (X, F ) = F (X) canonically.
Theorem 3.7.6. Given any short exact sequence of sheaves
0
/K
/F
/G
/0
(1.5)
/ Hn (X, F )
/ Hn (X, G )
/ Hn+1 (X, K )
/ Hn (K )
/ Hn (F )
/ Hn (G )
/ Hn+1 (K )
where all the maps but are the obvious induced ones, and is canonically constructed.
20
3.8
Applications
is the set of homogeneous forms of degree k that vanish on C1 C2 . There is the following
exact sequence, which is actually a locally free resolution of I (k).
0
/ O 2 (k m n)
P
/ O 2 (k m) O 2 (k n)
P
P
/ I (k)
/0
Here, (c) = (F2 c, F1 c) and (a, b) = F1 a + F2 b. The theorem we are trying to prove
simply asserts the surjectivity of on global sections. Taking global sections, we get
0
/ (O(k m n))
/ (I (k))
/ H1 (O(k m n))
We will see that H1 (Pn , OPn (k)) = 0 for all n > 1 and all k Z. This yields the result.
Example 3.8.2 (Exponential Sequence). Let X be a complex analytic manifold, for example
/Z
/ OX exp / O
X
/1
where Z is the constant sheaf and exp(s) = e2is . The obstructions to lifting sections from
OX
to OX lie in the first cohomology group H1 (X, Z), which does not vanish in general.
In fact, in cohomology we get an exact sequence
...
/ H1 (X, Z)
/ H1 (X, OX )
/ H1 (X, O )
X
/ H2 (X, Z)
/ H2 (X, OX )
/ ...
(1.6)
If X is complete (i.e. compact in this case), then the global sections H0 (X, OX
) is C , and
3.9
Cech
cohomology
(0 , . . . , n ) 7 (f (0) , . . . , f (m) )
U
f (0)
f (m)
U
n
0
/ F (U U )
n
0
H(U,
F ) := H (X , B)
Definition 3.9.2. A covering U is called F -acyclic if Hi (U0 Un , F ) = 0 for all
0 , . . . , n I and i > 0.
The following result allows one to establish acyclicity of some coverings.
Theorem 3.9.3 (H.Cartans criterion). Let A be a class of open subsets of a topological
space X such that
(a) A is closed under finite intersections, i.e.
U1 , . . . , Un A U1 Un A
(b) A contains arbitrary small open subsets, i.e. for any open U there is V ( U such that
V A.
Suppose that for any U A and A-covering U = {Ui } of U , Hi (U, F ) for all i > 0.
Then any A-covering is F -acyclic. In particular, for any A-covering of the space X there is
isomorphism
H(U,
F ) ' H (X, F ).
23
24
Chapter 2
Group cohomology
1.1
Recall that given a (discrete) group G, we define (BG)n = Gn , the n-fold cartesian product of
G with itself. For f : [m] [n], we define BG(f ) : Gn Gm by (g1 , . . . , gn ) 7 (h1 , . . . , hm ),
where
(Q
f (i1)<j6f (i) gj if f (i 1) 6= f (i)
hi =
1
otherwise
This is a simplicial set. Next, given a representation of G in an abelian group A (i.e. a
homomorphism G Aut(A)), define a cohomological system of coefficients:
Bx = A
for all x BG
B(f, x)(a) = ha
where for x = (g1 , . . . , gn ) Gn , we set h =
Qf (0)
j=1
Definition 1.1.1. With the above notation, we define C (G, A) = C (BG, B), and define
the cohomology of G with coefficients in A to be H (G, A) = H [C(G, A)].
dn
Explicitly, C (G, A) has C 0 (G, A) = A, and for n > 1, C n (G, A) = HomSet (Gn , A), with
: C n C n+1 defined by
n
X
(df )(g1 , . . . , gn+1 ) = g1 f (g2 , . . . , gn+1 )+ (1)i+1 f (g1 , . . . , gi gi+1 , . . . , gn+1 )+(1)n+1 f (g1 , . . . , gn )
i=1
1.2
Interpretation of H1 (G, A)
We would like to interpret H1 (G, A) and H2 (G, A) in terms of more familiar objects. Recall
that an extension of G by N is an exact sequence of groups:
/N
/E
/G
/1
/G
/1
/G
/1
/ E0
/N
1
i
/E
An extension 1 N
E
G 1 is split if it splits on the right, i.e. there is a
homomorphism s : G E such that s = idG .
Lemma 1.2.1. Let A be an abelian group with is also a G-module. Then any split extension
of G by A is equivalent to the canonical one:
0
/A
/AoG
/G
/ 1.
Recall that AoG, the semidirect product of G and A, is AG as a set, with (a, g)(b, h) =
(a + gb, gh). The first cohomology group H1 (G, A) classifies splittings up to A-conjugacy.
Every splitting s : G A o G is of the form g 7 (dg, g), where d is some map G A. The
fact that s is a group homomorphism forces (dg, g) (dh, h) = (dg + gdh, gh) = (d(gh), gh).
Thus we need d(gh) = dg + gdh. It would be natural to write dg h + g dh, but A is not
a G-bimodule. If it were, then this condition would require d : G A to be a derivation.
26
1.3
Interpretation of H2 (G, A)
/ A of G
/G
/0
extensions 0 A
E
G1
{normalized 2-cocycles}
with a (set-theoretic) normalized
Hp (Z/2, Z) = Hp (RP , Z) = 0
Z/2
if p = 0
if p 0 (mod 2) and p > 2
otherwise
/Z
Pn1
i=0
/ Z[G] 1t / Z[G]
/Z
/0
/ Z[G]
/ Z[G]
/ Z[G] 1t / Z
/0
One can check that this is a projective resolution of Z as a Z[G]-modules, and it yields
Z
Hp (Z/n, Z) = 0
Z/n
if p = 0
if p 0 (mod 2) and p > 2
otherwise
28
Hochschild (co)homology
2.1
For the rest of this section, let k be a field, and A be a associative, unital k-algebra. Also, let
M be an A-bimodule (also called a two-sided module), i.e. we have (am)b = a(mb). Define
enveloping algebra of A by Ae = A k Ao , where A denotes the opposite algebra of A. It
is easy to see that the category of A-bidmodules is equivalent to the categories of left and
right Ae -modules. Indeed, we define
(a b )m = amb
m(a b ) = bma
Example 2.1.1. Consider M = Ae as a module over itself. It is naturally a Ae -bimodule
in two different (commuting) ways. We can compute explicitly:
(a b )(x y) = ax b y = ax yb
outer structure
(x y)(a b ) = xa yb = xa by
inner structure
e A to be the complex
Consider the multiplication map m : A A A. Define B
[
e A :=
B
/ A3
/ A2
/A
/0
n1
X
(1)i a0 ai ai+1 an
i=1
/ B2 A
/ B1 A
/ B0 A
/0
We now compute
0
(h b )(a0 an ) =
n1
X
(1)i 1 a0 ai ai+1 an
i=0
(b0 h)(a0 an ) = 1 a0 a1 an +
n1
X
(1)i+1 1 a0 ai ai+1 an
i=0
2.2
Definition 2.2.1. A chain differential graded (DG) algebra (resp. cochain DG algebra)
over a field k is a Z-graded k-algebra, equipped with a k-linear map d : A A1 (resp.
d : A A+1 ) such that
1. d2 = 0
2. d(ab) = (da)b + (1)|a| adb for all a, b A with a homogeneous.
Here |a| denotes the degree of a in A . We call the second L
requirement the graded Leibniz
rule. (Recall that a graded algebra is a direct sum A = iZ Ai such that 1 A0 and
Ai Aj Ai+j .)
A DG algebra A is called non-negatively graded if Ai = 0 for all i < 0. In addition, if
A0 = k then A is called connected. We let DGAk denote the category of all DG k-algebras,
and DGA+
k denote the full subcategory of DGAk consisting of non-negatively graded DG
algebras.
Example 2.2.2 (Trivial DG algebra). An ordinary associative algebra A can be viewed as
DG algebra with differential d = 0 and grading A0 = 0, Ai = 0 for i =
6 0. Hence the category
Algk of associative algebras over k can be identified with a full subcategory of DGAk .
Example 2.2.3 (Differential forms). Let A be a commutative k-algebra. The
L de Rham
algebra of A is a non-negatively graded commutative DG algebra (A) = n0 n (A)
defined as follows. First, we set 0 (A) = A and take 1 (A) to be the A-module of K
ahler
differentials. By definition, 1 (A) is generated by k-linear symbols da for all a A (so
d(a + b) = da + db for , k) with the relation
d(ab) = a(db) + b(da), a, b A.
It is easy to show that 1 (A) is isomorphic (as an A-module) to the quotient of A A
modulo the relations ab c a bc + ca b = 0 for all a, b, c A. Then we define n (A)
using the exterior product over A by
V
n (A) := nA 1 (A)
30
Thus n (A) is spanned by the elements of the form a0 da1 dan , which are often
denoted simply by a0 da1 . . . dan and called differential forms of degree n.
The differential d : n n+1 is defined by
d(a0 da1 . . . dan ) := da0 da1 . . . dan
The product on the space (A) is given by the formula
(a0 da1 . . . an ) (b0 db1 . . . bm ) = a0 b0 da1 . . . dan db1 . . . dbm
This makes (A) a differential graded algebra over k. If X is a complex variety and
A = O(X) is the algebra of regular functions, then (A) = (X), where (X) is
the algebra of regular differential forms on X. However, if M is a smooth manifold and
A = C (M ), then the natural map (A) (M ) is not an isomorphism. Indeed, David
Speyer pointed out that if f, g are algebraically independent in A, then df and dg are linearly
independent in (A). (see the discussion before Theorem 26.5 in [Mat89]). Since ex and 1
are algebraically independent, d(ex ) and d(1) = dx are linearly independant over A = C (R)
in 1 (A). But certainly d(ex ) = ex d(1) in 1 (R).
Example 2.2.4 (Noncommutative differential forms). The previous example can be generalized to all associative (not necessarily commutative) algebras.
Suppose A is an (associative) algebra over a field k. First we define noncommutative
K
ahler differentials 1nc (A) as the kernel of multiplication map m : A A A:
0
/ 1 (A)
nc
/AA
/A
/0
d(a0 a
1 a
n ) = 1 a
0 a
1 a
n
Here for a A we denote by a
the element d0 (a) = a 1 1 a 1nc (A).
Actually, there is more conceptual way of defining Kahler differential and noncommutative
forms. Consider the functor Der(A, ) : A-bimod Sets associating to a bimodule M the set
of all derivations Der(A, M ). This functor is representable. Precisely, we have the following
31
(even)
(odd)
Lemma 2.2.8. Any derivation d (even or odd) is uniquely determined by its values on the
generators of A as a k-algebra. In other words, if S A is a generating set and d1 (s) = d2 (s)
for all s S, then d1 = d2 .
Proof. Apply iteratively the Leibniz rule.
Corollary 2.2.9. If d : A A is an odd derivation and d2 (s) = 0 for all s in some
generating set of A, then d2 = 0 on all of A.
Proof. If d is an odd derivation, then d2 is an even derivation. Indeed, d2 = 12 [d, d]+ , or
explicitly
d2 (ab) = d((da)b + (1)|a| adb)
= (d2 a)b + (1)|da| dadb + (1)|a| dadb + (1)|a|+|a| ad2 b
= (d2 a)b + a(d2 b)
The result follows now from the previous lemma.
Definition 2.2.10. If (A , d) is a DG algebra define the set of cycles in A to be
Z (A, d) := {a A : da = 0}
Notice that Z (A, d) is a graded subalgebra of A.
32
2.3
Why DG algebras?
/ (A A) o
Tk (A )
! d
A A
(2.1)
can
/ Tk (A )
In this diagram, : A (AA) is the linear map dual to , the map A A (AA) is
given by f g 7 [x y 7 f (x)g(y)] and it is an isomorphism because A is finite-dimensional.
By Lemma 2.2.8, any linear map A Tk (A ) determines a derivation d : T (A ) T (A ):
precisely, there is a unique d : T (A ) T (A ) such that
(1) d|A =
(2) deg(d) = +1
(3) d satisfies the graded Leibniz rule
Conversely, if d : T (A ) T (A ) satisfies (2) (3), then restricting d|A : A A A
and dualizing d : [A A ] ' A A A we get a linear mapping A a A.
Thus, if A is finite-dimensional, giving a bilinear map A A A is equivalent to giving
a derivation of degree 1 on T (A ).
Remark 2.3.1. For notational reasons, one usually takes = d, so that : A (A A)
is given by ()(x y) = (xy), A , x, y A.
Lemma 2.3.2. The map : A A A is associative if and only if 2 = 0 on T (A ).
33
2.4
da = 0,
d = 1
This makes Ahi a DG algebra. Notice, that since d(a) = 0 and |a| = 0 for a A, then d
e A, b0 ). Indeed, we have
is A-linear. DG algebra Ahi is isomorphic as a complex to (B
d(a0 a2 . . . an ) = a0 d()a1 a2 . . . an a0 d(a1 . . . an )
= a0 a1 a2 . . . an a0 a1 d()a2 . . . an + a0 a1 d(a2 . . . an ) =
= a0 a1 a2 . . . an a0 a1 a2 . . . an + a0 a1 d(a2 . . . an )
= ...
n
X
=
(1)i a0 . . . ai ai+1 . . . an ,
i=0
34
which exactly maps to the differential b0 (a0 a2 an ) via the identification map .
Notice that 1 = d, so 1 = 0 in H (Ahi), hence H (Ahi) = 0. Under our identification of
e A with Ahi, the homotopy h is just u 7 u. Indeed, we can check that for all u Ahi,
B
(dh + hd)(u) = (du + (1)1 du) + du = 1 u du = u
Given an A-bimodule M , define M Ae B A to be the complex
M Ae A(n+2)
Note that M Ae A(n+2) ' M Ae Ae An ' M k An via the map
m Ae (a0 an+1 ) 7 an+1 ma0 (a1 an )
The induced differential b : M k An M k An1 turns out to be
m1
X
m(a1 an ) 7 ma1 a2 an +
i=1
2.5
where
HomA (M, N )n = {f HomA (M, N ) : f (Mi ) Ni+n for all i Z}
is the set of A-module homomorphisms M N of degree n.
Warning In general, HomA (M, N ) 6= HomA (M, N ), i.e. not every A-module map f :
M N can be written as a sum of homogeneous maps.
L
Example 2.5.3. Let A = k be a field, N = k and M = V = nZ Vn a graded k-vector
space such that dim Vn > 1 for all n. Let f : V k be such that f (Vn ) 6= 0 for infinitely
many n. Then f 6 Homk (V, k). Why?
35
Exercise Prove that if M is a finitely generated (as A-module) then HomA (M, N ) =
HomA (M, N ).
Definition 2.5.4. For A a ring and M, N chain complexes over A, define
dHom : HomA (M, N )n HomA (M, N )n1
by
f 7 dN f (1)n f dM
Note that this is well-defined because (dN f )(Mi ) dN (Ni+n ) Ni+n1 and (f
dM )(Mi ) f (Mi1 ) Ni1+n . We claim that d2Hom = 0. Indeed, we have
d2Hom (f ) = dN (dN f (1)n f dM ) (1)n1 (dN f (1)n f dM ) dM
= d2N f (1)n dN f dM + (1)nN f dM + (1)2n1 f d2M
=0
Definition 2.5.5. Let A be a k-algebra, M an A-bimodule. The Hochschild cochain complex
of M is
C n (A, M ) = HomAe (B A, M )n
where M is viewed as a left Ae -module via (a b )m = amb, and we view M as a complex
concentrated in degree zero.
Explicitly, we have
C n (A, M ) = HomAe (A(n+2) , M )
= HomAe (An+2 , M )
= Homk (An , M )
A map f Homk (An , M ) is identified with : A(n+2) M , where (a0 an+1 ) =
a0 f (a1 , . . . , an )an+1 . The differential is dnHom = (1)n+1 b0 , or in terms of f : An M ,
n
X
(d f )(a1 , . . . , an+1 ) = a1 f (a2 , . . . , an+1 )+ (1)i f (a1 , . . . , ai ai+1 , . . . , an+1 )+(1)n1 f (a1 , . . . , an )an+1
n
i=1
2.6
Example 2.6.1 (Center). Unpacking the definition, we get C 0 (A, M ) = Homk (A0 , M ) =
Homk (k, M ) = M . The differential d0 : M C 1 (A, M ) = Homk (A, M ) sends m to the
function
d0 (m)(a) = am am
We have HH0 (A, M ) = Ker(d0 ) = {m M : am ma = 0} = Z(M ), the center of the
bimodule M .
Example 2.6.2 (Derivations). We have d1 : C 1 (A, M ) C 2 (A, M ), defined by
(d1 f )(a1 a2 ) = af (a2 ) f (a1 a2 ) + f (a1 )a2
One checks that Ker(d1 ) = Derk (A, M ) = {f Homk (A, B) : f (ab) = af (b) + f (a)b}. The
map d0 : M C 1 (A, M ) sends m to the inner derivation adm : a 7 [a, m]. Thus we have
an exact sequence
0
/ Z(M )
/M
ad
/ Derk (A, M )
/ HH1 (A, M )
/0
2.7
Extensions of algebras
/M
/R
/A
/0
/E
/0
/A
/0
/ E0
/M
/A
2.8
Crossed bimodules
/0
/ C1
/ C0
/0
Explicitly, C0 is an algebra, C1 is a bimodule over C0 , and C12 = 0. The Leibniz rule implies
that for all a C0 , b C1 , we have (ab) = a(b) and (ba) = (b)a, i.e. : C1 C0 is
38
/ C1
d1
/ C0
/0]
/ coker(d2 )
d1
/ C0
/0]
/ C2
C = [
/M
/ C1
/ C0
/A
/0
/M
/ C1
/ C0
/A
/0]
2.9
L
Let A = ( p0 Ap , d) be a DG algebra. Consider the graded vector spaces
C1 := Coker(d)0 [1] C0 := Ker(d)
Note that C0 is a graded subalgebra of A while C1 is a graded C0 -bimodule. The
differential d on A induces a graded map
: C1 C0
(2.2)
which makes (2.2) a graded cross-bimodule. The cokernel of is the algebra H (A), while
the kernel of is the H (A)-bimodule whose underlying (graded) vector space is H1 (A)[1].
The right multiplication on H1 (A)[1] is given by the usual multiplication in H (A), while
the left multiplication is twisted by a sign:
a
s(
x) = (1)|a| s(
ax
),
where a
H (A) is homogeneous, x
H1 (A)[1].
Definition 2.9.1. By Theorem 2.8.6 the crossed bimodule
0
/ H (A)
/ H1 (A)[1]
/0
represents an element A HH3 (H (A), H1 (A)[1]), which is called the characteristic class
of A.
This class is secondary (co)homological invariant of A. It is naturally related to Massey
triple products. In more detail, let
E := [ 0
/M
/ C1
/ C0
/B
/0]
where E is the Hochschild 3-cocycle associated to the crossed extension E in the proof of
Theorem 2.8.6.
Finally, if the crossed extension E comes from a DG algebra A , i.e. A = [E] in
HH3 (H (A), H1 (A)[1]), then we recover the classical definition of triple Massey products
for homology classes a, b, c H (A) of a GD algebra (see [GM03] for details).
Remark 2.9.2. One useful application of characteristic classes of DG algebras is concerned
with realizability of modules in homology:
, we say
Given a DG algebra A with homology H (A) and a graded H (A)-module M
. Here, by DG
is realizable if there is a DG module M over A such that H (M ) ' M
that M
module we mean a graded module M over the DG algebra A endowed with a differential
dM : M M1 satisfying d2M = 0 and dM (am) = dA (a) m + (1)|a| a dM (m).
It turns out that the characteristic class A of the DG algebra A provides a single
. In particular, if A = 0, then any graded H (A)-module is
obstruction to realizability of M
realizable. For details, see [BKS03].
Deformation theory
The main reference for the main part of this section is the survey by Bertrand Keller [Kel03].
3.1
Motivation
In classical mechanics, one starts with the phase space, which is a symplectic manifold (e.g.
the cotangent bundle T X). The ring of smooth functions C (M ) has extra structure: the
Poisson bracket {, } : C (M ) C (M ) C (M ), and the Hamiltonian H C (M ).
Locally, the equations of motion (i.e. the Hamilton equations) are, for coordinates pi , qi M :
pi = {H, pi }
qi = {H, qi }
where f =
df
dt .
pi qi
pi qi
i=1
Note that {pi , pj } = 0, {qi , qi } = 0, and {pi , qj } = ij . In this context, the Hamilton
equations are
H
qi
H
qi =
pi
pi =
41
3.2
Formal deformations
Let k be a commutative ring (later a field of characteristic zero). Let A be a fixed unital
associative k-algebra. Write kt = kJtK for the ring of formal power series in t, and let
At = AJtK = A k kJtK. Elements of At look like
X
u=
an tn
, an A.
n>0
t70
AA
/ At
t70
/A
Basically, we are thinking of At as a topological kt -algebra, and taking tensor product in the
category of topological kt -algebras. By continuity, is determined uniquely by its restriction
to AA. For all a, b A, we can write ab = ab+B1 (a, b)t+B2 (a, b)t2 + +Bn (a, b)tn + ,
where the Bi : A A A are bilinear maps and B0 = m.
Let Gt = Autkt (At ) be the group of all kt -linear automorphisms of At such that
such that : At At satisfies (a) a (mod t) for all a A. In other words, we have
(a) = a + 1 (a)t + 2 (a)t2 + where the i are k-linear.
Definition 3.2.2. Two products and 0 are equivalent if there exists Gt such that
(u v) = (u) 0 (v) for all u, v At .
Definition 3.2.3. A Poisson bracket on A is a k-bilinear map {, } : A A A such
that for al a, b, c A
42
/ {Poisson brackets on A}
F G F G
.
x y
y x
43
Then A is given by
X
n F n G tn
F t G =
xn y n n!
n=0
{F, G} =
(x1 , . . . , xn )
xi xj
xi xj
i<j
It means that there are (unique) smooth functions ij C (M ), 1 i < j n such that
X
ij
{, } =
xi xj
i<j
The functions (ij ) are actually components of a tensor field of type (2, 0) on M which is
called Poisson bivector.
3.3
The general idea is any deformation problem is controlled by a dg Lie algebra. A bit
more precisely, suppose we hav a category A and an object A Ob A. We can define a
deformation functor Def A (A, ) : R Set, where R is some category of test commutative
algebras (or cocommutative coalgebras). For example, R could be the category of artinian
local algebras over a field (which can be of characteristic p > 0). The functor Def A (A, )
assigns to some R the set of all deformations of A parameterized by Spec(R), modulo
equivalence. We say that a dgla LA controls this deformation problem if there is a natural
isomorphism
Def A (A, R) ' MC(LA , R)
where MC(LA , R) is the set of Maurer-Cartan elements in LA R.
Assume from now on that k is a field of characteristic zero. Let A = Algk be the category
of associative unital k-algebras. Let R = Artk be the category of local Artinian k-algebras
with residue field k. That is, R Ob R if and only if R is a local commutative k-algebra
with finite-dimensional maximal ideal m R, such that R/m = k. This clearly implies
mn = 0 for all n 0. A good example is R = k[t]/(tn ).
Given R R, write AR for A k R.
Definition 3.3.1. An R-deformation of A is an associative R-linear map : AR R AR
AR such that the following diagram commutes:
AR R AR
eR e
R
m
AA
44
/ AR
eR
/A
where
eR = 1 R and R : R R/m is the canonical projection.
We say that two R-deformations and 0 are equivalence if there is an R-module
isomorphism g : AR AR such that
AR
/ AR
commutes, and such that g(u v) = g(u) 0 g(v) for all u, v AR . Note that by R-linearity,
: AA Am
is determined by its restriction to A A. Moreover, is determined by e
because = m + e
.
Definition 3.3.2. The deformation functor Def(A, ) : R Set is given by
Def(A, R) = {R-deformations of A}/equivalence.
Definition 3.3.3. If R = k[t]/t, then R-deformations of A are called infinitesimal.
Lemma 3.3.4. There is a natural bijection Def(A, k[t]/t2 ) = HH2 (A, A).
Proof. By definition, for R = k[t]/t2 , an R-deformation is determined by
: A A A[t]/t2 = A k[t]/t2
which will be of the form a b 7 ab + B1 (a, b)t, where B1 : A A A is some k-linear
map. The associativity of is equivalent to B1 being a Hochshield 2-cocycle. Indeed,
(a b) c = (ab)c + (B1 (a, b)c + B1 (ab, c))t
a (b c) = a(bc) + (aB1 (b, c) + B1 (a, bc))t
Since A is associative, the two are equal exactly when
(d2Hoch B1 )(a, b, c) = aB1 (b, c) B1 (ab, c) + B1 (a, bc) B1 (a, b)c = 0
Moreover, if , 0 are equivalent, then there is g : A A[t]/t2 such that g(a b) = g(a) 0 g(b),
which is equivalent to
B10 (a, b) B1 (a, b) = ag1 (b) g(ab) + g(a)b
Thus 0 if and only if B10 B1 = dg1 .
Remark 3.3.5. One might hope that Def Algk (A, k[t]/tn ) ' HHn (A, A). Unfortunately, this
is not true in general if n > 3. However, it is true that Def Alg (k) (A, k[t]/tn ) = HHn (A, A),
where |t| = 2n, and Alg (k) is the category of A -algebras (also called strongly homotopyassociative algebras) over k. Note that A -algebras are not in general associative.
45
Definition 3.3.7. Let L be a differential graded Lie algebra. The space of Maurer-Cartan
elements in L is
1
MC(L) = x L1 : dx + [x, x] = 0
2
If L1 is finite-dimensional, then MC(L) is actually a variety (in fact an intersection of
quadrics in L1 ). For x MC(L), set
Tx MC(L) = v L0 : dv + [x, v] = 0 .
This is precisely the Zariski tangent space to MC(L). The action of L0 on L1 fixes MC(L).
If L0 is a nilpotent Lie algebra, we can define the reduced Maurer-Cartan space to be
MC(L) = MC(L)/ exp(L0 ).
We will see that there exists a dgla LA such that Def(A, R) ' MC(LA mR ).
We continue to assume that k is a field of characteristic zero. Let L = L be a dgla over
k. The Maurer-Cartan space
1
1
MC(L) = x L : dx + [x, x] = 0
2
can be regarded as a subscheme of L1 , where L1 is viewed as affine space over k. We would
like to compute the Zariski tangent space of MC(L). Recall that if X is an affine variety (or
scheme) over an algebraically closed field k, and x X is a closed point, then there is (by
the Nulstellensatz) a k-algebra homomorphism : O(X) k with Ker() = mx .
Definition 3.3.8. The Zariski tangent space of X at x is
Tx X = Der(O(X), x) = { : O(X) k : (f g) = f g + gf }
There is a canonical identification of Tx X with (mx /m2x ) . For more details, see any
good book on algebraic geometry.
46
/ Tx f 1 (y)
/ Tx V
dfx
/ Ty W
1
d(x + vt) + [x + vt, x + vt]
2
t=0
= dv + [x, v].
This allows us to make the following definition even if L1 is not locally finite-dimensional.
Definition 3.3.12. The (Zariski) tangent space to MC(L) at x is
Tx MC(L) = {v L1 : dx v = 0}.
where dx = d + adx .
47
From here on out, assume that L0 is a nilpotent Lie algebra, i.e. for all L0 , the
endomorphism ad End(L0 ) is nilpotent. Moreover, we assume that the adjoint action
of L0 on L1 is nilpotent. Consider the group Aff(L1 ) of affine linear transformations of L1 .
This is just the semidirect product GL(L1 ) n L1 . Let aff(L1 ) be the Lie algebra of Aff(L1 ).
We have aff(L1 ) = gl(L1 ) n L1 , and there is an anti-homomorphism L0 aff(L1 ) given by
7 dx = d + [x, ]. Exponentiation gives an anti-homomorphism exp(L0 ) Aff(L1 ),
which yields a right action of exp(L0 ) on L1 . This action restricts to an action of exp(L0 )
on MC(L).
Definition 3.3.14. MC(L) = MC(L)/ exp(L0 )
Remark 3.3.15. For x MC(L), we can consider the orbit Ox = x exp(L0 ) MC(L). It
turns out that Tx MC(L)/Tx Ox = Ker(dx )/ Im(dx ) = H1 (L, dx ).
Let R = Art be the category of local commutative k-algebras with finite-dimensional
(hence nilpotent) maximal ideal and residue field k. For any dgla L and for any R R with
maximal ideal m, the Lie algebra L m is nilpotent.
Definition 3.3.16. MC(L, R) = MC(L mR ) and MC(L, R) = MC(L mR ).
48
/ MC(L2 , R).
/ (L2 m, d).
3.4
Let A be an associative k-algebra, where kLis a field of characteristic zero. Recall that the
p
p
Hochschild complex of A is C (A, A) =
pZ C (A, A), where C (A, A) = 0 for p < 0,
p
p
p
p+1
C (A, A) = Homk (A , A) for p > 0, and d : C C
is
p1
X
(1) (df )(a0 , . . . , ap ) = a0 f (a1 , . . . , ap )+ (1)i f (a0 , . . . , ai ai+1 , . . . , ap )+(1)p1 f (a0 , . . . , ap1 )ap
p
i=0
p
X
i=0
Notice that the operation is not associative in general. We can use to define the
following bracket.
Definition 3.4.2. The G-bracket (Gerstenhaber bracket) [, ]G : C p C q C p+q1 is
defined by
[f, g]G = f g (1)|f ||g| g f .
Let LAS (A) = C (A, A)[1]. That is, LpAS = C p+1 (A, A). The Gerstenhaber bracket
induces a bracket [, ]G : LpAS LqAS Lp+q
AS of degree zero.
Lemma 3.4.3 (Gerstenhaber). (LAS (A), [, ]G , d) is a dg Lie algebra.
Proof. This can be checked directly, though checking the (super) Jacobi directly in this way
is quite tedious. However, it follows from the following.
Since is not associative we define the associator
A(f, g, h) = (f g) h f (g h)
One can check that A is (super)symmetric in g and h, i.e. A(f, g, h) = (1)(|g|1)(|h|1) A(f, h, g).
This symmetry formally implies the Jacobi identity.
49
Remark 3.4.4. The Lie algebra (LAS (A), [, ]G ) only depends on A as a vector space.
The multiplication on A enters in this picture as follows. We have m C 2 (A, A) = L1AS (A),
and it is easy to check that for all f C p (A, A), we have [m, f ] = df , where d is the
Hochschild differential. We can rewrite this as d = adm .
We define the cup-product on C (A, A) by
(f ` g)(a1 , . . . , ap+q ) = f (a1 , . . . , ap )g(ap+1 , . . . , ap+q )
where f C p , g C q . It is easy to check that d(f ` g) = df ` g + (1)|f | f ` dg.
By the lemma 3.4.3, (HH (A, A)[1], [, ]G ) is a graded Lie algebra. In fact, (HH (A, A), [, ]H , `)
is a (graded commutative) Gerstenhaber algebra.
Definition 3.4.5. A Gerstenhaber algebra G is a graded commutative algebra with product
and bracket {, } : Gp Gq Gp+q1 such that
1. (G [1], {, }) is a graded Lie algebra
2. {a b, c} = a {b, c} + (1)(|a|1)|b| {a, b} c
Example 3.4.6 (Gestenhaber algebra from differential geometry). Let M be
Vpa smooth
manifold over R, and let T M be the tangent
P bundle of M . Set p (M ) = (M,L T M ). In
coordinates, an element of p (M ) looks like 1 p . We can set (M ) = p>0 p (M ).
This is a graded commutative algebra with respect to the exterior product . The algebra
p (M ) has a Schouten bracket {, }S : p q p+q1 , defined by
{1 p , 1 , q } =
p X
q
X
(1)i+j [i , j ] 1 bi 1 bj q
i=1 j=1
3.5
Stasheff construction
The following is a construction by Jim Stasheff, 1993. Recall that an associative algebra is a
vector space A with a map m : A A A such that the following diagram commutes:
AAA
m1
1m
m
AA
/AA
/A
/C C
C C
/C C C
(x1 , . . . , xn ) =
n1
X
(x1 , . . . , xi ) (xi+1 , . . . , xn )
i=1
3.6
Recall that for an algebra A, (LAS (A) = (C (A, A)[1], [, ]G ) is a dg Lie algebra that
controls deformations of A in the sense that
MC(LAS (A), R)) ' Def(A, R)
for all R R.
Definition 3.6.1. Two dg Lie algebras L and L0 are called homotopy equivalent if there is
a sequence of dg Lie algebras L1 , . . . , Ln with quasi-isomorphisms
L = L0 L1 Ln = L0
(the arrows can be in either direction).
Lemma 3.6.2 (Goldman-Milson). Let L and L0 be dg Lie algebras. The following are
equivalent:
1. L and L0 are homotopy equivalent
2. There exists L0 such that that there are quasi-isomorphisms
L0
L0
3. There is an L homomorphism L L0
We think of L0 from the Goldman-Milson lemma as being a generalized morphism
from L to L0 .
Let A = C (M ) be the ring of functions on a smooth manifold M . Recall that in
Lemma 3.2.4 we have constructed a natural map A from formal deformations of A to
Poisson brackets on A. Kontsevich proved that A is surjective. There is a dg Lie algebra
LPois (A) that controls Poisson deformations of A with the trivial bracket. Kontsevichs
theorem can be interpreted as saying that LAS (A) and LPois (A) homotopy equivalent (i.e.
there is an L -morphism LAS (A) LPois (A). The existence of this homotopy is known as
the formality theorem. To be more precise, we need the following definitions.
Definition
V2 3.6.3 (Chevally-Eilenberg (co)homology). Let g be a Lie algebra with bracket
[, ] :
g g. Let V be a g module (i.e. there is a homomorphism of Lie algebras
g End V ). The Chevally-Eilenberg complex C (g, V ) has C p (g, V ) = 0 if p < 0, and
C p (g, V ) = Homk (
52
Vp
g, V )
with differential
(1)p df (X0 , . . . , Xp+1 ) =
ci , . . . , X
cj , . . . , Xp+1
(1)i+j+1 f [xi , xj ], X1 , . . . , X
i<j
X
ci , . . . , Xp+1 )
+
(1)i Xi f (X1 , . . . , X
i
Define [f, g]C = f g = (1)(|f |1)(|g|1) g f . Then (LCE , [, ]CE ) is a dg Lie algebra.
Theorem 3.6.5 (Kontsevichs formality theorem). For a smooth manifold M and A =
C (M ), gA = (A, {, } = 0), there is a homotopy equivalence
LAS (A) LCE (gA )
Its very interesting to study the space of L equivalences between LAS (A) and LCE (gA ).
d . The group
This group admits a faithful action of the Grothendieck-Teichm
uller group GT
d
d
GT contains Gal(Q/Q) as a subgroup, and Grothendieck conjecture that GT = Gal(Q/Q).
d on L -equivalences between LAS (A) and LCE (gA )
There is a conjecture that the action of GT
is simply transitive!
3.7
Another source of motivation for deformation theory comes from algebraic number theory.
Let k be a field of characteristic not 2 or 3. Recall that an elliptic curve over a field k is the
subset of P2k given by a homogeneous equation
y 2 z = x3 + axz 2 + bz 3
where a, b k are such that = 16(4a3 + 27b2 ) 6= 0. Let E be an elliptic curve. One can
show using the Riemann-Roch theorem that E naturally has the structure of an abelian
variety (projective group variety) with unit (0 : 1 : 0) in projective coordinates. For each
integer N > 5, there is a smooth projective curve X0 (N ) over Q that parameterizes elliptic
53
et
Q
As a Zp -module, Tp ' Z2
p , so the action of GQ = Gal(Q/Q) on Tp gives us a representation
E,p : GQ GL(2, Zp )
On the other hand, each modular form f of level N gives rise to an ideal If End(J0 (N )),
and the quotient J0 (N )/If is an elliptic curve. We denote the representation J0 (N )/If ,p by
f,p . It is a theorem that E is modular if and only if E,p is isomorphic to f,p for some
modular form f .
For a special class of elliptic curves, it was already known that the mod p representation
E,p : GQ GL(2, Zp ) GL(2, Fp )
was modular. What Andrew Wiles did is consider the category R whose objects are finite
local Zp -algebras with residue field Fp , and define the functor DE,p : R Set by
DE,p (R) = { : GQ GL(2, R) : E,p
(mod mR )} /
There are many other places in number theory where one attempts to classify deformations
of object X0 defined over Fp . One does this by defining a functor assigning to each R R
some class of lifts of X to R, and then hoping that the deformation functor is representable.
Unlike the situation of this course, where the fact that our deformation functor is representable
(by a dgla) is trivial, in the number-theoretic context it is often very difficult to show that a
given deformation problem is representable.
55
56
Chapter 3
Category theory
1
1.1
Definition of categories
58
is a quiver. Any quiver Q can be thought of as a category Q, whose objects are vertices
of Q, and morphisms between vertices vi and vj are all paths from vi to vj . Here
path means a sequence of arrows f1 , . . . , fn such that f1 starts at vi and fn ends in
vj . If vi = vj = v then we also include the identity morphism idv into Hom(v, v).
There is another category associated to a quiver. Namely, denote by Qcom the category
with Ob(Qcom ) = Ob(Q). Set of morphisms HomQcom (vi , vj ) contains unique element
if HomQ (vi , vj ) is nonempty, and HomQcom (vi , vj ) = otherwise. Intuitively this
means that all paths between vi and vj define the same morphism in Qcom .
1.2
F (f )
F (Y )
/ G(X)
G(F )
/ G(Y )
Example 1.2.7. Consider functor GLn () : ComRings Gr with GLn (R) being the group
of n n invertible matrices with coefficients in the ring R. Also, consider the functor
() : ComRings Gr which associates to a ring R the subgroup R of its units. Then
taking determinant of a matrix defines a natural transformation det : GLn () () .
Example 1.2.8 (Convolution of morphims of functors). Suppose there are categories C, D
and E with functors F : C D and G, G0 : D E, along with a natural transformation
: G G0 . We write
C
/D
6E
G0
6D
/E
F0
/ G(X)
F (f )
F (Y )
G(f )
/ G(Y )
/ F (X)
G(f )
G(Y )
F (f )
/ F (Y )
1.3
Equivalences of categories
Example 1.3.3. Let D = Vectnk be the category of n-dimensional vector spaces over a field
k. Let C be the full subcategory of D consisting of a single object k n . The inclusion functor
i : C D is an equivalence of categories, even though i is not a bijection at the level of
objects. A quasi-inverse of C arises from choosing a basis for each V Ob Vectnk . To each
f : V W , we assign the matrix representation of f in terms of our chosen bases of V and
W.
This example is typical. Equivalent categories may have different objects, but the same
isomorphism classes of objects. Also, the construction of a quasi-inverse typically requires
the axiom of choice.
Recall that F is fully faithful if the maps HomC (X, Y ) HomD (F (X), F (Y )) are
bijections, and F is essentially surjective if for all Y Ob D, there exists X C for which
F (X) ' Y .
Theorem 1.3.4 (Freud). A functor F : C D is an equivalence of categories if and only
if it is fully faithful and essentially surjective.
Proof. We will prove a bit later more general result that will imply this theorem.
If C is a category, a skeleton sk C of C is a full subcategory of C with one object in each
isomorphism class. The theorem shows that sk C , C is an equivalence of categories.
Example 1.3.5 (Groupoids). A groupoid is a (small) category in which all morphisms are
isomorphisms. So a group is just a groupoid with one object. We say that a groupoid is
connected if any two objects can be connected by arrows (possibly in both directions). It
is easy to see that the skeleton of a connected groupoid is a group. The main example is
the fundamental groupoid (X) of a topological space X. Objects of (X) are points in X,
and Hom(X) (x, y) is the set of homotopy classes of paths x y. If X is connected, the
choice of a point x X gives rise to an equivalence of categories 1 (X, x) , (X).
Example 1.3.6. There are several equivalences of categories relating algebra and geometry.
For example, if k is a field, the category of finitely generated commutative reduced k
algebras is anti-equivalent to the category of affine varieties over k via the functor A 7 Spec A.
More generally, the category of all commutative k-algebras is anti-equivalent to the category
of affine schemes over k, once again via Spec. In both cases, Spec has a quasi-inverse, namely
X 7 OX (X).
The category of all associative k-algebras does not have a good geometric analogue. Of
course, one can define the category of non-commutative affine schemes to be the opposite
of the category of associative k-algebras, but this is not reasonable, as is seen by the next
section.
1.4
, g 7 f g = f g.
hf (Y )
/ hX (Y )
2
hX1 (Z)
hf (Z)
/ hX (Z)
2
The upper path sends g : Y X1 to (f g) s, while the lower path sends g to f (g s).
The two are equal by associativity. It is easy to see that hf g = hf hg , because this is
equivalent to (f g) = f g .
b Then the map Hom b(hX , F ) F (X) given by
Theorem 1.4.2 (Yoneda). Let F Ob(C).
C
7 (X)(idX ) is a bijection.
Proof. Let y : HomCb(hX , F ) F (X) be the map 7 (X)(idX ). We show that y is a
bijection by constructing an explicit inverse. Given x F (X), we want a morphism of
functors i(x) : hX F . This would consist of morphisms i(x)(Y ) : hX (Y ) F (Y ) for
each Y . Given f hX (Y ) = Hom(Y, X), we have a map F (f ) : F (X) F (Y ). We define
i(x)(Y )f = F (f )(x). It is not difficult to show that i(x) actually is a morphism of functors.
We will show that i is an inverse to y.
First we show that i is a right inverse to y. For x F (x), i(x) is defined by i(x)(Y )f =
F (f )(x), so y(i(x)) = i(x)(X)idX = F (idX )(x) = x.
Now we show that i is a left inverse for h. Given : hX F , let x = y() = (X)(idX ).
We need i(x) = , i.e. (Z) = i(x)(Z) for all Z Ob C. This is just the claim that
for all f : Y X, we have (Z)(f ) = F (f )(x). Apply the definition of is a natural
transformation to f : Y X. We get a commutative diagram
hX (X)
(X)
hX (Y )
(Y )
63
/ F (X)
F (f )
/ F (Y )
The commutativity of this diagram is the fact that F (f )((X)s) = (X)(s f ) for any
s : X X. If we choose s = idX , then we get F (f )((X)idX ) = (X)(f ), i.e. i(x)f =
F (f )(x) = (X)(f ).
Corollary 1.4.3. Let X, Y Ob C. Then h : HomC (X, Y ) HomCb(hX , hY ) is a bijection.
Proof. Let F = hY . The Yoneda lemma says that the map Hom(hX , hY ) F (X) =
Hom(X, Y ) given by 7 (X)(idX ) is a bijection. It is easy to check that this map is the
inverse to h : Hom(X, Y ) Hom(hX , hY ), so h is a bijection.
In light of the corollary, we can use h to regard C as a full subcategory of Cb consisting
b is representable, then the representing object X is
of representable functors. If F Ob(C)
determined uniquely (up to canonical isomorphism). By this we mean the following. Suppose
b and isomorphisms : F hX , : F hY . Thus 1 : hX hY
we have F Ob(C)
is a natural isomorphism. The Yoneda lemma gives an isomorphism HomC (X, Y )
HomCb(hX , hY ) Applying h1 to 1 gives a canonical isomorphism X Y . We can
We are asking if Repn (A) is corepresentable. The answer is yes! Define n A = (A k Mn (k))Mn (k) ,
where denotes free product. In other words,
n
A = {A A k Mn (k) : [a, m] = 0, m Mn (k)}.
Let eij Mn (k) be the elementary matrix with 1 in the (i, j)-th coordinate. Then eij form
a canonical k-basis for Mn (k). For a A k Mn (k), let
aij =
n
X
eki a ejk .
k=1
n
We claim that [aij , Mn (k)] = 0, and that
A is spanned by the aij . Let An = n A/h[ n A, n A]i
be the abelianization of the algebra n A. It turns out that Repn (A) is represented by An .
This construction is due to Bergman [Ber74].
Example 1.4.5 (Hilbert scheme). Let k be an algebraically closed field of characteristic
zero, and let X be a projective variety over k. Define the functor HilbX : Schk Set by
U 7 {Z U X closed subscheme such that U : Z U is flat}
That is, HilbX (U ) is the set of families of closed subschemes of U X parameterized by U .
One can prove that HilbX is representable. However, the representing object is a scheme,
not a variety. One remedies this by stratifying HilbX via Hilbert polynomials.
For u U , let Zu = 1 (u), and define the Hilbert polynomial of Z at u by PZ,u (m) =
(OZu OX (m)), where denotes Euler characteristic. It is a theorem that PZ,u actually
is a polynomial that is independent of u if U is connected. For some polynomial P , let
HilbPX (U ) = {Z U X : Z is a flat subscheme with PZ = P }.
It is a major theorem of Grothendieck that HilbPX is representable by a projective variety.
Example 1.4.6 (PDEs). Let n > 1, and let U Cn be an open subset. Consider a
differential operator
X
P =
a (z1 , . . . , zn )z
||6m
1.5
Adjoint functors
It is tedious but straightforward to check that this construction actually makes G a functor.
Example 1.5.4 (Free objects). Consider forgetful functor for : Gr Set that associates
to each group itself, but viewed as a set (i.e. it forgets the group structure). Then
operatornamef or is right adjoint to the functor free : Set Gr that associates to each set S
free group free(S), generated by S. The same is true with forgetful functors from Ab, Ring,
ComRing, Algk et.c.
Example 1.5.5 (Matrix representations continued). Recall we had a natural isomorphism
HomComAlgk (An , B) ' HomAlgk (A, Mn (B)).
Now we realize that the functor ()n : Algk ComAlgk , A 7 An = n A/[ n A, n A], is left
adjoint to the functor Mn : ComAlgk Algk . It turns out that we have an adjoint pair
n
: Algk Algk : Mn ().
Example 1.5.6 (Tensor-Hom adjunction). Let R and S be associative unital rings. Let
C = Mod(R), D = Mod(S) be the categories of right modules over R and S respectively. Let
B be an (R, S)-bimodule. Then we have a functor () R B : Mod(R) Mod(S) has right
adjoint HomS (B, ). That is, there is a natural isomorphism
: HomS (M R B, N )
HomR (M, HomS (B, N )).
This is easy to check. Given f : M B N , we can define the map (f )(m) = f (m ).
The map has inverse 1 ()(m b) = (m)b.
Let f : R S be any ring homomorphism. Then we have an adjoint triple of functors
f!
Mod(S)
The functor f is restriction of scalars via f . We have f (N ) = N S S, with right adjoint
f ! = HomS (S, ). The left adjoint of f is f = R S.
As a concrete example, suppose f : R R/I is the canonical surjection, where I R is
a two-sided ideal. Then f assigns to an R-module the largest quotient killed by I, f ! sends
an R-submodule to the largest submodule which is also a submodule over S, and f sends
an R-module to its quotient by I.
Let (F, G) be an adjoint pair, and let X,Y : HomD (F (X), Y ) HomC (X, G(Y )) be
a natural isomorphism witnessing this adjunction. If we take Y = F (X), we get a map
: X GF (X) corresponding to idF (X) . The maps X = x,F (X) (idF (X) ) define a
morphism of functors : idC GF called the unit of the adjunction. Dually, if we
1
take X = G(Y ), we can define : F G idD by Y = G(Y
),Y (idG(Y ) ). One calls the
counit of the adjunction. Note that we can define the convolutions F : F F GF and
G : F GF F .
67
/ G F (X)
F (X)
/ E F (X) E(a) / EF (Y ) Y / Y
1.6
Let J be a fixed category, called the index category. We assume that J is finite, or at least
small. For an arbitrary category C, we write C J for the functor category Fun(J, C). It is
helpful to think of C J as the category of diagrams of shape J in C.
For example, J could be a category with no non-identity morphisms (such categories are
called discrete). Another very useful example is J = { }. Objects of C J are called
pullback data in C. Both of these are special cases of when J is a poset, where we treat J as
a category via
(
{} if i 6 j
HomJ (i, j) =
otherwise
There is an obvious functor : C C J that sends X Ob C to the constant diagram
j 7 (X)(j) = X, with (X)(i j) = idX . Given a morphism : X Y in C, define
() : (X) (Y ) by letting (j : (X)(j) = X (Y )(j) = Y be itself. We call
the diagonal (or constant) functor. It is natural to ask whether has a left or right adjoint.
Fix F : J C (i.e. F Ob(C J )) and define Fe : C Set by Y 7 HomC J ((Y ), F ).
Definition 1.6.1. If Fe is representable, we call the representing object X the limit of F ,
written X = lim F .
Any such natural isomorphism comes from a morphism s : (lim F ) F . We can make the
vertex Y to get a come with vertex Y 0 . It is natural to ask if there is a terminal cone, i.e. a
cone that induces all cones by pullback. This happens precisely when F has a limit. That
is, X = lim F with s : (X) F is the closest (to C) cone over F .
can be extended to a functor lim : C J C. The functor lim is the right adjoint of , i.e.
We call a category C complete if all limits of shape J exist in C for all small categories J.
Dually, we can define colimits, which (if they exist) are left adjoint to . That is, there
is an adjoint pair
/C:
lim : C J o
The terminology is slightly confusing. A limit is often called a projective (or inverse)
limit, especially if the index category is a poset and F is contravariant. Dually, colimits are
called inductive (or direct) limits, especially if J is a poset and F is covariant. The notation
colim F is sometimes used instead of lim F .
Example 1.6.3 (Initial and terminal objects). If the index category J = is the empty
category, then there exists unique functor F : J C. The colimit and limit of F are initial
and terminal objects in C respectively.
Example 1.6.4 (Products and coproducts). A (small) category J is called discrete if for
any X, Y Ob(C), HomC (X, Y ) = if X =
6 Y and HomC (X, X) = {idX }. Schematically we
have J = { . . . }. Then a functor
F
:
J C is just a collection Q
of objects {Xj }jJ in C.
`
Then lim F is just the coproduct
Xj and lim F is the product
Xj .
jJ
jJ
Example 1.6.5 (Pushouts). If J = { }, then functors F : J C are just
diagrams X Y Z, also known as pushout data. The colimit of F corresponding to
X Y Z is the pushout of X and Z over Y , denoted X tY Z.
Let I be a small category. If we have a functor F : I C, we think of F as a diagram
of shape I. For i I, write Xi for F (i). We defined the limit of F , denoted lim F , to
/ Xi
fj
F (f )
Xj
69
and such that (X, ) is terminal with respect to this property. If G : C D is any functor,
we can apply G to (X, ) to get computable morphisms G(i ) : G(lim F ) G(Xi ), and
exists.
Dually, given F : I C, the colimit X = lim F has morphisms : Xi X. Applying
Theorem 1.6.8. For any object X in a category C, the functor hX = HomC (X, ) preserves
limits and the functor hX = HomC (, X) maps colimits to limits.
Proof. The fact that hX preserves limits is easy to check. Just construct an inverse to F .
Moreover, the fact that hX maps colimits to limits follows trivially from the fact that hX is
limit-preserving. Treat hX : C Set as a covariant functor on the opposite category C . In
70
adjointness
by Theorem 1.6.8
by Theorem 1.6.8.
2.1
Example 2.1.1. Let Ab be the category of abelian groups. For two groups A, B, the set
HomAb (A, B) is not only a set it naturally has the structure of an abelian group in which
composition is bilinear.
Definition 2.1.2. A category A is additive if
AB1 For all X, Y Ob A, the set HomA (X, Y ) has the structure of an abelian group, and
the composition maps
HomA (Y, Z) HomA (X, Y ) HomA (X, Z)
are bilinear.
AB2 There is a (unique) object 0 for which Hom(0, X) = Hom(X, 0) = 0 for all X Ob C
AB3 Binary products and coproducts exist (and coincide) in A
If a category C satisfies only the Axiom AB1, we say that C is a preadditive category. We
also call preadditive categories Z-categories, thinking of them as categories enriched over Z.
Axiom 3 means that for any X1 , X2 Ob A, there is an object Y with morphisms
Xo
i1
p1
/ Y o i2 / X
2
p2
/ X1
/ X1
p2
X2
/0
X2
i2
i1
/Y
2.2
(r) = (r 1) = r (1) = rm. Moreover, for any element m M there exist unique
morphism R M with 1 7 m. Hence, morphism R is completely defined by R (1) R.
By the definition of natural transformation, for any : R R we have the following
commutative diagram
R
R
/R
R
/R
If (1) = s R, then commutativity of the diagram implies rs = sr. Since this is true for
all s R, r = R (1) is an element of the center Z(R).
We want to show that is completely defined by R . Take any R-module M and any
m M . We want to prove that M (m) M is completely defined by R (1). Again, from
the definition of natural transformation, we have commutative diagram
R
R
: 17m
/M
M
: 17m
/
From this diagram we have M (m) = M ((1)) = (R (1)) = rm. This proves the claim.
The opposite direction is easy. Whenever we have an element r Z(R) we can define
r : M M , r (m) = rm. Since
a natural transformation r : idR-Mod idR-Mod by M
M
r Z(R), M will be indeed a morphism of R-modules since (sm) = rsm = srm = s(m).
Any diagram
N
N
/M
/M
will be commutative since for all n N we have N (n) = (rn) = r(n) = M (n).
Example 2.2.3 (Bernstein trace formula). Let V be a finite-dimensional vector space over
a field k, a Endk (V ). We define the trace trV (a) k ' Homk (k, k) to be the composite
k
/ Endk (V )
/ V V a1 / V V h,i / k.
Here End V V V is the inverse of the canonical map vf 7 [x 7 f (x)v]. It is not hard
to check that this agrees with the usual definition of the trace. Let M be any k-vector space
(not necessarily finite-dimensional). Define a linear map trV : Endk (M V ) Endk (M ) by
letting trV (a) be the composite
M
/ M End V
/M V V
73
aidV
/ M V V idM h,i / M .
trV (aM )
/M
trV (aN )
/
/D
Then ( ) F = ( F ) ( F ).
74
/
EE
Since g-Mod ' U(g)-Mod, we have EndFun(A) (idA ) ' Z(U(g)) By the Poincare-BirkhoffWitt theorem, we can write U(g) = U(h) (n+ U(g) + U(g)n ), so we have a canonical
projection : U(g) U(h) = C[h ]. Here h is a Cartan subalgebra of g. Denote by W the
corresponding Weyl group.PThen W acts on h via the dot action, i.e. (w, ) 7 w =
w( + ) , where = 12 R+ is one-half the sum of the positive roots.
Theorem 2.2.7 (Chevalley). The map |Z(g) : Z(U(g)) C[h ]W is a ring isomorphism.
For the next theorem, we need to set up some notation. We set P (V ) = { h : V =
6 0},
FV
P (V ) (f )
.
(3.1)
hand,
from Lemma 2.2.4 it equals Tr(z|V V ). It is well-known that V V is isomorphic to
L
V+ , where sum is taken over weights P (V ) with multiplicities. Hence
Tr(z|V V ) =
Tr(z|V+ ) = c
f ( + ) = c (P (V ) (f ))().
2.3
Morita theory
For an associative unital ring A, Mod(A) denotes the category of all right (unital) A-modules,
and A-Mod denotes the category of all left A-modules.
Definition 2.3.1. Let A and B be (possibly noncommutative) rings. We call A and B
M
2.4
Sh(Z)
j!
/ Sh(X)
[
/ Sh(U )
i!
lim F (V ).
V f (U )
open
j! j
/ id
/ i i
/ i i!
/ id
/ j j
/0
Later on, we will extend these sequences to exact triangles in the derived category.
Example 2.4.1 (projective plane). Let X = P2C be the projective plane over C. Write
X = Proj C[x, y, z], and let Z = {z = 0} = Proj C[X, Y ] be the line at infinity. The
injection i : Z , X is induced by the graded homomorphism C[x, y, z] C[x, y] that sends
f for some graded C[x, y, z]-module
z to 0. If F is a quasicoherent sheaf on X, then F = M
M = M/z.
] The sequence M i i M 0 is obviously exact,
M . It turns out that i F = ig
so sheafifying we get the exact sequence F i i F 0.
Definition 2.4.2. An additive category A is said to be a recollement of A0 and A00 if there
exist six functors
i
AW 0
j!
/A
W
/ A00
i!
2.5
Kan extensions
/E
?
D
A typical example is as follows. Let S Mor C and T Mor E be classes of morphisms. We
can construct categories C[S 1 ] and E[T 1 ], called the localizations of C and E at S and T .
79
/GK
F
O
LanK F K
A left Kan extension of F : C E along K : C D represents the functor E C (F, K) :
Fun(D, E) Set that sends G : D E to the set HomFun(C,D) (F, G K). In other words,
there is a natural isomorphism
HomFun(D,E) (LanK F, G) = HomFun(C,E) (F, G K).
Let Cat be the category of all categories. A functor K : C D is just a morphism in
Cat. It induces, for any E, a functor K : E D E C given by G 7 G K. The functor
LanK : E C E D is the left adjoint to K .
Dually, we can define right Kan extensions.
Definition 2.5.2. The right Kan extension of F along K, written RanK F , is a functor
RanK F : D E with a natural transformation : RanK (F ) K F which is universal
among all pairs (G : D E, : G K F ), in the sense that for any such pair there exists
a unique : G RanK F such that the following diagram commutes:
GK
/ RanK F K
'
/ Top
;
||
Set
It is a good exercise to check that LanY F = ||. By Lemma 2.5.3, the natural transformation
F | | Y is an isomorphism, i.e. n = |n |.
81
82
Chapter 4
Abelian categories
Additive categories
Recall a category A is called a pre-additive (or Z-) category if it satisfies the axiom AB1
below. We say that A is additive if A is preadditive, and also satisfies AB2 and AB3.
[AB1] each HomA (X, Y ) is given the structure of an abelian group in such a way that
compositions
HomA (X, Y ) HomA (Y, Z) HomA (X, Z)
are bilinear;
[AB2] A has initial object , terminal object with = ;
[AB3] A has finite products.
We call the zero object, denoted 0. (Note that we can define A = lim( A), where
Xi = lim{I A}
(product)
Ai = lim{I A}
(coproduct)
iI
a
iI
The simplest case is when I = {0, 1}. One obtains products X Y and coproducts X t Y .
Lemma 1.1.1. Let A be an additive category. If X Y exists, then so does X t Y and
X Y ' X t Y (canonically).
83
1.2
Non-additive bimodules
Let R be an associative unital ring, and let R-Mod be the category of left R-modules. Let
R-Bimod be the category of bimodules over R. Both R-Mod and R-Bimod are abelian
categories. Define F(R) to be the full subcategory of R-Mod consisting of objects isomorphic
to Rn for some n N. Better yet, we could consider P(R), the category of finitely generated
projective R-modules.
Definition 1.2.1. A non-additive bimodule is just a functor T : F(R) R-Mod.
Set F (R) = Fun(F(R), R-Mod). For example, the inclusion F(R) , R-Mod is a nonadditive bimodule. There is a canonical functor : R-Bimod F (R) defined by the rule
M 7 (M : Rn 7 M n ). Note that F (R) is an abelian category.
85
Theorem 1.2.2. The functor is fully faithful, with essential image the full subcategory
of F (R) consisting of additive functors F(R) R-Mod.
Proof. We give a construction showing that the image of consists of additive functors. If
T F (R) is additive, define M = T (R). By definition, M is a left R-module. The right
R-module structure on M is defined by a ring homomorphism : R EndR (M ). Note
that R = EndR (R), where R is treated as a left R-module. If we identify R with EndR (R),
then the fact that T is a functor yields a homomorphism
EndR (R) EndR (T (R)) = EndR (M )
hence M is an R-bimodule.
Remark 1.2.3. It turns out that Hochschild (co)homology can be extended to the category
of non-additive bimodules, yielding topological Hochschild (co)homology. Details can be
found in [BL04].
Theorem 1.2.4. The functor : R-Bimod F (R) has both a left and right adjoint.
Proof. Write for the left adjoint, and ! for the right adjoint of . We will construct
and ! directly. Given X Ob(R-Mod), define six natural morphisms i : X X X,
di : X X X for i {0, 1, 2}. We have
0 (x) = (0, x)
1 (x) = (x, x)
2 (x) = (0, x).
Assume X Ob(F(R)). Then i and di are elements of Mor(F(R)). Define, for any
T Ob(F (R)),
X (T ) = T ( 0 ) T ( 1 ) + T ( 2 ) : T (X) T (X X)
dX (T ) = T (d0 ) T (d1 ) + T (d2 ) : T (X X) X
It is easy to check that for any T ,
X 7 Ker( X (T ))
X 7 Coker(dX (T ))
are additive functors. Tate X = R, viewed as a left R-module. We define
! (T ) = Ker( R (T ) : T (R) T (R) T (R))
(T ) = Coker(dR (T ) : T (R) T (R) T (R)).
We claim that this definition actually gives left and right adjoints to . That is, for any
M R-Mod and T Ob(F (R)), we have isomorphisms
HomR-Bimod ( (T ), M ) ' HomF (R) (T, M R ) = HomF (R) (T, M )
HomR-Bimod (M, ! (T )) ' HomF (R) (M R , T ) = HomF (R) ((M ), T ).
86
/ / (T )
1.3
Abelian categories
Recall that additive categories are categorized by some basic axioms (see 1.1).
Definition 1.3.1. An additive category A is abelian if it satisfies an extra axiom AB4. We
say A is a Grothendieck category if in addition it satisfies AB5 (see below).
87
Before we can state the extra axioms, we need to define kernels and cokernels in arbitrary
additive categories. Let A be an additive category, : X Y a morphism in A. Consider
the functor Ker() : A Ab defined by
Z 7 KerAb ( : HomA (Z, X) HomA (Z, Y )).
If Ker() is representable, then its representing object is called the kernel of denoted by
Ker(). If Ker() exists, we have a (by definition) an exact sequence
0
/ HomA (Z, X)
/ HomA (Z, Y ).
A similar definition for Coker() does not work (the obvious analog of the functor above
is not representable). If we take Z = K, then (idK ) is a morphism k : K X. Hence
Ker() is represented by the pair (K, k : K X). The kernel has a much easier definition.
Let I be the category { }. A diagram of shape I is just a diagram
X0
0
1
// X .
1
/ X0
0
1
// X .
1
such that 0 f = 1 f . The equalizer of the diagram is the limit of the corresponding functor
I A. It is a good exercise to check that our definition of the kernel is equivalent to letting
Ker() be the equalizer of the diagram
// Y .
A naive definition of the cokernel would be to look at the functor Z 7 Coker(hX (Z)
hX (Z)). But this does not agree with the classical definition for abelian groups. In fact, this
functor is not usually representable. Indeed, suppose this functor is represented by some C.
Then we would have exact sequences
Hom(Z, X)
/ Hom(Z, Y )
/ Hom(Z, C)
/0
for all Z Ob A. These sequences are not all exact even if A = Ab. For example, let
X = Y = Z, and let be multiplication by n. If C = Z/n, then the sequence is
Hom(Z, X)
/ Hom(Z/n, Z)
/ Hom(Z/n, Z)
/0
which is not exact. (Strictly speaking, this only shows that C = Z/n does not work.)
So let : X Y be a morphism. The correct definition is the following.
88
Definition 1.3.2. Coker() is the representing object (if it exists) of the functor Coker() :
A Ab defined by
/
c
/Z
O
K0
// Y .
The axiom defining an abelian category is due to MacLane and Grothendieck. We say
an additive category A is abelian if it satisfies
[AB4] Every : X Y can be decomposed in the following way:
K
/X
/I
/Y
k0
/ K0
where
1. = j i
2. (K, k) = Ker() and (K 0 , k 0 ) = Coker()
3. (I, i) = Coker(k) and (I, j) = Ker(k 0 ).
Lets see what this axiom requires in the case A = Ab. Let : A B be a homomorphism
of abelian groups. Then we can decompose as
Ker()
k /
X
/Y
k0
/ / Coker()
So essentially, axiom AB4 requires that the first isomorphism theorem holds in A.
It was Grothendiecks insight that AB4 is equivalent to the combination of the following
two axioms:
[AB4.1] Ker() and Coker() exist for all ;
[AB4.2] If : X Y is such that Ker() = Coker() = 0, then is an isomorphism.
89
/X
/I
/ I0
/Y
k0
/ K0
Theorem 1.3.7. If C is a small category, A any abelian category, then Fun(C, A) is abelian.
Proof. Let F, G : C A be functors, and let : F G be a natural transformation.
We define Ker() as a pair (K : C A, k : K F ) directly. For X Ob(C), set
K(X) = Ker((X) : F (X) G(X)). On morphisms, given f : X Y , consider the
following commutative diagram:
K(X)
kX
/ F (X) X / G(X)
K(f )
K(Y )
kY
F (f )
G(f )
/ F (Y ) Y / G(Y )
`X
/ I 0 (X) = Ker(Coker X )
1.4
In what follows, we will deal mostly with abstract abelian categories. How are we to think
of these? The basic idea is that any general statement involving only finitely many objects
and morphisms is true in any abelian category, if and only if it is true in a module category.
This is justified by the following theorem:
Theorem 1.4.1 (Mitchell). If A is an abelian category, there is an associative unital ring
R and a fully faithful exact functor F : A Mod(R).
Remark 1.4.2. In other words, Theorem 1.4.1 says that every abelian category can be
thought of as being a full exact subcategory of some module category. Though, we prefer
not to think about abelian categories this way. For example, the category Qcoh(X) of
quasi-coherent sheaves on a projective scheme is abelian, but there is no obvious way to
embed it into Mod(R) for some R. Moreover, this is not the way one usually thinks about
91
sheaves. Nevertheless, Mitchells theorem can be rather useful in proving facts about general
abelian categories, because viewing objects of abelian categories as modules allows to pick
elements.
Let A be an abelian category. We can define (co)chain complexes and (co)homology
in A, just as in module categories. It is not obvious that the notion of cohomology makes
sense. Suppose we have a chain complex (C , d ), and look at a piece:
Coker(dn )
O
bn+1
Cn
dn
an
/ C n+1
O
$
%
/ C n+2
dn+1
Ker(dn+1 )
The arrows an , bn are uniquely determined by the universal properties of Ker(dn+1 ) and
Coker(dn ). There is a canonical morphism Coker(an ) Ker(bn+1 ), which is an isomorphism
by axiom AB4. Thus we can define Hn (C) = Coker(an ) ' Ker(bn+1 ).
Thus for any abelian category A, we can define the category of complexes in A, written
Com(A), and it is an easy exercise to show that Com(A) is an abelian category.
Definition 1.4.3. For all n Z, the n-th cohomology is the functor Hn : Com(A) A
defined as above.
We can define homotopies between morphisms in Com(A) the same way we did that
earlier for complexes of abelian groups. Homotopic morphisms induce the same morphism
on cohomology.
Remark 1.4.4. There is much more general notion of a homotopy between two morphisms
in a category. Namely, a notion of a homotopy in model categories. We will discuss model
categories later.
Let A , B , C be objects in Com(A). It is easy to check that a sequence
0
/ A
/ B
/C
/0
is exact if and only if it is point-wise exact. Let Exc(A) be the category of short exact
sequences in Com(A). Objects of Exc(A) are short exact sequences as above, and morphisms
are commutative diagrams:
0
/ A
1
/ A
2
f2
/ B
1
/ B
2
92
/ C
g2
/0
/ C
2
F n ( , , ) = Hn ( )
Gn (A B C ) = Hn+1 (C )
Gn ( , , ) = Hn ( )
..
.
/ An
fn
/ Bn
dn
A
..
.
/ Cn
gn
dn
B
/0
dn
C
n+1
n+1
/ B n+1 g
/ C n+1
/ An+1 f
dn+1
A
/ An+2
f n+2
dn+1
B
/ B n+2
g n+2
dn+1
C
/ C n+2
..
.
/0
/0
..
.
..
.
We use the Mitchell embedding theorem and work as through everything were modules. Choose
c C n such that dcC (n) = 0 (i.e. [c] Hn (C )). Since g n is epi, there exists bn B n
such that g n (b) = c. Note that g n+1 (dnB b) = dnC g n (b) = 0, so dn b Ker g n+1 , which is the
image of f n+1 . It follows that there exists a An+1 such that f n+1 (a) = b. We claim that
n+1 b) = dn+1 (b) = 0, whence dn+1 a = 0 since f n+2
da = 0. Indeed, f n+2 (dn+1 a) = dn+1
B (f
n
is injective. Set (c) = a.
Theorem 1.4.6. For any short exact sequence of complexes
0
/ A
/ B
/ C
/0
()
the sequence
/ Hn (A )
Hn (f )
/ Hn (B )
Hn (g)
/ Hn (C )
n (f ,g )
/ Hn+1 (A )
()
is exact.
The long exact sequence () is functorial in (). So ()7() is a functor Exc(A)
Com(A). Theorem 1.4.6 has a number of useful consequences.
93
/ X1
/Y1
f1
/ X2
f2
g1
/ Z1
/Y2
/0
g2
/ Z2
/0
Assume the rows are exact. Then there are two exact sequences with a connecting snake
(natural in the sequences):
0
Ker
Ker
Ker
Coker
Coker
Coker
Proof. Add zeros and think of the vertical sequences as complexes as in:
0
/ X1
f1
/ X2
f2
/Y1
/ Z1
/0
/Y2
g1
g2
/ Z2
/0
f1
Y1
/ X2
f2
/ Y2
/ X3
f3
/ Y3
/ X4
f4
/ Y4
/ X5
f5
/ Y5
Assume the rows are exact, f1 is epi, f5 is mono, and that f2 and f4 are isomorphisms.
Then f5 is an isomorphism.
94
/ X1
/ Y1
/ Z1
/0
/ X2
/ Y2
/ Z2
/0
/ X3
/ Y3
/ Z3
/0
If the columns and middle row are exact, then if either the first or last row is exact, so is
the other.
Proof. This is also an easy consequence of Theorem 1.4.6.
1.5
Exact functors
/Y
/Z
/ 0,
/ FX
/ FY
/ FZ
We say F is right exact if the analogous sequence (with 0 on the right) is exact. We say F
is exact if it is both left and right exact.
As motivation, consider the classical Riemann-Roch problem. Let X be a topological
space, and let F be an interesting sheaf on X. One is usually interested in the global
sections of F . Often the sheaf F can be decomposed via a short exact sequence:
0
/ F1
/F
/ F2
/0
If the functor F 7 (X, F ) = F (X) were exact, we would have a short exact sequence of
abelian groups:
/ F1 (X)
/ F (X)
/ F2 (X)
/0
0
In the classical setting, F is a sheaf of complex vector spaces, and we are only interested
in dimC (X, F ). The short exact sequence of global sections would give dim (X, F ) =
dim (X, F1 ) + dim (X, F2 ). Unfortunately, the functor (X, ) is almost never exact.
95
1.6
/ Hom(X 00 , G(Y ))
/ Hom(X, G(Y ))
/ Hom(F X 00 , Y )
/ Hom(X 0 , G(Y ))
/ Hom(F X, Y )
(F f )
/ Hom(F X 0 , Y )
Y 7 Ker(Hom(F X, Y ) Hom(F X 0 , Y ))
hence (F X 00 , F g) ' Coker(F f ). By definition, this means the sequence
F X0
/ FX
/ F X 00
/0
In other words, left adjoints are right exact, and right adjoints are left exact.
Corollary 1.6.2. If F : A B is an additive functor between abelian functors that has
both left and right adjoints, then F is exact.
The converse of this theorem is not true there are functors that are exact but have no
adjoints.
Example 1.6.3. Let X be a topological space, U X an open subset. Let Sh(X) be the
category of abelian sheaves on X, and consider the functor (U, ) : Sh(X) Ab given by
(U, F ) = F (U ). We claim that (U, ) is exact. Indeed, let PSh(X) be the category of
presheaves on X. Recall that the forgetful functor i : Sh(X) , PSh(X) has a left adjoint
denoted ()+ : F 7 F + , were
P + (U ) = {s : U Et(P ) : s = idU }
Here Et(P ) is the total space of P and : Et(P ) X is the canonical projection.
Next, define ZU Ob(PSh(X)) by
(
0 if V U =
ZU (V ) =
Z if V U 6=
The restriction maps are the obvious ones. For any presheaf P on X, we have P (U ) =
HomPSh(X) (ZU , P ). It follows that (U, ) == HomPSh(X) (ZU , ) i, so (U, ) is the
composite of two left-exact functors, hence (U, ) is left-exact.
Example 1.6.4. Let X be a topological space, Z X a closed subspace, and let U = X \ Z
be the complement of Z. Write i : Z , X and j : U , X for the canonical inclusions.
Recall that there is a diagram:
i
Sh(Z)
j!
/ Sh(X)
[
/ Sh(U )
i!
where (i , i = i! , i! ) and (j! , j ! = j , j ) are adjoint triples. It follows that i! and j are left
exact, i and j! are right exact, and i , j are exact.
Example 1.6.5. Let R, S be rings, A = Mod(R), B = Mod(S). Let B be a (R, S)-bimodule.
Then we have an adjoint pair:
R B : Mod(R) Mod(S) : HomS (B, ).
so R B s right exact. The converse is also true.
97
Theorem 1.6.6 (Watt). Let R, S be rings, and let f : Mod(R) Mod(S) be an additive
functor that is right exact and commutes with direct sums. Then f (R) has the structure of
an (R, S)-bimodule, and there is a natural isomorphism f ' R f (R).
0
Suppose C, C 0 are
}I
` categories with coproducts and F : C C is a functor. For {X`
Ob C, recall that X = limI X. This coproduct`comes with maps i : X X .
F (X )
F (i )
/F
`
We say that F commutes with coproducts if F (i ) is an isomorphism in C 0 for all collections
{X }I . (Clearly this construction works for arbitrary colimits.)
Proof. Write B = f (R). By definition, B is a right S-module. For each x R, define
x : R R, a 7 x a. The map gives us a ring homomorphism
: R HomMod(R) (R, R).
Since f is additive, the following composite is also a ring homomorphism:
R
/ Hom(R, R)
L
2. Take any free R-module F = I R = RI , where I is a (possibly infinite) index
set. Since f commutes with direct sums,
f
tF : F R B ' B I
/ f (B)I ' f (B I ).
/ F0
F1
/0
where F0 and F1 are free. Then in the following commutative diagram, the top and bottom
rows are exact (by the right-exactness of f and R B), and the latter two vertical arrows
are isomorphisms.
/ F0 R B
F1 R B
tF1
tF0
/ f (F0 )
f (F1 )
/ M R B
/0
tM
/ f (M )
/0
2
2.1
Finiteness conditions
AB5 categories
We would like to define a class of abelian categories which are sufficiently large to have
arbitrary direct sums, but still satisfy some finiteness properties.
Definition 2.1.1 (AB5). We say that an abelian category A satisfies AB5 if
AB5 A has exact (filtered) colimits.
This axiom merits some explanation. Recall that a direct system is just a diagram
Y
Z
indexed by a category that is actually a poset. Let {Xi , X
ij }iI , {Yi , ij }iI and {Zi , ij }iI
be three direct systems, and suppose we have compatible exact sequences
/ Yi
/ Yi
/ Zi
/0
/ lim Xi
/ lim Yi
/ lim Zi
/0
Theorem 2.1.2. For any ring R, the category Mod(R) satisfies AB5.
99
Example 2.1.3. The axioms for an abelian category are self-dual, so for any ring R, the
category Mod(R) is abelian. However, the category Mod(R) is almost never AB5 because
colimits in Mod(R) are just limits in Mod(R), and these are not necessarily exact.
Indeed, let R be a commutative local domain of dimension one with maximal ideal
m R. Consider the exact sequence of inverse systems:
0
...
/ m3
/ m2
/m
/ R/m3
/ R/m2
/ R/m
T
with (unique) maximal ideal m, then
mn = 0.
n=1
fi
/ Yi
gi
/ Zi
/0
/ lim Xi
/ lim Yi
/ lim Zi
/0
provided (Xi , ij ) satisfies the Mittag-Leffler condition: for every n I, there exists n0 > n
such that for all i, j > n0 , we have Im(in ) = Im(jn ).
If an abelian category A does not satisfy AB5, then some pathologies may occur. An
object generated by simple subobjects may not be a direct sum of its simple subobjects (i.e.
it may not be semisimple).
100
2.2
Grothendieck categories
2.3
Let R be a right Noetherian ring. Let mod(R) Mod(R) be the full subcategory of right
Noetherian modules. Similarly, if X is a Noetherian scheme, we have the category coh(X)
as a subcategory of Qcoh(X). It is natural to ask if X and coh(X) determine each other
in a categorical way. Similarly, one could ask if R and mod(R) determine each other. The
answer to this involves the inductive closure of a category.
Definition 2.3.1. A coindex category J is a small nonempty category such that
1. J is connected
2. For all j 0 i j, there is a k with a commutative diagram:
i
/j
/k
j0
For any object X, the functor Hom(X, ) is left exact (see example 1.5.2). So P is
projective if and only if Hom(P, ) is right exact. There is an equivalent, but very useful
definition of projective objects. An object P is projective if and only if for all surjections
: X X 0 and morphisms : P X 0 , there exists a lift : P X of as in the
following diagram:
P
/ X0
/0
/ HomA (P, X)
/ HomA (P, K)
/ HomA (P, X 0 )
/ 0.
/ X0
{
/0
Since F is free,
e has an extension to F , and its restriction |P is the desired extension
of to P .
Finally, we show that if P is projective, then there exists Q such that F = P Q is
free. Indeed, choose a free module F with a surjection : F P . Consider the following
diagram:
P
s
/P
F
id
/0
2.4
Finiteness conditions
i,
/ HomA (X,
X ) .
HomA (X, X )
/ Hom (X,
X ) .
Definition 2.4.4. The object X is coherent if X is finitely presented and every finitely
generated subobject of X is finitely presented.
Example 2.4.5. Let R = khx1 , . . . , xn i be the free algebra on n generators over a field k.
Clearly R is finitely generated (by the unit) as an R-module. There are ideals I R that are
not finitely generated, and the quotient R/I is finitely generated but not finitely presented.
For many purposes, coherent modules over non-Noetherian algebras are the correct
substitute for finitely generated modules over a Noetherian ring.
Lemma 2.4.6. If A is an abelian category satisfying AB5, then every finitely generated
object is compact.
Proof. This is a good exercise.
The converse is false. There is a counter-example due to Rentschler: there exists a
commutative integral domain R such that the field of fractions K of R is compact but not
finitely generated. (Need reference here!)
Lemma 2.4.7. Let A be an AB5 abelian category, and let P Ob(A) be projective. Then
P is finitely generated if and only if P is compact, if and only if P is finitely presented.
104
Proof. We will prove that if P is compact and projective, then P is finitely presented. First,
observe the following. Let I be a directed set, and {(X , f )}I be a direct system of
objects in A. Then limI X is the colimit of the diagram X : I A. Let |I| be the
L
underlying (discrete) set of I. Then lim|I| X = X by definition, with canonical
L
embeddings i : X , X . We have canonical
X ; these
L morphisms j : X lim
patch together to yield a canonical morphism j : X lim X . The morphism j is
(,)S
where S = {(, ) I I : 6 }, f : X X and X(,) = Im i f i .
Assume P is compact projective, and apply HomA (P, ) to the sequence above. We get
L
L
/ Hom P, lim X
/0
/ Hom P,
Hom P, (,)S X(,)
I X
O
O
O
(,)S
Hom P, X(,)
Hom(P, X )
/ lim Hom(P, X )
/0
The first two vertical arrows are isomorphisms because we assumed P is compact. Since P is
projective, rows are exact. Then 5-lemma implies the third vertical arrow is an isomorphism.
So Hom(P, ) commutes with arbitrary direct limits, hence P is finitely presented.
Projective objects are analogs of vector bundles. This analogy can be made somewhat
precise.
Theorem 2.4.8 (Swan). Let X be a para-compact topological space, and let A = C(X)
be the ring of continuous C-valued functions on X. Then the category Vect(X) of vector
bundles on X is equivalent to the category Proj(A) of projective A-modules via the functor
(X, ).
Theorem 2.4.9. Let X = Spec(A) is a Noetherian affine scheme. Then the category
Vect(X) of vector bundles on X is equivalent to the category Proj(A) of projective A-modules
via the functor (X, ).
Lemma 2.4.10. Let F : A B be an exact functor between abelian categories. Then F is
faithful if and only if F (X) 6= 0 whenever X 6= 0.
Proof. Suppose f is faithful (injective on hom-sets). Then X 6= 0 implies idX =
6 0, which
implies F (idX ) 6= 0, hence F (X) 6= 0. (In other words, arbitrary faithful functors send
nonzero objects to nonzero objects.)
Conversely, suppose that for all X 6= 0 we have F (X) 6= 0. Let f be a nonzero morphism
in A. Then Im(f ) 6= 0. Since F is exact, it commutes with taking kernels, cokernels and
images. So F (Im f ) = Im F (f ) 6= 0, hence F (f ) 6= 0.
105
3.1
/ X0
/X
/ X 00
/ 0,
/ Hom(X 00 , E)
/ Hom(X, E)
/ Hom(X 0 , E)
/0
We already know it is exact on the left, so the only nontrivial condition is that Hom(, E)
is right-exact.
Definition 3.1.2. An abelian category A has enough injectives if every X Ob(A) is
isomorphic to a subobject of an injective object in A.
Tautologically, E is injective in A if and only if E is projective in the dual category A .
So it would appear that injective and projective objects are dual in some sense. However,
in practice we usually work with Grothendieck categories, and for A Grothendieck, A is
Grothendieck if and only if A is the zero category. Proving this is a good exercise.
We will see that any Grothendieck category has enough injectives. On the other hand, a
Grothendieck category does not necessarily have enough projectives (or any projectives at
all). A typical example is A = Qcoh(X) for X a projective scheme. This category has no
projective objects.
If A is a Noetherian abelian category, then A as a rule does not have enough invectives.
It is possible for such categories to have enough projectives just take the category of finitely
generated modules over a semisimple ring. In any case, even if A does not have enough
injectives, its inductive closure Ae always have enough injectives.
106
Theorem 3.1.3. For any ring R, the categories Mod(R) and R-Mod have enough injectives.
Q
Lemma 3.1.4. Let {E }I be a set of injectives in A. Assume that the product E = E
exists. Then E is injective.
Q
Proof. Note that E = limI E . For all X in A, we have
Y
Hom(X, E) '
Hom(X, E ).
I
The latter is a product of exact functors, hence exact. It follows that E itself is exact.
Definition 3.1.5. Let R be an associative unital ring. An element x R is called right
regular if the right R-module map r 7 xr is injective on R, i.e. if xr = 0 implies r = 0.
Definition 3.1.6. Let M be a right R-module. We say M is divisible if M x = M for any
right regular x R.
m0
In other words, M is divisible if for any m M and right regular x R, there exists
M such that m0 x = m.
X is
Example 3.1.10. Let R = Z. Then we claim that Mod(Z) = Ab has enough injectives.
Indeed, Q is obviously divisible, and hence injective. By Lemma 3.1.4, any product of copies
of Q is injective. Moreover, Baers theorem shows that any quotient of injectives is injective.
In particular, Q/Z is injective. But any Z-module can be identified with a subquotient of a
product of copies of Q, via
L
Q
/
MZ
MQ
M.
Thus M is isomorphic to a submodule of an injective.
107
Mod(R)
/ Mod(S)
f!
3.2
The following constructions are motivated by topology, but they actually make sense in
much greater generality (i.e. for model categories, as we will see later in the course).
Suspension
Let A be an abelian category and Com(A) be the category of (cohomological) chain complexes
(X , d ).
Definition 3.2.1. Let k Z. The k-th suspension functor [k] : Com(A) Com(A) is
i+k
defined by X 7 X [k], where X [k] = X i+k and diX[k] = (1)k dX
.
Lemma 3.2.2. The functor [k] is an auto-equivalence of categories, with inverse [k].
Moreover, [k] [m] = [k + m].
108
Cone
Well construct the functor cone explicitly. There is a more functorial construction, and we
will describe it later when we will be talking about model categories. Given f : X Y ,
set cone(f ) = Y X [1]. The complex cone(f ) looks like
k
/ Y k X k+1 dcone / Y k+1 X k+2
where
dkcone
=
dkY
0
/ ,
f k+1
.
dk+1
X
k
It is easy to check that dk+1
cone dcone = 0 if and only if
k
dk+1
Y dY = 0
k+1
dk+2
X dK = 0
f k = f k dkX .
dk+1
Y
Proposition 3.2.3. A morphism f : X Y is a quasi-isomorphism if and only if
cone(f ) is acyclic.
Proof. Recall there is a short exact sequence
/Y
/ cone(f )
/ X [1]
/0
/ Hi1 (X[1])
Hi (X)
/ Hi (Y )
8
/ Hi (cone f )
fi
/ Hi (X[1])
Hi+1 (X)
/ Hi+1 (Y )
8
f i+1
The arrows f i and f i+1 are isomorphisms, which is easily seen to imply Hi (cone f ) = 0.
Exercise Let Comb (A) be the full subcategory of Com(A) consisting of bounded complexes
(complexes X with X i = 0 for all |i| 0). The category A naturally embeds into Comb (A).
Show that Comb (A) is generated by A in the sense that every X Comb (A) can be
obtained by taking iterated suspensions and cones of objects in A.
Cylinder
Definition 3.2.4. Let f : X Y be a morphism of complexes. The cylinder of f is
p [1]
X
Cyl(f ) = cone(cone(f )[1]
X ).
109
Explicitly,
Cyl(f ) = Y X X [1]
with differentials
dkY
=0
0
dkcyl
0
dkX
0
f k+1
idX k+1
dk+1
X
/ X
/0
By Proposition 3.2.3, the morphism iX is a quasi-isomorphism if and only if f is a quasiisomorphism, which happens if and only if cone(f ) is acyclic.
The other short exact sequence associated to f : X Y is
/Y
/ Cyl(f )
/ cone(idX )
/ 0.
/Y
/ cone(f )
/ X[1]
/ cone(f )
/0
/0
/ X
/ X
/ Cyl(f )
f
/Y
/ Z
/0
i0
/X I
/ Cyl(f )
3.3
/X
/ I1
/ I0
/ I2
/ P 2
/ P 1
/ P0
/X
/ 0,
X2
fe
/ I
2
/X
111
/0
Reversing arrows, we see that an object I is injective exactly when lifts exist in the following
diagram.
/X
/ X0
0
~
/ Ri F (X 0 )
/ Ri F (X)
/ Ri F (X 00 )
/ Ri+1 F (X 0 )
By Theorem 3.3.6, injective objects are universally acyclic, i.e. they are F -acyclic for
any left-exact functor F .
Definition 3.3.8. An F -acyclic resolution of X Ob(A) is an exact complex X , J ,
where J consists of F -acyclic objects.
Proposition 3.3.9. If X J is an F -acyclic resolution, then Ri F (X) ' Hi (F (J )) for
all i.
Example 3.3.10. Flabby sheaves are acyclic relative to the global sections functor. Thus
we can compute sheaf cohomology using flabby resolutions.
We can define (classical) left-derived functors similarly. Given a right-exact additive
functor F : A B, define its left-derived functors Li F : A B as follows. Given X Ob(A),
choose a projective resolution P X, and define
Li F (X) = Hi (F (P )).
3.4
-functors
Classical derived functors satisfy a universal mapping property. Let A, B be abelian categories.
Definition 3.4.1. A (covariant) -functor is a collection T = (T i )i>0 of additive functors
A B given together with natural transformations i : T i T i+1 , where here T i and T i+1
are functors Exc(A) B via T i (0 X 0 X X 00 0) = T i (X 00 ) and T i+1 (0 X 0
X X 00 0) = T i+1 (X 0 ). We require that for any short exact sequence 0 X 0 X
X 00 0, the following sequence be exact:
0
/ T 0 (X 0 )
/ T 0 (X)
/ T 0 (X 00 )
/ T 1 (X 0 )
Proof. We will prove this theorem by induction. Given any -functor (T, ) : A B and
f0 : T0 E0 , assume we have constructed fi : Ti Ei for i < k.
We want to construct fk : Tk Ek . For this we need to define fk (X) : Tk (X) Ek (X)
for each X A. For each X, choose p : Y X, s.t. Ei (Y ) = 0 for all i > 0. Let
X 0 = Ker(p), so that we have a SES in A
0 X0 Y X 0
Since (T, ) and (E, ) are both -functors, by applying T and E we get
/ Tk (X)
/ Tk1 (X 0 )
fk
/ Ek (X)
k1
k1
k1
fk1 (X 0 )
/ Ek1 (X 0 )
k1
/ Tk1 (Y )
/ Tk1 (X)
fk1 (Y )
/ Ek1 (Y )
/ ...
fk1 (X)
/ Ek1 (X)
/ ...
/ X0
2
/ Y1
/ X1
/ Y2
/0
/ X2
/0
Tk (X1 )
Tk1 (0 )
Tk ()
z
/ Tk (X 0 )
2
Tk (X2 )
fk (X1 )
fk1 (X20 )
Ek (X1 )
fk (X2 )
/ Ek (X 0 )
1
Ek1 (0 )
Ek ()
Ek (X2 )
fk1 (X10 )
z
/ Ek (X 0 )
2
114
In this diagram top and bottom faces commute since T and E are -functors. The
front and back faces commute by construction of the maps fk . The right face commutes
by induction. So we only need to show that the left face commutes. Since the front
face commutes, we get fk (X2 ) Tk () = Ek () fk (X1 ). But E(Y2 ) = 0, so
: Ek (X2 ) Ek1 (X20 ) is injective. Hence fk (X2 ) Tk () = Ek () fk (X1 ).
3.5
The goal of this section is to prove the existence of classical derived functors. Let A be an
abelian category. We assume A has enough projectives, and as before write Com(A) for
the category of chain complexes in A. Recall that if X is a complex in A, its projective
resolution (or approximation) in Com(A) is a quasi-isomorphism P X , where P is
(pointwise) projective.
Lemma 3.5.1. If X Ob(Com(A)) is bounded from above (X i = 0 for i 0), then X
admits a projective resolution.
Proof. The proof is by induction. First, if X i = 0 for i > i0 , put P i = 0 and f i for i > i0 .
Now assume that we have already constructed P i and f i for i > k + 1:
P k+1
/ X k1
/ Xk
dkX
dk+1
P
/ P k+2
f k+1
/ X k+1
dk+1
X
dk+2
P
f k+2
/ X k+2
f k dk+1
P
//Y
/ X k P k+1 .
f k+1
dk+1
P
fk
dk+1
P
115
=0
k
which occurs if and only if dkX f k f k+1 dk+1
= 0 and dk+1
P
P dP = 0. Hence, we can extend
f : P X by
dkP
Pk
/ P k+1
fk
f k+1
/ X k+1
Xk
P
/ Q
/Y
dkP
f k k
dkQ
Qk
dkP
/ P k+1
"
f k+1
/ Qk+1
dk+1
Q
/ P k+2
/ Qk+2
k+1 k+1 k
k
k
k
Put k = f k+1 dkP . Then dk+1
dP = f k+2 dk+1
Q = dQ f
P dP = 0, so Im( )
k
Ker(dk+1
Q ) = Im(dQ ). Hence we have
Pk
fk
Qk
{ dkQ
/ Im(dk )
Q
/ 0.
By the projectivity of P k , there exists f k such that dkQ f k = k = f k+1 dkP . This finishes the
induction.
116
/ P k1
f k1
g k1
}
/ Qk1
dkP
/ Pk
fk
hk
gk
}
/ Qk
dk1
Q
/ P k+1
/ Qk+1
hk+1
dkQ
k
k+1 dk . Put h0 = 0. We argue by induction.
such that for all k, f k g k = dk1
P
Q h +h
Assume that we already have constructed hk+1 , hk+2 , . . . and we want to construct hk .
Consider k = f k g k hk+1 dkP , and note that dkQ k = (f k+1 g k+1 dkQ hk+1 )dkP =
k
hk+2 dk+1
P dP = 0. Consider
Pk
hk
Qk1
/ Im(dk1 )
Q
/0
3.6
/ T1 (A)
/ T1 (A00 )
/ T0 (A0 )
/ T0 (A)
/ T0 (A00 )
/0
Definition 3.6.2. Let F : A B be a right-exact additive functor. The classical leftderived functor of F is a pair (LF, ) consisting of a left -functor LF = (Li F, i )i>0 and
117
an isomorphism of functors L0 F
F , which is universal among all left -functors in the
following sense. For all left -functors (T, ) and morphisms f0 : L0 F T0 , there is a
unique sequence of morphisms fi : Li F Ti extending f0 .
Remark 3.6.3. It follows from the definition that (LF, ) is determined up to unique
isomorphism, if it exists. The functor LF is only determined up to automorphisms of F .
F . Then
Indeed, given (LF, ) and (L0 F, 0 ), by definition : L0 F
F and 0 : L00 F
f0 = (0 )1 is an isomorphism L0 F L00 F , which extends uniquely to an isomorphism
f : LF L0 F .
Previously we defined derived functors using injective (resp. projective) resolutions.
Here, we prove that under various assumptions, this agrees with the above definition.
Theorem 3.6.4. Let F : A B be a right-exact additive functor. Assume A has enough
projective objects. For any X Ob(A) choose a projective resolution P X in Com(A).
Define, for i > 0,
Li F (X) = Hi (F (P ))
Then (Li F )i>0 is the classical left-derived functor of F in the sense of definition 3.6.2.
Proof. Our main tools will be lemmas 3.5.1 and 3.5.2. In the proof we need to do the
following steps.
1. prove that the Li F are actually functors;
2. check independence of Li F on the choice of resolutions (up to isomorphism);
3. prove that Li F actually form a -functor. Namely, we need to show that there exist
connecting morphisms i : Li+1 F (A00 ) Li F (A0 ) satisfying the definition 3.6.1;
0X
Y
Z 0 in A. Replace X and Y by projective resolutions. By Lemma 3.5.2,
/ Q
/ R
/ P [1]
/0
()
of complexes in A.
Remark 3.6.5. Instead of writing the sequence () we could have written an exact distinguished triangle
P
/ Q
/ R
/ ...
Q
i
P o
[1]
/ H1 (Q )
/ H1 (R )
/ H1 (P [1])
o
/ H0 (Q )
/ H0 (R )
/ H1 (P [1])
o
/Y
o
/Z
/ L0 F (X)
/ L0 F (Y )
/ L0 F (Z)
/ L1 F (X)
P1
/ R
h
f1
/ Q
g1
/ R
/X
/Y
/Z
}
/ X1
/ Q
}
/ Y1
f1
/0
g1
/ Z1
/0
We claim that this gives us a morphism of exact triangles. Namely, we have the following
diagram:
P
/ Q
P1
f1
/ R
/ Q
1
i1
/ P [1]
[1]
p1
/ R
1
/ P [1]
1
We need to check that the diagram above commutes up to homotopy. We have already seen
that the first square commutes up to homotopy. Lets check the second square.
Qk
Qk1
ik
/ Qk P k+1
q_k
/ (q k , 0)
_
ik
/ Qk P k+1
1
1
k (q k )
/ ( k (q k ), 0)
(4.1)
/ Li+1 F (Z)
/ Li F (X)
/ Li F (Y )
/ Li F (Z)
/ Li+1 F (Z)
/ Li F (X)
/ Li F (Y )
/ Li F (Z)
120
This exactly means functoriality on short exact sequences. Thus, (Li F, ) is a left
-functor.
Next we need to prove existence of natural isomorphism L0 F ' F . Indeed, choose any
X A. Replace X by resolution P X and look at the first two terms
P 1
/ P0
/X
/0
Applying right exact functor F to this exact sequence, we get an exact sequence
F (P 1 )
F (d)
/ F (P 0 )
/ F (X)
/0
Then L0 F (X) := H0 (F (P )) = Coker F (d) ' F (Coker(d)) ' F (X). Notice that here we
used right exactness of F again, which gave Coker F (d) ' F (Coker(d)).
Finally, we need to check the universal property.
Lemma 3.6.6. If A has enough projectives then the functor Li F, is coeffaceable.
Proof. Given X A we can replace it by a projective cover P X for some projective
P A. Then we claim that Li F (P ) = 0, i > 0. But this is obvious, since P is projective,
so we can choose its projective resolution to be the complex P = [0 P 0].
Having proved lemma 3.6.6 we can apply theorem 3.4.4 to automatically get the universality of classical derived functors.
3.7
The functors Exti are classical derived functors of Hom : (X, Y ) 7 Hom(X, Y ). We need to
make this precise, because we can think of Hom(, ) as a functor in either variable. We
will assume that A is an abelian category with enough injectives and enough projectives.
Then, for X, Y Ob(A), consider the functors
Hom(, Y ) : A Ab
Hom(X, ) : A Ab
Both these functors are left-exact, so we can define
Exti1 (X, Y ) = Ri Hom(, Y )(X) = Hi (Hom(P , Y ))
Exti2 (X, Y ) = Ri Hom(X, )(Y ) = Hi (Hom(X, I )),
where P X is a projective resolution of X and Y I is an injective resolution of Y .
Theorem 3.7.1. There is a natural isomorphism Exti1 (X, Y ) ' Exti2 (X, Y ).
121
We prove this by giving another (more explicit) construction of Ext, due to Yoneda.
Lets start with i = 1. Fix X and Y in A, and consider the set E 1 (X, Y ) of all short exact
sequences
= (0 Y Z X 0)
Define an equivalence relation on E 1 (X, Y ) by setting 0 if there is a morphism
: Z Z 0 in A such that the following diagram commutes:
/Y
/Z
/X
/0
/X
/0
/Y
/Z
The snake lemma forces to be an isomorphism. Define Ext1 (X, Y ) = E 1 (X, Y )/ . The
additive structure on Ext1 (X, Y ) is defined by + : Ext1 (X, Y ) Ext1 (X, Y ) Ext1 (X, Y ).
Given 1 = (0 Y Z1 X 0) and 2 = (0 Y Z2 X 0), we define 1 + 2
as follows. There is an exact sequence 1 2
0
/Y
O
/Z
O
/X
/0
e
/Z
/X
/0
/Y Y
/Y Y
/ Z1 Z2
/X X
/0
e = X XX (Z1 Z2 ). We define Z
Here Ze is the pullback in the lower-right square, i.e. Z
to be the pushout in the upper-left square, i.e.
Z = Y tY Y Ze = (X XX (Z1 Z2 )) tY Y Y .
The sequence 1 + 2 is 0 Y Z X 0. We need to check that + is compatible
with . The unit is the trivial split-exact sequence 0 Y Y X X 0.
In general, for i = k > 1, we define E k (X, Y ) to be the set of exact sequences of the form
= 0 Y Zk Zk1 Z1 X 0 .
{z
}
|
Z
/Y
/Y
/ Zk
/ Zk1
/ Z0
k
/ Z0
k1
k1
122
/ Z1
/X
/0
/X
/0
/ Z0
1
(n)
Z0 .
Z Z Z
More formally,given X, Y Ob(A) and k > 1, define the following category Extk (X, Y ) with
objects complexes of length k having homology
Y if i = k
Hi (Z ) = X if i = 1
0 otherwise
Morphisms in Extk (X, Y ) are morphisms : Z Z0 of chain complexes satisfying H1 ( ) =
idX , H1 ( ) = idY . We could then define Extk (X, Y ) = E k (X, Y ) = 0 (Extk (X, Y )).
(the definition using Ext does not work when k = 1.)
By convention, Ext0 (X, Y ) = HomA (, Y ).
Lemma 3.7.2. The Yoneda construction of Extk defines a bifunctor, covariant in Y and
contravariant in X, i.e. Extk : A A Ab.
p1
/Y
/ Z0
k
/ Z0
2
/ Z0
1
/ X0
/0
/Y
/ Zk
/ Z2
/ Z1
p1
/X
/0
/ V
/ Z0
1
Z0
/0
/ V
/ Z1
/X
/0
/ Extk1 (X 00 , Y )
k1
/ Extk1 (X 0 , Y ) / Extk (X 00 , Y )
/ Extk1 (X, Y )
00
Given Z0 E k (X 0 , Y ), define Z100 = X, and Zi+1
= Zi0 for i = 1, . . . , k 1. We claim that
Z00 E k (X 00 , Y ), and that this assignment is compatible with the equivalence relation on
E k (X 0 , Y ).
Exercise Let A = Ab. Prove that ExtiA (X, Y ) = 0 for all i > 2, where X and Y are
arbitrary abelian groups.
Section 3 in the Appendix A demonstrates an application of the classical derived functors
to representation theory of quivers.
124
Chapter 5
Derived categories
1
1.1
Localization of categories
Motivation
We start with an example motivating the use of derived categories. Let k be an algebraically
closed field of characteristic zero. Consider the projective line P1k over k, and let A = coh(P1k )
be the category of coherent sheaves on P1k . Let E be the object OP1 OP1 (1) in A. One
usually calls E the tilting sheaf.
It is easy to compute EndA (E ):
EndA (E ) = HomA (O O(1), O O(1))
HomA (O, O)
HomA (O, O(1))
'
HomA (O(1), O) HomA (O(1), O(1))
k k 2
'
0 k
' k[ ].
Write B = EndA (E ) for this ring.
Note that if F coh(P1 ), then HomA (E , F ) is naturally a right B-bimodule, with the
action of B = HomA (E , E )
HomA (E , F ) HomA (E , E ) HomA (E , F )
coming from composition. Thus Hom(E , ) induces a functor
F : coh(P1 ) Mod(kQ) ' Repk (Q ),
where Q is the quiver and Repk (Q ) is the category of its representations (see Section
3 in the Appendix A on quivers).
125
The functor F has a right-adjoint G, induced by tensoring with E . One might hope
that the functor F is an equivalence of categories, but this is not the case. For example, let
F = OP1 (1). Then F (F ) ' (X, O(1)) (X, O(2)) = 0, but F 6= 0.
Lets compute Ext1 (E , F ). This is, by definition Ext1 (E , O(1)), which is isomorphic
by Grothendieck-Serre duality (Reference here?) to
Ext1 (E (1), O(2)) ' H0 (X, E (1)) = (O(1) O) ' (O) ' k 6= 0.
So we see that even though F (OP1 (1)) = 0, the derived functor RF (OP1 (1)) = Ext1 (E , O(1))
is non-zero, allowing us to retain some more information.
Theorem 1.1.1 (Beilinson, [Bei78]). There is an equivalence of (triangulated) categories
R Hom(E , ) : Db (Coh P1 ) Db (Repk Q ).
Here Db (Repk Q ) denotes the derived category of the category of modules over the
path algebra kQ. The derived category of an abelian category A can be defined as a
category obtain from the category Com(A) of complexes in A by formally inverting all
the quasi-isomorphisms. More precisely, we have the following definition.
Definition 1.1.2. Let A be an abelian category. Let S be the class of all quasi-isomorphisms
in Com(A). Then the derived category D(A) of A is the localization D(A) = Com(A)[S 1 ]
of A at S.
To explain what this definition actually means, we now begin systematic study of
localization of categories.
1.2
Definition of localization
Remark 1.2.8. The two-out-of-three property can be strengthened to give the so-called
two-out-of-six property (see [DHKS04]) which is defined as follows. Given any triple f, g, h
of composable morphisms, then in the diagram
hgf
gf
hg
'/
if gf, hg S, then f, g, h and hgf are in S. As an exercise, try showing that the collection
of quasi-isomorphisms in Com(A) satisfies this property.
1.3
Calculus of fractions
with s S. We call such diagrams left S-fractions (resp. right S-fractions.) We think
of the diagram on the left as an avatar for s1 f , and the diagram on the right as an avatar
for f s1 , and whenever it wont lead to confusion we will denote these diagrams simply
by s1 f and s1 f .
Consider a chain
f1
> Z 1 ` s 1 f2 > Z 2 ` s 2
Y1
Y2
fn
; Z n a sn
Yn1
Yn
1
1
which we formally denote by s1
n fn s2 f2 s1 f1 . Such diagram is called composit
S-fraction.
Definition 1.3.1. Call two composite fractions elementary equivalent if one can be obtained
from the other by one of the following rules
(E1) adding (inserting inside the chain) s1 s or s s1 with s S
(E2) replacing g id1 f by g f
f
> Z _ id g > Y
gf
X
128
>Y
1
1
(E3) replacing s1
1 id s2 by (s2 s1 )
X`
s1 id
? Z ` s2
X`
s2 s1
We say that two composite fractions are equivalent if there is a chain of elementary equivalences relating them.
Definition 1.3.2. Let C be a category, S Mor(C) a saturated class of morphisms. We
define C[S 1 ] to be the category with the same objects as C, and with
HomC[S 1 ] (X, Y ) = {equiv. classes of compos. S-fractions from X to Y }.
Composition is induced by concatenation.
Remark 1.3.3. There is a subtle problem with this definition. Namely, HomC[S 1 ] (X, Y )
might be a proper class and not a set.
Proposition 1.3.4. If C is a small category and S Mor(C), then the category C[S 1 ]
defined in 1.3.2 is indeed the localization of C at S.
Proof. Define a functor Q : C C[S 1 ] that is the identity on objects, and sends a morphism
f : X Y to the formal fraction id1
Y f . It is easy to see that Q is a functor; we have
id1 (f g) (id1 f ) (id1 g).
Suppose we have a functor F : C D with F (S) Iso(D). We define Fe : C[S 1 ] D
by Fe(QX) = F (X) and
1
1
1
Fe(s1
F (f1 ).
n fn sn1 s1 f1 ) = F (sn )
f1
> Z 1 ` s1 f 2 > Z 2 ` s2
Y1
129
Y2
=Z a t
> Z 1 ` s 1 f2 > Z 2 ` s 2
Y1
Y2
(ts2 )1 (tf2 ) s1
1 f1
= (ts2 )1 (gs1 ) s1
1 f1
(ts2 )1 (g id1 f1 )
(ts2 )1 (gf1 ).
Remark 1.3.6. If C has a null object (i.e. an object 0 which is both initial and terminal)
then the corresponding object 0 in C[S 1 ] is a null object in the localized category. To see
this, we need to show that HomC[S 1 ] (0, ) and HomC[S 1 ] (, 0) are both constant functors
with value {}. This is easy.
Example 1.3.7 (Universal (Cohn) localization of rings). Let R be a unital ring. Let be
a set of morphisms : P Q, where P, Q range over some finitely-generated projective
(left) R-modules. The universal localization of R at is a ring R[1 ] together with
a homomorphism : R R[1 ] such that that the corresponding induction functor
: R-Mod R[1 ]-Mod sends all to isomorphisms.
Theorem 1.3.8 (Bergman). If R is hereditary (see definition 3.4.16 in the Appendix A),
then all the universal localizations R[1 ] are also hereditary. Moreover, universal localization
1
1
is pseudo-flat in the sense that TorR
1 (R[ ], R[ ]) = 0.
1.4
Ore localization
Definition 1.4.1. We say that S Mor(C) satisfies the left Ore conditions if
(LO1) S is saturated;
(LO2) for any right S-fraction g t1 : X Y , there exist a left S-fraction s1 f such
that f t = sg;
(LO3) if f s = gs with f, g : X Y and s : Z X, s S, then there exists t S such
that tf = tg
Z
/X
f
g
// Y
130
// Y
/W
f1
> Z1 `
> Z2 `
f2
s1
s2
f1
> Z1 h
/6 Z 2
`
s2
s1
f2
Remark 1.4.4. A more standard way to define an equivalence relation on left fractions
(under the assumption that S is multiplicatively closed, but not necessarily saturated) is via
the following commutative diagram:
f
f1
>Z `
4 Z2 `
> Z1 j
s2
s1
f2
If S is saturated, then this definition is equivalent to the definition 1.4.3 above. All
interesting examples of localizations one meets in real life are localizations over a saturated
class of morphisms, so our definition 1.4.3 is usually enough. For example, quasi-isomorphisms
in the category of complexes in an abelian category is a saturated class.
131
/ Hom 1 (X, Y )
C[S ]
{left S-fractions}/ `
We begin by showing that the map defined in the obvious way as a map
{left S-fractions X Y } HomC[S 1 ] (X, Y )
actually descends to the quotient {left S-fractions}/ ` .
1
If s1
1 f1 ` s2 f2 , then there exists t S such that s2 = ts1 and f2 = tf1 . We then
compute:
1
1
s1
f1
1 f1 s1 id id
1
s1
t id1 f1
1 id t
(ts1 )1 (tf1 )
= s1
2 f2 .
Now we construct the map inverse to . Take a composite fraction a = s1
n fn
1
f1 . Using (LO2), we replace each right fraction fi s1
with
s
f
i1,i . Similarly,
i1
i1,i
1
1
we can replace fi,i+1 si1,i with si1,i+1 fi1,i+1 . For n = 4, the following commutative
diagram illustrates the described process:
s1
1
f1,4
f1,3
f1,2
f1
> Z1 b
Z
< 1,3 b s1,3 f2,4
Z
< 1,2 c s1,2 f2,3
f2
s1
W1
Z
< 1,4 b s1,4
; Z2 c
Z
< 2,4 b s2,4
Z
; 2,3 c s2,3 f3,4
f3
s2
W2
; Z3 c
Z
; 3,4 b s3,4
f4
s3
W3
< Z4 `
s4
In general, let f be the composite f1,n f1,n f1 and let s = s1,n sn1,n sn .
We would like to define by (a) = s1 f .
132
g
.
We
have
f
s
=
s
f
and
g
s
=
t
f
.
Consider
the
right
fraction
s1,2 t1
1,2
1,2
1
1,2 2
1,2 1
1,2 2
1,2
1,2
and apply (LO2) to find a left fraction u1 v = s1,2 t1
.
Then
us
=
vt
.
Since
S
is
1,2
1,2
1,2
saturated, us1,2 = vt1,2 implies v S. Multiply g1,2 s1 = t1,2 f2 on the left by v. We get
vg1,2 s1 = vt1,2 f2
= us1,2 f2
= uf1,2 s1 .
By (LO3), there exists w S such that wvg1,2 = wuf1,2 . Hence
(s1,2 s2 )1 (f1,2 f1 ) ` (wus1,2 s2 )1 (wvf1,2 f1 )
` (wvt1,2 s2 )1 (wvg1,2 f1 )
` (t1,2 s2 )1 (g1,2 f1 ).
1 (tf ) does not change the value
It remains to check that replacing s1
i
i fi with (f si )
of , and also that and are actually inverses of each other. The proof is a (tedious)
calculation similar to the one weve done above. We leave it as an exercise to the interested
reader.
2.1
Serre quotients
0
0
(M ,N )
the limit being taken over all pairs (M 0 , N 0 ) such that M 0 M and N 0 N satisfying
M/M 0 , N 0 Ob(T ).
Lemma 2.1.10. This definition of the Serre quotient makes sense.
Proof. Let I = {(M 0 , N 0 ) : M 0 M, N 0 N, M/M 0 , N 0 Ob(T )}. We put (M 0 , N 0 ) 6
(M 00 , N 00 ) if M 00 M 0 and N 0 N 00 . The set (I, 6) is directed because whenever
(M10 , N10 ), (M20 , N20 ) I, the pair (M10 M20 , N1 + N2 ) is also in I. The HomA (M 0 , N/N 0 )
form a directed system as follows. If i : M 00 , M 0 , j : N 0 , N 00 are injections, we get
Hom(M 0 , N/N 0 )
/ Hom(M 00 , N/N 0 )
/ Hom(M 00 , N/N 00 )
135
/ Qcoh(X).
Write Tails(A) for the quotient category appearing in this theorem. If M Tails(A) is
finitely-generated, then it turns out that one has HomTails (M, N ) = lim Hom(M/M>n , N )
for any N .
There is a projection functor : A A/T , that assigns to M A the same object M ,
and that sends f to the induced morphism f .
Lemma 2.1.12. An object M is zero in A/T if and only if it M Ob(T ).
Proof. If M = 0 in A/T , then HomA/T (M, M ) = 0, whence idM = 0. Thus there exist
id
/L
/ L
/M
/ M
136
/N
/0
/ N
/0
2.2
Injective envelopes
Let A be an abelian category. Recall (see definition 3.1.1 in Chapter 4) that an object E in
A is injective if the functor Hom(, E) : A Ab is exact. Alternatively, arrows lift as in
the following diagram:
0
/M
/N
~
The category A has enough injectives if for any M in A, there is a monic M , E with E
injective.
For convenience we will think of the category A as a subcategory of modules over some
ring. Mitchells theorem 1.4.1 in Chapter 4 guaranties that we are not losing generality.
Definition 2.2.1. Let M N be objects in A. We say that M is essential in N if for all
N 0 M with N 0 6= 0, the intersection N 0 M 6= 0.
We say that a monic f : M N is essential if Im(f ) N is essential.
Theorem 2.2.2 (Eckmann-Schopf,[ES53]). An object E is injective if and only if E has no
proper essential extensions.
Definition 2.2.3. An injective envelope of an object M Ob(A) is any essential injective
extension of M .
Example 2.2.4. Let A be a simple Noetherian hereditary domain (for example, a Dedekind
domain or the Weil algebra khx, yi/([x, y] = 1)). Goldies theorem gives existence of the
field of fractions Q = Frac(A), the localization of A at the (Ore) set A r 0. It turns out that
A , Q is an injective envelope of A.
Theorem 2.2.5 (Baer,[Bae40](not sure if the correct reference)). Let E be an injective
envelope of M with respect to the inclusion i : M , E. Then
1. for any essential monic f : M N , there exists monic g : N , E such that gf = i
2. for any injective extension j : M , E 0 , there exists g : E E 0 such that j = gi.
Corollary 2.2.6. An injective envelope of M is a maximal essential extension of M .
137
2.3
Localizing subcategories
The Serre subcategory T appearing above need not be closed under taking arbitrary (infinite)
direct sums. For example, one can consider Vectfd
k as a subcategory of Vectk .
Definition 2.3.1. Call a Serre subcategory T A localizing if the quotient functor :
A A/T has a right adjoint.
One traditionally denotes the adjoint by , and call the local section functor. We
will see later that this setup captures coherent sheaf cohomology.
Lemma 2.3.2. For a Serre subcategory T A, the following are equivalent:
1. every M Ob(A) has a largest torsion subobject
2. the inclusion i : T , A has a right adjoint (called the torsion functor)
3. the direct sum of torsion modules is torsion
Example 2.3.3. Let A = k[x0 , . . . , xn ] with the standard grading. Let A = GrMod(A) and
T = Tors(A). Recall we defined Tails(A) = A/T . Serres theorem tells L
us that Qcoh(Pnk ) '
0
n ) GrMod(A) is M 7
A/T . Geometrically, our adjoint : Qcoh(P
nZ H (X, M (n)),
L
which has right derived functors Ri ' nZ Hi (X, (n)). Along the same lines, induces
the local cohomology M 7 H0m (M ), where m A is the augmentation ideal.
Theorem 2.3.4. Assume A has injective envelopes, and that T A is closed under essential
extensions (so in particular T is closed under injective envelopes). Then the following are
equivalent:
1. T is a localizing subcategory
2. the inclusion functor T A ha a right adjoint (the torsion functor) : A T
In this case, (A, T ) is called a stable torsion pair.
Corollary 2.3.5. Under the assumptions of Theorem 2.3.4,
1. A/T has enough injectives (so both and have right derived functors)
138
/ idA
/ R1
/ 0,
/ Hom(A, M )
/ Hom(A>n , M )
/ Ext1 (A/A>n , M )
/0
M = lim Hom(A/A>n , M );
L
M = n0 H0 (X, M (n));
L
Ri M = n0 Hi (X, M (n)).
It is a good exercise to derive carefully these isomorphisms.
3
3.1
Derived categories
Definition and basic examples
Example 3.1.2. Lets call the complex (C , d ) cyclic if all di = 0. Let Com0 (A) be the
full subcategory of Com(A) consisting of the cyclicQ
complexes. Let i : Com0 (A) , Com(A)
0
be the canonical inclusion. Notice that Com (A) ' nZ A[n], where A just denotes a copy
of A that corresponds to n Z.
Define the functor H : Com(A) Com0 (A) in the opposit direction, which sends a
complex (C , d ) to the cyclic complex (H (C), 0). Notice that H sends quasi-isomorphisms
to isomorphisms.
Hence, by the universal property of localizations, H factors through Q
H
Com(A)
%
/ Com0 (A)
9
D(A)
Recall that an abelian category A is called semisimple if every short exact sequence in
A splits. Equivalently, A is semisimple if Exti (X, Y ) = 0, i > 0 and for any X, Y A. For
example, categories Vect of vector spaces and kGfdMod of finite dimensional representations
of a finite group G are semisimple.
is an equivalence of categories.
Theorem 3.1.3. If A is semisimple, then H
Proof. Define for any complex (C , d) Com(A) two morphisms
fC : (C , d) (H (C), 0)
: (H (C), 0) (C , d)
gC
as follows.
n1
/ Cn
/ . . . ] we put B n = Im(dn1 ), Z n = Ker dn and
/ C n1 d
For C = [ . . .
H n = B n /Z n for each n Z. Then we have the following exam=ct sequences
/ Zn
/ Cn
/ B n+1
/0
(5.1)
/0
(5.2)
Zn
Cn
B n+1
Since A is semisimple, for each n we can choose splittings of the above sequences. This
will give us C n ' Z n B n+1 ' B n H n B n+1 . With these identifications, the map
dn : C n C n+1 will be given by the map B n H n B n+1 B n+1 H n+1 B n+2 which
sends (bn , hn , bn+1 ) to (bn+1 , 0, 0).
as the canonical projection B n H n B n+1 H n
Then we can define the maps fC and gC
n
n
n
n+1
and embedding H , B H B
respectively.
0
We can define the functor L : Com (A) D(A) to be the composition
Com0 (A)
/ Com(A)
140
/ D(A)
. Namely,
It is a straight forward calculation to show that L is mutually inverse to H
}
the families of morphisms {fC }CCom and {gC
CCom defined above induce isomorphisms of
Remark 3.1.4. In general, D(A) captures much more information than just cohomology
functor. How much more? is a very delicate question. We call A and B derived equivalent if
D(A) ' D(B), i.e. their derived categories are equivalent (as triangulated categories).
For example, take A = Mod(A) and B = Mod(B). Define global dimension (which can
be ) of A as gldim(A) = n s.t. Extn (X, Y ) 6= 0 for some X, Y A, but Extn+1 (X, Y ) =
0, X, Y A. Then gldim is not a derived invariant, but the finiteness of gldim is.
Unfortunate fact: The class Qis is a saturated class, but it does not satisfy any other
Ore conditions.
Hence we cant really conclude anything about the derived category D(A). We dont
even know if it is additive. So we need a bit different construction of D(A). We will give a
construction using the homotopy category K(A).
3.2
Definition 3.2.1. Let Hot Qis be the subclass of homotopy equivalences (see 1.0.8 on page
2). Then the homotopy category K(A) is defined as the localization K(A) = Com(A)[Hot1 ].
We denote by Qh the corresponding localization functor.
This category can be equivalently defined as follows. To fix the notation, if f, g : X Y
are homotopy equivalent morphisms of complexes, we write f h g. Now define the category
K0 (A) by Ob(K0 (A)) = Ob(Com(A)), and
HomK0 (A) (X, Y ) = HomCom(A) (X, Y )/ h
Key fact: Nilhomotopic morphisms form a two-sided ideal in complexes. It means, if
f h 0 then g f h 0 and f l h 0 for any morphisms g, l that are composable with f .
Lemma 3.2.2. The natural quotient functor h : Com(A) K0 (A) maps homotopy equiva : K(A) K0 (A). The induced functor
lences to isomorphisms, and so it induces a functor h
is an equivalence of categories.
h
The description
So from now on we will identify K(A) with K0 (A) via the functor h.
of the homotopy category as a quotient turns out to be more convenient than the
description K(A) as a localization. We will use the notation K(A) for the homotopy category.
K0 (A)
141
Lemma 3.2.3.
1. The localization functor Q : Com(A) D(A) factors through the
quotient functor h : Com(A) K(A):
Q
Com(A)
/ D(A)
;
Q0
K(A)
2. Functor Q0 factors through the the localization functor Qh
Q0
K(A)
/ D(A)
7
'
Qh
K(A)[h(Qis)1 ]
Moreover, the induced functor G is an equivalence of categories.
Proof. Since Qh h maps Qis to isomorphisms, there exists a functor F : D(A) K(A)[h(Qis)1 ]
s.t. Qh h ' F Q:
h
Com(A)
/ K(A) Qh / K(A)[h(Qis)1 ]
7
$
D(A)
We construct the inverse functor to F as follows. First observe that the functor
Q : Com(A) D(A) maps all nilhomotopic morphisms f to zero.
Indeed, suppose f : X Y , X, Y Com(A) and suppose f h 0. This means that there
is a map h : X Y of degree 1 s.t. f = dh + hd. Consider cone(idX ) = X X[1] and
define a map c : cone(idX ) Y by cn = (f n , hn+1 ). It is a straightforward calculation to
verify that since h is a homotopy, the map c will actually be a morphism of complexes, and
f factors through it.
f
/Y
:
c
cone(idX )
Applying the functor Q to this diagram gives Q(f ) = Q(c)Q(i). Now, since idX is obviously
a quasi-isomorphism, cone(idX ) is acyclic (see 3.2.3 on page 109), and so Q(cone(idX )) = 0.
Therefore, Q(f ) = 0.
But then Q must factor through h : Com(A) K(A), since h is universal among all the
functors L : Com(A) K(A) such that L(f ) = 0 whenever f h 0. Hence Q ' Q0 h.
142
This implies Q0 (h(Qis)) = Q(Qis) Iso. But if Q0 sends h(Qis) to isomorphisms, it must
factor through Qh . This implies that Q0 = G Qh for some functor G : K(A)[h(Qis)1 ]
D(A).
To summarize, we have obtained the following diagram:
/ K(A)
Com(A)
Q0
D(A) o
Qh
/ K(A)[h(Qis)1 ]
3.3
Verdier theorem
To simplify the notations we will denote the image h(Qis) of Qis under the functor
h : Com(A) K(A) again by Qis.
Question (Yuri): Is h universal among all (not necessarily additive) functors F : Com(A)
C s.t. F (f ) = 0C whenever f h 0.
Theorem 3.3.1 (Verdier, [Ver96]). The class of morphisms Qis in K(A) is both left and
right Ore.
Proof. We want to check the Ore axioms 1.4.1 and 1.4.2. Saturatedness of the localizing
family of morphisms Qis is obvious.
First lets prove the axiom (LO2). Suppose we are given some maps t : W X and
g : W Y , t Qis. We want to find f : X Z and s : Y Z with s Qis.
>Z `
s
X`
>Y
g
W
We are thinking of morphisms in the homotopy category K(A) as honest morphisms of
complexes, since any morphism in K(A) is represented by some morphism in Com(A).
Define = t g : W X Y . Take Z = cone(). Then we have the following exact
sequence
W
/X Y
/ cone()
143
/ W [1]
(5.3)
iX
/X Y
/ cone()
iY
/X Y
/ cone()
We claim that Z, f, s are exactly what we are looking for. So first of all we want to prove
that the diagram
>Z `
f
X`
>Y
g
W
commutes up to homotopy. We have 0 h i = (f, s) (t, g) = f t + sg, and so
sg h f t.
Next we want to show that s is actually a quasi-isomorphism. To see that, first note
that the exact sequence 5.3 induces the following long exact sequence
...
/ Hk (W )
Hk ()
/ Hk (X) Hk (Y )
/ Hk (Z)
/ Hk+1 (W )
/ ...
Hk (Y )
Hk (t)
Hk (s)
/ Hk (X)
Hk (f )
/ Hk (Z)
being cartesian and cocartesian. Then since Hk (t) is an isomorphism, Hk (s) also must be
an isomorphism. So the map s is a quasi-isomorphism, and we have proved the axiom
(LO2).
144
Chapter 6
Triangulated categories
1
The basics
1.1
Definitions
X`
[1]
/Y
/W
/ X[1] .
u
u(1)
then so is Y
Z
X[1] Y [1].
145
/Y
e
/Y
u
e
e
X
/Z
e
/Z
v
e
/ X[1]
w
e
f [1]
e
/ X[1]
W
Y [1], we can construct the following commutative diagram
X
id
e
u
/Y
/Z
e
/Y
e
/Z
e
v
id[1]
/ X[1]
/W
Y [1]
idW
/ X[1]
v[1]
/ Z[1].
1.2
Example 1.2.1. Let D = Vectk be the category of vector spaces over a field k. Put the
shift functor [1] to be just [1] = idD . The class E of exact triangles is defined by the following
rule:
u
v
w
E = {U V
V W
W U
U V}
for any triple U, V, W of vector spaces. The maps u, v, w are
u = U V V , V W
v = V W W , W U
w = W U U , U V
It is a good exercise to check that this actually gives Vectk the structure of a triangulated
category.
146
Example 1.2.2. The standard examples are K (A) and D (A), for A an abelian category
and {, b, +, }. In these examples, [1] : C 7 C is the usual shift of complexes. The
exact triangles are diagrams isomorphic to
X
/Y
iu
/ cone(u)
pu
/ X[1]
We will postpone verification that this actually is a triangulated structure to the next section.
Example: vector bundles on projective spaces
This example is due to [BGG78], and is exposited in [OSS11]. Let k be V
a field of characteristic
L
Vi
not 2. Let E be a fixed E-vector space of dimension n + 1. Let = k (E) = n+1
(E)
i=0
be the Z-graded exterior algebra of E. Let M() be the category of graded left -modules,
and let Mb () be the full subcategory of graded modules which are finite-dimensional over k.
Let F Mb () be the full subcategory of free modules over . It turns out that here F
actually coincides with both subcategories of projective and injective modules.
Lets call f : V V 0 in Mb () equivalent to zero if f = V F V 0 for some
F Ob(F). Obviously the class I of morphisms in Mb () which are equivalent to zero is a
b
b
two-sided ideal. Define the stable module category M () to be M () = Mb ()/I. Objects
b
of M () are the same as those in Mb (), and hom-sets are
HomMb () (X, Y ) = Hom (X, Y )/I.
Theorem 1.2.3 (Beilinson, Gelfand, Gelfand). There is a natural structure of a triangulated
b
category on M () in which the shift functor [1] is defined by
V 7 ((n) k V )/i(V )(n)
where W (m) =
iZ W
im ,
(n = dim E 1),
I what follows we will try to explain where this triangulated structure comes from.
Observe that every graded -module V can be viewed as a family of complexes
Le (V ) =
i1
/ V i1 de / V i
die
/ V i+1
In the opposite direction, given a rigid complex L , define V (L ) M() by the rule
V (L ) =
V (L )i =
iZ
P(E), L i (i) .
iZ
iZ (P, OP (i))
is
Lemma 1.2.5. The equivalence from Lemma 1.2.4 restricts to the equivalence Rigf ' Mb ().
1.3
Lemma 1.3.1. Let D be a category with an auto-equivalence [1] : D D. Then there exists
f:D
e with an automorphism [1]
eD
e and F : D D
e such that the following
a category D
diagram commutes:
D
[1]
/D
e
D
f
[1]
e
/D
and n : Xn [1]
Xn+1 . We define HomDe ((Xn , n )n , (Yn , n )n ) to be the set of tuples
(fn HomD (Xn , Yn ))n such that for all n, the following diagram commutes:
Xn [1]
fn [1]
/ Yn [1]
Xn+1
fn+1
/ Yn+1
u[2]
v[1]
T
w
u[1]
v
149
A morphism of exact triangles extends to a double helix, and such a double helix is
uniquely determined (by TR3) up to isomorphism by any of the two arrows in the exact
triangle.
For an integer n > 0, let [n] be the n-fold composition of [1], and for n 6 0, let [n] be
the n-fold composition of [1]. Also, to simplify notation we write (X, Y ) = HomD (X, Y ).
u
/ (U, Z[1])
/ (Y [1], U )
/ (U, Y [1])
/ (Z[1], U )
Proof. It suffices to prove that the top sequence is exact at (U, Y ). First we show that
u v = 0. For f : U X, define g = u f = f u, and consider the following diagram.
/0
/Y
/ U [1]
h
/Z
f [1]
/ X[1]
The first row is exact by TR1 and TR2. By TR3, there exists h : 0 Z such that
vg = h 0 = 0, which implies v u f = 0 for all f .
Now assume v g = 0 for some g : U Y and consider the following diagram with exact
rows:
/ U [1] id[1] / U [1]
/0
U
g
/Z
fe
/ X[1]
u[1]
g[1]
/ Y [1]
By TR3, there exists fe : U [1] X[1] such that g[1] = u[1] fe. Applying [1], we see
that g = u fe[1], hence g is in the image of u . This proves exactness of the first sequence.
For the second one the proof is similar.
Corollary 1.3.3 (5-lemma). If the morphisms f, g in axiom TR3 are both isomorphisms,
then so is h.
Proof. Fix some extension h, and choose an arbitrary U Ob(D). Consider the following
sequence (notation as in Proposition 1.3.2):
(U, X)
e
(U, X)
u
e
/ (U, Y )
e)
/ (U, Y
v
e
f [1]
g[1]
u
e[1]
e we / (U, X[1])
e
e [1])
/ (U, Z)
/ (U, Y
150
By Proposition 1.3.2, both rows are exact, so by the usual 5-Lemma for abelian groups,
e is an isomorphism for all U . By the Yoneda Lemma, h : Z Ze is also
h : (U, Z) (U, Z)
an isomorphism.
u
/Y
/Y
u
v
w
u
e
v
e e w
e e
e
Corollary 1.3.6. Let X
Y
Z
X[1] and X
Ye
Z
X[1]
be exact triangles.
e
Assume there exists g : Y Y such that vegu = 0. Then there exists f and h making the
following diagram commute:
X
/Z
e
X
/Y
u
e
e
/Y
v
e
e
/Z
w
e
/ X[1]
f [1]
e
/ X[1]
e
Moreover, if HomD (X, Z[1])
= 0, then f and h are uniquely determined b g.
e yields
Proof. Consider gu : X Ye . Since ve (gu) = 0, Proposition 1.3.2 applied to U = X
e such that gu = u
f :XX
e (f ) = u
e f . Similarly one shows that h exists. The uniqueness
of f and h follows from the long exact sequence.
Remark 1.3.7. For any u : X Y , the axioms TR1 and TR3 yield an exact triangle
u
X
Y Z X[1]. In analogy to the situation with categories of complexes, we write
Z = cone(f ). We could even hope that cone : Mor(D) D is a functor. Unfortunately,
in general one does not always have cone(u v) = cone(u) cone(v). In other words, for
arbitrary triangulated categories, the cone construction is not functorial. Non-functoriality
of cone causes a lot of inconvenience, and later we will see possibilities how to deal with
that.
151
2
2.1
Further properties
Abstract cone and octahedron axiom
/Y
/Z
e
X
e
/Y
u
e
v
e
/ X[1]
e
/Z
w
e
f [1]
e
/ X[1]
The problem is that C(u2 u1 ) = C(u2 ) C(u1 ) does not follow from the axiomatics. What
can we say about C(u2 u1 ) just using the axioms for a triangulated category in particular,
TR4?
Suppose we are given two morphisms u1 : X Y and u2 : Y Z. We can complete
u1
v1
w1
u2
v2
w2
these to exact triangles X
Y
C(u1 )
X[1] and Y
Z
C(u2 )
Y [1].
v1 [1]
2
Define w to be the composite C(u2 )
Y [1] C(u1 )[1].
w
Lemma 2.1.1. C(u2 u1 ) ' C C(u2 )
C(u1 )[1] [1].
u2 u1
/Z
/ C(u2 u1 )
C(u2 )
C(u2 )
Y [1]
v1 [1]
/ C(u2 )[1]
152
/ X[1]
1
2
Corollary 2.1.2. In D(A), any X
Y
Z gives rise to
2.2
diagram X
Y
Z
X[1] is exact in K(A) if and only if it is isomorphic to a diagram
of the form
pu
iu /
u /
/ X[1].
X
Y
cone(u)
u
Y
Z 0 is an exact sequence in Com(A) such
Lemma 2.2.2. Suppose 0 X
n
that for every n, the sequence 0 X Y n Z n 0 splits (this is not the same as
0 X Y Z 0 splitting). Let sn : Z n Y n and pn : Y n X n be the right
(resp. left) inverses of v n (resp. un ). Define w : Z X[1] by wn = pn+1 dnY . Then
u
v
w
X
Y
Z
X[1] is an exact triangle in K(A).
Lemma 2.2.3. Every exact triangle in K(A) is isomorphic to one as in Lemma 2.2.2.
Lemma 2.2.4. The axioms of a triangulated category hold for triangles as in Lemma 2.2.2.
2.3
Let D be a triangulated category and S Mor(D). Let D[S 1 ] be the (abstract) localization
of D at S. We know that if S is Ore, then D[S 1 ] is additive. We are interested in extra
conditions on S that force D[S 1 ] to be triangulated.
Verdier found such conditions. Assume that S, in addition to being Ore, satisfies the
following two properties:
O4 S is closed under [1], i.e. s S if and only if s[1] S.
O5 If f, g S are as in TR3, then h can be chosen to be in S.
Proposition 2.3.1. If S satisfies O1-O5, then D[S 1 ] has a natural triangulated structure.
Proof. Recall that if S is Ore, then morphisms in D[S 1 ] can be represented by equivalence
classes of left S-fractions. It suffices to define the shift functor on these fractions. Put
(s1 f )[1] = s[1]1 f [1]. Say that a triangle in D[S 1 ] is exact if and only if it is isomorphic
to the image of an exact triangle in D.
All that remains is to check the axioms TR3 and TR4. This is pretty tedious direct
verification.
153
Theorem 2.3.2. Let A be an abelian category. Then the derived category D(A) is triangulated.
Proof. First we use the fact that D(A) ' K(A)[Qis1 ]. Next, check that Qis in K(A) satisfies
O1-O5. The axiom O4 is obvious because Hk (S[1]) = Hk+1 (S). On the other hand, showing
O5 requires some use of TR4. We will omit this part.
2.4
Exact functors
/ Y
_
/X
/ cyl(f )
f
/ cone(f )
/ X[1]
/ cone(f )
/0
/0
/Y
such that the rows are exact and , are homotopy equivalences. Moreover, = idY and
h idcyl(f ) .
Corollary 2.4.5. Every exact triangle in K(A) (and also D(A)) is isomorphic to one of
the form
X
/ cyl(f )
/ cone(f )
154
pf
/ X[1]
/Y
O
/ cone(f )
/ X[1]
/ cone(f )
/ X[1]
/ cyl(Y )
Proposition 2.4.6. Every short exact sequence in Com(A) can be completed to an exact
triangle in D(A). More precisely, every short exact sequence of complexes 0 X Y
Z 0 is quasi-isomorphic to
X
if
/ cyl(f )
pf
/ cone(f )
/ X[1].
/X
/Y
O
/X
/0
/Z
O
/ cyl(f )
/ cone(f )
/0
Define : cone(f ) Z by (xk+1 , y k ) := g(y k ). One checks that the diagram above
commutes in Com(A), and that is a quasi-isomorphism. Indeed, since g is surjective, so is
, and Ker() = X[1] Ker(g) = X[1] Im(f ) ' X[1] X. If we show that the kernel of
is acyclic, it will imply that H () is an isomorphism. Now, Ker() is acyclic because one
can exhibit an explicit homotopy idKer() h 0. The homotopy h : Ker() Ker()[1] is
given by (xk+1 , xk ) 7 (xk , 0).
As a conclusion of the results above, we might forget about short exact sequences and
work instead with exact triangles.
2.5
Verdier quotients
2.6
Exact categories
Example 2.6.5. Let B be any additive category. Then B is exact if we let E consist of split
exact sequences X X Y Y .
i
Y
Z
Example 2.6.6. Let A be an abelian category, B = Com(A). Let E consist of X
in
pn
Definition 2.6.13. Let B be a Frobenius category, and let X, Y Ob(B). Define I(X, Y )
g
/ I0
/Y0
where v is not unique on the nose, but any two v are the same in HomB (Y, Y 0 ). This allows
u
Su
/ IX
/ SX
/C
!w
/ SX
with the leftmost square co-cartesian. There exists a unique w : C SX making the
diagram commute. Exact triangles in B are exactly those of the form X Y C SX
as above.
Let (B, E) be an exact category such that B has enough injectives. Let IB be the class
of injectives in B. Recall that we defined the stable category B, whose objects are objects
in B, and has morphisms Mor(B)/I, where I consists of morphisms which factor through
injectives. The category B is additive, but not exact.
For every Ob(B), choose SX Ob(B) such that there is X IX SX in E with
IX in IB . Given
X
i0
/I
/ I0
158
p0
/Y
/Y0
the injectivity of I 0 yields u making the diagram commute. Exactness gives v making the
diagram commute. Suppose we also have extensions u
e and ve of idX : X X.
Lemma 2.6.17. Given two extensions (u, v), (e
u, ve), we have v = ve in HomB (Y, Y 0 ).
Proof. By commutativity of the diagram, we get ui = i0 = u
ei, whence (u u
e)i = 0. Similarly
(e
v v)p = 0. The first tells us that Ker(u u
e) Im(i) = Ker(p), whence u u
e factors
through I/ Ker(p) as in
p
I/ Ker(p)
/Y
ue
u
p0
I0
ve
v
/Y
We have v ve = p0 (u u
e)p1 , whence the result.
Corollary 2.6.18. The assignment X 7 SX extends to an endofunctor S : B B.
Proof. Suppose we have f : X Y . Then as above, we have an extension
/ IX
X
/ SX
/ IY
/ SY
iX
/Y
/ IX
pX
/Z
/ SX
d0
P
I
where P
X is a projectives resolution, X
I 0 , and P 0 I 0 is P I . It is easy to
check that [1] ' S , and that sends exact triangles to exact triangles.
3.1
The same triangulated categories can be identified with the derived categories of completely
different abelian categories. In other words, a triangulated category D can be equivalent to
D(A) and D(B) for unrelated abelian categories A, B. We would like to have some kind of
extra structure that allows us to recover A and B from D. More generally, we would like to
be able to catch abelian subcategories of a triangulated category D. Our main example is
D = Mod(DX ), where X is a variety and DX is the sheaf of differential operators on X.
Example 3.1.1. Let A be an abelian category, D = D (A) for {, +, , b}. Define
D>n (A) = {X D : Hi (X) = 0 for all i < n}
D6n (A) = {X D : Hi (X) = 0 for all i > n}.
The natural inclusion i : A D sending X to 0 X 0 is fully faithful, and
Im(i) = D60 (A) D>0 (A). The general notion of a t-structure is an axiomatization of this
situation.
Definition 3.1.2. Let D be a triangulated category. A t-structure on D is a pair (D60 , D>0 )
of strictly full subcategories satisfying
TS1 D60 D61 and D>1 D>0
TS2 For all X Ob(D60 ) and Y Ob(D>1 ), HomD (X, Y ) = 0.
TS3 For all X Ob(D), there is an exact triangle A X B A[1] in D with
A Ob(D60 ) and B Ob(D>1 ).
Here, we are using the standard notation
D6n = D60 [n]
D>n = D>0 [n].
160
3.2
Definition 3.2.1. Let D be a triangulated category with t-structure (D60 , D>0 ). The core
of this t-structure is the full subcategory D60 D>0 . We will sometimes denote it by D .
Lemma 3.2.2. The core of the standard t-structure on D (A) is A.
Theorem 3.2.3. The core of any t-structure is an abelian category.
During the rest of this subsection we will sketch the main part of the proof. The proof
(which follows the original paper [BBD82]) is based on the following key lemma.
Lemma 3.2.4. Let D be a triangulated category equipped with a t-structure (D60 , D>0 ).
1. For each n Z, the inclusion functors
i6n : D6n D
i>n : D>n D
have right (resp. left) adjoints
6n : D D6n
>n : D D>n
called truncation functors.
2. For all X Ob(D), there is an exact triangle of the form
60 (X) X >1 (X) 60 (X)[1].
Moreover, any two triangles A X B A[1] with A Ob(D60 ) and B Ob(D>1 )
are canonically isomorphic.
Proof. Lets prove the existence of 60 and >1 . The proofs for general n are similar. Using
the axioms for a t-structure, for all X Ob(D), we can choose A X B A[1] with
A D60 and B D>1 . Define
>0 (X) = A
>1 (X) = B
f
B 0 [1]
fe
/ A0
/X
f
/Y
161
/B
/ A[1]
/ B0
/ A0 [1]
6m
6n 6m
>n
>n >m
>m 6n
6n >m
we write [m,n] for the last functor.
We can now use these relations to construct kernels and cokernels in the core A = D . Let
f
f : X Y be a morphism in A D. Extend this to an exact triangle X
Y Z X[1]
in D. As before, we denote Z = cone(f ), and call Z the (abstract) cone of f . Recall that
the cone construction is not functorial in this setting, even though the object Z = cone(f )
is unique up to a (non-canonical) isomorphism. Define
K = 61 (Z)[1]
C = >0 (Z)
There are exact triangles
k
K
Y
/ Z[1]
/Z
/X
c
/C
One checks that K is the categorical kernel of f , and that C is the categorical cokernel of f .
Moreover, Coker(k) = Ker(c), so A is an abelian category.
3.3
Cohomological functors
Recall that if D = D (A), then for all i Z, there are cohomology functors Hi : D A,
defined by Hi (X) = H0 (X[i]). The functor H0 () is just X 7 >0 60 X = [0,0] X.
162
nZ
/ F (X)
/ F (Y )
/ F (Z)
/ F (X[1])
/ ...
is exact.
Theorem 3.3.4.
3.4
t-exact functors
e
/A
F =R
e
/D
3.5
Question: Take a triangulated category D with a t-structure, take its core A, and take
its derived category D(A). How are D and D(A) related?
Answer: In general there is no relation between them, unless we have some extra conditions.
Indeed, suppose we have some functor F : Com(A) D such that it restricts to the
f
/X
/Y
/ 0] we
embedding A , D. Then for a complex of the form C = [0
should have F (C) = conecD (f ). But the cone construction is not functorial in general. So
there is no such functor F .
We will now look at a situation when such F does exist and we will give a criterion when
F is an equivalence of categories.
(6.1)
3.6
Gluing t-structures
/D
/ D/N
(6.2)
N 60 = N D60
E >0 = Q(D>0 )
E >0 = Q(D>0 )
3.7
Examples of gluing
the correspondent derived categories. Then we have the following diagram (see subsection
2.4 in the Chapter 3)
i
DZX
j!
/ DX
X
/ DU
i!
e by putting p D60 =
Define perverse (or twisted) t-structure on D := Db (coh(X))
{B D60 | H0 (B) T} and p D>0 = {B D>1 | H1 (B) F}. Let p A =p D60 p D>0 be
the core of this perverse t-structure. Then we have the following theorem.
Theorem 3.7.2. If we denote Q : D D/N ' Db (coh(X)) (using notations above) the
canonical quotient functor, then we have equivalence of categories Q(p A) ' coh(X).
3.8
We have already breafly discussed the ideas below in the subsection 2.4 of the Chapter 3,
and also in the example 3.7 above. Here we will look at it more carefully.
Definition 3.8.1. Given three triangulated categories D0 , D, D00 (notice: we do not mention
any t-structures at all) we say that D is recollement (or gluing) with respect to D0 and D00
if we have six functors
i
DZX
j!
/ DX
X
/ DU
(6.3)
i!
i i idD0 i i!
j j idD00 j j!
R4 There are exact triangles
i i! X X j j ! X i i! X[1]
j! j X X i! i X j! j X[1]
Remark 3.8.2. Notice that the axiom R3 implies that functors i , i! , j , j! are full embeddings.
167
3.9
We have already discussed this example when we were considering gluing theorems for
t-structures, see subsection 3.7.
Recall that we have the following setup. We have a topological space X, Z X is a
closed subset, and U = X \ Z is its complement, and i : Z , X and j : U X are the
natural inclusions.
Consider the categories AX , AZ , AU of abelian sheaves on the spaces X, Z, U respectively.
We denote by D = Db (AX ), D0 = Db (AZ ) and D00 = Db (AU ) the correspondent derived
categories.
Then we will have the recollement diagram (6.3) (where we need to replace functors j!
and j by Rj! and Rj respectively).
We need to say a few words what these functors from the diagram (6.3) actually are.
Functors i , Rj are push-forwards of sheaves, and i, j are pull-backs.
Recall that a continuous map f : X Y is called proper if for any compact subset
K Y the set f 1 (K) X is also compact. For any such map f we can define functor f! as
follows. For any sheaf F AX we define f! F f F to be the sub-sheaf of the push-forward
sheaf, such that
f! F(U ) = {s F(f 1 ) | f |supp(s) : supp(s) , U is proper}
Notice that if f is a closed embedding, then f! = f . If f is open embedding, then f! is
extension by zero functor.
Later we will need the following table
Functor
f
f!
f
Hom
Hom
Exactness
left
left
left
left
left
Acyclic classes
injective, flabby
inj., flabby
inj., flabby, soft
inj.
inj.
Total derived
R
Rf
Rf!
R Hom
RHom
right
flat
Classical derived
H (X, ) (sheaf cohomology)
Ri f
Ri f!
Exti
Exti
i
T ori
3.10
Let be a k-algebra, and assume that it has finite global dimension. Recall that global
dimension is defined to be the supremum over all -modules M of lengths of minimal
projective resolutions of M . Assuma also that we fixed a 2-sided ideal I and denote by
D the quotient D = /I. Denote by i the canonical projection /I = D.
168
Mod(D)
/ Mod()
i!
Lets describe what the functors above are. First of all, i denotes the restriction of
scalars, given by i M = MD D D . Using the usual tensor-hom adjunction, we get
Hom (MD D D , N ) ' HomD (M, Hom(D D , N )).
Thus, using this adjunction we can define the right adjoint functor i! to the functor i
simply by i! = Hom (D D , ).
From the other hand, we might as well define the restriction of scalars i as i =
HomD ( DD , MD ). Again, using the standard tensor-hom adjunction we get the left adjoint
functor i given by i = D. So weve got an adjoint triple (i , i , i! ). We can pass to
the derived categories to obtain the adjoint triple
Li
Db [Mod(D)]
/ Db [Mod()]
Ri!
Db (D)
Lj!
/ Db ()
[
Ri!
/ Db (U )
Rj
In this diagram, functors Ri! , i and Li were defined above. Moreover, functor j is
L
For an application of the above general construction see [BCE08]. The following example
describes the situation the paper is dealing with.
Example 3.10.2. Suppose X is a smooth variety, G is a finite group acting on X. Denote
by D(X) the ring of differential P
operators on X, and consider the algebra = D(X) o G.
1
Define an idempotent e = |G|
g. Note that Mod(D(X) o G) ' Mod(D(X)G ). this
gG
observation is very useful when one is interested in some invariants of D(X)G which are
invariant under Morita equivalence. Indeed, algebra D(X) o G is much easier to understand
than D(X)G .
170
Chapter 7
In this section we will discuss application of the gluing theorems for t-structures. First we
will discuss abstract recollement, and then we will start talking about perverse sheaves. Our
main references on the subject are [BBD82], [Rei10], [GMV96] and [MV88].
1.1
Stratifications
1.2
Equivalently,
p
Theorem 1.2.2. For any (S, p), the pair (p D60 , D>0 ) is a t-structure on D(X), which
induces a t-structure on D (X) for all {b, +, }.
Definition 1.2.3. We put Pervp (X) = p D60 (X) p D>0 (X); this is the (abelian) category
of p-perverse sheaves on X.
If p = 0, then Pervp (X) = ShX . We can replace Z by any constant sheaf O of rings on
X, and replace ShX by the category of sheaves of O-modules. We write Pervp (S, O) for the
resulting category.
The most common setting is where each Si is over C, and p = 21 dimR S = dimC S.
This is called middle perversity.
1.3
Middle extension
>0
Theorem 1.3.1 (GM). Let Z
X
U be the open-closed stratification. Let (D60
U , DU )
>0
b
b
and (D60
Z , DZ ) be t-structures on D (U ) and D (Z). Let AU = DU and AZ = DZ be the
corresponding cores. Then for any F AU there exists a unique G AX such that
1. j G = F
2. i G D61
Z
172
3. i! G D>1
Z .
Proof. Recall that j j ' idU and (j! , j , j ) is an adjoint triple. Thus for any F1 , F2 DU ,
HomD (j! F1 , j F2 ) = HomDU (F1 , j j F2 )
' HomDU (F1 , F2 ).
NF
DW 0
j!
/D
W
/ D 00
i!
Here, (i! , j ! = j , j ) is an adjoint triple, and we require that j j ' idD00 ' j j! .
For any F, G Ob(D0 ), we have
HomD (j! F, j G) ' HomD00 (F, j j G) ' HomD00 (F, G).
Setting F = G, we get for any F Ob(D00 ) a morphism N (F ) : j! F j F . This yields a
natural transformation (called the norm) N : j! j .
Assume D0 , D, D00 are derived categories of abelian categories.
Definition 1.3.2. The (Goresky-MacPherson) middle extension is j! = Im(N : j! j ).
So j! is a functor D00 D. [Note that it is not clear a priori if this makes sense
the image may not exist.]
Definition 1.3.3. Let (X, S) be a stratified space with perversity function p : S Z. Let
R be a noetherian ring, and D = D(X) be the derived category of sheaves of R-modules on
X. Let iS : S , X be the inclusion for S S.
p
Hence, if P Pervp (U ) is a perverse sheaf, then Rj (P) Ob(p D(X)>0 ) and j! (P)
Ob(p D(X)60 ).
Definition 1.3.5. Let the situation be as above. Define functors p j , p j! : Pervp (U )
Pervp (X) by
p
Here p 60 and p >0 are the truncation functors with respect to the p-perverse structure.
The norm transform N : j! Rj induces a perverse norm p j ! p j , which a natural
transformation between functors Perv(p, U ) Perv(p, X).
Definition 1.3.6. The (perverse) Goresky-MacPherson extension is
p
Proposition 1.3.7. For any union of open strata U , if P Pervp (U ), then F = p j ! (P)
Pervp (X), as an extension of P to X (i.e. j F = P), is characterized uniquely by the
following conditions:
Hn (iS F ) = 0
(i!S F )
=0
for all S X r U .
1.4
Constructible complexes
Let (X, S) be a stratified space. We assume that all strata are good, i.e. equidimensional,
with t each iS : S , X having finite cohomological dimension. Let R be a noetherian ring.
Definition 1.4.1. A sheaf F Sh(X, R) is constructible with respect to stratification
S if iS F =: F |S is locally constant (i.e. a local system) for each S S. A complex
F Com(ShX (R)) is (cohomologically) constructible if each Hn (F ) is a local system.
Definition 1.4.2. We let DS (X, R) D(X,S
R) be the full subcategory of constructible
complexes with respect to S. Let Dc (X, R) = S DS (X, R), where S ranges over all good
stratifications of X.
One reason to introduce Dc (X, R) is Verdier duality. Recall that for f : X {} the
= f ! R; this is called the dualizing complex.
unique map, we define DX
Theorem 1.4.3 (Poincare-Verdier duality). Let R = k be a field. Then the dualizing complex
DX is in Dc (X, k). If we define D = RH (, DX ), then DX : Dc (X, k) Dc (X, k) is an
equivalence of categories.
om
174
There is a duality theory for perverse sheaves. Define p : S Z by p (S) = p(S)dimR (S).
Notice that
Hn (iS or ! F ) = 0
Hn (iS or ! DX F ) = 0
n > p(S)
n > p (S).
Again, assume R = k is a field. Then the duality functor DX : Pervpc (X, k) Pervpc (X, k)
is an equivalence of categories. Assume all S S are even-dimensional, and put p1/2 (S) =
21 dimR (S). This is known as the middle perversity. Poincare-Verdier duality restricts
p
to an autoduality on Pervc 1/2 (X, k). In general, we get a duality between Pervpc (X, k) and
1.5
Refining stratifications
Fix a space X. Assume that a stratification T is obtained from S by refining strata in that
each S S is a union of strata in T . There is a natural inclusion DS (X, R) , DT (X, R).
Moreover, we assume that S and T come with compatible perversities p, q, i.e.
p(S) 6 q(T ) 6 p(S) + dim S dim T
for all S S, T T with T S.
Theorem 1.5.1. The t-structure on DT (X, R) of perversity q induces the t-structure of perversity p on DS (X, R), and an exact embedding of abelian categories Pervpc (X) , Pervqc (X).
For any j : U , X, the corresponding functors q j , q j ! , q j ! are compatible with those for p,
in the sense that e.g. the following diagram commutes:
/ Pervq (X)
c
Pervpc (X)
q j
pj
/ Pervq (U )
c
Pervpc (U )
n 6 m.
()
Using Theorem 1.5.1, define a t-structure of perversity p on Dbc (X, C). Let Pervpc (X) be the
core of this t-structure.
175
Proposition 1.5.3. For F Dcb (X, C). The following are equivalent:
1. F Pervpc (X)
j
i > p(dimR S)
i < p(S).
Theorem 1.5.4. Under the assumption (), the category PervcS (X) = Pervp (X) DS (X, C)
is Artinian (and Noetherian) and its simple objects are of the form
L (S, E ) = (p iS )! E [p(S)]
where E is a locally constant sheaf on S S. In particular, if all S are simply connected,
then there is a bijection between simple objects and strata in S.
In general, the category Pervp (X) is not Artinian, but it is Noetherian. If p = p1/2 ,
then autoduality + Noetherian Artinian. For this reason, one often restricts to the
(finite-length) category Pervp1/2 (X).
1.6
Gluing
Theorem 1.6.2. Pervc 1/2 (C, Can ) is equivalent to the category of quadruples (V, W, A : V
W : B) such that AB + 1 and BA + 1 are invertible.
This equivalence comes from the functors vanishing cycles and nearby cycles.
This corresponds to the quiver Q = . From this, we can form the deformed
p
preprojective algebra q,} (Q), and Pervc 1/2 (C, Can ) ' q,} (Q)-Mod.
176
Matrix factorizations
2.1
Introduction
Matrix factorization was first introduced by Eisenbud [Eis80] to describe the stable homological features of modules over a hypersurface singularity. The ideas and phenomena
described in that paper was later formalized in modern language ([Orl04] [Buc86]) as exact
equivalences between the triangulated categories
[MF(R, w)] ' MCM(S) ' Dsg (S),
where
1. R is a regular commutative ring with finite Krull dimension, and w R is a non-unit
non-zerodivisor.
2. The category [MF(R, w)] is the homotopy category associated to the DG category of
matrix factorizations
3. The category MCM(S) is the stable category associated to the Frobenius category of
maximal Cohen-Macaulay modules over the (Gorenstein) ring S
4. The category Dsg (S) is the singularity category of X = Spec(S), defined by the Verdier
quotient Dsg (X) = Db (coh X)/perf(X)
By its construction, The category [MF(R, w)] has a natural enrichment MF(R, w), which
is a Z/2-graded DG category. (i.e., its morphism are 2-periodic complexes). Dycherhoff[Dyc]
then set out to study this DG-category, viewing it as a noncommutative space in the sense of
[KKP08]. Indeed, showing that a suitable enlargement of M F (R, w) has a compact generator,
Dycherhoff has shown that MF(R, w) is derived Morita equivalent to a DG-algebra, in the
sense of Toen [Toe].
Therefore, the noncommutative space in question is DG-affine, and we can use that to
compute the Hochschild cohomology of this DG-category, which can be thought of as Hodge
theory on the noncommutative space.
2.2
In this section, we describe the three triangulated categories mentioned in the introduction,
and show that they are exact equivalent.
Definition 2.2.1. Let R be a regular domain with finite Krull dimension, and w R is a
nonzero non-unit. The category MF(R, w) of matrix factorization is defined by:
1. An object of this category consists of a Z/2-graded R-module P = P 0 P 1 with
two R-linear maps d0 : P 0 P 1 and d1 : P 1 P 0 satisfying d1 d0 = w idP 0 and
177
d0 d1 = w idP 1 . We will often call such an object a matrix factorization, and denote
it by
d1
1
0
*
P ) P .
d0
2. Between two objects (P, dP ) and (Q, dQ ) there is a Z/2-graded morphism complex
Hom(P, Q) defined as the set of R-linear homomorphism from P to Q with obvious
grading. This complex has the differentials d0 : Hom0 (P, Q) Hom1 (P, Q) and
d1 : Hom1 (P, Q) Hom0 (P, Q) both defined by the formula df = dQ f (1)|f | f dP .
3. Composition of morphism complexes is defined by composition of the R-module maps.
It is then straightforward to check that this is a Z/2-graded DG-category. That is,
its morphism complexes actually satisfy d2 = 0, and its composition maps Hom(Q, R)
Hom(P, Q) Hom(P, R) are chain maps.
Given any Z/2-graded DG-category C, we can define the category [C] = H0 (C) by
Ob([C]) = Ob(C)
Hom[C] (X, Y ) = H0 (HomC (X, Y )) .
In our case C = MF(R, w) and we see that [C] is the category of chain maps between matrix
factorizations modulo homotopy.
Given objects X and Y in MF(R, w), and a chain map f : X Y , (i.e. f
Z 0 (Hom(P, Q))), we can define the cone of f in exactly the same way as the standard
cone construction of chain complexes. The collection of triangles isomorphic to the associated standard triangles will make [C] into a triangulated category. For details, see
[Orl04].
Definition 2.2.2. Let X be a Noetherian scheme of finite Krull dimension which is either
affine or is quasiprojective over a field. Then we define its singularity category as the Verdier
quotient
Dsg (X) = Db (coh X)/perf(X),
where perf(X) denotes the triangulated subcategory of Db (coh X) consisting of complexes
that are isomorphic in Db (coh X) to a bounded complex of locally free sheaves.
The following proposition explains why it is called the singularity category.
Proposition 2.2.3. X is nonsingular if and only if Db (coh X) = perf(X).
Proof. By Serres theorem, coh(X) has enough locally free sheaves. Thus, for any bounded
complex M of coherent sheaves, there is a quasi-isomorphism P M from a bounded
above complex of locally free sheaves. Then, Coker(dn
P ) is always locally free if and only if
the local ring at every stalk has finite global dimension, if and only if X is regular.
178
/ Mp /(a1 , . . . , ai1 )
ad
/ Mp /(a1 , . . . , ai1 )
/ Mp /(a1 , . . . , ai )
/ 0.
ext (Mp , Sp ) = 0.
3 4. Use the fact that R Hom(, S) is a duality functor on Db (S), we see that such a
functor restricts to a duality functor on the class of finitely generated modules such that 3
holds, and is simply the functor () there. Thus, we can obtain a free resolution of M ,
and then apply () again.
4 1. This is obvious.
179
Definition 2.2.5. We call a module satisfying any of the equivalent conditions a maximal
Cohen-Macaulay module over S, and they form a full subcategory MCM(S) of Mod(S).
Proposition 2.2.6. The category M CM (S) can be made into an exact category by endowing
it with sequences that are short exact in Mod(S), Then, an object is projective injective
projective as an S-module.
Proof. By Proposition 2.2.4, a maximal Cohen-Macaulay module has enough projective
module in both sides. Hence, if M is either injective or projective in the exact category
MCM(S), then it must be a direct summand of a projective S-module, and hence is itself a
projective S-module. The other direction is easy.
Proposition 2.2.6 tells us that MCM is a Frobenius category in the sense of [Kela]. Thus
we can form its associated stable category.
Definition 2.2.7. Write MCM(S) for the stable category associated to the Frobenius category
MCM(S).
Now, let R be a regular Noetherian domain of finite Krull dimension n, and w R is a
nonzero non-unit. Let S = R/w.
We claim that the triangulated categories [MF(R, w)], MCM(S) and Dsg (Spec S) are
exact-equivalent.
Definition 2.2.8. Given an object X in MF(R, w), we denote by X the complex of Smodules
. . . X0 X1 X0 0
where the rightmost nontrivial module is at degree 0.
Proposition 2.2.9. For any object X in MF(R, w), the complex X is exact except possibly
at degree 0.
Proof. This is a straightforward diagram chasing, merely using the fact that multiplication
by w is injective on finitely generated projective modules over the regular ring R.
Given an object X in MF(R, w), we note that Im(d1 ) Im(d1 d0 ) = Im(w), and hence
Coker(d1 ) is an S-module.
We know that Coker(d1 ) is an arbitrarily high syzygy, and is therefore a maximal CohenMacaulay module over S. One can show that the construction X 7 Coker(d1 ) actually
induces a functor
Coker : [MF(R, w)] MCM(S).
Theorem 2.2.10. The functor Coker : [MF(R, w)] MCM(S) is an equivalence of categories.
180
Proof. The fact that Coker is fully-faithful is a diagram chasing exercise. (It is useful to note
that a map of S-modules factor through a projective iff it factors through a free module)
To show that Coker is essentially surjective, suppose we are given a maximal CohenMacaulay module M , then apply the Auslander-Buchsbaum formula to the localizations of
M at primes p ann(M ), considered as an Rp -module, we have:
pdRp (Mp ) = depth(Rp ) depth(Mp ) = m (m 1) = 1.
Hence, there is an exact sequence
d1
0 X 1 X 0 M 0
where X 1 and X 0 are projective R-module. Then use projectivity of X 0 to show that
w : X 0 X 0 lifts to an R-linear map d0 : X 0 X 1 , as in the following commutative
diagram:
X0
d0
/ X1
/ X0
d1
!
/M
/0
i.e., we have
d1 d0 = w
Hence, we have
d1 d0 d1 = w d1 = d1 w
Then we can use the injectivity of d1 to show that d0 d1 = w as well. This shows essential
surjectivity.
Now, we shall describe the functor
: MCM(S) Dsg (S).
By definition, MCM(S) is a full subcategory of mod(S). We claim that the inclusion
MCM(S) , mod(S) induces as follows:
MCM(S)
MCM(S)
/ mod(S)
/ Dsg (S)
Indeed, since every projective S-module is a perfect complex in Dsg (S), hence every morphism
that factors through a projective S-module is zero in Dsg (S)
Proposition 2.2.11. The functor : MCM(S) Dsg (S) is an equivalence of categories.
181
Proof. See [Orl04, Prop 1.21] for a proof that is fully faithful.
To show that is essentially surjective, suppose we are given a bounded complex M of
S-modules
M = [0 M e . . . M 0 0]
which we assume, without loss of generality, to be bounded above at degree 0.
d1
P = [. . . P e . . . P 0 0]
By the characterization of Proposition 2.2.4(1), we see that Im(den1 ) is maximal
Cohen-Macaulay since it is the n-th syzygy of Im(de1 ). Then the good truncation
en1 (P ) = [0 Im(den1 ) P en P 0 0]
is still quasi-isomorphic to M . Moreover, in Db (S), we have exact triangles
en (P ) en1 (P ) Im(den1 )[e + n + 1] en (P )[1]
which translate to an equivalence
en1 (P ) Im(den1 )[e + n + 1]
in Dsg (S). This shows that M ' Im(den1 )[e + n + 1] in Dsg (S).
Use the characterization of Proposition2.2.4(4), we can syzygy down the MCM module
Im(den1 ), and apply the same argument as above, to show that
Im(den1 )[e + n + 1] ' N
in Dsg (S) for some MCM module N . This shows essential surjectivity.
Theorem 2.2.12. The equivalences Coker and are exact equivalences.
Proof. The exactness of is straight-forward checking.
For the exactness of Coker, it will be easier to show that coker is exact. Indeed, by
Proposition2.2.9, this composition is just the functor
X 7 X.
To show that Coker is exact, it suffices to note that X and X[1] only differ by a projective
module; and that cone(f ) and cone(f ) also only differ by a projective module.
More precisely, we have maps in Db (S):
182
X
f
X
f
X1
/ cone(f )
/ cone(f )
/ X 1 [1]
X1
/ X
/ X[1]
/ X 1 [1]
where the first long column is the image under Coker of an exact triangle in [MF(R, w)];
the second long column is an exact triangle in Db (S); the horizontal maps are also exact
triangles in Db (S) This diagram, when read in Dsg (S), shows that Coker is exact.
2.3
In the next several sections, we will survey the results from [Dyc], showing that a suitable
enlargement of the triangulated category [MF(R, w)] has a compact generator in the case of
an isolated singularity.
Let R be a regular domain of finite Krull dimension, w R a nonzero nonunit, and
S = R/w. Let (x1 , ..., xm ) is a regular sequence in R. Suppose w is in the ideal (x1 , ..., xn ),
then L := R/(x1 , ..., xm ) is an S-module. We will obtain a description of the object in
[MF(R, w)] that correspond to L in Dsg (S).
Let V be the free R-module V = Re1 . . . Rem , and suppose
V V has homological
degree 1. It is easy to see that on the graded commutative algebra
V , there isVa unique
DG-algebra
structure
with
differential
(e
)
=
x
of
degree
+1.
(Here,
ei V = 1 V and
i
i
V0
xi R =
V ) We
V denote this differential by s0 , which is a degree +1 differential on the
graded complex
V :
0
^m
0
0
V
. . .
^1
0
V
^0
V 0
^m
1
1
V
. . .
^1
183
0
V
^0
V .
wi xi . Then we can
by Leibniz rule of s0
= w x s1 s0 (x)
V
So s0 s1 + s1 s0 = w id on V .
Thus, letting s = s0 + s1 , we have s s = w id, and hence
Vodd
s Veven
stab
*
L
=
V
V .
)
2.4
Definition 2.4.1. Let R be a regular domain with finite Krull dimension, and w R
is a nonzero non-unit. We define MF (R, w) as the category whose objects are matrix
factorization
1
P P0
as in Definition 2.2.1, but this time we do not require P 0 and P 1 to be finitely generated.
Morphism complexes are defined exactly as in Definition 2.2.1.
Proposition 2.4.2. Let R, w, S be as above.
1. MF (R, w) is a Z/2-graded DG category, and MF(R, w) MF (R, w) is a full DGsubcategory.
184
Y )
is an isomorphism, where the coproduct on the left is taken in the category of abelian groups.
Definition 2.4.4. Let T be a triangulated category. We call an object X T a generator
if the smallest strictly full triangulated subcategory of T containing X is T itself.
Proposition 2.4.5. Let T be a triangulated category with arbitrary coproducts, and X T
be a compact object. Then X is a generator if and only if for all Y T ,
HomT (X[i], Y ) = 0 for all i Z
Y = 0 in T .
2.5
(a) Surjectivity on Hom complexes: For any objects x, y T , the map of complexes
HomT (x, y) HomS (f (x), f (y))
is surjective.
(b) The induced map [f ] : [T ] [S] lifts isomorphisms. That is, if y 0 f (x) is an
isomorphism in [S], then there exists an isomorphism x0 x that maps to that
given isomorphism.
Proof. See the paper [Tab05].
In order to clarify the definitions that will follow, we first make the following definitions.
Give k[u ] the structure of a DG-algebra with deg(u) = 2. THe category of all modules over
k[u ] is denoted by C(k[u ]). Its objects are simply 2-periodic k-complexes, and morphisms
are 2-periodic cochain maps. Let C(k) be the category of all k-complexes. There are obvious
DG-enrichments of these categories, which we denote by Cdg (k[u ]) and Cdg (k), respectively.
Thus Cdg (k[u ]) is a Z/2-graded DG-category such that Z 0 (Cdg (k[u ])) = C(k[u ]), and
Cdg (k) is a DG-category with Z 0 (Cdg (k)) = C(k).
Write dgcatk for the category of al (small) DG-categories over k. Similarly, write
dgcatk[u ] for the category of all (small) Z/2-graded DG-categories over k. The reason for
the notation dgcatk[u ] is that we can think of a Z/2-graded DG-category as a category
enriched over the monoidal category C(k[u ]). We will write Ho(dgcatk ) and Ho(dgcatk[u ] )
for the respective homotopy categories associated to these two model categories.
Definition 2.5.3. Given a DG categories T , a (left) T -module is a DG functor T
Cdg (k). Similarly, given a Z/2-graded DG-category T , a (left) T -module is a DG functor
T Cdg (k[u ])
For the rest of this subsection, statements and definitions will be given for DG-categories
only. There is obvious counterpart in the Z/2-graded case which we shall not repeat.
Example 2.5.4. If x Ob(T ), then we have a DG-functor
hx : T Cdg (k)
y 7 HomT (y, x).
Note that this is really a DG functor because the map
HomT (y, z) HomCdg (k) (HomT (z, x), HomT (z, x))
is a map of complexes. Thus, hx is a left T -module.
Given two DG functors F, G : T S between two DG-categories. It is straightforward
to define the notion of a natural transformation of given homogenous degree from F to G; it
is likewise straightforward to define the differential of such natural transformation. Then
the class of such natural transformations naturally form a complex. (cf [Kelb] for details.)
187
By the previous paragraph, the class of all DG functors T S can be given the structure
of a DG-category, which we denote by Hom(T , S). In particular, the class of all T -modules
is naturally a DG-category, which we denote by mod(T ), i.e.
mod(T ) = Hom(T , Cdg (k)).
A map between T -modules is defined to be a morphism in Z 0 (mod(T )). By definition,
this means a natural transformation between two DG functors from T to Cdg (k) which is
required to be degree zero and commute with differentials.
Theorem 2.5.5. The DG-category mod(T ) has the structure of a model category, where the
fibrations and weak equivalences are defined levelwise using the model structure on C(k[u ]).
We denote by Int(mod(T )) the full subcategory of fibrant cofibrant objects in this model
category. Write
Tb = Int(mod(T )).
An object in Tb is called perfect or compact if it is perfect or compact as an object in the
triangulated category [Tb ] ' D(T ), and we denote by Tbpf the full subcategory of perfect
objects of Tb .
It is easy to check that the image of the Yoneda functor h : T mod(T ) actually lies in
Tbpf , i.e. h induces a functor T Tbpf . Moreover, there is an enriched Yoneda lemma which
asserts that this functor is always fully faithful.
Definition 2.5.6. A DG category T is called triangulated if the DG functor
h : T Tbpf
is a quasi-equivalence. The DG category Tbpf is called the triangulated hull of T .
For any C(k)-model category M (i.e., a model category with C(k)-enrichment which is
compatible with its model structure), we have
[Int(M)] ' Ho(M).
In particular, for M = mod(T ), we have
[Tb ] ' D(T )
[Tb ]pf ' Dperf (T ).
For more details, refer to [Toe11].
If T is a triangulated DG category, it is clear that the followings hold:
1. For any object x T , the T -module hx [1] is represented by an object, say x[1].
f
3. The shifts and cone above makes [T ] into an idempotent complete triangulated category.
It seems to be a folklore that the converse is also true, but the author knows of no statement
explicitly written down.
Definition 2.5.7. A DG functor F : T S induces (by T S) an extension functor of
modules. These fit into a commutative diagram
T
Tbpf
/S
Fbpf
/ Sbpe
x, y Tbpf
x, y Tbpf
/ H0 (Sbpf )
/ Dperf (mod(S ))
Dperf (mod(T ))
hence the term derived Morita equivalence.
2.6
actually restricts to
f : T Int(S op mod) , x 7 T (, x)|S
and this restricted functor is a quasi-equivalence.
Furthermore, f induces a quasi-equivalence
Tpe
= Int(S op mod)pe = Sbpe
Proof. We proceed in several steps:
1. The first claim that each hx |S is fibrant cofibrant in S op mod is straight-forward.
2. Check that f commute with coproducts.
3. Show that f induces isomorphism on H 0 () of the complexes. i.e., Show by the
following steps that:
Hom[T ] (x, y)
HomD(S op ) (f (x), f (y)
()
Note that [M F (R, w)] has arbitrary coproducts, and hence is idempotent complete (cf
[Kra07]). Therefore, by the folklore mentioned above, we see that M F inf ty (R, w), defined
in Section 3.2, is a triangulated DG category,
Hence, by Theorem 2.4.7, we may apply Theorem 2.6.1 to the cases:
1. T = M F (R, w) and S = M F (R, w)
2. T = M F (R, w) and S = {k stab }
and we have:
190
Theorem 2.6.2. Let (R, w) has isolated singulated at a rational point. Denote A as the
full subcateogry {k stab } M F (R, w). (Alternatively, A can be viewed as the Z/2-graded
endomorphism DG algebra HomM F (R,w) (k stab , k stab ).) Then there exists quasi-equivalences:
2.
\
M F (R, w)
M F
(R, w)
b
M F (R, w)
A
3.
b
\
MF
(R, w)pe
A
pe
1.
Thus, in particular, (3) says that M F (R, w) is derived Morita equivalent to the Z/2graded DG algebra A.
2.7
Hochschild cohomology
Warning: The author is not completely familiar with the material in this subsection.
Therefore, the material presented might be inaccurate. But the author has tried to make
sure that what is written here is at least morally right.
Recall that, for two Z/2-graded DG categories T ,S, there is a Z/2-graded DG categories
Hom(T , S).
This gives a functor:
Hom : dgcatop
dgcatk[u,u1 ] dgcatk[u,u1 ]
k[u,u1 ]
This functor can be derived to:
RHom : Ho(dgcatk[u,u1 ] )op Ho(dgcatk[u,u1 ] ) Ho(dgcatk[u,u1 ] )
Definition 2.7.1. Given Z/2-graded DG categories T , the image under
Hom(T , T ) RHom(T , T )
of the identity natural transformation idT is an object of RHom(T , T ), which we still denote
by idT . Then the Hochschild complex of T is defined as the complex
Hom(RHom(T ,T )) (idT , idT )
Hochschild cohomology of T is defined to be the cohomology of the Hochschild complex:
HH (T ) := H (Hom(RHom(T ,T )) (idT , idT ))
which is a Z/2-graded algebra.
It turns out that Hochschild cohomology is invariant under derived Morita equivalence
(cf. [Toe11]). Thus, to compute the Hochschild cohomology of M F (R, w), it suffices to
compute that of M F (R, w). To do that, we need to do two things:
191
(7.1)
Assume from now on that (R, w) has an isolated critical locus at a rational point. Denote
by w
= w 1 + 1 w R R.
Let I R R be the ideal I = (x 1 1 x), hence the module := R R/I is the
diagonal module. (i.e., I is the ideal of the diagonal of Spec(R) Spec(R).)
By an analogy to Fourier-Mukai theory, the following proposition, whose proof we refer
the reader to [Dyc], should not be a surprise.
Proposition 2.7.5. The stabilization stab M F (R R, w)
of the diagonal module corresponds to the identity functor under the quasi-equivalence 2.7.3.
Before we proceed to the explicit calculation of the Hochschild cohomology, we quote
the following result which allows us to pass to the completion of R.
b
Proposition 2.7.6. If (R, w) has an isolated singularity at a rational point, then R R
induces a quasi-equivalence
b w)
M F (R, w)
M F (R,
(stab , stab )
=Hom
M F (RR,w)
[
[
stab ,
stab )
(
=HomM F (RR,
\ w)
[
stab = stab R
b is also the stabilization of the R
\
where
R-module Rb
R
\
R must be isoBy the Cohen structure theorem, the regular local complete ring R
morphic to k[[x1 , ..., xn , y1 , ..., yn ]]. Moreover, the diagonal module is given by Rb =
k[[x1 , ..., xn , y1 , ..., yn ]]/I where I is generated by the regular sequence (1 , ..., n ), i =
x i yi .
We denote w
= w 1 + 1 w as w(x,
y) = w(y) w(x). Since w
I, we have
w(x,
y) = wi i for some wi = wi (x, y)
Hence, the matrix factorization stab
has the form
b
R
stab
:=
b
R
Vodd
V o
/ Veven
where:
\
\
V =R
Re1 ... R
Ren
s0 = DG algebra differential on
Lemma 2.7.7. Let X and Y be objects in M F (R, w) and assume tat X has finite rank.
Let A be the endomorphism DG algebra of X and that Y is the stabilization of the S-module
L. Then there exists a natural isomorphism
Z/2
HomM F (R,w) (X, Y )
= HomR (X, L)
Z/2
\
R/I)
=Hom \ (stab , R
RR
Vodd
V o
/ Veven
\
where () denotes RR
\ (R R/I)
This last complex is just the Koszul complex with respect to the elements (w1 , ..., wn ) in
\
R
R/I.
\
To compute this complex as an R-complex, it suffices to determine what wn ) R
R/I
corresponds to under the isomorphism
x7x1 \
b
\
R
R R R
R/I
Thus, we calculate:
w(x i ) w(x)
i
w(x1 , ..., xi1 , yi , ..., xn ) w(x)
= lim
i 0
i
i , ..., n )
w(y1 , ..., yn ) w(x)
(mod 1 , ...,
= lim
i 0
i
w(x,
y)
i , ..., n )
= lim
(mod 1 , ...,
i 0
i
= wi (x, y)
i w(x) = lim
i 0
Here, the computation is taken inside the power series ring k[[x1 , ..., xn , y1 , ..., yn ]]. Therefore, the limit operation makes sense.
Theorem 2.7.8. The Hochschild cochain complex of the Z/2-graded DG category M F inf ty (R, w)
is quasi-isomorphic to the Z/2-graded Koszul complex of the regular sequence 1 w, ..., n w
in R. In particular, the Hochschild cohomology is isomorphic, as an algebra, to the Jacobian
algebra
HH (M F (R, w))
= R/(1 w, ..., n w)
concentrated in even degree.
194
Since Hochschild cohomology is invariant under derived Morita equivalence, we have the
following:
Corollary 2.7.9. The Hochschild cohomology of the Z/2-graded DG category M F (R, w) is
isomorphic to the Jacobian algebra
HH (MF(R, w)) ' R/(1 w, . . . , n w)
concentrated in even degree.
Some physics
Well begin with a standpoint that emphasizes physical intuition, at the expense of mathematical rigor. This will be centered on a specific quantum field theory, due to Jones and
Witten.
3.1
Physical motivation
Well approach gauge invariants via an example: the group U (1). The Maxwell equations
are
B =0
E =
B
t
From these equations we know that B = A for some potential A, but A is not unique
it can be replaced by any A + A0 .
Let f = A A , where =
, and dA = 12 f dx dx . Here, the Maxwell
equations are
df = 0
df =0
These equations are invariant under the operation A 7 A .
In quantum mechanics, one starts with a stateR vector, i.e. a function (x, t). A
property E is identified with the functional hEi = d3 x E. If we replace by ei ,
nothing is changed. On the other hand, if we replace by ei(x) , then the operator will not
send to the same value. To account for this, we replace with the covariant derivative
D = + ieA . This plays the role of a connection.
We know have D ei(x) (x) = ei(x) D . This is a type of parallel transport. It can be
defined as
R
R(c; A) = ei c A dX
where c is some curve. This integral is not gauge-invariant, in the sense that if A 7 A0 =
A , then R(c; A0 ) = ei(P2 ) R(c; A)ei(P2 ) , where P1 and P2 are the endpoints of c.
This construction is motivated by the Bohm-Aharonov effect.
195
Lets move to a gauge theory over a general Lie group. In the Lie algebra, we have
a
[Ta , Tb ] = ifabc Tc , and put U = ea T , where the a R. Define A (x) = Aa Ta , where x
lives in spacetime, and the Aa are some functions on spacetime.
Define a parallel transport operator
R(x + dx, x; A) = 1 idx A (x).
Partition the curve, and get
R(c; A) = lim
N
Y
(1 ix` A (x` )) = P e
A(x)dx
`=1
where is the start and is the end of the curve. This is not gauge invariant it transforms
by U ()R(c; A)U 1 (). The Taylor expansion becomes
A0 = U A U 1 + i( U )U 1 .
We want to compute R(; A), where is a small rectangle at x with side-lengths dx and
x . We have
R(; A) = R(x, x + dx)R(x + dx, x + dx + x)R(x + x)R(x),
which is not gauge invariant. Its trace is
tr R(; A) = Wc (A) = tr(P ei
A(x)dx
).
The operator Wc is Wilson loop. The result of this holonomy calculation is eiF dx
Working out the details, we get a formula for this curvature, which is
F = A A i[A , A ].
In the language of differential forms,
1
dA AA = F dx dx .
2
Lets
switch to some Feynman integrals. Start with the Lagrangian L = T V , and let
R
S = dtL. The Lagrangian L depends only on x and x.
We have the equation
L
L
= 0.
t x
x
Over a three-dimensional manifold whose metric has signature , +, +, we put
Z
1
L=
d3 x| |2 .
2 M
196
We also put
Lguage
1
=
2
d3 x|F (A)|2 ,
3.2
Cobordism
/M o
! }
M0
`
Cobordisms M0 : 0 1 and M1 : 1 2 can be composed: M1 M0 = M1 1 M0 . It
is not immediately obvious that the composite has a smooth structure, but this is in fact
the case. One uses Morse functions fi : Mi [i, i + 1], such that f0 is regular on [1 , 1]
and f1 is regular on [1, 1 + ] for some > 0.
The composite of cobordisms to be independent of the isomorphism classes of the
individual cobordisms. We can define the n-cobordism category cob(n), whose objects
are closed oriented (n 1)-manifolds, and whose morphisms are differmorphism classes
of cobordisms. It is an easy exercise to check that the cob(n) are in fact categories. The
category cob(n) has a monoidal structure coming from disjoint union. The identity element
for this monoidal structure is the empty (n 1)-manifold. In cob(n), there are natural
isomorphisms : 1C t
, etc. Even better, cob(n) is a symmetric monoidal category.
3.3
S=
L ,
x
M
.
Definition 3.3.1. A topological quantum field theory is a monoidal functor cob(n) Vectk
for some field k.
If we are thinking of physical intuition, one should set n = 4. To each , associate to
the space Fun(O()), i.e. the space of all (not necessarily linear) functions O() R. To
each cobordism M : 0 1 , associate an integral operator ZM : O(0 ) Fun(O(1 ))
defined via the kernel
Z
KM (1 , 2 ) =
O(M )
|0 =1
|1 =2
eiS() .
K=
O(M )
KM (0 , 1 ) =
|0 =0
|1 =1
F (M )
1
.
# Aut()
Moreover,
ZM (f ) =
KM (0 , 1 )f (0 ).
0 F (0 )
198
3.4
Chern-Simons
3.5
S
A
Atiyah-Segal axioms
Scholze introduced perfectoid fields (and more generally, perfectoid spaces) in his paper
[Sch12], in which he proved a wide range of special cases of Delignes Weight Mondromy
Conjecture for p-adic fields.
4.1
Perfectoid fields
Definition 4.1.1. A perfectoid field is a complete valued field k with respect to a nondiscrete rank-one valuation, with residue characteristic p > 0, such that the Frobenius
Fr : k /p k /p is surjective.
A typical example of a perfectoid field is Qp (p ) , the completion of Qp (pn : n > 1)
with respect to the p-adic topology. Similarly, Qp (p1/p ) and Cp = Qp are perfectoid.
Fr
i>0
It is not too difficult to check directly that k [ is a valuation ring, and we put k [ = Frac(k [ ).
There is a canonical map ()] : k [ k defined by
n
(x0 , x1 , . . . )] = lim x
fn p ,
n
where x
fn is an arbitrary lift of xn k / to k . The map ()] is not additive unless k already
has characteristic p, in which case k = k [ . In general, ()] extends to an isomorphism
of multiplicative groups k [ k , and we can use this to define a valuation on k [ by
|x|k[ = |x] |k . See Lemma 3.4 of [Sch12] for a proof that ()] has the claimed properties,
and that k [ is a perfectoid field with the same value group as k.
Example 4.1.3. Let N be a lattice (i.e. a finite free Z-module) and let NR be a strongly
convex polyhedral cone. Let NR be its dual. If we put M = N , then the spectra of
algebras of the form k[] = k[ M ] form affine charts for toric varieties over k. There is
a perfectoid version of this. Write khi = kh M [ p1 ]i for the ring
1
k [ M Z[ ]]
k.
p
Then khi is a perfectoid algebra over k. (Note: khi is not, in this context, a noncommutative polynomial algebra over k.
If A is a perfectoid k-algebra, define A[ in much the same way, via
!
A[ =
lim A /
Fr
200
k[ k [ .
It turns out that if A is perfectoid, then A[ is also perfectoid, and we have the following
deep theorem:
Theorem 4.1.4 (Scholze). The functor ()[ : Perf(k) Perf(k [ ) is an equivalence of
categories.
In fact, much more can be shown, e.g. ()[ induces equivalences of categories between
FEt(A) and FEt(A[ ) for all A, where FEt(A) is the category of finite etale algebras over A
(it turns out that such algebras are perfectoid).
Example 4.1.5. If A = khi = kh M [ p1 ]i as in Example 4.1.3, then A[ = k [ hi =
k [ h M [ p1 ]i.
Theorem 4.1.4 is proved without introducing much heavy machinery in Section 3.6 of
[KL13]. The basic idea is that the map ()] : k [ k induces a ring homomorphism
: W (k [ ) k , where W () is the ring of p-typical Witt vectors. The inverse to the
functor ()[ is A] = W (A ) W (k[ ) k. Scholzes proof is more conceptual, and passes
through a diagram
Perf(k)
/ Perf(k a )
/ Perf(k a /)
/ Perf(k [a )
/ Perf(k [a / [ )
()[
Perf(k [ )
in which most of the categories have yet to be defined. The superscript ()a should be read
almost, following the almost mathematics initially created by Faltings, and developed
systematically in [GR03].
4.2
Almost mathematics
Almost mathematics was first introduced by Faltings in [Fal88], where he proved a deep
conjecture of Fontaine on the etale cohomology of varieties over p-adic fields. We follow the
treatment in Section 2.2 of [GR03].
Let V be a valuation ring with maximal ideal m. Throughout this section, we assume
that the value group of V is non-discrete. This implies m2 = m. In fact, much of the theory
works for any unital ring V with idempotent two-sided ideal m, but we have no need to
work at this level of generality. The reader should keep in mind the example V = k for a
perfectoid field k.
Let Mod(V ) be the category of all V -modules, and let Ann(m) be the full subcategory
consisting of those modules killed by m.
Lemma 4.2.1. Ann(m) is a Serre subcategory of Mod(V ).
Proof. We need to show that if 0 M 0 M M 00 0 is an exact sequence of V -modules
for which M 0 and M 00 are killed by m, then M is also killed by m. We trivially have m2 M = 0,
but m2 = m, whence the result.
201
Definition 4.2.2. The category of V a -modules is the Serre quotient Mod(V a ) = Mod(V )/Ann(m).
Write ()a : Mod(V ) Mod(V a ) for the quotient functor. Even though V a does
not exist as a ring we will write HomV a (, ) for hom-sets in Mod(V a ). While in general,
hom-sets in quotient categories can be difficult to describe, the category Mod(V a ) is relatively
easy to understand via the following theorem.
Theorem 4.2.3. There is a natural isomorphism HomV a (M a , N a ) = HomV (m M, N ).
It follows that Mod(V a ) is naturally a Mod(V )-enriched category. We define two functors
()! , () : Mod(V a ) Mod(V ):
M = HomV a (V a , M )
M! = m M .
Theorem 4.2.4. The triple (()! , ()a , () ) is adjoint. Moreover, these adjunctions induce
natural isomorphisms (M )a = M = (M! )a .
This is suggestive of the situation in which j : U , X is an open embedding of topological
spaces. The restriction functor j : Sh(X) Sh(U ) fits into an adjoint triple (j! , j , j ) in
which j j = 1 = j j! . So we should think of Mod(V a ) as the category of quasi-coherent
sheaves on some subscheme of Spec V , even though there is no such subscheme. In fact,
since Mod(V a ) does not contain enough projectives, it should be thought of as some kind of
non-affine object.
The category Mod(V a ) inherits the structure of a tensor category from Mod(V ). In fact,
we have internal tensor and hom defined by
M a N a = (M N )a
Homa (M a , N a ) = Hom(M, N )a
There is a tensor-hom adjunction Hom(L N, M ) = Hom(L, Homa (M, N )). This allows us
to speak of algebra objects in Mod(V a ) as commutative unital monoid objects in the tensor
category (Mod(V a ), ). We call such objects V a -algebras. It is easy to check that they are
all of the form Aa , for A some V -algebra.
If A is a V a -algebra, we can form the category Mod(A) of A-modules in the obvious way.
This is also an abelian tensor category, so it makes sense to speak of flat objects in the
usual way, i.e. an A-module M is flat if the functor M A is exact.
Definition 4.2.5. Let k be a perfectoid field. A perfectoid k a -algebra is a flat, -adically
complete k a -algebra A for which Fr : A/ 1/p A/ is an isomorphism.
Note that even though 1/p may not exist as an actual element of k , the ideal ( 1/p )
is well-defined, so it makes sense to write M/ 1/p if M is any k - (or k a )-module. Write
Perf(k a ) for the category of perfectoid k a -algebras.
Perf(k a ).
Idea of proof. See the first part of Section 5 in [Sch12]. Besides some technicalities, one has
the existence of an inverse functor A 7 A k k.
202
4.3
We have seen in the last section that localization induces an equivalence between the category
of perfectoid k-algebras and the category of perfectoid k a -algebras. Since k and k [ are
canonically isomorphic via ()] , also write for the element of k [ corresponding to k.
It is easy to check that k / = k [ /. It easily follows that Mod(k a /) = Mod(k [a /).
The idea of this section is to pass from Perf(k a ) to a suitable category of perfectoid
k a /-algebras.
Definition 4.3.1. Let k be a perfectoid field. A k a /-algebra A is perfectoid if it is flat
and Fr : A/ 1/p A is an isomorphism.
Let Perf(k a /) denote the category of perfectoid k a /-algebras. We will show that the
functor A 7 A/ from Perf(k a ) to Perf(k a /) is an equivalence of categories by introducing
an almost version of the cotangent complex. Lets start by recalling the classical theory:
Theorem 4.3.2. There is a functorial way of assigning to a flat map A B of (commutative, unital) rings an object (the cotangent complex) LB/A of D60 (B). This complex satisfies
the following properties:
e A is a
1. There is a functorial way of assigning to a square-zero extension 0 I A
square-zero extension an obstruction class
o(I) ext2 (LB/A , IB )
e
e such that B
e e A = B.
such that o(I) = 0 if and only if there is a flat A-algebra
B
A
e a deformation of B to A.)
e
(One calls such a B
e exists, then the set Def Ae(B) of deformations of B to A
e is
2. If a deformation of B to A
A
a
ext1 (LB/A , IB )-torsor.
e and B
e 0 are deformations of A-algebras B, B 0 to A,
e and if f : B B 0 is an
3. If B
A-algebra map, then there is a functorial way of assigning to f an obstruction class
o(f ) ext1 (LB/A , IB 0 )
e
e e
e0
such that the set Def A
A (f ) of isomorphism classes of lifts of f to f : B B is nonempty
if and only if o(f ) 6= 0.
Perf(k a )
Perf(k a /).
Proof sketch. Let An = k a / n+1 . We content ourselves with showing that objects and
morphisms in Alg(A0 ) lift uniquely to each Alg(An ). Let B0 be a perfectoid A0 -algebra.
Theorem 4.3.3 tells us that LaB0 /A0 = 0, so B0 and any morphisms from B0 to other perfectoid
A0 -algebras lift uniquely to A1 by the almost version of Theorem 4.3.2. All that remains
for the induction to work is to show that LaBn /An = 0 implies LaBn+1 /An+1 = 0. There is an
exact sequence
n
/ B0 / Bn+1
/ Bn
/ 0,
LaB0 /A0
Bn+1 /An+1
/ La
Bn /An
/.
Since LaB0 /A0 = 0 by Theorem 4.3.3 and LaBn /An = 0 by assumption, we get that LaBn+1 /An+1 =
0.
5.1
Construction
The following is originally from Delignes section La formalisme des cycles evanescents in
[DK73].
Consider the diagram
Z
{0}
/Xo
/Co
204
? _U o
? _ C o p
f
X
f
C
f
X X by .
f , C) : The functor is
Recall that we have the adjunction : Sh(X, C) Sh(X
exact, and is left-exact, so we have an adjunction at the level of derived categories:
f ) : R .
: D(X) D(X
Definition 5.1.1. The nearby cycles functor is the functor f : D(X) D(Z) defined by
f (F ) = i R (F ).
f ) = U X, the functor f factors through D(U ). In other words,
Note that since (X
there exists f : D(U ) D(Z) such that f = f j . In other words, f (F ) only
depends on F |U . Moreover, f comes with a natural transformation
: i f
defined by applying i to the adjunction transformation 1 R = f
Definition 5.1.2. The vanishing cycles functor is the functor f : D(X) D(Z) given by
f (F ) = cone(F : i (F ) f (F )).
Proposition 5.1.3. Both f and f restricts to functors on the corresponding subcategories
of perverse sheaves:
f : Perv(U ) Perv(Z)
f : Perv(X) Perv(Z).
5.2
Monodromy
/ R R
/ f
/ f
/ f
/ i [1]
/ i [1].
i (F )
/ f (F ) canF / f (F )
T id
/ f (F )
/ i (F )[1]
varF
f (F )
/0
It follows that
var can = T id.
5.3
Gluing problems
V
.
C
The functor f acts via (V, A) 7 V and T(V,A) : f (V, A) f (V, A) is the operator A.
Theorem 5.3.2 now says that
Perv(X) = {(W, V, E : V W, F : W V ) : F E + id is invertible} .
There is a general way of assigning an algebra q,h (Q) (called the multiplicative preprojective
algebra) to a quiver Q such that
Perv(X) = q,h ( ).
Example 5.3.4. Let n > 1, let X = Cn , and let S = {XI }I[n] , where
XI = {(x1 , . . . , xn ) Cn : xi = 0 if and only if i I}.
Let Perv(X) be the category of perverse sheaves on X with respect to S. Combinatorialists
are interested in perverse sheaves because of the following theorem. We first define a category
A(n) whose objects are tuples
{VI }I[n] , EI,i : VI VI{i} for i
/ I, FI,i : VI{i} VI for i
/I
such that
1. EI{j},i EI,j = EI{i},j EI,i for i
/I
2. FI,j FI{i},i = FI,i FI{i},j for i 6= j both not in I
3. FIr{j},i IFIr{j},j = FI{i}r{j} FI,i for i
/ I and j I
4. FI,i EI,i + id is invertible for all i
Morphisms are {fI : VI VI0 } commuting with the Es and F s.
Theorem 5.3.5. With X as above, there is an equivalence of categories Perv(X) ' A(n) .
Proof. See [GMV96].
Example 5.3.6. Let X = C2 and {0} {xn = ym} C2 be our stratification (we assume
n 6 m). Then there is an equivalence of categories Perv(X) ' A(n,m) , where A(n,m) consists
of tuples (A, B1 , . . . , Bn , C) with arrows
Ao
qk
pk
/ Bk o
Bk+m
sk
rk
k
(mod n)
207
/C
5.4
Monadology
Recall that an exact category (in the sense of Quillen) is an additive category A together
p
i
with a special class E of exact pairs, consisting of diagrams of the form A
B
C such
that (A, i) = Ker(p) and (C, p) = Coker(i). These exact pairs are required to satisfy certain
axioms. [reference earlier section for these]
Example 5.4.1. If A is an abelian category, then we can let E be the class of all short exact
sequences. More interestingly, we could restrict to the class of all split exact sequences.
Example 5.4.2. The category of finitely-generated projective modules over a unital ring is
exact (when given the obvious class of exact pairs).
Example 5.4.3. Let A = Com(B) for some abelian category B. We can set E to be the
class of termwise-split short exact sequences in A.
In an exact category, there is are notions of admissible monic (a morphism which fits
into the first half of an exact pair) and admissible epic (a morphism which fits into the
second half of an exact pair). We write , for admissible monics and for admissible epics.
Fix an exact category A.
+
Proof. We consider the equivalence Ae ' Ae1 . To a sequence P P P+ associate the
0
sequence P Ker(+ ) , P , and in reverse send P1 , P0 ,
P1 to the sequence
0 1
P1 P1 P1 /P0 .
For Ae1 ' Ae2 , consider
(0 ,can)
[P1 P0 P1 ] 7 [P0 P1 P0 /P
1 /P1 ]
...
The motivation for all of this is quite classical. Recall the following theorem in [Bar77].
Let M(n, r) be the moduli space of framed torsion-free coherent sheaves on P2C of rank r
and Chern class c2 = n. Let i : ` , P2 be the inclusion of the line at infinity. Elements
of M(n, r) are pairs (F , ), where F is a rank-r torsion free sheaf on P2C with c2 (F ) = n,
and : i F
OPr
1 is a framing at infinity. There is a natural way to give M(n, r) the
structure of a scheme.
Theorem 5.4.6 (Barth). There is an isomorphism between M(n, r) and the moduli space
of isomorphism classes of triples (B1 , B2 , i, j), where B1 , B2 Mn (C), i : Cr Cn and
j : Cn Cr , satisfying [B1 , B2 ] + ij = 0. Also, the triples are stable in the sense that there
are no S Cn such that B (S) S for {1, 2} and Im(i) S.
The linear-algebraic data in this theorem can be encoded in terms of the quiver [couldnt
live-TEXthe quiver: it has i : W V , j : V W and Bi : V V ]
Sketch of proof. We will construct a functor in one direction. To a tuple (B1 , B2 , i, j), assign
a
b
F = H V OP2 (1)
(V OP2 )2 W OP2
V OP2 (1) .
We can identify P2C with the proj of C[z0 , z1 , z2 ], where we can regard the zi as global
sections of OP2 (1). Even better, we think of the zi as maps OP2 (i) OP2 (i + 1) for all i.
The maps a and b are given by
z0 B z1
a = z0 B 2 z 2
z0 j
b = (z0 B2 z2 ) z0 B1 z1 z0 i
One can check that ba = 0 if and only if [b1 , b2 ] + ij = 0.
Now we define the operation F 7 (B1 , B2 , i, j). Define a rank-two vector bundle E on
P2 by the Euler exact sequence
0 OP2 (1) OP3
2 E 0.
209
We put
F 7 H1 (P2 , F (2)) OP2 (1) H1 (P2 , F (1) E ) OP2 H1 (P2 , F ) OP2 (1) .
There is an example from [BW02]. Let r = 1 and n be any integer. Let Cn be the space
of isomorphism classes of tuples (X, Y, i, j) such that X, Y Mn (C), i Cn and j (Cn )
such that [X, Y ] + 1 = ij. This is generally known as the Calosero-Moser space.
`
There is a bijection between n>0 Cn and the set of isomorphism classes of right ideals
in A = A1 (C) = ChX, Y i/([X, Y ] 1).
5.5
Reflection functors
/A
/ C+
/
+
/ / C+
AB
where + = 0. Finally, let A]2 be the category of special exact sequences, i.e. those of the
form
+
2
+
2
1
2
+
( , , )
/ A B1 B2
/ / D+
D
i ) is mmonic for each i, and such that ( , i ) is epic for each i.
such that ( ,
+ +
Proposition 5.5.2. There are canonical exact equivalences A] ' A]1 ' A]2 .
Proof. We first construct a functor A] A]1 . Send a dyad to the monad
C
( , )
/AB
210
+
+
/ C+ .
A key observation is that on A]2 , here is a natural auto-equivalence r : A]2 A]2 , sending
Q2 to r(Q2 ), which is
D A B 2 B 1 D+ .
Our equivalences transport this to an auto-equivalence of A] , which we call reflection.A
dyad
/
A
C
B
is sent by r to
Ker(+ )
/ C+
/A
H(Q1 )
/ Coker( )
0
+
where
0
Exercise 1. Check that any Barth or C-M monad is naturally special. Compute the
reflection functor on the corresponding dyads.
Exercise 2. Recall the moduli space M(n, 1) of torsion-free coherent sheaves on P2C of
/ (O 2 V )2 O 2
P
P
/ O 2 (1) V.
P
Via the restriction functor j : coh(P2C ) coh(C2 ) ' mod(C[X, Y ]), this corresponds to
0
/ A0 V
/0
/ (A0 V )2 A0
/ A0
211
/ A0 V
/V
/0
/0
/AV
/ (A V )2 A
/AV
/A
/0
/0
/V
/ 0.
The claim is that there is such a quasi-isomorphism in the category mod (A) of A -modules
over A.
This is worked out in the paper [BC07]. It reflects the fact (from commutative algbra)
that any `
ideal in C[x, y] is isomorphic to a unique ideal of finite codimension. So we can
think of n>0 Hilbn (C2 ) as the moduli space of all ideals in C[x, y]. Similarly, we could
`
L A
interpret n Cn as the space of isomorphism classes of right ideals in
2 ~ .
6.1
The setup
%
/ DU
9
/ DX
i! M D0
Z
i M D0
Z
Definition 6.1.2. M PervX is called a tilting perverse sheaf if
i M PervZ
and
i! M PervZ
Lemma 6.1.3. Given any M PervX , then M is tilting perverse if and only if both of the
followings hold:
212
is surjective
j! j M M
is injective
Proposition 6.1.4. Let MU PervU be such that j (MU ) PervX and j! (MU ) PervX .
Then there exists a tilting perverse sheaf M PervX such that j M
= MU .
Sketch of proof. Define A, B PervX by the following exact sequence:
M
U
0 A j! MU
j MU B 0
/ Glue(Z, U )
diads
where:
213
/ diads
Q(M ) :=
j! j M
(j M )
can
can
can
can
%%
j9 j M
(MU )
8
can
+
!
R(MU , , , ) := i i ((MU ))
&
i
6.2
&&
i i (MU )
8
can
a = a, we have an involution
() : k(Q) k(Q)
a 7 a
214
1
:aQ
1 : a Q
{~a }aQ
Y
X
2.
(aa + ~a )(a)
qv ev = 0
vI
aQ
Lemma 6.2.2. Such q,~ (Q) always exists for any quiver Q and any sets of parameters
q = {qv }vI , ~ = {~a }aQ .
In [CBS06], they defined the preprojective algebra only for the case ~a = 1 for all a Q.
Example 6.2.3. Consider the one-loop quiver. (i.e., the quiver Q = Qloop with one vertex
v and one arrow a).
Write qv = q and ~a = ~.
Then the multiplicative preprojective algebra is:
k(Q) q,~ (Q) =
khx, y, (~ + xy)1 i
(xy qyx ~)
khx1 , y 1 i
(xy qyx)
khx, y, (1 + xy)1 i
(xy qyx 1)
q,~ (Q) =
e2i = ei )
(e +
a a
k(Q)
= q e , e0 + aa = q0 e0 )
q,~ (Q )
e 70/ /
k((Qloop ))
/ / q,~ (Qloop )
khx1 ,y 1 i
(xyqyx)
Db (Modq,~ (Qloop ))
l
/ Db (Modq,~ (Q ))
,
/ Db (ModUq,~ (Qloop ))
2
Rep q,~ (Q ), n = (1, n) //GL(n)
n=0
216
Appendix A
Miscellaneous topics
1
Let A be an associative unital ring. Consider the (discrete) group GLn (A) = Mn (A) of
invertible n n matrices with coefficients in A. We call GLn (A) the general linear group
over A. We have embeddings GLn (A) GLn+1 (A) given by
0
7
.
0 1
We define
GL(A) = GL (A) = lim GLn (A).
Put
H (GL(A), k) =
i>0
where BGL(A) is the classifying space of GL(A) (see 2.5.3). The diagonal map GLn (A)
GLn (A) GLn (A) gives H (GL(A), k) the structure of a coassociative cocommutative
coalgebra over k.
Note that H (GL(A), k) has also the structure of a graded-commutative k-algebra. There
is a natural map taking direct sum : GLn (A) GLm (A) GLn+m (A), defined by
A 0
AB =
, A GLn (A), B GLm (A)
0 B
It induces a map : GL(A) GL(A) GL(A), which in turn induces a map BGL(A)
BGL(A) BGL(A). The latter map turns out to be associative and commutative up to
homotopy. It follows that H (GL(A), k) naturally has the structure of a graded k-algebra.
We want to define a functor of points GrComAlgk Set for which the representing
object is H (GL(A), k). First we need to extend the notion of representation. Let G be a
(discrete) group. We can think of a group G as a category (actually, groupoid) with one
object.
217
Let A be any (small) additive category. Let Isom(A) be the groupoid of isomorphisms
in A. That is, the only morphisms in A which we keep in Isom(A) are isomorphisms. Let
S = 0 (Isom(A)) be the set of isomorphism classes of objects in A.
Definition 1.0.8. A representation of G in A is a functor : G Isom(A).
We denote by Isom(G, A) the groupoid Isom(Fun(G, A)) of isomorphism classes of representations of G in A. That is,
a
Isom(G, A) =
Hom(G, Out(Ps )),
sS
There is a canonical map : limn [G, GLn (A)] [G, GL(A)]. Concretely, let E : G
Aut(P ) represent an element of St(G, A). Choose Q with P Q ' An . We let E be the
composite
E
G
GLA (P ) GLA (P Q) = Aut(An ) = GLn (A) , GL(A).
Remark 1.0.13. Our construction is parallel to the topological situation. Namely, let X be
a paracompact topological space. We replace representations of G in A by complex vector
bundles on X. Let VB(X) be the set of`isomorphism classes of vector bundles on X. There
is an obvious decomposition VB(X) = n VBn (X), where VBn (X) consists of isomorphism
classes of n-dimensional vector bundles on X. In fact, VBn (X) = [X, BUn ], where BUn is
e 0 (X) := [X, BU ],
the classifying space of the group Un . Instead of St(G, A), we think of K
the (reduced) topological
K-theory. The analogue of our map : St(G, A) [G, GL(A)]
`
e 0 (X), which is an isomorphism if X is compact or a
is the map [X, n VBn (X)] K
finite-dimensional CW complex.
L
Let M = i>0 Mi be a graded abelian group. For an arbitrary group G, define
Y
H0 (G, M ) =
Hi (G, Mi ),
i>0
f : Isom(, M 0 ) H0 (, M )
H0 (, M 0 ). One says that f is induced from by f .
The theorem asserts that there is a natural bijection between stable exponential characteristic
classes in A with coefficients in M and algebra maps H (Isom(A)) M .
Generalized manifolds
The ideas in this example come from Freed and Hopkins paper [FH13].
Let Man be the category of smooth (finite-dimensional) manifolds and smooth maps.
A generalized manifold is a sheaf on Man. That is, a generalized manifold is a functor
F : Man Set such that whenever M Ob Man has an open cover {U }I , the following
diagram is an equalizer:
F (M )
F (U )
//
F (U U )
,I
The main idea is to try to extend differential geometry to the category of generalized smooth
manifolds. One example is differential forms. We have theVfunctor : Man Set, which
assigns to a manifold M the de Rham complex (M ) = 1 (M ).
g allows us to treat any manifold as a generalized
The Yoneda embedding Man , Man
manifold. Recall that the Yoneda lemma shows that there is a natural bijection
HomMan
g (hX , F ) ' F (X).
Definition 2.0.17. For any generalized manifold F , we can define the de Rham complex
of F by
(F ) = HomMan
g (F , ).
220
The Yoneda lemma tells us that this definition agrees with the usual one if F is
representable. Can we compute the de Rham cohomology
L of F ? We have a direct-sum
decomposition of functors on generalized manifolds, = q q . Let F = 1 : M 7 1 (M ).
It is a highly nontrivial theorem that (1 ) is isomorphic to
R
/R
/R
I
I
I
2.0.17, (X ) = HomMan
g (X , ). Having a morphism of functors (X ) is the same
as having a compatible family of maps N : X I (N ) (N ). But this is exactly how Chen
defined differential forms on X I . So the two constructions are coherent.
Now take forms 1 , . . . , k on X of degrees s1 , . . . , sk respectively. For any smooth map
f : N P (X) these forms define pull-back forms f (1 ), . . . , f (k ) on N I. Each of
these pull-backs can be written as f (i ) = i0 (n) + i00 (n)dti . Then we can integrate the
form 100P
(n1 ) . . . k00 (nk )dt1 . . . dtk over the simplex k1 = {(t1 , . . . , tk ) I k | t1
tk and
ti = 1} to obtain a (s1 1) + + (sk 1)-form on N N = N k . Then
i
RR
One of the main results in Chens paper is that
induces a morphism of Hopf algebras
c
3.1
Basic definitions
First we recall (see 1.1.6) basic definitions about quivers. Also, it will fix the notations.
Definition 3.1.1. A quiver is just a directed graph with finitely many vertices. More
formally, a quiver is a quadruple Q = {Q0 , Q1 , s, t}, where Q0 is the (finite) set of vertices,
Q1 is the set of arrows, and s, t : Q1 Q0 are the incidence maps assigning to an arrow
its source (resp. target).
For example, a quiver may look like
Let Path(Q) be the set of paths in Q. We can naturally extend s and t to maps
s, t : Path(Q) Q0 by
s() = s(m )
t() = t(1 ).
Definition 3.1.6. The length function ` : Path(Q) Z+ is a map where `() is the number
of arrows in . We put `(e) = 0 if e is a vertex.
3.2
Definition 3.2.1. Let Q be a quiver. The path algebra of Q over k, denoted kQ, is the
k-vector space with basis Path(Q). The product on kQ is given by
(
if t() = s()
=
0
otherwise
Example 3.2.2. Let the vertices of Q be {1, 2, 3}, with arrows : 1 2 and : 2 3.
Then Path(Q) = {e1 , e2 , e3 , , , }, so kQ is the 5-dimensional k-algebra spanned by the
above elements. Some basic computation yields the following presentation of kQ:
kQ = khe1 , e2 , e3 , , , : eij = ij ei , e1 = e2 = , e1 = e2 = e3 = e3 = = 0i
It is easy to see that e1 + e2 + e3 = 1 in kQ.
Example 3.2.3. Let Q be the quiver with a unique edge and vertex. Then kQ ' k[x],
where the unique vertex corresponds to q, and the edge corresponds to x. Similarly, let Qn
be the unique quiver with a single vertex and n edges. One has kQn = khx1 , . . . , xn i.
Example3.2.4. In general, if A and B are k-algebras and M is an (A, B)-bimodule, then
A M
we write
for the algebra, which as a vector space is isomorphic to A M B,
0 B
and with multiplication given by
a1 m1
a2 m2
a1 a2 a1 m2 + m1 b2
=
.
0 b1
0 b2
0
b1 b2
A 0
.
N B
k 0
If now Q is the Kronecker quiver 1 2, then kQ '
. The isomorphism is
k 2 k
given by
1 0
0 0
0
0
0
0
e1 7
, e2 7
, 1 7
, 2 7
,
0 0
0 1
(1, 0) 0
(0, 1) 0
Similarly we define for (B, A)-bimodule N an algebra
223
Example 3.2.5. Let Q be a quiver with at most one path between any two vertices. Write
n = #Q0 . Then kQ ' {A Mn (k) : Aij = 0 if there is no path j i}. For example, if
Q = 1 2 n, then kQ is the algebra of n n lower-triangular matrices.
Example 3.2.6. Finally, let Q be the quiver
k[x] k[x]
Then kQ '
. This is called a framed 1-loop quiver. Path algebra Path(Q) gives
0
k
an example of the ring which is right Noetherian, but not left Noetherian.
L
Let Q be
a
quiver.
We
write
kQ
=
0
eQ0 ke kQ. This is a semisimple k-algebra.
L
Let kQ1 = :ij k be the span of all arrows in Q.
Lemma 3.2.7. Let Q be a quiver. Then kQ1 is naturally a kQ0 -bimodule, and kQ '
TkQ0 (kQ1 ).
Proof. Recall that if S is a k-algebra and M is an S-bimodule, then the tensor algebra
TS M satisfies the following universal property: given any k-algebra map f0 : S A and
an S-bimodule map f1 : M A, there is a unique f : TS M A such that f |S = f0 and
f |M = f1 . We apply this to the case S = kQ0 , M = kQ1 . These both embed into kQ, giving
a map TS M kQ. Surjectivity follows from the definition of kQ, and injectivity follows
from induction on the grading.
3.3
First of all, note that {ei : i Q0 } is a complete set of orthogonal idempotents in kQ, i.e.
(
X
ei if i = j
ei = 1,
ei ej =
0 otherwise
i
Moreover, if A = kQ, then Aei is the span of all paths starting at i, while ej A is the span of
all paths ending at j. L
Thus ej Aei is the span of all paths starting
L at i and ending at j. As
a left A-module, A = i Aei , and as right A-modules, A = i ei A. This implies that the
Aei are projective left ideals and the ei A are projective right ideals.
Lemma 3.3.1. Let Q be a quiver and put A = kQ.
(a) For any left A-module M and right A-module N ,
HomA (Aei , M ) ' ei M and HomA (ej A, N ) ' N ej .
(b) If 0 6= a Aei and 0 6= b ei A, then ab 6= 0.
(c) If ej Aei A, then i = j.
224
3.4
Representations of quivers
Definition 3.4.1. For any quiver Q and a field k, define the category Repk (Q) to be just
the category Fun(Q, Vect) of functors from the quiver Q viewed as a category to the category
of vector spaces.
Explicitly, objects of Repk (Q) are assignments i 7 Xi Vectk , together with k-linear
maps X : Xs() Xt() for any edge Q1 . A morphism : X X 0 is given by a
collection of k-linear maps {i : Xi Xi0 }iQ0 such that for all edges Q1 , the following
diagram commutes.
Xs()
/ Xt()
s()
0
Xs()
t()
/ X0
t()
Proposition 3.4.2. There is a natural equivalence of categories kQ-Mod ' Repk (Q). Moreover, Mod(kQ) ' Repk (Q ), where Q is the opposite quiver of Q, with the same vertices
and edges as Q, but s = t, t = s.
225
Proof. There is an obvious functor F : kQ-Mod Repk (Q), which assigns to a kQ-module
X, the Q-representation F (X) with F (X)i = ei X. Given a morphism f : X X 0 of
kQ-modules, we let F (f ) : ei X ei X 0 be the restrictionL
to ei X of f : X X 0 .
Conversely, given {Xi } Ob Repk (Q), define X =
iQ0 Xi . There are canonical
projection and injection maps i : Xi , X and i : X Xi . Give X a kQ-module structure
by
(1 m ) x = t(1 ) X1 Xm s(m ) x
The similar proof works for the category of right kQ-modules Mod(kQ).
If A is a k-algebra, write Irr(A) for the set of isomorphism classes of irreducible Amodules. Let Ind(A) be the set of isomorphism classes of indecomposable projectives over
A.
Theorem 3.4.3. Assume Q has no oriented cycles. (Equivalently, A = kQ has finite
dimension over k.)
1. The following assignments are bijections:
Q0 Irr(A)
i 7 S(i)
Q0 Ind(A)
i 7 Aei ,
where S(i)j = k ij .
2. There is a natural linear isomorphism for every i, j Q0 :
Ext1A (S(i), S(j)) ' Span Q(i, j)
where Q(i, j) is the set of all arrows from i to j.
3. ExtkA (M, N ) = 0 for all k > 2 and all A-modules M, N .
Remark 3.4.4. Part 1 of the theorem obviously fails if Q has oriented cycles. For example,
let Q consist of a single loop. Then kQ = k[x], and Irr(A) = Spec(A) k, which is much
larger than Q0 .
Before proving Theorem 3.4.3 we will need some other lemmas and facts.
Proposition 3.4.5. For any Q and any left A-module X, the following sequence is exact.
M
M
f
g
/
/
/X
/0
0
Aet() es() X
Aei ei X
()
Q1
iQ0
Remark 3.4.6. Note that () is a projective resolution because each Aei k V is a direct
summand of the free module A k V , where V = kQ1 . Moreover, if we put X = A, then
we get a L
projective bimodule resolution of A. Indeed, Aei ej A is a direct summand of
A A = i,jQ0 Aei ej A, which is a rank one A-bimodule.
Proof. (of Prop.3.4.5) First we show that g is surjective. This can be easily seen from the
fact that any element of X can be written as
X
X
X
x=1x=
ei x =
ei x = g
ei ei x .
iQ0
iQ0
iQ0
Ln
i=1 Aei
ei X can be written
i=1 paths a
s(a)=i
where all but finitely many of xa es(a) X are zero. Let deg() be the length of the longest
path a such that xa 6= 0. If a is a nontrivial path, we can factor it as a = a0 , with
s(a0 ) = t() and a0 a single edge. We have a0 xa = a0 es(a) es(a) x = a0 et() es(a) . By
definition,
f (a0 xa ) = a0 xa a0 xa = a xa a0 xa
We claim that for any , the set +Im(f ) contains elements of degree zero. For, if deg() = d,
then
n
X
X
f
a0 xa
i=1 s(a)=i
`(a)=d
has degree strictly less than d. The claim follows by induction on `(a) = d.
Let
now Ker(g), and take 0 + Im(f ) an element of degree zero. In other
P
P words,
n
0) =
0 = i=1 eL
x
.
If
g()
=
0,
then
because
g
f
=
0,
we
get
g()
=
g(
xei , an
i
ei
n
element of i=1 ei X, which can be zero if and only each xei = 0. But each xei = 0 implies
0 = 0, i.e. Im(f ). This finishes the proof that Ker(g) = Im(f ).
Finally, we prove that Ker(f ) = 0. Suppose f () = 0. Then we can write
=
Q1
path a
s(a)=t()
a x,a = a x,x + ,
227
XX
a x,a
XX
i i
iQ0
t() s() .
Q1
/ HomA (X, Y )
HomA (Aei ei X, Y )
iQ0
/ Ext1 (X, Y )
A
Q0
M
M
Homk (ei X, ei Y )
iQ0
Q1
/Y
/Z
/X
/0
()
/0
Proof. Suppose () splits. Then there exists : X Z such that = idX . Then (X) Z
is a submodule 6= (Y ) because (X) = X 6= (Y ) = 0.
Conversely, suppose K is a proper submodule of Z and K 6= (Y ). Define = |K : K
X. Since X is simple, is either 0 or surjective. If = 0, then K Ker = Im = (Y ),
a simple module. This implies K = (Y ), a contradiction. Thus is surjective. If k K
is killed by , then K (Y ) 6= 0, so K = (Y ), a contradiction. We have shown that
: K X is an isomorphism. We define the splitting by = 1 .
Lemma 3.4.12. Let X, Y be two non-isomorphic, simple A-modules. Then
a) If A is commutative, then every extension of X by Y splits.
b) If z Z(A) such that z AnnA (Y ) \ AnnA (X), then every extension of X by Y
splits.
Proof. a) Note that if A is commutative and 0 6= x X, then the map A X given by
a 7 ax is surjective, with kernel a maximal ideal mx . After choosing nonzero y Y , then
X ' Y iff mx = my . Let Z be an extension of X by Y . Choose a my \ mx , and define
b
a : Z Z by z 7 az. Then b
a(Y ) = 0 and X ' Z/ Ker(b
a) ' X. This lets us create a
splitting of this exact sequence.
b) Define b
a : Z Z as in part (a). Everything else works.
We return to the proof of Theorem 3.4.3.
Proof. The only non-trivial part of the Theorem 3.4.3 is about Exts.
Lemma 3.4.13. Let Q be a quiver, and let i, j be vertices in Q. Then there exists an arrow
: i j if and only if there is a nonsplit extension
0
/ S(j)
/X
/ S(i)
/ 0.
/ S(j)
/X
/ S(i)
/ 0.
0 1
if 0 =
X0 =
0 0
0
otherwise
The opposite direction is obvious.
229
(A.1)
=R
/ R/I
A-bimodule M,
n > 2
Proof. Let 1 (A) = Ker(A A A). Then there is a short exact sequence of bimodules:
0
/ 1 (A)
/AA
/A
/ 0.
The fact that A is formally smooth tells us that 1 (A) is a projective bimodule. Tensor by
M for some left A-module M , to get
0
/ 1 (A) A M
/AM
/M
/ 0.
1
(we have exactness on the left because TorA
1 (A, M ) = 0 since A is projective.) Since (A)
is projective, so is 1 (A) M , whence the result.
231
232
Appendix B
Exercises
1
Exercise 1. For a morphism of sheaves : F G show that the presheaf Im() defined
in section 3.3 is actually a sheaf.
Exercise 2. Show that if X is an irreducible topological space, then any constant sheaf on
X is flabby.
Exercise 3. Suppose G is a group and A is G-module. Check that the set Z 1 (G, A) =
{f C 1 (G, A) : d1 f = 0} of 1-cocycles is exactly the set of derivations d : G A, i.e.
Z 1 (G, A) = Der(G, A). Moreover, the set B 1 (G, A) = Im(d0 ) of 1-coboundaries is exactly
the set of inner derivations.
Exercise 4. Let A be a k-algebra for a commutative ring k, and (M , dM ), (N , dN ) be
two complexes of left A-modules. In 2.5 we defined graded Hom-space HomA (M, N ). Prove
that if M is a finitely generated (as A-module) then HomA (M, N ) = HomA (M, N ).
Exercise 5. Show that any associative star product (see 3.2) on At = A k kJtK is unital,
and that for any , there exists 0 such that 10 = 1A .
Exercise 1. Let A be a ring, M an A-module. Prove the dual basis theorem 2.3.6. Moreover,
show that M is a generator for the category Mod(A) in the sense of definition 2.3.5 if and
only if M M = A.
Exercise 2. Show that if A and B are commutative rings, then A and B are Morita
equivalent if and only if they are isomorphic.
Exercise 3. Prove the properties 2.4 of functors (i , i , i! ) and (j! , j , j ) from the yoga
of six functors.
233
// Y .
Exercise 13. If A is an abelian category show that Com(A) is also an abelian category.
More generally, show that for any small category I the category Fun(I, A) is naturally an
abelian category.
234
i0
/X I
/ Cyl(f )
3.1
Commensurable subspaces
Exercise 1. Let A and B be subspaces of V . Prove that the following conditions are
equivalent:
(i) dim (A + B)/(A B) <
(ii) dim A/(A B) < and dim B/(A B) <
(iii) dim (A + B)/A < and dim (A + B)/B < .
Call the subspaces A and B commensurable (and write A B) if they satisfy the above
conditions.
Exercise 2. Prove that commensurability is an equivalence relation on the set of subspaces
of V . (Hint: prove first that if A B and B C then dim (A + B + C)/(A B C) < .)
Exercise 3. Let A, B, A0 and B 0 be subspaces in V such that A A0 and B B 0 . Prove
then that
A + B A0 + B 0 and A B A0 B 0 .
235
Exercise 4. Given two commensurable subspaces A and B in V , define the relative dimension
of A and B by
[A|B] := dim A/(A B) dim B/(A B).
Prove that if A, B and C are pairwise commensurable, then
[A|B] + [B|C] = [A|C].
Exercise 5. Let A, B, A0 and B 0 be subspaces in V such that A A0 and B B 0 . Prove
that
[A|A0 ] + [B|B 0 ] = [A B|A0 B 0 ] + [A + B|A0 + B 0 ].
3.2
Traces
g1 , g2 Endfin (V ).
1 , 2 k.
3.3
Residues
Exercise 14. Let f, g End(V, A). Let U be a subspace of V containing A and invariant
V (f, g) = U (f, g). (Because of this property we can drop
under f and g. Prove that ()
()
V .) Check that, if A is invariant under f and g, then
the superscript V in ()
() (f, g) = 0.
Exercise 15. Let A0 be another subspace of V , and let 0 : V A0 be a projection onto
A0 . Assume that A A0 . Then 0 Endfin (V, A), and
(0 ) (f, g) () (f, g) = trV ( 0 )[f, g] .
(Hint: Use property 11 of 3.2.)
Exercise 16. Using 15, prove that, if f, g End(V, A) are such that [f, g] = 0, then
() (f, g) is independent of the choice of the projection : V A and depends only on the
commensurability class of A. (We will write A (f, g) instead of () (f, g) when [f, g] = 0.)
Exercise 17. Prove that if f, g End(V, A) commute and if dim A < or dim V /A < ,
then A (f, g) = 0.
Exercise 18. Let A and B be subspaces of V . Choose projections A : V A, B : V B
and AB : V A B such that A + B AB is a projection A+B of V onto A + B
(this is clearly always possible). Prove that, for any f, g End (V, A) End (V, B), the
following formula holds
A (f, g) + B (f, g) = A+B (f, g) + AB (f, g).
Conclude that, if f, g End (V, A) End (V, B) are such that [f, g] = 0, then
A (f, g) + B (f, g) = A+B (f, g) + AB (f, g).
This last formula is called the (abstract) Tate Residue Formula.
237
3.4
We denote the Lie algebras End(V, A), Endfin (V, A) etc. by gl(V, A), glfin (V, A) etc. respectively, with the usual (commutator) bracket.
Exercise 19. Check that () is a 2-cocycle on the Lie algebra gl(V, A). Hence it defines a
canonical central extension
/k
e
/ gl(V,
A)
/ gl(V, A)
/0
Exercise 20. Show that the cohomology class cA := [() ] H2 (gl(V, A), k) is independent
of the choice of and depends only on the commensurability class of A. (Hint: Use 15 from
3.3.)
Exercise 21. Show that cA induces a cohomology class cA H2 (gl(V, A)/glfin , k).
3.5
The above formalism comes from algebraic geometry. In what follows, we briefly outline
a classic construction of residues of differential forms on curves due to Tate [Tat68]. This
material requires familiarity with basic algebraic geometry.
Let X be a smooth connected algebraic curve over k, which we now assume to be
algebraically closed. Let K = k(X) be the field of rational functions on X. For a (closed)
point x X, let Ox denote its local ring. Write Ax = Ox for the completion of Ox and Kx
for the field of fractions of Ax . Note that Ax is canonically a subspace of Kx , so we can
consider End(Kx , Ax ) defined as in 3.2.
Now, choose a local parameter (coordinate) t at x and identify Ax ' kJtK and Kx ' k((t))
in the usual way, where kJtK and k((t)) are the rings of formal power and Laurent series in t,
respectively. In addition, identify the elements of Kx with the corresponding multiplication
operators on Kx : this gives us the embedding
: Kx , End(Kx )
f 7 [f : g 7 f g].
Kx
K
A
(f, g) = O(S)
(f, g) f, g Kx .
x
xS
238
xS
Ox K. Prove that
(
VS =
(fx )
xS
(The elements of VX are called adeles on the curve X.) Note that K embeds diagonally in
VS . Using the short exact sequence of sheaves
0
/ OX
/ F0
/ F1
/ 0.
239
240
Bibliography
[AZ94]
[Bae40]
Reinhold Baer. Abelian groups that are direct summands of every containing
abelian group. Bull. Amer. Math. Soc., 46:800806, 1940. 137
[Bar77]
[BBD82]
[BC07]
[BCE08]
[Bei78]
A. A. Beilinson. Coherent sheaves on Pn and problems in linear algebra. Funktsional. Anal. i Prilozhen., 12(3):6869, 1978. 126
[Ber74]
[BGG78]
[BKS03]
[BL04]
Societe
[BO95]
[BRT13]
Yuri Berest, Ajay Ramadoss, and Xiang Tang. The Picard group of a noncommutative algebraic torus. J. Noncommut. Geom., 7(2):335356, 2013. 214
[Buc86]
[BW02]
Yuri Berest and George Wilson. Ideal classes of the Weyl algebra and noncommutative projective geometry. Int. Math. Res. Not., (26):13471396, 2002. With
an appendix by Michel Van den Bergh. 210
[CB92]
[CBS06]
[Che77]
Kuo-Tsai Chen. Iterated path integrals. Bulletin of Am. Math. Soc., 83, Number
2:831879, 1977. 221
[CQ95]
[DHKS04] William G. Dwyer, Phillip S. Hirschhorn, Daniel M. Kan, and Jeffry H. Smith.
Homotopy limit functors on model categories and homotopical categories, volume
113 of Mathematical Surveys and Monographs. American Mathematical Society,
Providence, RI, 2004. 128
[DK73]
[DS95]
[Dyc]
[EH00]
David Eisenbud and Joe Harris. The geometry of schemes, volume 197 of Graduate
Texts in Mathematics. Springer-Verlag, New York, 2000. 17
[Eis80]
[ES53]
[Fal88]
Gerd Faltings. p-adic Hodge theory. J. Amer. Math. Soc., 1(1):255299, 1988.
201
[FH91]
William Fulton and Joe Harris. Representation theory, volume 129 of Graduate
Texts in Mathematics. Springer-Verlag, New York, 1991. A first course, Readings
in Mathematics. 75
[FH13]
Daniel S. Freed and Michael J. Hopkins. Chern-Weil forms and abstract homotopy
theory. Bull. Amer. Math. Soc. (N.S.), 50(3):431468, 2013. 220, 221
[FL98]
Daniel R. Farkas and Gail Letzter. Ring theory from symplectic geometry. J.
Pure Appl. Algebra, 125(1-3):155190, 1998. 43
[Gab62]
Pierre Gabriel. Des categories abeliennes. Bull. Soc. Math. France, 90:323448,
1962. 134
[Gin05]
[GM03]
[GMV96] Sergei Gelfand, Robert MacPherson, and Kari Vilonen. Perverse sheaves and
quivers. Duke Math. J., 83(3):621643, 1996. 171, 207
[GR03]
Ofer Gabber and Lorenzo Ramero. Almost ring theory, volume 1800 of Lecture
Notes in Mathematics. Springer-Verlag, 2003. 201
[Hak72]
[Ill71]
[Kela]
[Kelb]
[Kel03]
[KKP08]
[KL13]
Kiran Kedlaya and Ruochuan Liu. Relative p-adic hodge theory, i: Foundations.
2013. arXiv:1301.0792. 201
[Kra07]
[LS75]
[Mat89]
[Maz73]
Barry Mazur. Notes on etale cohomology of number fields. Ann. Sci. Ecole
Norm.
Sup. (4), 6:521552, 1973. 79
[MLM92] Saunders Mac Lane and Ieke Moerdijk. Sheaves in geometry and logic. Universitext. Springer-Verlag, New York, 1992. A first introduction to topos theory.
15
[MV88]
[Nee]
[Oka50]
Kiyoshi Oka. Sur les fonctions analytiques de plusieurs variables. VII. Sur
quelques notions arithmetiques. Bull. Soc. Math. France, 78:127, 1950. 17
[Orl04]
D. O. Orlov. Triangulated categories of singularities and D-branes in LandauGinzburg models. Tr. Mat. Inst. Steklova, 246(Algebr. Geom. Metody, Svyazi i
Prilozh.):240262, 2004. 177, 178, 182
[OSS11]
[Qui73]
[Rei10]
[Sch12]
[Ste75]
Bo Stenstrom. Rings of quotients, volume 217 of Die Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, New York, 1975. 134
[Tab05]
[Tat68]
[Toe]
Bertrand Toen. The homotopy theory of dg-categories and derived morita theory.
arXiv:math/0408337. 177, 186
[Toe11]
[Ver96]
245