You are on page 1of 42

1 Groebner Bases

Diane Maclagan
Hill 240 maclagan@math
Office Hour M1-2
Text: Eisenbud, recommended: Atiyah-Macdonald
Outline:
Gröbner Bases
Localization
Associated Primes
Integral Dependence
Blow-ups/filtrations
Flatness
Completion
Dimension
CM rings
Gröbner Basics
Definition 1.1 (Affine Variety). Let S = k[x1 , . . . , xn ] and let I be an ideal of
S. Fact: S is Nötherian so I = hf1 , . . . , fk i.
The affine variety V (I) = {v = (v1 , . . . , vn ) ∈ k n : f (v) = 0∀f ∈ I} = {v ∈
k n : fi (v) = 0, 1 ≤ i ≤ k}.
For example, S = k[x, y] and I = hx − yi gives the line y = x and V (y 2 −
x3 + x) is an elliptic curve.

Definition 1.2 (Projective Space). Projective space, Pnk is the set of lines
through the origin in k n+1 . We write a point in Pn as v = (v0 : . . . : vn )
and v ∼ v 0 if v = λv 0 for some λ ∈ k ∗ .
Definition 1.3 (Projective Variety). If we set S = k[x0 , . . . , xn ], and grade
S by deg(xi ) = 1 for 0 ≤ i ≤ n, then a polynomial is homogeneous if every
monomial has the same degree. In fact, if f ∈ S is homogeneous of degree d,
we can ask if f (~v ) = 0 for ~v ∈ Pn , because f (λ~v ) = λd f (~v ).
A Projective Variety is V (I) = {~v ∈ Pn : f (~v ) = 0∀f ∈ I homogeneous }
So, we ask the question, given I ⊂ k[x1 , . . . , xnP ] = S, f ∈ S is f ∈ I? ie, if
r
I = hf1 , . . . , fr i, are there h1 , . . . , hr ∈ S with f ∈ i=1 hi fi .
2 2
e.g., is x + 7 ∈ hx − 4x + 3, x + x − 2i ⊂ k[x]? Well, k[x] is a PID, so
I = hf i for some f ∈ k[x], and x + 7 ∈ I iff f |(x + 7). So we use the Euclidean
Algorithm.
The Euclidean Algorithm gives x − 1 as the gcd, so I = (x − 1), so x + 7 ∈ /I
as x − 1 6 |x + 7.
How about in several variables? Is x + 3y − 2z ∈ hx + y − z, y − zi, ie, is
(1, 3, −2) ∈ h(1, 1, −1), (0, 1, −1)i? We can use Gaussian Elimination to say no.
Now, is xy 2 − x in hxy + 1, y 2 − 1i? Naive attempts at division don’t work,
but xy 2 − x = x(y 2 − 1), so it IS in the ideal.

1
Definition 1.4 (Gröbner Basis (Vague)). A Gröbner Basis for an ideal I is a
generating set for which long division decides the ideal membership problem.
Definition 1.5 (Term Order). A term order is a total order on the monomials
in S = k[x1 , . . . , xn ] such that

1. 1 < xu := xu1 1 . . . xunn for all u ∈ Nn \ {0}


2. xu < xv ⇒ xu+w < xv+w .

Examples include lexicographic ordering, that is, xu < xv if the first entry
of v − u is positive, for example y 7 < xz 2 < x2 .
Graded lex, xu < xv if deg(xu ) < deg(xv ) or deg(xu ) = deg(xv ) and xu <lex
v
x .
Reverse Lex (degree revlex), xu < xv if deg(xu ) < deg(xv ) or the last nonzero
element of v − u is negative. eg, xz < y 2 < z 3 .
Definition 1.6 (Lead Term, Initial Ideal). The leading term of a polynomial
f ∈ I is the largest monomial appearing in it with respect to a term order, e.g.
f = 3x2 − 7xy + 8z 2 with lex gives in< (f ) = x2 .
The initial ideal in< (I) = hin< (f ) : f ∈ Ii. WARNING: if I = hf1 , . . . , fr i,
then in< (I) ⊇ hin(f1 ), . . . , in(fr )i but they are not, in general, equal.
eg, I = hxy + 1, y 2 − 1i, x + y ∈ I, y(xy + 1) − x(y 2 − 1), so if < is lex,
in(x + y) = x ∈ in(I) 6= hxy, y 2 i.
Definition 1.7 (Gröbner Basis). A Gröbner Basis for I ⊂ S is a generating
set G = {g1 , . . . , gs } for I for which in(I) = hin(g1 ), . . . , in(gS )i.
Point: We can define a division algorithm. Order the Gröbner basis and
divide the polynomial by multiplying elements of the Gröbner basis to cancel
the leading term of f if possible, otherwise pass to the next monomial, etc.
If G is a Gröbner basis, then division by G will have remainder 0 if and only
if the polynomial is in the ideal.
Facts: Every ideal in k[x1 , . . . , xn ] has a (finite) Gröbner basis, and there
exists an algorithm called the Buchberger Algorithm to compute it.
Definition 1.8 (Division Algorithm). Input: f , {g1 , . . . , gk }
Output: Remainder on dividing f by {g1 , . . . , gk }.
Set f 0 = f , r = 0. While in(f 0 ) ∈ hin(g1 ), . . . , in(gk )i, let j be the smallest
index for which in(f 0 ) = xu in(gj ). Set f 0 = f 0 − lc(f 0 )/lc(gj )xu gj
If f 0 = 0, return r otherwise r = r + lc(f 0 )in(f 0 ) and f 0 = f 0 − lc(f 0 )in(f 0 ),
and return to the while loop.
Note: this algorithm terminates because a term
P order has no infinite descend-
ing chains. Also, this algorithm writes f = hi gi + r for some polynomials
hi ∈ S with in(hi gi ) ≤ in(f ).
Proposition 1.1. If G = {g1 , . . . , gs } is a Gröbner basis for I then the remain-
der on dividing f by G is 0 iff f ∈ I.

2
P
Proof. If r = 0 then f ∈ I since f = hi gi .
Conversely, if f ∈ I, then in the while lop, f 0 ∈ I and we only leave it when
f 0 = 0, so r = 0.
lcm(in(f ),in(g))
Definition 1.9 (S-Pair). If f, g ∈ S their S-pair is S(f, g) = lc(f )in(f ) f −
lcm(in(f ),in(g))
lc(g)in(g) g.
x2 y x2 y
eg, if f = 3x2 −7y 2 and g = 8xy+z 2 , so S(f, g) = 2 2
3x2 (3x −7y )− 8xy (8xy+
z 2 ) = − 37 y 3 − 18 xz 2 .
The S is for syzygy.
Algorithm 1 (Buchberger). Input: {f1 , . . . , fs } generating I, and a term order
<.
Output: A Gröbner basis for I wrt <.

1. Current={(fi , fj ) : 1 ≤ i < j ≤ s}, G = {f1 , . . . , fs }


2. While Current6= ∅, do Pick (f, g) ∈Current, Current=Current\(f, g), r =remainder
on dividing S(f, g) by G . If r 6= 0, then G = G ∪{r} and Current=Current∪{(r, f ) :
f ∈ G }.
3. Output G

Corollary 1.2. If {g1 , . . . , gs } ⊂ I and hin(g1 ), . . . , in(gs )i = in(I) then {g1 , . . . , gs }


generate I.
Definition 1.10 (Minimal Gröbner Basis). A GB is minimal if each in(gi )
is a minimal generator of in(I) and each minimal generator appears once in
{in(gi )}.
Definition 1.11 (Reduced Gröbner Basis). A Gröbner basis is reduced if for
all g ∈ G remainder on dividing g by G \ {g} is g.
Example: There is a unique reduced GB for each term order.
Proof. We must check that Buchberger’s Algorithm terminates and gives the
correct answer.
At stage i of the algorithm, set Ii = hin(g) : g ∈ G i. Note that Ii+1 ) Ii .
If the algorithm did not terminate, we would get an infinite ascending chain
I1 ( I2 ( I3 ( . . . which would contradict the fact that S is Nötherian.
P
If the output is not a Gröbner basis, then there is f ∈ I, f = hi gi with
gi ∈ G , hi ∈ S with in(f ) ∈ / hin(g) : g ∈ G i. Write m = max in(hi gi ), we will
assume that m is minimal for such a counterexample. Let Im = {i : in(hi gi ) =
m}, since m ∈ h∈ (g) : g ∈ G i, we must have in(f ) < m. Thus, |Im | ≥ 2.
Second assumption, Im is minimal forP such expressions.
Pick i, j ∈ I . Write S(gi , gj ) = pj gk , pk ∈ S. Since in(gi ), in(gj )|m,
lcm(in(gi ), in(gj ))|m, so there exit c ∈ k, m0 monomial such that in(cm0 `i gi ) =
in(hi gi ) = in(cm lj gj ), so cm`i gi = cm (`j gj + pk gk ), where in(cm0 pk gk ) =
0 0
P
m0 in(pk gk ) < m.

3
Replace hi gi + Phj gj by (hi − cm0 `i )gi which has initial term < m as does
0
(hj + cm `j )gj , so cm0 pk gk has initial term < m.
This gives either a set with smaller |Im | or smaller m, contradicting our
minimality assumption.
Applications of the Division Algorithm
1. Given I ⊂ k[x1 , . . . , xn ], compute I ∩ k[x2 , . . . , xn ] = I 0 . V (I 0 ) ⊂ k n−1
is the closure of the projection of V (I) to k n−1 . The algorithm is to
compute a lex GB for I with x1 largest and take the polynomials without
x1 in them.
2. As V (I ∩ J) = V (I) ∪ V (J), we may want to compute I ∩ J. Compute
K = tI + (1 − t)J ⊂ S[t], then compute K ∩ S.
3. I : J = hf |f g ∈ I for all g ∈ Ji. This is, geometrically, the closure of
V (I) \ V (J). I : J = ∩(I : fi ) where J = {f1 , . . . , fs }, and I : f is
computed by computing I ∩ (f ) and then dividing the generating set by
f.

2 Hom, Tensor and Localization


Definition 2.1 (homR (M, N )). homR (M, N ) is the set of R-module homo-
morphisms from M to n, ie, ϕ : M → N is a group homomorphism with
ϕ(rm) = rϕ(m). It is, in fact, a group with (φ + ψ)(m) = φ(m) + ψ(m), and
has an R-module structure by (rφ)(m) = r(φ(m)).
This means homR (M, −) is a covariant functor from R-mod to R-mod. The
map on objects takes N to homR (M, N ). If α : N → N 0 is an R-module
homomorphism, then homR (M, α) : homR (M, N ) → homR (M, N 0 ) by φ 7→
α ◦ φ. Similarly, homR (−, N ) is a contravariant functor from R-mod to R-mod.
α β
Proposition 2.1. hom is left exact. That is, if 0 → A → B → C then
homR (M,α) homR (M,β)
0 → homR (M, A) → homR (M, B) → homR (M, C).
WARNING: If 0 → A → B → C → 0, we don’t expect 0 → homR (M, A) →
homR (M, B) → homR (M, C) → 0 to be exact.
x2
Example: 0 → Z → Z → Z/2 → 0, apply hom(Z/2, −). This is where Ext
comes from.
In general, 0 → A → B → C → 0 and if F is a functor, you don’t expect
0 → F (A) → F (B) → F (C) → 0 to be exact. If f is left (right) exact we get
derived functors from taking cohomology.
Tensor Product
Definition 2.2 (Tensor). If M, N are R-modules then M ⊗R N is the abelian
group that is the quotient of the free abelian group on the symbols {m ⊗ n : m ∈
M, n ∈ N } modulo the relations (am + bm0 ) ⊗ (cn + dn0 ) = acm ⊗ n + adm ⊗
n0 + bcm0 ⊗ n + bdm0 ⊗ n0 . It is an R-module by r(m ⊗ n) = rm ⊗ n = m ⊗ rn.

4
Remember, not everything in M ⊗ N is of the form m ⊗ n.
Example: What are the elements in k mn = k m ⊗k k n of the form m ⊗ n?
The answer is the rank 1 matrices. These can be written as uv T where u ∈ k m
and v ∈ k n .
Z/2 ⊗Z Z/3 = 0 since a ⊗ b = 3a ⊗ b = a ⊗ 3b = a ⊗ 0 = 0.
Change of coefficients: k[x] ⊗k R ' R[x] when R is a k-algebra.
Proposition 2.2. The tensor product satisfies the following universal property:
if ϕ a bilinear map ϕ : M × N → P that is bilinear, then there exists a unique
ϕ̃ such that
M × ..N .................................. M ⊗ N
..... ...
.....
..... ...
..... ...
..... ...
.....
.....
..... ϕ ...
...
.....
.....
.....
...
...
ϕ̃
..... ...
..... ...
.....
..... ........
.
.......

P
Geometric Interpretation
Given X ⊆ k n , we can form I(X) = {f ∈ k[x1 , . . . , xn ] : f (x) = 0∀x ∈ X}.
Hilbert’s
√ Nullstellensatz says that if k is algebraically closed, then I(V (I)) =
I.
Coordinate ring of a variety V is S/I(V ), and maps of varieties correspond
to maps of rings in the other direction (see algebraic geometry for details).
If M is the coordinate ring of V , N is for U and R is for W . If f : V → W
and g : U → W , then the fiber product of V and U over W is F = {(u, v) ∈
U × V : f (u) = g(v)}. The universal property it satisfies is that ϕ : Z → U × V
such that f ◦ π1 ◦ ϕ = g ◦ π2 ◦ ϕ then
Z
... .......
... ......
.. ......
... .. ...
.... ..... .....
... ... ...
... ... ...
... ... ....
... ... ....
...
....
∃! ... ...
... .......
... .....
... ... ...
... ... ...
... ....
..
... U ×V ...
... .....
.
... ..... ...
... . ....
.... ............. ...
. . ....
... ... ...
..
...
.... ..
. π ...
... 2
πF . . ...
... ..
... .... ....1 ..... ....
... ... ...... ..... ...
..... ...
.
. . .
.. ...... .
..... .......
........ .............
. . ......... ............

U V
... ..
...
... ...
... ...
... ..
.
...
...
...
f ...
.
...
...
... .
...
..
.
g
... ...
.
.......... .....
......

W
Point: the map V → W gives a map R → M which makes M an R-module
by rm = ϕ(r)m.
Claim: M ⊗R N is the coordinate ring of the fiber product f , eg, k[x1 , . . . , xn ]⊗k
k[y1 , . . . , ym ] ' k[x1 , . . . , xn , y1 , . . . , ym ].

5
Claim: − ⊗R M is a right exact functor from R-mod to R-mod.
If ϕ : M → M 0 , then ϕ ⊗ N : M ⊗R N → M 0 ⊗R N comes from the bilinear
map ψ : M × N → M 0 ⊗ N by ψ(m, n) = ϕ(m) ⊗R n.
Now we check right exactness. Suppose A → B → C → 0 is exact. We
α⊗1 β⊗1
want to show that A ⊗R N → B ⊗R N → C ⊗R N → 0. To see that β ⊗ 1
is surjective, note that if c ⊗ n ∈ C ⊗ N then there is b ∈ B with β(b) = c, so
β ⊗ 1(b ⊗ n) = c ⊗ n.
To show that the other step is exact, we’ll show that C ⊗ N ' B ⊗R N/α ⊗
1(A ⊗R N ). We’ll do this by checking that B ⊗ N/A ⊗ N satisfies the universal
property of C ⊗ N . Suppose that ϕ : C × N → P is a bilinear map with
ϕ(rc, n) = ϕ(c, rn) = rϕ(c, n). Define ψ : B × N → P by ψ(b, n) = ϕ(β(b), n).
Then ψ is bilinear. Thus, there exists a unique ψ̃ : B ⊗R N → P an R-module
homomorphism. We now show that α ⊗ 1(A ⊗ N ) ⊆ ker ψ̃. ψ̃(α ⊗ 1(a ⊗ n)) =
ψ̃(α(a) ⊗ n) = ϕ(β ◦ α(a), n) = ϕ(0, n) = 0.
Thus, we get a unique induced R-mod homomorphism ψ̄ : B⊗R N/α⊗1(A⊗R
N ) → P so by the universal property, B ⊗R N/α ⊗ 1(A ⊗R N ) ' C ⊗R N .
Warning: Tensor is not always left exact! 0 → Z → Z → Z/2 → 0, tensor
with Z/2, and get Z ⊗ Z2 → Z ⊗ Z/2 → Z/2 ⊗ Z/2 → 0 Specifically, Z ⊗ Z/2 '
Z/2, but the map given by multiplication by 2 is the zero map, so it is not
injective.
Definition 2.3 (Flat Module). An R-module M is flat iff − ⊗R M is exact.
That is, if P → P 0 is an injection, so is P ⊗ M → P 0 ⊗ M .
Localization
Motivation: We put the Zariski Topology on k n , the closed sets are of the
form V (I) for some I. We ask: what are the rational functions defined every-
where on k n \ V (f )? Well, they’re of the form p/f i where p ∈ S, i ≥ 0, that is,
elements of S[f −1 ].
Definition 2.4 (Localization). Let U ⊂ R be a multiplicatively closed set
(u, u0 ∈ U ⇒ uu0 ∈ U , 1 ∈ U ).
For an R-module M , we set M [U −1 ] = {(m, u) : m ∈ M, u ∈ U }/ ∼ where
(m, u) ∼ (m0 , u0 ) iff ∃v ∈ U such that v(u0 m − um0 ) = 0. We write (m, u) as
m/u.
If M = R, then R[U −1 ] is a ring, with (r/u)(r0 /u0 ) = rr0 /uu0 .
Example: R = M = Z, U = Z \ {0}, then R[U −1 ] = Q.
Check: If M is an R-module, then M [U −1 ] is also an R-module by r(m, u) =
(rm, u) and (m, u) + (m0 , u0 ) = (u0 m + um0 , uu0 ).
Example: If R is a domain, then U = R \ {0} gives R[U −1 ] is the field of
fractions, or quotient ring, of R. In general, let U = {nonzero divisors in R},
then K(R) = R[U −1 ] is the total quotient ring of R.
Example: R = k[x], U = {xi : i ≥ 0}, then R[U −1 ] = k[x, x−1 ] is the ring of
Laurent Polynomials.
Warning: The localization of a nonzero module can be zero!
Example: R = Z, M = Z/5. Let U = {5i : i ≥ 0}. M [U −1 ] = 0 as
(m, u) ∼ (0, 1) for all m ∈ Z/5 as 5(1m − u0) = 0.

6
Proposition 2.3. Let U be a multiplicatively closed set of R and M an R-
module, let ϕ : M → M [U −1 ] be an R-module homomorphism ϕ(m) = m/1.
Then ϕ(m) = 0 iff there is u ∈ U with um = 0, and if M is a finitely
generated R-module, then M [U −1 ] is zero iff there is a u ∈ U that annihilates
M.
Proof. m/1 = 0/1 iff ∃u ∈ U with u(1m − u0) = um = 0.
Suppose M is finitely generated by {m1 , . . . , ms }. If there exists u ∈ U
such that um = 0 for all m ∈ M , then m/u0 = 0/1 for all m/u0 ∈ M [U −1 ].
Conversely, suppose that M [U −1 ] = 0, then
Q mi /1 = 0 for all i, so there exists
ui such that ui mi = 0 for each i, let u = ui . Then uM = 0.
Notation: For any U ⊂ R, we’ll denote by R[U −1 ] the localization R[Ũ −1 ]
where Ũ is the multiplicative closure of U .
The most important example is if P is a prime ideal of R and U = R \ P .
Notation: R[(R \ P )−1 ] = RP , and M [(R \ P )−1 ]P .
The residue class field κ(P ) = RP /PP , where PP is ϕ(P )RP .
If R = Z, then Z0 = Q, κ(0) = Q. P = (p), so ZP = {a/b : p 6 |b}, and
PP = {a/b : p|a, p 6 |b}. Then κ(P ) = ZP /PP ' Z/P .
RP is an example of a local ring.
Definition 2.5 (Local Ring). A ring R is local if it has a unique maximal ideal.

Proposition 2.4. Let ϕ : R → R[U −1 ] be the map r 7→ r/1.


For any ideal I ⊆ R[U −1 ], we have I = ϕ−1 (I)R[U −1 ]. Thus that map
I 7→ ϕ−1 (I) is an injection on the set of ideals in R[U −1 ] to the set of ideals of
R. This preserves inclusions and intersections, and takes primes to primes.
An ideal J is of the form ϕ−1 (I) for some ideal I ⊂ R[U −1 ] iff J =
ϕ (JR[U −1 ]) iff for u ∈ U , ur ∈ J ⇒ r ∈ J for r ∈ R. In particular,
−1

I 7→ ϕ−1 (I) gives a bijection between the primes in R[U −1 ] and the primes of
R not meeting U .
Proof. I 7→ ϕ−1 (I) gives an injection {primes of R[U −1 ]} → {primes of R}.
Suppose that J is ϕ−1 (I). Then ur ∈ J implies r ∈ J for u ∈ U , r ∈ R, so
J ∩ U = ∅, as otherwise u ∈ J ∩ U , then u1 ∈ J, so 1 ∈ J, so J = R.

Corollary 2.5. RP is a local ring.


Note: If ϕ : R → S is a ring homomorphism with ϕ(u) a unit of S for all u ∈
U , then if v(u0 r−r0 u) = 0, ϕ(v(u0 r−ur0 )) = 0, ϕ(v)(ϕ(u0 )ϕ(r)−ϕ(u)ϕ(r0 )) = 0.
Since ϕ(u) is a unit, ϕ(u0 )ϕ(r)−ϕ(u)ϕ(r0 ) = 0 can be written as ϕ(r)ϕ(u)−1 =
ϕ(r0 )ϕ(u0 )−1 . So we can define ϕ̃ : R[U −1 ] → S by ϕ̃(r/u) = ϕ(r)ϕ(u)−1 .
In fact, we have a universal property of localization: If ϕ : R → S is a ring
homomorphism with ϕ(u) a unit for all u ∈ U , then ∃!ϕ̃ such that the following
diagram commutes:

7
ϕ .
R .. ....................................................... S
..... .
.....
..... .........
.....
..... ....
..... ..
..... ...
..... ...
.....
..... ∃!ϕ̃ ...
.....
..... ...
..... ...
...... ...
....... ..

R[U −1 ]
If ϕ : M → N is an R-module homomorphism, then ϕ̃ : M [U −1 ] → N [U −1 ]
defined by ϕ̃(m/n) = ϕ(m)/u is an R[U −1 ] homomorphism.
Check well defined: m/u = m0 /u0 implies ∃v with v(mu0 − um0 ) = 0, so
v(u ϕ(m) − uϕ(m0 )) = 0.
0

Lemma 2.6. The map of R-modules α : R[U −1 ] ⊗R M → M [U −1 ] by α(r/u ⊗


m) = rm/u is an isomorphism.
Proof. We must first check that α is well defined. Define α̃ : R[U −1 ] × M →
M [U −1 ] by (r/u, m) 7→ rm/u. This is bilinear and α̃(rs/u, m) = α̃(r/u, sm) =
rsm/u, so we get a map on the tensor product.
Now define and inverse map β : M [U −1 ] → R[U −1 ] ⊗R M by m/u 7→
1/u ⊗R m. This is well defined, as m0 /u0 = m/u means that for some v ∈ U ,
v(um0 − u0 m) = 0, so β(m/u) = u1 ⊗R m = vu0 /vuu0 ⊗R m = vuu 1 0
0 ⊗R vu m =
0 0
β(m /u ).
Check that it is an R-module homomorphism.
Finally, check that β = α−1 . β ◦ α(r/u ⊗ m) = β(rm/u) = 1/u ⊗ rm =
r/u ⊗ m.
α ◦ β(m/u) = α(1/u ⊗ m) = m/u.
Lemma 2.7. R[U −1 ] is a flat R-module.
α
Proof. Suppose that 0 → A → B → C → 0 is exact. It is enough to show that
α⊗1
A ⊗ R[U −1 ] → B ⊗ R[U −1 ] is injective. These are isomorphic, by the previous
lemma, to A[U −1 ] → B[U −1 ] with α̃(a/u) = α(a)/u is injective.
Suppose α̃(a/u) = 0 so there exists v ∈ U with v α̃(a/u) = 0, vα(a)/u) = 0
in B[U −1 ], so there is v ∈ U with v 0 (vα(a) − u0) = 0, so v 0 v(α(a)) = 0, so
α(v 0 va) = 0, α is injective, so v 0 va = 0, so a/u is 0 in A.
Exercise: M1 , M2 ⊆ M are R-modules, show that (M1 ∩ M2 )[U −1 ] =
M1 [U −1 ] ∩ M2 [U −1 ].
Hint: 0 → M1 ∩ M2 → M → M/M1 ⊕ M/M2 is exact.
Next: True Locally often implies True Globally
Lemma 2.8. Let R be a ring, M an R-module

1. If m ∈ M then m = 0 if and only if , /1 = 0 in each localization of M at


a maximal prime m of R.
2. M = 0 iff Mm = 0 for each maximal ideal m of R.

8
Proof. m/1 is zero in Mm if any only if ∃u ∈
/ m with um = 0. ie, the annihilator
of m in M , I = {r ∈ R : rm = 0} is not contained in m. So if m/1 = 0 in every
localization of M at a maximal prime, then I is not contained in any maximal
ideal of R, which is a contradiction, so m = 0.

Corollary 2.9. Let α : M → N be an R-module homomorphism. Then α is


injective, surjective or isomorphism iff αm : Mm → Nm is injective, surjective
or isomorphism for all maximal ideals m of R.
α s
Proof. We have 0 → ker α → M → N → coker α → 0. So 0 → ker(α)m →
Mm → Nm → coker(α)m → 0 so if αm is injective for all m, then ker(α)m = 0
for all m so ker α = 0, similarly for coker α.
Next: (S is an R-algebra) S ⊗R hom(M, N ) → homS (S ⊗R M, S ⊗R N ) by
s ⊗R ϕ 7→ s ⊗ ϕ = s(1 ⊗ ϕ). Check that this is well-defined.
Define α̃ : S ×homR (M, N ) → homS (−, −) by α̃(ϕ) = s(1⊗ϕ) ∈ homS (S ⊗R
M, S ⊗R N ) is bilinear and respects the R-action, so α is well defined. We’ll see
that it is an isomorphism if S is a flat R-module and M is finitely presented.
α
Recall: M is finitely generated iff ∃α such that R2 → M → 0 is exact. It
finitely presented if ker α is also finitely generated.
Fact: If R is Nötherian then finitely generated implies finitely presented.
A corollary of this is that if M is finitely presented, then homR (M, N ) lo-
calizes, that is, homR[U −1 ] (M [U −1 ], N [U −1 ]) ' homR (M, N )[U −1 ].

Lemma 2.10. If R is Nötherian and M is finitely generated, then every sub-


module of M is finitely generated. Thus, every finitely generated module is
finitely presented.
ϕ
Proof. If M is f.g., then ∃ Rs → M → 0, and ker ϕ is a submodule of the finitely
generated R-module Rs , so the second sentence follows from the first.
Now suppose that N is a submodule of M , where M is gen by m1 , . . . , ms .
The proof is by induction on s. If s = 1, then M ' Rm1 , we let I = ker(R →
M ), so M ' R/I. Thus, ϕ−1 (N ) is an ideal in R, so ϕ−1 (N ) is finitely generated
by n1 , . . . , nk , since R is Nötherian, so ϕ(n1 ), . . . , ϕ(nk ) generate N .
Now suppose that the lemma holds for all s0 < s. Consider Ñ ⊆ M/Rm1
is generated by m2 , . . . , ms , so by induction Ñ is generated by {ḡ1 , . . . , ḡ` :
gi ∈ N, 1 ≤ i P ≤ `}. Also, N ∩PRm1 is finitely generated by Ph1 , . . . , hP
r . So for
n ∈ N , n̄ = ri ḡi and n − ri gi ∈ N ∩ Rm1 , so n − ri gi = rj hj so
g1 , . . . , g` , h1 , . . . , hr generate N .
This is NOT true if R is not Nötherian. eg R = k[x1 , x2 , . . .], M = R,
N = (x1 , . . . , xn , . . .) is an infinitely generated submodule of M .
Next: Move towards localization of hom.
Proof. We first prove the proper when M = R. αR : S ⊗R homR (R, N ) →
homS (S ⊗R R, S ⊗R N ) by s ⊗ ϕ 7→ s(1 ⊗ ϕ).
S ⊗ ϕ(1) → s((1 ⊗ ϕ)(1 ⊗ 1)) = s ⊗ ϕ(1).

9
Similarly, if M = Rs , then αRs : S ⊗R N s → (S ⊗ N )s , so finally, we take
M finitely presented.
As such, we have R` → Rs → M → 0. We write −0 for S ⊗ −, and so we
apply this and get (R` )0 → (Rs )0 → M 0 → 0, we apply homR (−, N ) and then
S ⊗ −, and we get the correct diagram and apply the five lemma.

3 Primary Decomposition
Consider V (xy) ⊆ C2 . Then V (xy) = V (x) ∪ V (y), so we can break it into
irreducible varieties.

Proposition 3.1. A variety V (I) ⊆ k n can be written uniquely as V1 ∪ . . . ∪ Vk


where Vi are irreducible subvarieties and no Vi ⊂ Vj for i 6= j.
Proof. We first show that such a decomposition exists. Let S be the set of
all varieties V that do not have such a decomposition into irreducibles. Since
k[x1 , . . . , xn ] is Nötherian, S has a minimal element, V (I). Since V (I) is in S ,
it is not irreducible, so we can write V (I) = V1 ∪V2 for V1 , V2 proper subvarieties.
One of these must not have an irreducible decomposition, else V (I) would, but
this contradicts minimality of V (I)
For uniqueness, suppose that V (I) = V1 ∪ . . . ∪ Vk = V10 ∪ . . . ∪ V`0 . So
V1 = V10 ∩ V (I) = (V10 ∩ V1 ) ∪ . . . ∪ (V10 ∩ Vk ). V10 is irreducible, and each of these
0

is a subvariety of V10 , so as V10 is irreducible, we must have some Vi such that


V10 ∩ Vi = V10 so V10 ⊆ Vi . Then Vi = (V10 ∩ Vi ) ∪ (V20 ∩ Vi ) ∪ . . . ∪ (V`0 ∩ Vi ), so there
is a j such that Vi = Vj0 ∩ Vi , so Vi ⊆ Vj0 , so V10 ⊆ Vi ⊆ Vj0 , so j = 1. Consider
Z = V20 ∪ . . . V`0 = V1 ∪ . . . ∪ V̂i ∪ . . . ∪ Vk = V (I) \ V10 , and then induction.

Note: If I = I and V (I) is irreducible, I is prime.
Definition 3.1 (Associated Prime). Let R be a ring and M an R-module. A
prime P of R is associated to M if it is the annihilator of an element of M .
We write AssR (M ) for the set of all associated primes of M as an R-module.

Notation: If I ⊆ R is an ideal, write AssR (I) for AssR (R/I). We can get
away with this, because AssR (I) is rarely interesting, eg, if R is a domain, then
AssR (M ) = {(0)} for M = I.
eg, if P is prime, AssR (P ) = {P }.
eg, if V = V1 ∪ . . . ∪ Vk , is the irreducible decomposition of a variety V , then
I(Vi ) will be the associated primes of I(V ).
eg, R = Z, AssZ (n) = AssZ (Zn ) =set of prime factors
Lemma 3.2. If R is Nötherian then R[U −1 ] is Nötherian.
Proof. Let I be an ideal in R. Then I = ϕ−1 (I)R[U −1 ] where ϕ : R → R[U −1 ]
has ϕ(r) = r/1. Then ϕ−1 (I) is an ideal of R, so finitely generated and, ϕ of
those generators must generate I.

10
This shows that ”Nötherian” is ”better” than ”finitely generated over a
field”, ie, quotient of a polynomial ring.
Lemma 3.3 (Prime Avoidance). Suppose that I1 , . . . , In , J are ideals of R and
J ⊆ ∪nj=1 Ij . If R contains an infinite field, or if all but two of the Ij are prime,
then J is contained in one of the Ij . If R is Z-graded and J is generated by
homogeneous elements of deg > 0, and all the Ij are prime, then it is enough
to assume that all homogeneous elements of J are contained in ∪j Ij .
Proof. If R contains an infinite field k, then J is a k-vector space, also each
J ∩ Ij is a k-vector subspace. If J 6⊆ Ij for all j, then J ∩ Ij is a proper subspace
of J, but J = ∪nj=1 J ∩ Ij , and we cannot write a vector space over an infinite
field as a finite union of proper subspaces. If one Ij ⊆ J, then quotient by it
and repeat.
Now consider a general R, but all but two of the Ij are prime. The proof is
by induction on n. If n = 1, then J ⊆ I1 .
If n = 2, J ⊆ I1 ∪ I2 , if J 6⊂ I1 , J 6⊂ I2 we can find x1 ∈ J ∩ I1 \ I2 and
x2 ∈ J ∩ I2 \ I1 . But then x1 + x2 ∈ J1 , so x1 + x2 ∈ I1 ∪ I2 , but x1 + x2 ∈ / I1 , I2 ,
a contradiction.
Suppose n > 2. J ⊆ ∪Ij , J 6⊂ Ij . So again we take xi ∈ J ∩Ii \∪ Qi6= j Ij . After
n
reordering, we may assume that I1 is prime. Consider f = x1 + j=2 xj ∈ J.
x1 ∈ I1 \ ∪nj=2 Ij , the product is in ∩nj=2 Ij \ I1 , since I1 is prime. So f ∈ ∪Ii but
f∈ / Ij for any j, a contradiction.
Finally, assume R is graded. The Proof is almost the same. In the n = 2
case, we need to consider xk1 +x`2 for some k, ` to make xk1 +x`2 homogeneous.
Proposition 3.4. Let R be a ring and M an R-module. If I is maximal among
all ideals of R that are annihilators of elements of M then I is prime and
so belongs to AssR (M ). Thus, if R is Nötherian, AssR M is nonempty and
∪P ∈AssR M P = 0 ∪ {zero divisors of M }. (that is, r ∈ R nonzero such that
∃m 6= 0 with rm = 0)
Proof. Let P be such an ideal maximal with respect to the property of an-
nihilating an element of M . Let rs ∈ P for r, s ∈ R. Let m ∈ M have
P = AnnR (m). Then if sm = 0, we have s ∈ P . Otherwise (rs)m = r(sm) = 0
so r ∈ AnnR (sm). But also if p ∈ P , psm = s(pm) = 0, so r + P ⊆ AnnR (sm).
Since sm 6= 0, AnnR (sm) 6= R, so AnnR (sm) = P . So r ∈ P . This shows
that P is prime. We now check that ∪P ∈AssR (M ) P = 0 ∪ {zero divs}. If
0 6= p ∈ P ∈ AssR (M ), then ∃m ∈ M with pm = 0, so p is a zero divisor
on M . This shows ⊆. Conversely, if r 6= 0 is a zero divisor, then ∃m ∈ M
with rm = 0, so r ∈ AnnR M . Let P be an ideal containing AnnR M that is
maximal with respect to annihilating some element of M . Then P is prime so
P ∈ AssR M , so r ∈ ∪P .
Corollary 3.5. Suppose that M is a module over a Nötherian ring R.

1. If m ∈ M , then m = 0 iff m/1 = 0 ∈ MP for all maximal associated


primes of M .

11
2. If K is a submodule of M , then K = 0 iff KP = 0 for all maximal
P ∈ Ass M .
3. If ϕ : M → N is an R-module homomorphism, then ϕ is an injective iff
ϕp : Mp → Np is an injection for each P ∈ Ass M .

Proof. If 0 6= m ∈ M , then ∃P ∈ Ass M containing Ann m. Then some P ∩


Ann m = {0}, we get m/1 6= 0 in MP . If Ann M = 0, then any P works. We
take P maximal to get the result. Part 1 implies 2 and 2 implies 3.
Q: How can we find ALL associated primes?
Lemma 3.6. If R is a Nötherian ring and M is a finitely generated R-module,
then M has a filtration 0 = M0 ( M1 ( . . . ( Mn = M with Mi+1 /Mi ' R/Pi
for some prime Pi of R.
Proof. If M 6= 0 then there is an associated prime P1 ∈ Ann(m1 ) for m1 ∈ M .
Set M1 = Rm1 ' R/P1 . If M1 6= M , consider M/M1 . This has an associated
prime P2 = Ann(m̄2 ), set M2 = M1 + Rm2 . By construction, M2 /M1 ' R/P2 .
Continue in this fashion, this must terminate with some Mi = M , else we would
have an infinite ascending chain of submodules of M . This is impossible as M
is finitely generated over a Nötherian ring.
Lemma 3.7. M is an R-module.

1. If M = M 0 ⊕ M 00 , then AssR (M ) = AssR (M 0 ) ∪ AssR (M 00 )


i p
2. If 0 → M 0 → M → M 00 → 0 is a s.e.s. of R-modules, then AssR (M 0 ) ⊆
AssR (M ) ⊆ AssR (M 0 ) ∪ AssR (M 00 ).

Proof. 1. Follows from 2.


2. Suppose m ∈ M 0 with AnnR (m) = P prime. Then AnnR (i(m)) = P , so
O ∈ AssR (M ).
Now suppose P ∈ AssR (M ) \R (M 0 ). P = AnnR (m) for m ∈ M . So
Rm ' R/P . Now for all r ∈ R with rm 6= 0, we have AnnR (rm) = P ,
since s ∈ P ⇒ srm = 0 and if srm = 0 then sr ∈ P , and r ∈
/ P (since
rm 6= 0) so s ∈ P . This means that Rm ∩ i(M 0 ) = {0}.
We now claim that AnnR (p(m)) = P since Rp(m) = p(Rm) ' Rm =
R/P , so P ∈ AssR (M 00 ).

Corollary 3.8. If R is Nötherian, M finitely generated, and 0 = M0 ( M1 (


. . . ( Mn = M with Mi /Mi−1 ' R/Pi , then AssR (M ) ⊆ {P1 , . . . , Pn }, so
AssR (M ) is a finite nonempty set.

12
Proof. When n = 1, M = R/P1 , which has AssR (R/P1 ) = {P1 }.
For n > 1, 0 → M1 → M → M/M1 → 0 is ses, both ends have filtrations,
the first 0 ( M1 , the second 0 ( M2 /M1 ( M3 /M1 ( . . . ( M/M1 . By
induction, AssR (M1 ) ⊆ {P1 }, and AssR (M/M1 ) ⊆ {P2 , . . . , Pn }, so AssR (M ) ⊆
{P1 , . . . , Pn }.
eg, k[x, y]/(x2 y, xy 2 ) has 0 ( Ry 2 /(x2 y, xy 2 ) ( R{x2 , y 2 }/(x2 y, xy 2 ) (
R{xy, x2 , y 2 }/(xy 2 , x2 y) ( M .
Now Ry 2 /(x2 y, xy 2 ) = R/(x) and R{x2 , y 2 }/(xy 2 , x2 y)/Ry 2 /(xy 2 , x2 y) =
Rx /(x2 y, xy 2 ) ' R/(x) and M3 /M2 ' R/(x, y). Set M4 = R{x, y 2 }/(xy 2 , x2 y),
2

M4 /M3 ' R/(x, y). Set M5 = R{x, y}/(x2 y, xy 2 ), M5 /M4 ' R/(x, y), and
M = R{1}/(x2 y, xy 2 ) and M/M5 ' R/(x, y). So we now have a filtration
0 ( M1 ( M2 ( M3 ( M4 ( M5 ( M , so P1 = (x), P2 = (y), P3 = P4 = P5 =
(x, y).
Warning: There is not always a filtration with all Pi associated! e.g. R =
k[x, y, z, w], I = (x, y) ∩ (z, w) = (xz, xw, yz, yw).
Claim: AssR (I) = {(x, y), (z, w)}. Claim 2: There is no filtration 0 ( M1 (
M with M1 ' R/(x, y), M/M1 ' R/(z, w).
Theorem 3.9. R Nötherian, M is finitely generated.

1. Associated primes commutes with localization, ie AssR[U −1 ] (M [U −1 ]) =


{P R[U −1 ]|P ∈ AssR (M ) with P ∩ U = ∅}
2. AssR (M ) contains all primes minimal over AnnR (M ).

Proof. If P ∈ AssR (M ), then there exists m ∈ M with P = AnnR (m). So


Rm ' R/P , and we get an inclusion R/P → M . As localization is exact,
(R/P )[U −1 ] → M [U −1 ] is an inclusion, so if P ∩ U = ∅, P R[U −1 ] is prime in
R[U −1 ], then P R[U −1 ] ∈ AssR[U −1 ] (M [U −1 ]).
Conversely, suppose that Q ∈ AssR[U −1 ] (M [U −1 ]), then Q = P R[U −1 ] for
some prime P of R with P ∩ U = ∅. Since R is Nötherian, we know that R[U −1 ]
is, so P R[U −1 ] is finitely generated. Thus, R[U −1 ]/P R[U −1 ] is finitely pre-
sented. Thus, homR[U −1 ] (R[U −1 ]/P R[U −1 ], M [U −1 ]) ' homR (R/P, M )[U −1 ],
thus the inclusion ϕ : R[U −1 ]/Q → M [U −1 ] must be f /u for some g ∈ homR (R/P, M )
since ϕ is injective, so is f , and so R/P → M is an injection, so P ∈ AssR (M ).
For part 2, we consider the RP module MP with P a minimal prime over
AnnR (m). AssRP (MP ) 6= ∅ and if Q ∈ AssRP (MP ), then PP0 = Q ( PP , but
if PP0 ( PP then P 0 6⊇ AnnR (M ), so there is r ∈ AnnR (M ) with r ∈ / P 0 so
that r/1 ∈ / PP0 and r/1 ∈ AnnRP (MP ). Thus, AnnRP (MP ) 6( PP0 , so PP0 is not
associated.
This means that AssRP (MP ) = {PP }, so P ∈ AssR (M ).
Fact: If I is radical then I = ∩P for P a prime containing I or for P primes
minimal wrt containing I.
Lemma 3.10. If R is commutative, U is multiplicatively closed and I is max-
imal among ideals not meeting U , then I is prime.

13
Proof. Suppose that f g ∈ I. If f, g ∈ / I, then I + (f ) ∩ U 6= ∅, I + (g) ∩ U 6= ∅, so
there exists i, i0 ∈ I, r, r0 ∈ R with i + rf, i0 + r0 g ∈ U so (i + rf )(r0 + r0 g) ∈ U ,
is equal to ii0 + ir0 g + rf 0 i + rr0 f g ∈ I, contradicting that I ∩ U = ∅.
So either f ∈ I or ∈ I, so I is prime.

Corollary 3.11. I ⊆ R is an ideal, then I √ = ∩P over primes containing I.
In particular, the intersection of all primes is 0, the set of nilpotents.

Proof. I ⊆ ∩P is straightforward. √ √
i
Suppose that f ∈ √ ∩P \ I. U = {f : i ≥ 0}, then I ∩ U = ∅. Let J be
an ideal containing I max wrt J√∩ U = ∅, then J is prime which contains I,
so f ∈ J, contradiction, so ∩P = I.

Corollary 3.12. If I = I, then I = ∩P , P minimal over I, and AssR (I) =
{P |P is minimal over I}.

√ time, I = ∩Qi , Qi ∈ AssR (I).
Next
As I = ∩P , P prime √ and I ⊆ P . So ∩P over all primes are the nilpotent
elements, and ∩P ∈Ass P = AnnR M .
First, IOUs.

1. If R is f.g. over a field, R[U −1 ] doesn’t have to be. k[x](x) , for example.
Suppose that k[x](x) as a k-algebra by f1 /g1 , . . . , fr /gr . Assume k is al-
gebraically closed, look at the factors of the gi . Only finitely many x − α
show up, but k is infinite, so these cannot generate everything.
2. We looked at = (x, y) ∩ (z, w) = ∩P . Suppose that P1 ∩ P2 = P3 ∩ P4 ∩
. . . ∩ Pn irredundant, all Pi necessary. So P1 ∩ P2 ⊆ P3 , so P1 P2 ⊆ P3 , so
P1 ⊆ P3 or P2 ⊆ P3 , and so by irredundancy, P3 = P1 and P4 = P2 .

Lemma 3.13. If I = I, then Ass(I) = {P minimal over I}.
This is because I = ∩P over the primes minimal over I.
In general, we replace primes by ideals with only one associated prime.
Ass(R/P ) = {P }.
Definition 3.2 (P -primary). Let R be a Nötherian ring and M a finitely gen-
erated R-module.
A submodule N ⊆ M is P -primary if AssR (M/N ) = {P }.
Proposition 3.14. Let P be a prime ideal in R. Then TFAE

1. AssR (M ) = {P }
2. P is minimal over AnnR M and every element not in P is a non zero
divisor on M .

3. A power of P annihilates M and every element not in P is a non zero


divisor.

14
Proof. 1 ⇒ 2 : If P is not minimal over AnnR M , then there exists P 0 minimal
over AnnR M with AnnR M ⊆ P 0 ( P . So P 0 ∈ AssR M . Also, every zero
divisor is in some associated prime, so they all lie in P .
2 ⇒ 3 : Since elements outside of P are non zero divisors, we get M → MP
injective, so if a power of PP annihilates MP that power of P annihilates M .
But PP is minimal over AnnRP MP and every prime is contained p in PP . So ∩Q
over Q ∈ RP prime, minimal over AnnRP MP , is equal to PP = AnnRP MP .
Thus, ∃k such that PPk ⊆ AnnRP MP .
3 ⇒ 1 : Since P k ⊆ AnnR M ⊆ P , and P must be contained in any prime
containing AnnR M , so P is minimal over AnnP M , so P ∈ AssR M . But also,
every associated prime is contained in P , from the nonzero divisor assumption,
so P is the only associated prime.
Corollary 3.15. Let I be an ideal in R. TFAE

1. I is P -primary
2. I contains a power of P and for all r, s with rs ∈ I and r ∈/ I, we have
s∈P

3. I = P and for all r, s ∈ R with rs ∈ I, either r ∈ I or there is a k such
that sk ∈ I.

Proof. 1 and 2 are equivalent because they are 1 and 3 of the prop.
2 ⇒ 3 : If √
rs ∈ I, r ∈/√I, then s ∈ P , so ∃k with sk ∈ I. Also, P k ⊆ I for
some k, P ⊆ I. If f ∈ I \ P , then ∃` such√that g = f ` ∈ I \ P . But now
g · 1 ∈ I, 1 ∈ g ∈ P , contradiction. Thus, I ⊆ P .
/ I so √
k
3 ⇒ 1 : Since I = P , we know √ P ⊆ I for some k, and if rs ∈ I, r ∈ / I,
`
then ∃` such that s ∈ I, so s ∈ I = P .
Theorem 3.16. A proper submodule M 0 of M is the intersection of primary
submodules.
Definition 3.3 (Irreducible Submodule). A submodule N of M is irreducible if
it cannot be written as the intersection N = N1 ∩ N2 with N ( N1 and N ( N2 .
Lemma 3.17. If M 0 is a proper submodule of M , then we can write M 0 =
∩ki=1 Mi where each Mi is irreducible.
Proof. R Nötherian and M finitely generated implies that M is Nötherian.
So if the lemma is false, there exists a submodule N maximal with respect
to not having an irreducible decomposition. In particular, N is not irreducible,
so we write N = N1 ∩ N2 , with N ( N1 and N ( N2 . But N1 = ∩ki=1 Mi and
N2 = ∩`j=k+1 Mj , so N = ∩`i=1 Mi , contradiction.
Lemma 3.18. If N ⊆ M is irreducible, then N is primary.
Proof. Suppose P 6= Q ∈ AssR (M/N ). Then R/P ' Rm̄1 for some m̄1 ∈ M/N ,
and R/Q ' Rm̄2 for some m̄2 ∈ M/N .
Rm̄1 ∩ Rm̄2 = {0}, so (N + m1 ) ∩ (N + m2 ) = N .

15
Theorem 3.19. Let M 0 be a proper submodule of M and write M 0 = ∩Mi
where Mi is Pi -primary. Then

1. Every associated prime of M/M 0 occurs among the Pi .


2. If the decomposition is irredundant, then the Pi are precisely the associated
primes.

3. If the intersection is minimal, then each associated prime occurs exactly


once, and if P is a minimal associated prime, then Mi is the P -primary
component of M 0 .

Proof. We first note that if M 0 = ∩Mi with AssR (M/Mi ) = {Pi }, then 0 =
∩(Mi /M 0 ) ⊆ M/M 0 with AssR ((M/M 0 )/(Mi /M 0 )) = {Pi }. ie, we can replace
M by M/M 0 and M 0 by 0, so we will assume that M 0 = 0.

1. So 0 = ∩Mi , so M → ⊕M/Mi is an injection. So AssR (M ) ⊆ AssR (⊕M/Mi ) =


∪ AssR (M/Mi ) = {P1 , . . . , Pn }.
2. If the decomposition is irredundant, then ∩i6=j Mi 6= 0 for any j. So
∩i6=j Mi = ∩i6=j Mi /(∩i6=j (Mi ∩ Mj )) ' (∩i6=j Mi + Mj )/Mj ⊆ M/Mj . So
AssR (∩i6=j Mi + Mj ) ⊆ AssR (M/Mj ) = {Pj }, so AssR (∩i6=j Mi + Mj ) =
{Pj }, so Oj ∈ AssR (M ).
3. We first note that if N1 , N2 ⊆ M with Ass(M/Ni ) = {P } for i = 1, 2,
then M/N1 ∩ N2 → M/N1 ⊕ M/N2 is an inclusion, so Ass(M/N1 ∩ N2 ) ⊆
Ass(M/N1 ⊕ M/N2 ) = {P }, so Ass(M/N1 ∩ N2 ) = {P }. Thus if Pi = Pj
we can replace Mi , Mj by Mi ∩ Mj to get a primary decomposition with
fewer terms. So each Pi shows up at most once. But minimal implies
irredundant, so each Pi shows up exactly once.
Now suppose that Pi is minimal over AnnR M . We want to show that
Mi = ker(M → MPi ). Consider the diagram
MP i
..... γ
α .................
..
.....
.........
..... ......
....

M .. (M/Mi )Pi
.....
..... β
..... δ....................
.......... ....

M/Mi
We have Mi = ker β. So show that Mi = ker α, it suffices to check that
δ, γ are injections. δ is because Ass(M/Mi ) = {Pi }. Since ∩Mj = 0,
φ : M → ⊕M/Mj localizes to φPi : MPi → ⊕(M/Mj )Pi . γ is the ith
component of φPi . To see that γ is injective, it suffices to note that
(M/Mj )Pi = 0. If it weren’t zero, we would have AssRPi ((M/Mj )Pi ) =
{QRPi |Q ∈ AssR (M/Mj ) with Q ∩ (R \ P ) = ∅} = {QRPi |Q = Pj and
Q ⊆ Pi }, which is empty as Pi is minimal.

16
eg, I = (x2 y 4 , x3 y 3 , x4 y 2 ) = (x2 ) ∩ (y 2 ) ∩ (x4 , xy 4 , y 7 , x3 y 3 ) = (x2 ) ∩ (y 2 ) ∩
(x 0, x4 y, x3 y 3 , y 4 )
1

Proposition 3.20. Let A be a finitely generated torsion free abelian group and
let R = ⊕a∈A Ra and M be a graded R-module. If P = AnnR M for any
m ∈ M and P is prime, then P is homogeneous and P is the annihilator of a
homogeneous element.
Proof. A ' Zk for some k. Choose the isomorphism and w ∈ Rk sufficiently
general and set u ≺ u0 for u, u0 ∈ Rk if w · u < w · u0 . (sufficiently general means
that this is a total order).
Note that if u < u0 then u + v < u0 + v. If P is Pnot homogeneous, then there
exists f ∈ P and a ∈ A with fa ∈ /
P P where f = a∈A fa . We can assume in
fact that no fa ∈ P . Write m = ma . Then 0 = f m = f1 m1 + HOT where
deg f1 = min{ai |fai 6= 0} and deg m1 = {ai |mai 6= 0}. So f1 m1 = 0. So if
m = m1 done. Otherwise, this is the base case of an induction on the number
of nonzero components, b. P
Suppose that it is true for small b. Write f1 m = ma 6=m1 f1 ma , this has
fewer terms, so P ⊆ I = AnnR f1 m. If P = I, then P is homogeneous by
induction. Otherwise, there is g ∈ I \ P with gf1 m = 0 so gf1 ∈ P so f1 ∈ P
as required.
Lemma 3.21. If M is a noetherian R-module and M 0 ⊆ M with M 0 = ∩Mi
with Mi Pi -primary is minimal, let U be a multiplicatively closed set of R. Then
M 0 [U −1 ] = ∩Mi [U −1 ] over the submodules with Pi ∩U = ∅ is a minimal primary
decomposition of M 0 [U −1 ] as an R[U −1 ]-module.
Proof. M 0 [U −1 ] = ∩Mi [U −1 ], since localization commutes with intersection. If
Pi ∩ U 6= ∅, then Mi [U −1 ] = M [U −1 ] since (M/Mi )[U −1 ] has no associated
primes, it must be the zero module.
As M 0 [U −1 ] = ∩Mi [U −1 ] with P ∩U = ∅, and AssR[U −1 ] (M [U −1 ]/M 0 [U −1 ]) =
{P R[U −1 ]| where P ∈ AssR (M/M 0 ) and P ∩ U = ∅}. So since ∩Mi was mini-
mal, each Pi was in AssR (M/M 0 ) and showed up exactly once, so in this new
intersection.

4 Integral Dependence
Cayley-Hamilton Theorem
In linear algebra, pA (X) = det(A − xI)
Cayley-Hamilton says pA (A) = 0.
Slightly more abstractly, if V is an n-dimensional vector space over k, and
ϕ : VP→ V is a linear map, then there exists a0 , . . . , an−1 ∈ k such that
n−1
ϕn + i=0 ai ϕi = 0.
Theorem 4.1. Let R be a ring and M a finitely generated R-module that has
a generating set with n elements. Let ϕ : M → M be an R-module homomor-
phism. If ϕ(M ) = IM for an ideal I ⊆ R, then there exists a monic polynomial

17
p(x) = xn + p1 xn−1 + . . . + pn with pj ∈ I j for each j such that p(ϕ) = 0 as an
endomorphism of M .
Pn
Proof. Let m1 , . . . , mn be generators for M . Write ϕ(mj ) = i=1 aij mi . Let
A be the matrix A = (aij ) and m = (m1 , . . . , mn )T ∈ M n . Regard M as an
R[x]-module where x · m = ϕ(m).
P Then (Ix − A)m P = 0 for all m ∈ M , where Ix : M → M where Pn if m =
ri mi , Ix(m) = ri ϕ(mi ), then Ix(mi ) = ϕ(mi ), and Ami = j=1 aji mj .
Let A0 be the matrix of cofactors of Ix − A, recall from linear algebra that if
A = (aij ) then the cofactor matrix B is (bij ) with bij = (−1)i+j det(Aij ).
Then A0 (Ix − A) = det(Ix − A)I, that is, det(Ix − A)m = 0 for all m ∈ M ,
so det(Ix − A) ∈ R[x] and is in AnnR[x] M .
Let p(x) = det(Ix − A). p(x) has degree n and p(ϕ) = 0 and if p(X) =
Pn−1
xn + i=0 pi xn−i then pi ∈ I i .
More linear algebra: If ϕ : V → V is surjective, then V is injective, if V is a
finite dimensional vector space.
Corollary 4.2. Let R be a ring and let M be a finitely generated R-module.

1. If α : M → M is a surjective R-module homomorphism, then α is an


isomorphism.
2. If M ' Rn then every set of n elements that generates M forms a free
basis, in particular, the rank of M is well-defined.

Proof. 1. Regard M as a module over R[t] with t acting as α, so tm = α(m).


Set I = (t) ⊆ R[t], then IM = M since α is surjective. Apply Cayley-
Hamilton to the identity homomorphism, 1 : M → M so there exists
xn + p1 xn−1 + . . . + pn with pi ∈ I i and (1 + p1 1 + . . . + pn 1)m = 0, so
(1 + p1 + . . . + pn )m = 0, write this as 1 − tq(t), so there exists q(t) ∈ R[t]
with (1−tq(t))m = 0 for all m ∈ M . So (1−q(α)α)m = 0, so q(α)◦α = 1.
Thus α is injective.
2. A set of n generators for M corresponds to a surjection β : Rn → M .
Since M is free of rank n, there exists γ : M → Rn , then β ◦ γ : M → M
is surjective, and thus it is an isomorphism. So β = (βγ) ◦ γ −1 is an
isomorphism, so that the given generators form a free basis.
To finish we check that rank is well defined. 1) suppose that Rm ' Rn for
m < n, we extend our generating set of size m to one of size n by adding
n − m zeros. Then part 1 says that this is a free basis, but it contains
zero, so it is a contradiction.
A second proof is that we let P be a maximal ideal of R, then R/P ⊗R M '
(R/P )m ' (R/P )n

Note: This is not true for injections! eg α : Z → Z by α(x) = 2x.


Integral Dependence

18
Definition 4.1 (Integral over R). Let S be an R-algebra and let p(x) be a
polynomial in R[x]. We say that s ∈ S satisfies p if p(s) = 0. The element s is
called integral over R if it satisfies some monic polynomial.
The equation p(s) = 0 is the equation of integral dependence for s over R.
If every element of S is integral over R, we say S is integral over R.
√ √
eg S = Q(√ 2) with R = Z. 2 is integral over Z, as it satisfies x2 − 2 = 0.
In fact, a + b 2 for any a, b ∈ Z is integral over Z. (x − a)2 = 2b2 . √
Claim:

These are all the elements integral over Z. eg, K = Q( 5), then
1+ 5
x = 2 is integral over Z, as it satisfies (2x−1)2 = 5 which is 4x2 −4x−4 = 0,
so x2 − x − 1 = 0.
Definition 4.2 (Integral Closure). The collection of all elements of S integral
over R is called the integral closure, or normalization, of R in S.
Definition 4.3 (Number Field). A finite field extension of Q is called a number
field. The integral closure of Z in a number field is called the ring of integers in
that number field.
Definition 4.4 (Normal). If R is a domain, then it’s normalization is its inte-
gral closure in its field of fractions. If R is equal to its normalization, then we
say that it is normal.
eg, Z is normal, though proving this really proves the following:
Lemma 4.3. Any UFD is normal.
Proof. Consider r/s with r, s relatively prime. If (r/s)n +p1 (r/s)n−1 +. . .+pn =
0, then rn + p1 spn−1 + . . . + pn sn = 0, which contradicts the relatively prime
assumption.
e.g. k[x] is normal.
Theorem 4.4. Let R be a ring and let J ⊆ R[x] be an ideal. Let S = R[x]/J
and let s be the image of x in s.

1. S is generated by ≤ n elements as an R-module iff J contains a monic


polynomial of deg ≤ n. In this case, S is generated by {1, s, . . . , sn−1 }.
In particular, S is a finitely generated R-module iff J contains a monic
polynomial.
2. S is a finitely generated free R-module iff J can be generated by a monic
polynomial. Then S has a basis of the form {1, s, . . . , sn−1 }.

eg Z[x]/(2x + 1) is not finitely generated.


Proof. 1. The powers of x generated R[x] as an R-module, and so generate
S as well. So if J contains a monic polynomial, p, of degree n, then for
d ≥ n, we can write sd in terms of smaller powers of s using sd−n p(s) = 0,
so {1, . . . , sn−1 } generate S. Conversely, suppose that S is generated by

19
≤ n elements as an R-module. Multiplication by s is an endomorphism of
S, so Cayley-Hamilton says that there exists a monic p of deg ≤ n with
p(s) = 0, so p(x) ∈ J.
2. Suppose that J is generated by a monic polynomial p of degree n. Then
from a, {1, . . . , sn−1 } P
generates S. if these
P do not form a free basis, then
there are ai ∈ R with ai si = 0. Thus, ai xi ∈ J of degree n − 1. This
contradicts the fact that J is generated by p which has degree n. Thus, S
is a free R-module with free basis {1, . . . , sn−1 }. Conversely, we suppose
that S is a free R-module of rank n. Then S is finitely generated, so by
1, J contains a monic p of degree n. Suppose there is q ∈ J \ (p). Use
the division algorithm to write q = ap + r for a, r ∈ R[x] with deg r < n.
Then r ∈ J, but if r 6= 0, then this would contradict {1, . . . sn−1 } being a
free basis. So r = 0 and q ∈ (p), thus (p) = J.

Definition 4.5 (Finite). An R-algebra S is finite over R if S is a finitely


generated R-module.
Corollary 4.5. An R-algebra S is finite over R iff S is generated as an R-
algebra by finitely many integral elements.
Proof. If S is finite over R and s ∈ S, multiplication by s is an endomorphism
of s. So Cayley-Hamilton shows that there exists monic p with coefficients in R
and p(s) = 0, so s is integral over R¿
Thus, since S is a finitely generated R-algebra, it is generated by finitely
many integral elements.
Conversely, suppose that S is generated by t integral elements as an R-
algebra.
If t = 1, then S ' R[x]/J for some J, and by the theorem, J contains a
monic polynomial, so S is finite over R by the theorem.
We may assume that t > 1 and that the result is true for t = 1. Let S 0 be the
subalgebra of S generated by the first t − 1 generators of S. Then, by induction,
S 0 is finite over R, so S 0 is generated, as an R-modules, by {s1 , . . . , s` }. The
extra generator s for S is integral over R, so it is integral over S 0 , so S is finite
over S 0 . So S is generated as an S 0 -module by t1 , . . . , tm , but then {si tj : 1 ≤
i ≤ `, 1 ≤ j ≤ m} generates S as an R-module. So S is finite over R.
Corollary 4.6. If S is an R-algebra and s ∈ S, then s is integral over R iff
there exists an S-module N and a f.g. R-submodule M ⊆ N nor annihilated by
any nonzero element of S such that sM ⊆ M .
In particular, S is integral iff R[s] is a finitely generated R-module.
Proof. If s is integral over R, take N = S, M = R[s] ⊂ S is a finitely generated
R-module and not annihilated by anything in S, since s0 1 = s0 6= 0.
Conversely, if ∃M ⊆ N with these properties, then multiplication by s is an
R-module homomorphism. So by Cayley-Hamilton, there exists a monic p with
coefficients in R such that p(s)M = 0, then p(s) = 0 in S, so S is integral over
R.

20
Theorem 4.7. Let R be a ring and S be an R-algebra.. The set of all elements
of S integral over R is a subalgebra of S.
In particular, if S is generated by elements integral over R, then S is integral
over R.

Proof. Let S 0 be the set of elements of S integral over R. We need to show


that if s, s0 ∈ S 0 then ss0 and s + s0 are in S 0 . Let M = R[s], M 0 = R[s0 ],
which are finitely generated R-module. Let M M 0 = R{f g, f ∈ M, g ∈ M 0 },
then M M 0 is a finitely generated R-module. ss0 M M 0 = (sM )(s0 M 0 ) ⊆ M M 0 ,
and (s + s0 )M M 0 = (sM )M 0 + M (s0 M 0 ), so by the corollary, since M M 0 ⊆ S
is a finitely generated submodule not annihilated by any element of S, since
1 ∈ M M 0 , so ss0 and s + s0 are integral over R.
Corollary 4.8 (To Cayley-Hamilton). If M is a finitely generated R-module
and I is an ideal of R such that IM = M then ∃r ∈ I such that (1 − r)M = 0.
Proof. By CH, we get p(x) ∈ R[x] with xn + p1 xn−1 + . . . + pn with pj ∈ I J such
that p(id)M = 0, that is, (1+p1 +. . .+pn ) id M = 0, so set r = −(p1 +. . .+pn ) ∈
I. So (1 − r)M = 0.
Definition 4.6 (Jacobson Radical). The Jacobson radical of R is the intersec-
tion of all maximal ideals.
eg, R = Z, then the Jacobson Radical is (0). If R is local, then the Jacobson
Radical is the unique maximal ideal.
Corollary 4.9 (Nakayama’s Lemma). Let I be an ideal contained in the Ja-
cobson radical of R and let M be a finitely generated R-module.

1. If IM = M then M = 0.
2. If m1 , . . . , mn ∈ M have images in M/IM that generate M/IM as an
R-module, then m1 , . . . , mn generated M as an R-module.

Proof. 1. By the corollary, ∃r ∈ I such that (1 − r)M = 0, since (1 − r) is


not in any maximal ideal (as r is in all of them), we have 1 − r is a unit
of R, so M = 0.
P
2. Suppose that m̄1 , . . .P
, m̄n generate M/IM . Let N = M/( Rmi ). Then
N/IN = M/(IM + Rmi ) = M/M = 0, so IN = N , so N = 0.

Mostly, we use this in the case (R, P ) is local, then M/P M is an R/P -
module, and R/P is a field, so it is a vector space.
Application: If (R, P ) is local, then P M = M ⇒ M = 0. ie, if 0 = M/P M ,
then M = 0. ie, if M/P M = R/P ⊗R M = 0, then M = 0.
Corollary 4.10. If M and N are f.g. R-modules and M ⊗R N = 0 then
AnnR M + AnnR N = R, in particular, if R is local, then M = 0 or N = 0.

21
Proof. We first prove the local case. Suppose M ⊗R N = 0 but M 6= 0. Then
Nakayama says that M/P M 6= 0. Now M/P M is an R/P -vector space, so
there exists a surjection M/P M → R/P and thus there exists a surjection
M → R/P . So 0 = M ⊗R N → R/P ⊗R N is surjective, since ⊗ is right exact,
so R/P ⊗R N = 0, so N/P M = 0, thus P N = N so N = 0.
Now suppose that R is general. M ⊗R N = 0 but AnnR M + AnnR N 6=
R. Then there exists prime ideal P with P ⊇ AnnR M + AnnR N . Then
MP ⊗RP NP = 0, so WLOG, MP = 0, so M = 0 since AnnR M ⊆ P . But then
AnnR M = R, contradiction.
Recall that we shows that integral closure is an R-subalgebra of S. Next:
Integral closure commutes with localization.
Proposition 4.11. Let R ⊆ S be rings and let U be a multiplicatively closed
subset of R. If S 0 is the integral closure of R in S, then S 0 [U −1 ] is the integral
closure of R[U −1 ] in S[U −1 ].

Proof. Any element of S integral over R is integral over R[U −1 ], so S 0 is integral


over R[U −1 ] and thus, S 0 [U −1 ] is integral over R[U −1 ]. So we just need to show
that if s/u ∈ S[U −1 ] is integral over R[U −1 ], then there exists u0 ∈ U with
su0 ∈ S 0 (then s/u = su0 /uu0 ∈ S 0 [U −1 ]).
If (s/u)n + r1 /u1 (s/u)n−1 + . . . + rn /un = 0, multiply by (uu1 . . . un )n to
get (su1 . . . un )n + . . . + rn (uu1 . . . un )n = 0, so su1 . . . un is integral over R, and
u1 . . . un ∈ U .
Warning: If R is a Nötherian domain, then the integral closure of R in its
quotient field is not necessarily Nötherian.
It is Nötherian if the integral closure is a finitely generated R-algebra (⇒ R-
module). Also Nötherian if R is a finitely generated domain containing a field
or the integers (Nöther)
eg, R = k[x1 , . . . , xn ]/I, the normalization is of the form k[y1 , . . . , ym ]/J.
Next: relationship between the primes in the integral closure of R and the
primes in R.
Proposition 4.12 (Lying Over and Going Up). Suppose R ⊆ S is an integral
extension of rings, given a prime P ⊆ R¡ then there exists Q ⊆ S prime with
Q ∩ R = P . (Lying Over)
Also, Q may be chosen to contain any ideal Q1 ⊆ S with Q1 ∩R ⊆ P . (Going
Up)

eg, R = Z, S = Z[ 2] and P = (7).

Proof. Factor out Q1 and R ∩ Q1 , to see that we just need to find a prime Q in
S with R ∩ Q = P . Let U = R \ P , so R[U −1 ] = RP . If we show ∃Q0 ∈ S[U −1 ]
with Q0 ∩ RP = Pp , then since Q0 = Q[U −1 ] for some prime Q of S, we would
have Q ∩ R = P , so we can assume that R is local with maximal ideal P .
Then, if P S 6= S, any maximal ideal Q of S containing P S will have P ⊆
Q ∩ R, so P = Q ∩ R. So we just need P S 6= S.

22
P`
If P S = S, then 1 = i=1 si pi , and let S 0 be the R-algebra generated by
{s1 , . . . , s` }, then 1 ∈ P S 0 so P S 0 = S 0 and S 0 is a f.g. R-module, since it
is a finitely generated R-algebra over R, so by Nakayama, S 0 = 0, which is a
contradiction, so P S 6= S.

Proposition 4.13. Let R ⊆ S be domains, if K(S) is algebraic over K(R),


then every nonzero or S intersects R nontrivially.
If R ⊆ S is an integral extension of domains, then S is a field iff R is a
field. Equivalently, if S is an integral R-algebra, P a prime of S, then P is a
maximal ideal of S iff P ∩ R is a maximal ideal of R.

Proof. For the first statement, if sufficies to


Pnshow it for a principal ideal bS of S.
If b ∈ S, then there exist ai ∈ K(R) with i=0 ai bi P = 0. Clearing denominators
m
and dividing by a power of b if necessary, we get i=0 a0i bi = 0 with a00 6= 0,
0 0 0
ai ∈ R. Then a0 ∈ bS ∩ R, a0 6= 0, so the ideal generated by b intersects R
nontrivially.
If R ⊆ S is an integral extension P of domains, then K(S) is alg over K(R),
if s/u ∈ K(S),Pmthen ∃a i ∈ R with ai si = 0, so is alg over R (IOU) and
j j−m
P
bj u = 0 ⇒ i=1 bj u
Suppose R is a field. Let P be a maximal ideal in S. Then P ∩ R 6= {0}, so
P ∩ R = R, this contains 1, so P = S.
Suppose instead that S is a field. Let P be a prime of R. Then by Lying
Over, there is a prime Q of S with Q ∩ R = P . But Q must equal (0), so
P = (0) ∩ R = (0). So the only prime in R is (0), and R is a field.
Finally, take S/P and R/(P ∩R). The statement follows from the field state-
ment once we check that S/P is integral over R/(P ∩ R). This is because ”inte-
Pn−1
gral dependence persists mod P ”, ie, if xn + ai xi = 0, then x̄n + i=0 āi x̄i = 0
P
in S/P , and āi = ai ∈ R/(P ∩ R).

Corollary 4.14 (Incompatibility). Suppose R ⊆ S is an integral extension


of rings. Two distinct primes of S having the same intersection with R are
incomparable, ie, neither is contained in the other.
Without integrality, we have, for example, k ⊆ k[x, y], (0) = k ∩ (x) =
k ∩ (x, y), but (x) ⊆ (x, y).

Proof. If Q ⊆ Q1 ⊆ S with Q ∩ R = Q1 ∩ R = P ⊆ R, factor out P ⊆ R, and


Q ⊆ S. Then we get 0 in R/P equals 0 ⊂ Q1 /Q ⊂ S/Q. So we are in the case
where R0 and S 0 are domains. S 0 is still integral over R0 , so K(S 0 ) is alg over
K(R0 ), so if Q1 6= Q = 0, Q1 ∩ R 6= (0), contradiction. So Q1 = Q = 0 in S.

5 Blowup Algebra
Geometric Motivation
Bl0 C2 ”the blowup of C2 at 0”. We want to be able to take a curve and
separate out the strands going through the origin, we’ll cut out the origin and

23
glue in a P1 . Algebraically, we replace C[x, y] by C[x, y, u, v]/(xv − yu), so we
replace C2 by V (xv − yu) ⊆ C2 × P2 .
If (x, y) 6= (0, 0), then WLOG, x 6= 0, so if xv − yu = 0, then v = y/xu. So if
y 6= 0, (u : v) = (1 : x/y) = (x : y). If y = 0, then (u : v) = (1 : 0) = (x : y). So
if (x, y) 6= (0, 0), there is a unique (u : v) with (x, y) × (u : v) ∈ V (xv − yu). If
(x, y) = (0, 0), then there are no conditions, so any (0, 0) × (u : v) ∈ V (xv − yu).
Ie, consider the map π : Bl0 C2 → C2 by (x, y) × (u : v) → (x, y), this map
is 1-1 away from the origin, and π −1 (0, 0) = P1 , the ”exceptional divisor”.
Now C[x, y, u, v]/(xv − yu) ' C[x, y, xt, yt] ⊆ C[x, y, t], deg t = 1, deg x =
deg y = 0.

Definition 5.1 (Blow-Up Algebras). If R is a ring and I is an ideal, then the


blow-up algebra of I in R is the R-algebra BI (R) = R⊕I⊕I 2 ⊕. . . ' R[It] ⊆ R[t].
eg, B(x,y) C[x, y] is the coordinate ring of Bl0 C2 .
Set deg(r) = 0 for r ∈ R and deg(t) = 1. Then BI R is a Z-graded ring.
The ideal I ⊆ BI (R) is homogeneous, thus the quotient BI (R)/I is graded.
Definition 5.2 (Associated Graded Ring). grI R = R[It]/IR[It] is the assci-
ated graded ring of R.
[Can also do for any diltration R ) I1 ) I2 ) . . . with Ij Ik ⊆ Ij+k .]
For C[x, y, tx, ty]/(x, y) ' C[x, y, u, v]/(xv−yu, x, y) ' C[u, v] = gr(x,y) C[x, y],
the coordinate ring of P1 . We can also do this for modules:
Definition 5.3. Let R be a ring, M an R-module, and I an ideal of R, then
I : M = M0 ⊃ M1 ⊃ M2 ⊃ . . . is a filtration of R-modules. It is an I-filtration
if IMi ⊆ Mi+1 for all i > 0 and it is I-stable if ∃n > 0 such that ∀i ≥ 0,
IMn+i = Mn+i+1 = I i+1 Mn .
OWED RESULT: If R ⊂ S is an integral extension of domains, then K(S)
is algebraic over K(R). Suppose s1 /s2 ∈ K(S). Then there exist ai , bj with
Pm Pn
ai sii = 0, am = 1 and j=0 bj sj2 = 0 with bn = 1 and b0 6= 0. Then
Pi=0
n j−n n
= 0 = j=0 bj /b0 (1/s2 )n−j , this shows that 1/s2 is integral
P
j=0 bj /b0 s2
over K(R), so s1 /s2 = s1 ∗ 1/s2 is integral over K(R).
Definition 5.4. If I is an I-filtration, then grI M = M/M1 ⊕ M1 /M2 ⊕ . . .
and BlI M = M ⊕ M1 ⊕ . . ..
Note: grI R is an R/I-algebra. If I is finitely generated, then grI R is a
finitely generated R/I-algebra. If I is a maximal ideal, then grI R is a finitely
generated algebra over a field.
Also grI M is a graded grI R module.
Proposition 5.1. Let I be an ideal in R and let M be a finitely generated R-
module. If I : M = M0 ⊃ M1 ⊃ . . . is an I-stable filtration by finitely generated
R-submodules of M , then grI M is a finitely generated module over grI R.

24
Proof. Suppose IM1 = Mi+1 for i ≥ n. Then (I/I 2 )(Mi /Mi+1 ) = Mi+1 /Mi+2
for i ≥ n. So the union of any set of generateors for Mi /Mi+1 0 ≤ i ≤ n
generates grI M as a grI R-module.
Since the Mi are finitely generated R-modules, so are the Mi /Mi+1 , so we
get a finite set of generators.
Proposition 5.2. Let R be a ring, I ⊆ R an ideal, M a finitely generated
R-module with I-filtration I = M0 ⊇ M1 ⊇ . . . by finitely generated Mi .
Then the filtration is I-stable iff the BI R-module BI M is finitely generated.
Proof. If BI M is finitely generated, then its genertaors appear in the first n
steps for some n, so BI M is gneerated by M0 , . . . , Mn . So Mn ⊕. . . is generated,
as a BI R-module, by Mn . This means that Mn+1 = I i Mn for i ≥ 0 so I is
I-stable.
Conversely, if I is I-stable, then ∃n such that I i Mn = Mn+i for all i ≥ 0,
so a generating set for M0 , . . . , Mn generated BI M .
Lemma 5.3 (Artin-Rees). Let R be a Nötherian ring, I ⊆ R an ideal, and
let M 0 ⊂ M be finitely generated R-modules. IF M = M0 ⊃ M1 ⊃ . . . isw an
I-stable filtration, then the induced filtration M00 = M 0 ⊂ M10 = M1 ∩ M 0 ⊃ . . .
is also I-stable, so ∃n such that M 0 ∩ Mn+i = I i (M 0 ∩ Mn ).
Note: IF R is Nötherian, then so is BI R, since I is finitely generated. So
BI R is a finitely generated R-algebra, so is Nötherian.
Proof. Let I 0 = M 0 = M00 ⊃ M10 ⊃ . . .. BI 0 M 0 is naturally a BI R submod-
ule of BI M . If I is stable, then BI M is a finitely generated BI R-module,
so all submodules are finitely generated and, in particular, BI 0 M 0 is finitely
generated, so I 0 is I-stable.
Q: Can we have 0 6= I 5 = I 20 in a nice ring? If so, actually have I j = I 5 for
all j ≥ 5. So we’d get ∩j≥1 I j = I 5 .
Corollary 5.4 (Krull Intersection Theorem). Let I ⊂ R be an ideal in a
Nötherian ring. If M is a finitely generated R-module, then there is an r ∈ I
such that (1 − r)(∩∞ j
j=1 I M ) = 0. If R is a domain or a local ring, and I is a

proper ideal, then ∩j=1 I j = 0.

Proof. For any P , ∩j≥1 I j M = (∩j≥1 I j M ) ∩ I p+1 M , now Artin-Rees applied


to ∩I j M ⊂ M for the I-stable filtration Mj = I j M says that ∃p such that
I(∩j≥0 I j M ∩ I p M ) = (∩j≥0 I j M ) ∩ I p+1 M = ∩j≥0 I j M . ie, ∃p such that
I(∩j≥0 I j M ) = ∩j≥0 (I j M ). So ∃r ∈ I such that (1 − r) ∩j≥0 I j M = 0.

6 Flatness
A Flat Family: A ”family” of varieties is one that varies with parameters.
Eg: V (x2 − a2 ) for a ∈ k is {a, −a}.
Eg: V (x) ∪ V (y − ax) = V (x(y − ax)), two lines through the origin.

25
A way to think about V (x2 − a2 ) is as a union of two lines projected down
to a line.
A family, then, is a map of varieties π : Y → X with the fibers π −1 (x) ⊂ Y
for x ∈ X.
Corresponding map of rings R → S in the other direction. IE, k[a] →
k[x, a]/(x2 − a2 ). So a family over Spec(R) is an R-algebra S.
Definition 6.1 (Fiber). The fiber over any prime P ⊂ R is K(R/P ) ⊗R S
eg (a − 7) ⊆ k[a] has fiber k[a]/(a − 7) ⊗k[a] k[a, x]/(x2 − a2 ) ' k[a, x]/(x2 −
a2 , a − 7) ' k[x]/(x2 − 49) is the ring of {7, −7}.
eg, Z/2Z is a Z-algebra, so Spec(Z2 ) → Spec(Z) is a family. The fiber over
7 is Z7 ⊗Z Z2 ' 0. So the fiber over (2) is Z2 ⊗Z Z2 and is trivial elsewhere.
We replace locally trivial is flat. A family is nice if it is flat, that is, S is a
flat R-module.
Recall that if 0 → A → B → C → 0 is a ses of R-mods, then A ⊗ M →
B ⊗ M → C ⊗ M → 0 is exact for any R-module M . If 0 → A ⊗ M → B ⊗ M
is exact, then M is called a flat R-module.
eg R = k[x] and S = k[x, y]/(x − y). Check: S is flat (since S ' k[x] = R as
an R-module).
Free R-modules are always flat.
eg: S = k[x, t]/(tx − 1), R = k[t]. Then S ' k[t, t−1 ], and R[U −1 ] is flat for
all multiplicative sets U ⊂ R.
eg: R = k[a, b], S = R[a, b, x, y]/(ax + by), then S is an R-module but is not
flat!
Important Example: ”Gröbner Degeneration”. If S = k[x1 , . . . , xn ], and
I ⊂ S is an ideal. Let w ∈ Rn and consider ≤w where xu < xv if w · u < w · v
or (other condition)
P Gröbner theory studies the ideal ∈w (I). So we define, given f ∈ S, f =
cu xu , set f˜ = f (xi /tw1 , . . . , xn /twn )tb ∈ S[t] where b = maxcu 6=0 w · u. ie,
f˜ = cu xu tb−w·u .
P
So if S = k[x, y], w = (2, 3) and f = x2 +y 3 , then f˜ = x3 +y 2 , if f = x3 +3xy
then F̃ = x3 + 3xyt.
Define It ⊂ S[t] to be It = hf˜|f ∈ Ii, then It |t=1 = I and It |t=0 =∈w (I).
So S[t]/It is a k[t]-module defining a family. Check that, for all a 6= 0, the fiber
over t = a is isomorphic to S/I.
Lemma 6.1. S[t]/It is a free, and thus flat, k[t]-module.
Recall: B = {xu |xu ∈∈
/ w (I)} is a basis for S/I as a k-vector space.
Proof. We claim that B is a k[t] basis for S[t]/It . That is, S[t]/It ' ⊕k[t]xu
over xu ∈ B.
The key point is that if G = {g1 , . . . , gr } is a Gröbner basis for I, then
{g˜1 , . . . , g˜r } generate It and g̃i =∈Pw (gi ) + t(other stuff). Given f ∈ S[t],
dividing by {g̃i } gives a polynomial pu (t)xu .P
u
”linear P 0independence” as before. If f = pu (t)xP = 0 in S[t]/It , then
k u
f = t ( pu (t)x ) are not all divisible by G so equals q(x, t)g̃i .

26
Summary: S[t]/It is a free and thus flat R-module, so V (It ) ⊆ k n × k is a
flat family.
We say this gives a Gröbner Degeneration from V (I) to V (∈ (I)).
Check: S[t]/It ⊗k[t] k[t]/(t) ' S/ ∈w (I) and S[t]/It ⊗k[t] k[t]/(t − a) ' S/I.
eg: I = (xy − y 2 ) ⊂ k[x, y], w = (2, 1), so It = (xy − t2 y 2 ), we get the union
of the x axis and a line of slope 1/t2 , and ∈w (I) = (xy).
eg: I = (x2 + y 2 − 4) ⊆ C[x, y], then It = (x2 + t2 y 2 − 4t2 ) ⊆ C[x, y, t],
∈≤ (T ) = (x2 ).
Let R = k[a, b] and S = k[a, b, x, y]/(ax + by), and S is not flat. Look at
(a, b) ⊗ S → R ⊗ S.
Tor
Flatness says that 0 → A → B → C → 0 exact implies 0 → A ⊗ M →
B ⊗ M → C ⊗ M → 0, but we get all but the first anyway.
If M is not flat, we would like to define Tor1 (M, C) → ker(A⊗M → B ⊗M ).
’General Homological Algebra’
If F is a right exact functor from R-modules to R-modules, then given
0 → A → B → C → 0 we get . . . → L1 F B → L1 F C → F A → F B → F C → 0.
Definition 6.2. Let P : . . . → P2 → P1 → P0 → M → 0 be a projective
resolution of an R-module M (ie P is exact with each Pi projective).
If R = k[x1 , . . . , xn ] and M = R/I ten we can construct a free resolution
using Gröbner Bases by Schreyer’s Algorithm.

Definition 6.3. Given a projective resolution P , define F P to be . . . → F P2 →


F P1 → F P0 → F M → 0. This is still a chain complex, so we take homology,
and LFi = ker F ϕi / Im F ϕi+1 .
For us: Tori (M, N ) = LFi (N ) where F is −⊗R M . So to compute Tori (M, N ),
we compute a projective resolution of N , tensor with M , and take the ith ho-
mology.
Basic Facts:

1. It doesn’t matter what resolution we take.


2. Tori (M, N ) = Tori (N, M ).
3. Tor0 (M, N ) = M ⊗R N .

4. If M is projective, then Tori (M, N ) = 0 for all i > 1.


x
Examples: If x ∈ R is a nonzero divisor then 0 → R → R → R/(x) → 0.
Claim: This is a free resolution of R/(x). Thus Tori (R/(x), M ) = R/(x)⊗R M =
x
M/xM if i = 0, is ker(M → M ) = (0 : m x) = {m ∈ M : xm = 0} if i = 1 and
0 else.
eg. R = k[x, y] and M = N = k[x, y]/(x, y) ' k. What is TorR i (k, k). Then
(x,y) (y,−x)t
0 ← M ← R ← R2 ← R ← 0 is a free resolution.

27
So then Tori (k, k) = k if i = 0, k 2 is i = 1 and k if i = 2, 0 else. In general,
if R = k[x1 , . . . , xn ] and M = R/(x1 , . . . , xn ), then Tori (M, N ) = k βi for some
βi , and we call the βi the Betti numbers.
A long exact sequence: If 0 → A → B → C → 0 then we get Tori (A, M ) →
Tori (B, M ) → Tori (C, M ) → Tori−1 (A, M ) → . . . → Tor1 (C, M ) → A ⊗ M →
B ⊗ M → C ⊗ M → 0.
More facts about Tor: If S is a flat R-algebra, then S ⊗R TorR i (M, N ) =
TorSi (S ⊗R M, S ⊗R N ).
If we have a short exact sequence 0 → A → B → C → 0 we get a long exact
sequence → TorR R R R
i (A, M ) → Tori (B, M ) → Tori (C, M ) → Tori−1 (A, M ) →
. . . → Tor1 (B, M ) → Tor1 (C, M ) → A ⊗ M → B ⊗ M → C ⊗ M → 0.
Proposition 6.2. Let R be a ring and M an R-module. If I is an ideal of R
then the multiplicative map I ⊗R M → M is an injection iff TorR i (R/I, M ) → 0.
The module M is flat iff this condition is satisfied for all finitely generated
I.
Proof. Consider the ses 0 → I → R → R/I → 0. From this we get a long exact
sequence Tor1 (R, M ) → Tor(R/I, M ) → I ⊗ M → R ⊗ M = M .
As Tor1 (R, M ) = 0, I ⊗M → M is injective iff Tor1 (R/I, M ) = ker(I ⊗M →
M ) is zero.
Recall that M is flat iff M ⊗ N 0 → M ⊗ N is an inclusion for all inclusions
N0 ⊆ N.
First we assume this for all N 0 finitely generated I and N = R. We must
show that this implies the general condition
Ps for N 0 ⊆ N . First, let I be a general
ideal of R and x ∈ I ⊗ M . Then x = i=1 ri ⊗ mi . Let I 0 be the ideal generated
by < r1 , . . . , rs >. Then x ∈ I 0 ⊗ M → M is an inclusion, so x 6= 0 in M .
Now consider N 0 ⊆ N . BY the same argument we may assume that N is
finitely generated. ie, x ∈ N 0 ⊗ M , then x =
P
ni ⊗ mi , let Ñ be the submodule
generated by the ni and any necessary relations. We can thus assume that N
is finitely generated. So we can find a filtration N 0 = N0 ⊂ N1 ⊂ . . . ⊂ Nr = N
with each Ni /Ni−1 cyclic.
It suffices to show that Ni ⊗ M → Ni+1 ⊗ M is an inclusion, so we may
assume that N/N 0 ' Rx ' R/I. Now from 0 → N 0 → N → N/N 0 → 0 we get
Tor1 (N/N 0 , M ) → N 0 ⊗ M → N ⊗ M . Tor1 (N/N 0 , M ) = Tor1 (R/I, M ) = 0 by
hypothesis. So N 0 ⊗ M → N ⊗ M is an inclusion for arbitrary N 0 ⊆ N , so M
is flat.
The point was that I ⊗ M → M being an inclusion for all finitely generated
I implies that N 0 ⊗ M → N ⊗ M is an inclusion for all N 0 ⊆ N .
So to check flatness, it is enough to check finitely generated ideals.
Corollary 6.3. Let k be a field and R = k[t]/t2 , and M an R-module. Then M
is flat iff multiplication by t from M to tM induces an isomorphism M/tM →
tM .
Proof. The only nonzero ideal in R is (t). So M is flat iff (t) ⊗ M → M is
an injection. As (t) ' R/(t) as an R-module by t ↔ 1, we have (t) ⊗ M '

28
R/(t) ⊗ M ' M/tM , so the map M/tM → tM by m 7→ tm is the composition
R/t ⊗ M → t ⊗ M → M . So it is injective.
Corollary 6.4. If a ∈ R is a nzd in R and M is a flat R-module, then a is a
nzd on M .
If R is a PID, then the converse is true: M is flat iff M is torsion free.
Proof. Let a ∈ R be a nzd and M flat. I = Ra ' R by 1 7→ a. So we have
R ⊗ M → I ⊗ M → R ⊗ M by 1 ⊗ m → a ⊗ m → a ⊗ m, so m 7→ am. Simce
the map m 7→ am is injective, a is a nzd on M .
Suppose that R is a PID and M is torsion free, so no element of R annihilates
an element of M . Then for any a 6= 0 in R, Ra ⊗ M → M is an injection, since
0 ⊗ M → M is an injection as well, this means that I ⊗ M → M is an inclusion
for all finitely generated I, thus M is flat.
Definition 6.4 (Rees Algebra). The Rees Algebra of R with respect to I,
P∞
R[R, I] = R[t, t−1 I] ⊆ R[t, t−1 ]. It is n=−∞ I n t−n with I n = R for n ≤ 0.

If R is a k-algebra (k a field) we’ll see that R[R, I] is a flat k[t]-algebra.


Facts: R[R, I]/iR[R, I] = grI R = R/I ⊕ I/I 2 ⊕ I 2 /I 3 ⊕ . . ..
If a 6= 0 in R, then R[R, I]/(t − a)R[R, I] ' R.
So R[R, I] is a family over [t] with fiber over t = 0 grI R and fiber over
everything else R.

Lemma 6.5. If R is a k-algebra, then S = R[R, I] is flat over k[t].


If ∩∞ d
d=1 I = 0 then every element of the form 1 − ts with s ∈ S is a nzd on
R[R, I].
Proof. Since k[t] is a PID, it suffices to observe that S is torsion free as k[t]-
module. This is immediate from the fact that S = R[t, It−1 ] ⊆ R[t, t−1 ].
For the second statement, suppose first that p(1 − ts) = 0 ∈ S. This means
that p ∈ (t), p = qt. t is a nzd on S, so q(1 − ts) = 0 in S, so q ∈ t, so p ∈ t2 .
n
Pj that p j∈ t S for all n.
Continue to get
Now p = i=−j pi t . Since p ∈ tn S for all n, we must have pi ∈ I m for all
m.

What to take away: Flatness is a niceness property, and flat families preserve
a lot of properties (ie, dimension)

7 Completions
The basic idea is that the open sets in the Zariski topology are too big, so we
look for smaller neighborhoods.
eq if R = k[x1 , . . . , xn ] and m = (x1 , . . . , xn ), then Rm = {f /g : g(0) 6= 0}.
We replace this by R̂ = k[[x1 , . . . , xn ]] formal power series, and we get a natural
map Rm → R̂.
One advantage is that we get a version of the inverse function theorem.

29
Definition 7.1 (Inverse Limit). If {Gi |i ∈ N} is a sequence Q of abelian groups

with homomorphisms ϕi : Gi → Gi−1 . Then lim Gi = {g ∈ i=1 Gi |ϕi (gi ) =
←−
gi−1 } is the inverse limit, which is an abelian group under coordinatewise addi-
tion.

We will be interested in the case where we start with a ring R and a filtration
m1 ⊃ . . . mn ⊂ . . . of ideals and set Gi = R/mi . Write R̂ = lim R/mi .
←−
R̂ is a ring by coordinate multiplication. Most important case is mi = mi
for some ideal m ⊂ R. Notation is R̂m . eg R = k[x], m = (x), then R̂m =
lim k[x]/xi .
←−
Claim: R̂m = k[[x]].
Pi−1
ai xi 7→ bi = n=0 an xn .
P
Proof. ϕ : k[[x]] → R̂m , a → (b1 , b2 , . . .) by
This is a well-defined homomorphism, so we just need to check that it is iso.
For the inverse map, given b = (b1 , . . .) ∈ R̂m , each bi has a representation of the
Pi−1
form j=0 aij xj and if k < ` then akj = a`j for j < k. Define ψ : R̂m → k[[x]]
P∞
by b 7→ j=0 aij xj .

Definition 7.2 (Complete with respect to m). There is a natural map, R → R̂m
by r 7→ (r, r, r, . . .), if this is an isomrophism, then R is complete.
Theorem 7.1 (Cohen Structure Theorem). If R is a complete local ring con-
taining a field, then R = k[[x1 , . . . , xn ]]/I for some I.

eg, the p-adics, p ∈ Z, then Ẑp = lim Z/pn , with ϕ : Z/pn → Z/pn−1 by
←−
a 7→ a. P∞
Then we can write elements of Ẑp as i=0 ai pi for 0 ≤ ai < p with addition
is ”add with carrying”
In Ẑ2 , 1 + 2 + 4 + 8 + 16 + . . . is b = (1, 3, 7, . . .) = (−1, −1, −1, . . .), and
1 = (1, 1, 1, . . .).
If we have m1 ⊃ m2 ⊃ . . . we get an ideal m̂1 = {g = (g1 , . . .)|gj = 0∀j ≤
i} ⊆ R̂. When mi = mi , then m̂1 = m̂.
Lemma 7.2. When m is a maximal ideal in R, then R̂m is a local ring. For
any filtration, we have R̂/m̂i = R/mi .

Proof. If g ∈ R̂/m̂i , then map g + m̂i 7→ gi + mi is the ”projection homomor-


phism”
If ϕ(g+m̂i ) = 0 then gi ∈ mi , so gj ∈ mj for j < i. So g = (0, 0, 0, . . . , 0, −, . . .)+
m̂i , so g ∈ m̂i , so g = 0.
It is surjective, since R → R̂ → R/mi by r 7→ (r, r, r, r, . . .) → r + mi .
Thus R̂/m̂i ' R/mi . If m is a maximal ideal in R, then R/m is maximal.
Thus, R̂/m̂ is a field, and so m̂ is a maximal ideal of R̂. We now show that if
g ∈ R̂ \ m̂, then g is a unit.
Note that each R/mi is a local ring with maximal ideal m. So if gi ∈ R/mi \m,
then gi is a unit in R/mi . If g = (g1 , g2 , . . .) ∈ R̂m \ m̂, then g1 6= 0, so each
gi ∈/ mR/mi .

30
Since gi = gj mod m, then gi−1 = gj−1 mod m, so set g −1 = (g1−1 , . . .), and
note that gg −1 = (1, 1, 1, . . .).

eg k[[x]] is local with maximal ideal (x). Ẑp is local with maximal ideal (p).
Note: If m is a maximal ideal, R/mi ' (R/mi )m = Rm /mmi .
So we get the same R̂m if we localize at m first.
So if m is maximal, we say that R̂m is a complete local ring.
Note: If R is complete with respect to m, then ∩∞ j
j=0 m = 0.

Definition 7.3 (Convergence). A sequence a1 , a2 , . . . ∈ R̂ converges to an ele-


ment a ∈ R̂ if ∀n∃in such that a−aj ∈ m̂n for j ≥ in , ie ain = (a1 , a2 , . . . , an , other).
A sequence is Cauchy if ∀n∃in such that ai − aj ∈ m̂n for i, j ≥ in .
A sequence converges iff it is Cauchy.
This is the usual notion of convergence in the m-adic topology which has a
base of open sets {a + m̂i |a ∈ R̂, i ≥ 1}.
Cauchy implies Convergence: we set a = lim ai , ie, if {ai } is Cauchy, ai =
(ai1 , ai2 , . . .), set a = (b1 , b2 , . . .) with bn = ajn for any j > in .
This is well defined, since if j, k ≥ in , aj − ak ∈ m̂n so ajn = akn . Check
a ∈ R̂: If j > in , then bn = ajn and bn−1 = ajn−1 so bn = bn−1 mod mn−1 .
Check: {ai } converges to a. Given n, ∃in such that ∀i, j ≥ in , ai − aj ∈ m̂n .
By construction, for such an i, ai −a ∈ m̂n , so ∀i ≥ in , ai −a ∈ m̂n , so {ai } → a.
From analysis: we know that if ai → a and bi → b, then ai + bi → a + b and
ai bi → ab.
Application:
Proposition 7.3. If R is complete with respect to m, then U = {1 − a|a ∈ m}
are units in R.
Pi
Proof. If a ∈ m, set ai = j=0 aj = 1 + a + a2 + . . .. Then the sequence {ai } is
Cauchy. If i, j > n − 1, then ai − aj ∈ mn . So it converges to some b ∈ R. And
(1 − a)ai converges to (1 − a)b.
But (1 − a)ai = 1 − ai+1 , so it converges to 1.

We saw this with


Pi a = 2 in Ẑ2 , there 1 + 2 + 4 + 8 + . . ., so the limit of the
Cauchy sequence j=0 2i is −1.
Corollary 7.4. R is a local ring with maximal ideal P , then R[[x1 , . . . , xn ]] is
local with maximal ideal P + (x1 , . . . , xn ).

Proof. R[[x1 , . . . , xn ]] = R[x1\


, . . . , xn ](x1 ,...,xn ) . If f ∈ P + (x1 , . . . , xn ), then
f has constant term f0 ∈ / P , so f0 is a unit in R. So f0−1 f = 1 + g for
g ∈ (x1 , . . . , xn ).
So 1 + g is a unit, thus f is a unit as well.
Definition 7.4. Let R = m0 ⊃ m1 ⊃ . . . be a filtration of ideals, and let gr R
be the associated graded ring.
Given f ∈ R let m = max{j|f ∈ mj }. Define in(f ) = f + mm+1 ∈
mj /mj+1 ∈ gr R.

31
Proposition 7.5. Suppose that R is a ring complete with respect to m1 ⊃ m2 ⊃
. . .. Suppose I ⊂ R is an ideal, a1 , . . . , as ∈ I. If in(a1 ), . . . , in(as ) generate
in(I) = (in(f )|f ∈ I) ⊂ gr R. Then {a1 , . . . , as } generate I.
Proof. Let I 0 = (a1 , . . . , as ). We may assume that none of the ai are 0. Choose
d >> 0 so that ai ∈ / md for all i. Ps
Given f ∈ I with in(f ) of degree e, we can write in(f ) = j=1 Gj in(aj ) with
Gj ∈ grm R homogeneous of P degree deg(in(f )) − deg(in(aj )). Take g1 , . . . , gs
s
with in(gj ) = Gj . Then f − j=1 gj aj ∈ me+1 .
Repeat: we can get f ∈ I with f − f 0 ∈ md+1 . Now keep repeating, noting
0 0

that now deg(Gj ) ≥ e − d > 0.


P (0) P (1) P (n) (i)
gj ∈ me−d , so get f − j gj aj − j gj aj − . . . − j gj aj with gj ∈
me−d+i .
P∞ (i)
f n ∈ me+n+1 , so P the series i=0 gj converges, and we will call the limit
hj . Now look at f − hj aj . This is 0, since ∩mi = 0, and so f ∈ I 0 , thus the
ai generate I.
Theorem 7.6. Let R be a Nötherian ring and let m be an ideal of R.

1. R̂m is Nötherian.
2. m̂n = mn R̂m , so grm̂ R̂ = grm R.

Proof. Write gr R̂ for the associated graded ring of R with respect to m̂i . Then
since R̂/m̂i ' R/mi , we have gr R̂ = grm R. Since R is Nötherian, so is R/m.
The ring grm R is finitely generated as an R/m algebra (since m is a finitely
generated ideal). So grm R ' R/m[x1 , . . . , x` ]/J for some ideal `, J, thus grm R
is Nötherian by the Hilbert Basis Theorem.
So gr R̂ is Nötherian. So for any ideal I ⊆ R̂, the ideal in(I) ⊆ gr R̂ is finitely
generated by the initial forms of a1 , . . . , as , so a1 , . . . , as generate I, and so R̂
is Nötherian.
To show that m̂n = mn R̂m it suffices to show that both have the same
initial ideals in gr R̂. (This uses Nötherian, as we need ideals in gr R̂ to be
finitely generated to apply the prop). But both initial ideals consist of all terms
of degree ≥ n.
Theorem 7.7. Let R be a Nötherian ring. Let I is an ideal of R.

1. If M is a finitely generate R-module, then the natural map R̂ ⊗R M →


lim M/I j M = M̂ is an isomorphism.
←−
2. R̂ is a flat R-module.

Lemma 7.8 (Hensel’s Lemma). Let R be a ring that is complete with respect
to an ideal m. Let f (x) ∈ R[x]. If a is an approximate root of f in the sense
that f (a) ≡ 0 mod f 0 (a)2 m then there is a root b of f near a in the sense that
f (b) = 0 and b ≡ a mod f 0 (a)m.
If f 0 (a) is a nzd on R, then b is unique.

32
Most often used when f 0 (a) is a unit.
Application
Is 8 a square in Ẑ7 ? P∞
Take c ∈ Ẑp . Write c = pn b where p 6 |b. (ie b = i=0 bi pi , bP 0 6= 0). Then

c is a square iff n is even and b is a square. So we’ll assume c = i=0 ci pi has
c0 6= 0. P∞
If c = d2 , d = i=0 di pi , then c0 = d20 mod p.
Consider f (x) = x2 − c. (assume p > 2). Then f (d0 ) = d20 − c ∈ (p).
f (x) = 2x, so f 0 (d0 ) = 2d0 6= 0, so is a unit in Ẑp . So Hensel’s lemma says that
0

there is a root of f , so c is a square.


So as 8 = 1 + 7, and 1 is a square mod 7, 8 must be square.
Idea: use Newton’s method to construct a1 , a2 , a3 , . . . by ai+1 = ai −f (ai )/f 0 (ai ).
Claim: This is a convergent sequence with limit b and f (b) = 0. Recall Tay-
lor’s Theorem that f (x + y) = f (x) + f 0 (x)y + h(y)y 2 for some polynomial
h(y) ∈ R[x][y].
Fact 1: If f 0 (a1 ) is a unit, so is f 0 (ai ) for all i (assume that f (ai ) ∈ m).
bi = f 0 (ai+1 ) − ai = −f (ai )/f 0 (ai ) = f 0 (ai ) + f 00 (ai )bi + hai (bi )b2i is a unit
plus something in m. As such, f (ai + 1) is a unit.
Fact 2: If f (a1 ) ∈ mi , f (ai+1 ) ∈ m2i . f (ai+1 ) = f (ai + bi ) = f (ai ) +
0
f (ai )bi + hai (b1 )b2i = ha (b1 )b2i ∈ m2i .
f (ai+1 ) ∈ m2i implies that bi+1 ∈ m2i , so ai+1 − ai ∈ m2i , so {ai } is Cauchy.
Let b be the limit. Argue that f (b) = 0.
Question: What about better approximation? ie, can we replace m by
m2 , . . . , mn ?
Yes, and the same proof/statement works, because if R is complete wrt m,
it is complete wrt mn .

Proposition 7.9 (Universal Property of Inverse Limit). If {Gi } is a set of


abelian groups with ϕi : Gi → Gi−1 and H is an abelian group with ψi : H → Gi
such that ϕi ψi = ψi−1 , then there exists a unique map ψ : H → lim Gi such that
←−
everything commutes.
Note: If R is a ring and S is and R-algebra that is complete with respect to
n and f1 , . . . , fn ∈ n, then ∃! R-algebra homomorphism ϕ : R[[x1 , . . . , xn ]] → S
sending xi to fi for each i. Theorem 7.16.

8 Dimension
Definition 8.1 (Krull Dimension). The Krull dimension is the supremum of
the lengths of ascending chains of distinct prime ideals.
The motivation is: The dimension of a vector space is the length of the
longest (or any maximal) chain of subspaces.
eg. Let k be a field. Then dim k = 0.
dim k[x] = 1, as 0 ⊆ (x) and if 0 ⊆ (p) ⊆ (q) then (q) = (p).
This actually proves the following:

33
Lemma 8.1. If R is a PID, then dim R = 1.
In affine algebraic geometry, the dimension of a variety V (I) ⊆ An is the
length of the longest chain of subvarieties V (I) = Vr ⊃ Vr−1 ⊇ . . . ⊇ V0 , which
is dim k[x1 , . . . , xn ]/I.
For rings of the form R = k[x1 , . . . , xn ]/I every maximal chain of primes has
the same length (a ring with this property is called Catenary)
Properties of Dimension

1. dim R = supP dim RP over all P prime in R.


2. Nilpotents don’t affect dimension. ie, if I is a nilpotent ideal of R (so
I k = 0 for some k), then dim R = dim R/I.

3. Dimension is preserved by maps with finite fibers. If R ⊆ S are rings such


that S if a finitely generated R-module, then dim R = dim S.
4. Calibration: if k is a field, then dim k[x1 , . . . , xn ] = n.
5. If R is an affine domain over a field R ' k[x1 , . . . , xn ]/I for I prime, then
dim R = tr degk R.

6. If R is Nötherian local with maximal ideal m, then dim R is the minimum


n such that ∃n elements f1 , . . . , fn ∈ m not in any prime other than m.
7. If R is an N-graded ring R0 = k generated in degree 1, then there exists
a polynomial P such that P (n) = dimk Rn for n >> 0. Then dim R =
1 + deg P . This polynomial is called the Hilbert Polynomial.
eg, R = k[x], deg(x) = 1, then P (n) = 1 for all n, so dim R = 1. This
gives an algorithm to compute dimension for R = k[x1 , . . . , xn ]/I.

Fact: Flat implies that the fibers have the same hilbert polynomial
By convension: if I ⊆ R is an ideal, the dimension of I is the dimension of
the ring R/I.
If M is an R-module, then the dimension of M is the dimension of Ann M
(which is dim R/ Ann M ). Think about this in the same way as we do primary
decomposition, so we just ignore the dimension of I as an R-module.
If I is prime, then the codimension of I is the dimension of RI , that is, the
length of the longest chain of primes contained in I. As dim I = dim R/I and
codim I = dim RI , dim R ≥ dim I + codim I.
For general I, codim I = min of codim P where P is a prime containing I.
If M is an R-module, then codim M = codim Ann M .
Today: 0-dimensional Nötherian rings
If R has dimension zero, then all primes are maximal.
If R is a zero-dimensional domain, then 0 is a maximal ideal and so R is a
field.
Recall: A ring R is Artinian iff it satisfies the descending chain condition on
ideals.

34
Definition 8.2 (Composition Series). If M is an R-module, a composition
series for M is M = M0 ⊇ M1 ⊇ . . . ⊇ Mn = 0 with Mi /Mi+1 a nonzero
simple module (has no nontrivial submodules)
The length of M is the smallest length of a composition series for M or ∞
if M has no finite composition series.
eg: the length of Z as a Z-module if ∞.
eg: the length of R = k[x]/x2 as an R-module is R ⊃ (x) ⊇ 0, so 2.
Proposition 8.2. Let R be a ring and let M be an R-module. M has a finite
composition series iff M is Artinian and Nötherian.
In tis case, any filtration of submodules of M has length at most n and can
be refined to a composition series.
Proof. Suppose that M is Artinian and Nötherian. Then by the ACC, M has
a maximal proper submodule M1 , which has a maximal proper submodule M2 ,
etc. This gives a descending chain of proper submodules M ⊇ M1 ⊇ M2 ⊇ . . .
where each Mi /Mi+1 is simple. As M is Artinian, this chain must be finite, so
some Mn = 0.
Suppose now that M has a finite composition series M = M0 ⊇ M1 ⊇ . . . ⊇
Mn = 0. We first show that if M 0 ⊆ M is a proper submodule, then the length
of M 0 is less than the length of M .
Indeed, consider Mi0 = M 0 ∩ Mi . Then M 0 = M00 ⊇ M10 ⊇ . . . ⊇ Mn0 = 0.
And Mi0 /Mi+10
= (M 0 ∩ Mi )/(M 0 ∩ Mi+1 ) = Mi0 + Mi+1 /Mi+1 ⊆ Mi /Mi+1 . So
0 0
Mi /Mi+1 is either Mi /Mi+1 and is simple, or it is zero. There must be at least
one i for which Mi0 /Mi+1 0
= 0, because otherwise we would get M 0 ⊇ Mi for
all i by descending induction, Mn = 0 ⊆ M 0 , so Mi = Mi0 + Mi+1 ⊆ M 0 by
induction. Then, M 0 ⊃ M0 , which is a contradiction.
This gives a filtration of M 0 = M00 ⊇ M10 ⊇ . . . ⊇ Mn0 = 0 where we leave
out any repeated factor to get length < n. This is a composition series because
successive quotients are simple. Thus, if we start with a composition series for
M of minimal length, then M 0 has smaller length.
Now suppose that M = N0 ⊇ N1 ⊇ . . . ⊇ Nk is a chain of submodules of M .
By assumption, M has a composition series of length n. We will show that k ≤
n. When n = 0, M = 0, so k = 0. Assume now that for m < n a composition
series of lenght m implies that all filtrations of submodules have length ≤ m.
Now N1 is a proper submodule of M , and so has length < length(M ) ≤ n. So
this means that k − 1 < n, so k ≤ n.
Thus, every chain of submodules is finite, and so in particular every ascend-
ing chain and every descending chain is, so M is Artinian and Nötherian.

Corollary 8.3. If M has length n, then every composition series of M has


length n.
Corollary 8.4. If R is of finite length as an R-module, then R is Artinian and
Nötherian.
Theorem 8.5. TFAE

35
1. R is Nötherian and all primes are maximal.
2. R is of finite length as an R-module.
3. R is Artinian.

Proof. 1 ⇒ 2: Let R be Nötherian with all primes maximal. Suppose that R


is not of finite length as an R-module. Let I be an ideal maximal with respect
to the property that R/I is not of finite length as an R-module. We claim
that I is prime. If not, we can find a, b ∈ R \ I with ab ∈ I. Then consider
0 → R/(I : a) → R/I → R/(I, a) → 0. Both (I : a) and (I, a) are strictly
larger ideals, so R/(I : a) and R/(I, a) have finite length as R-modules. We
can now get a finite composition series for R/I from the ones for R/(I : a) and
R/(I, a), because length is additive in short exact sequences. This contradicts
our assumption on I, so I is prime. As I is prime, it is maximal, so R/I is a
field. Which has finite length. This contradiction means that I does not exist,
so R has finite length as an R-module.
2 ⇒ 3: follows from the previous proposition.
3 ⇒ 1: Suppose that R is Artinian. We’ll first show that 0 is a prime of
finitely many maximal ideals. Since R is Artinian, we may choose from all ideals
in R that are products of finitely many maximal ones a minimal one J. We want
to show that J = 0. For each maximal ideals M , we must have M J = J, so
J ⊂ M , so J is contained in the intersection of all maximal ideals. J 2 = J
If J is nonzero, we can find an ideal I minimal wrt not annihilating J. Since
(IJ)J = IJ 2 = IJ 6= 0, IJ ⊆ I. We must have IJ = I by minimality of I.
Also, ∃f ∈ I with f J 6= 0, so I = (f ). Since IJ = I, ∃g ∈ J with f g = f , so
f (g − 1) = 0, but g − 1 is in no maximal ideal, so it is a unit. Thus f = 0, and
so I = 0 and J = 0.
The above is LEMMA: If R has finite length as an R-module, then 0 =
M1 . . . Mt where Mi are maximal.
Consider the R-module Vs = M1 . . . Ms /M1 . . . Ms+1 for 0 ≤ s < t. This
is an R/Ms+1 -module, and so a vector space. Subspaces of Vs correspond to
ideals of R containing M1 . . . Ms+1 and contained in M1 . . . Ms . Since R is
Artinian, Vs must be finite dimensional, so has a finite composition series. We
now glue together the composition series for R/M1 , M1 /M1 M2 , . . . to get a finite
composition series for R.
So R has finite length as an R-modules, and is thus Nötherian. Now let P
be a prime in R. Then 0 = M1 . . . Mt ⊆ P , so at least one Mi ⊆ P . But since
Mi is maximal, Mi = P . Thus we have only a finite number of primes in R,
each of which is maximal.
Corollary 8.6. If R is Nötherian and dimension 0, then R is Artinian and
there are only finitely many primes. Also, (0) is the product of powers of these
primes.
Corollary 8.7. Let V = V (I) ⊆ k n be a variety over k = k̄. TFAE

1. V is a finite set.

36

2. k[x1 , . . . , xn ]/ I is a finite dimensional k-vector sapce whose dimension
is |V |.

3. k[x1 , . . . , xn ]/ I is Artinian.

Proposition 8.8. Let R be Nötherian and M a finitely generated R-module.


TFAE
Qs
1. Some finite produt of maximal ideals i=1 Pi annihilates M .
2. All primes that contain the annihilator of M are maximal.
3. R/ Ann(M ) is Artinian.
Q Q
Proof. Suppose that Pi annihilates M . Let P ⊃ Ann(M ). Then P ⊇ Pi .
So there exists i with Pi ⊆ P , and so Pi = P , as Pi is maximal.
R is Nötherian, so R/ Ann(M ) is, and every prime in R/ Ann(M ) is maximal,
so by the theorem, R/ Ann(M ) is Artinian. Q
QIf R/ Ann(M ) is Artinian and Nötherian, then 0/ Ann(M ) = (Pi / Ann(M )),
so Pi annihilate M .
Recall: If P ⊆ R is prime, codim P = dim RP . (if (R, m) is local, codim m =
dim R)
For general I, codim I = min codim P for P ⊃ I.
Goal: R Nötherian.
Theorem 8.9. If x1 , . . . , xc ∈ P and P is minimal among primes containing
x1 , . . . , xc , then codim P ≤ c.
This is a generalization of
Theorem 8.10 (Principle Ideal Theorem). If x ∈ R and P is minimal among
primes of R containing x, then codim P ≤ 1.
Proof. We will show that every prime Q properly contained in P has codim 0.
Since P is minimal over x, all primes in RP /x are maximal (since PP /x is
the only one) and RP /x is Nötherian, so RP /x is Artinian. Also, x ∈ / Q.
Look at the chain QQ + (x), Q2Q + (x) + . . . in RQ . It must stabilize, so
QnQ + (x) = Qn+1 Q + (x). So for any f ∈ QnQ , we can write it as f = ax + g with
g ∈ Qn+1Q and a ∈ RQ .
So ax ∈ QnQ , and thus a ∈ QnQ , since x is not.
Thus, QnQ = xQnQ +Qn+1 n+1
Q , so the finitely generated RQ /QQ -module QQ /QQ
n n+1

satisfies QnQ /Qn+1


Q = xQnQ /Qn+1
Q .
The Idea WOULD HAVE BEEN appeal to Nakayama to get that QnQ =
Qn+1 n
Q , and thus, QQ = 0 and thus codim Q = 0. Look in book, and attempt to
fill in details.
Theorem 8.11. If x1 , . . . , xc ∈ R and P is minimal over primes containing
x1 , . . . , xc , then codim P ≤ c.

37
Proof. Localize at P to assume that R has a unique maximal ideal. Then since
all primes in R/(x1 , . . . , xc ) are maximal, as P is the only one, we know that
Ann(R/(x1 , . . . , xc )) ⊃ P k for some k. Choose P1 prime with P1 ( P with no
primes between them. As R is Nötherian, this exists if codim P > 0.
We will show that P1 is minimal over an ideal generated by c − 1 elements,
which will, by induction that codim P1 ≤ c − 1, and so since P1 was arbitrary,
codim P ≤ c.
Since P is minimal over (x1 , . . . , xc ), there must be an xi , say x1 , with
x1 ∈/ P1 . Thus P is minimal over (P1 , x1 ), and so ∃s with P s ⊆ (P1 , x1 ).
So for 2 ≤ j ≤ c, we can write xsj = aj x1 + yj with yj ∈ P1 . Now we claim
that P1 is minimal over (y2 , . . . , yc ). Indeed, P is minimal over (x1 , y2 , . . . , yc ).
So codim P/(y2 , . . . , yc ) in R/(y2 , . . . , yc ) is ≤ 1 by PIT. So P1 /(y2 , . . . , yc )
has codim 0, and so P1 is minimal over (y2 , . . . , yc ).
Corollary 8.12. (x1 , . . . , xn ) ⊂ k[x1 , . . . , xn ] has codim = n.
Corollary 8.13. Prime ideals in a Nötherian ring satisfy the DCC with the
length of a chain of ideals descending from a prime bounded by the number of
generators of P
Corollary 8.14 (Converse to PIT). Any prime P of codim P = c is minimal
over an ideal generated by c elements.
eg, R = k[a, b, c, d] and I = (ad − bc, ac − b2 , bd − c2 ). dim R/I = 2.
Corollary 8.15. Any prime P of codim c is minimal over an ideal generated
by c elements. (actually of codim c)
Proof. The proof is by induction on c.
If c = 0, then P is minimal over 0, which has codim 0.
Now suppose that the corollary is true for smaller c and choose P1 prime,
maximal proper in P of codim c − 1.
Then, by induction, P1 is minimal over (x1 , . . . , xc−1 ) of codim c − 1. Each
prime Qi minimal over (x1 , . . . , xc−1 ) has codimension c − 1.
So P ( Qi for any Qi minimal over (x1 , . . . , xc−1 ), so by prime avoidance,
P 6⊆ ∪Qi . So we can find xc ∈ P \ ∪Qi . Then we must have P minimal
over (x1 , . . . , xc ), as if P 0 ( P contains it, then there exists P 00 minimal over
(x1 , . . . , xc−1 ) which has codim c − 1, so codim P > c, which is impossible.
Also, if Q is minimal over (x1 , . . . , xc ), then ∃P 0 with (x1 , . . . , xc−1 ) ⊆
0
P ( Q and codim P ≥ c − 1. So codim Q ≥ c. By PIT codim Q = c, so
codim(x1 , . . . , xc ) = c.
Corollary 8.16. Let (R, m) be a local ring. Then dim R = min{d|∃x1 , . . . , xd ∈
m with mn ⊂ (x1 , . . . , xd ) for n >> 0}.
Proof. If mn ⊂ (x1 , . . . , xd ) ⊂ m, then m is minimal over (x1 , . . . , xd ). So
codim m = dim R ≤ d by the PIT.
Conversely, it d = dim R = codim m, then by the converse to the PIT,
∃x1 , . . . , xd with m minimal over (x1 , . . . , xd ). But R/(x1 , . . . , xd ) has only the

38
one prime ideal, we have that all elements of m are nilpotent module (x1 , . . . , xd ).
So ∃k such that mk ⊂ (x1 , . . . , xd ), because the ring is Nötherian. (since if
m = (y1 , . . . , ys ), yiN = 0 for all i, then msN = 0)
Definition 8.3 (System of Parameters). A system of parameters for m ⊂
R with (R, m) local, is a sequence x1 , . . . , xd , d = dim R such that mn ⊂
(x1 , . . . , xd ) for n >> 0.
Parameters are a ”local coordinate system” (up to finite ambiguity). eg,
k[x, y](x,y) , parameters (x2 − y, y), so (up to the square) you are picking a
horizontal line and a parabola to determine a point (determines 2, but that’s
finite ambiguity).
π
Recall: A family of varieties is U → B. The fiber over b is ”π −1 (b)” Alge-
braically, this is a map R → S with fiber over P ⊆ R being S ⊗ R/P ' S/P S.
We expect dim U ≤ dim B+dim π −1 (b) for each b. Recall the blowup, π −1 (0)
has dimension 1 and dim B = dim U = 2 for the plane at the origin.
Theorem 8.17. If (R, m) → (S, n) is a map of local rings, then dim S ≤
dim R + dim S/mS.
Proof. Write d = dim R, e = dim S/mS. Then there exist x1 , . . . , xd , s with
ms ⊆ (x1 , . . . , xd ) and y1 , . . . , ye , t such that nt ⊆ (y1 , . . . , ye ) + mS.
Then, nst ⊆ ((y1 , . . . , ye )+mS)s ⊆ ms S+(y1 , . . . , ye ) ⊆ (x1 , . . . , xd , y1 , . . . , ye )S,
so n is minimal over na ideal generated by d + e elements, so codim n = dim S ≤
d + e = dim R+ ∼ S/mS.
Theorem 8.18. If (R, m) → (S, n) is a map of local rings and S is flat over
R, then dim S = dim R + dim S/mS.
We’ll need the following:
Lemma 8.19. If N is a flat R-module and R → T is any ring map, then
N ⊗R T is flat over T .
Lemma 8.20. Suppose ϕ : R → S is a map of rings with S flat over R. If
P ⊃ P 0 are primes of R, and Q is a prime of S with ϕ−1 (Q) = P , then ∃Q0 of
S contained in Q with ϕ−1 (Q0 ) = P 0 .
Corollary 8.21. If (R, m) is a local ring, then dim R̂m = dim R.
Proof. R̂m is flat over R and R̂/mR̂ ' R/m is a field, so dim R̂/mR̂ = 0.
Theorem 8.22. 1. If k is a field, then dim k[x1 , . . . , xn ] = n.
2. In general, dim R[x] = dim R + 1
3. If P is a prime of R, then ∃ a prime Q of R[x] with Q ∩ R = P , and
for a maximal such ideal, dim R[x]Q = 1 + dim RP , so codimR[x] Q =
1 + codimR P .
Proof. 1. Follows from 2 by induction on n.

39
2. If P1 ⊂ . . . ⊂ Pd of R, we get P1 R[x] ⊂ P2 R[x] ⊂ . . . ⊂ Pd R[x] ⊂
Pd R[x] + (x). (using P R[x] ∩ R = P and if P is prime then P R[x] is
prime). So dim R[x] ≥ dim R + 1.
Conversely, if Q is a maximal ideal of R[x], then Q is maximal among
primes meeting Q ∩ R, so by part 3, we get codim Q = 1 + codim Q ∩ R,
so dim R[x] ≤ 1 + dim R.
3. We first prove the case where R is a field and P = 0. Then Q ∩ R = 0 for
any proper prime of R[x], so we take Q to be any maximal ideal of R[x].
Then codim Q = 1 = 1 + dim R.
In general, P R[x] is a prime in R[x] with P R[x] ∩ R = P . Localize at P
to assume that R is local and P is maximal. Let Q be a maximal ideal of
R[x] containing P . Then Q ∩ R = P . So all the remains to be shown is
that codim Q = 1 + codim P .
If P0 ⊂ . . . ⊂ Pd = P is a chain of primes in P , then P0 R[x] ⊂ . . . Pd R[x]
is a chain in R[x]. Look at Q/P R[x] ⊆ R[x]/P R[x] ' (R/P )[x], then
codim Q/P R[x] = 1. So codim Q ≥ d + 1.
Then codim Q ≤ dim RP +dim R[x]Q /P R[x]Q = dim Rp +1 = codim P +1.

Therefore, dim k[x1 , . . . , xn ] = n.

Definition 8.4 (Regular Local Ring). A ring (R, m) of dimension d is a regular


local ring if m can be generated by d elements.
If R is a regular local ring, then Nakayama says that dimR/m m/m2 = d.
Thus, every generating set for m has size d.
A generating set of size d is called a regular sequence of parameters.

Definition 8.5 (Regular Sequence). In general, a regular sequence is a se-


quence x1 , . . . , xd such that (x1 , . . . , xd ) is proper and xi+1 is a nonzero-divisor
on R/(x1 , . . . , xi ) for all i.
Definition 8.6 (Cohen-Macaulay). A ring R is Cohen-Macaulay if there is a
regular sequence of length dim R.

Goal: deg PR + 1 = dim R.


Let R = ⊕i≥0 Ri be a Nötherian graded ring with R0 a field and R is gen-
erated as an R0 -algebra by R1 . This is sometimes called a standard-graded
algebra.
This means R = k[x1 , . . . , xn ]/I for some n and homogeneous I.
Why? Because R is Nötherian, R1 is a finite dimensional R0 -vector space.
Then k[x1 , . . . , xn ] surjects onto R, so take I to be the kernel, which is homo-
geneous as this is a graded homomorphism of rings.
Definition 8.7 (Hilbert Function). The Hilbert function of R is HR (t) =
dimk Rt .

40
In homework, we showed that there exists a polynomial PR (t) which is equal
to the Hilbert Function for sufficiently large t.
If M is a graded R-module, then HM (t) = dimk Mt . Again, it eventually
will agree with a polynomial.
Lemma 8.23. If 0 → M 0 → M → M 00 → 0 is a graded (ie, degree 0) short
exact sequence of modules, then HM (t) = HM 0 (t) + HM 00 (t).
Proof. The ses is the direct sum of ses 0 → Md0 → Md → Md00 → 0 of k-vector
spaces, so additivity follows.
Lemma 8.24. If x is a nonzero divisor on R, homogeneous of degree 1, then
HR (t) = HR (t − 1) + HR/x (t) so PR/x (t) = PR (t) − PR (t − 1) for all t, so
deg PR = deg PR/x + 1.
x
Proof. Consider the ses 0 → R[−1] → R → R/x → 0 where R[−1] is R with
graded R[−1]n = Rn−1 . In general, R[a]b = Ra+b .
So HR (t) = HR[−1] (t) + HR/x (t) = HR (t − 1) + HR/x (t).
Fact: If R is graded etcetera, then dim R = max{codim Q|Q is a homoge-
neous prime}.
Proposition 8.25. If deg(x) > 0 and x is a nonzerodivisor on R then dim R/x =
dim R − 1.
Proof. Since x is a nzd, x is not contained in any associated prime, so in par-
ticular, x does not lie in any minimal prime.
If P0 ⊂ . . . ⊂ Pd is a chain of primes in R/x, that is, a chain of primes in R
containing x, then ∃ prime P ( P0 so dim R > dim R/x.
It remains to show that dim R/x ≥ dim R − 1. Since R is graded and
dim R < ∞, there is a homogeneous prime Q with codim Q = dim R. Suppose
that dim R/x = e < dim R − 1 with d = dim R.
We know that x ∈ Q since (Q, x) is proper (since Q, x are homogeneous) so
contained in a maximal ideal. But Q is maximal. This menas that RQ /x has
dim ≤ e. So there are x1 , . . . , xe ∈ Q qwith QnQ ⊆ (x1 , . . . , xe ) + (x) for n >> 0.
So QnQ ⊆ (x1 , . . . , xe , x), and so codim Q = dim RQ ≤ e + 1, so d ≤ e + 1.
Theorem 8.26. dim R = deg PR + 1
Proof. The proof is by induction on dimension. If dim R = 0, then R is Artinian.
So finite length ` as an R-module, so every filtration has length ≤ `. If
R`+1 6= 0 R ⊂ R>0 ⊂ R>1 ⊂ . . . ⊂ R>` is a filtration of length ` + 1. If
R>j = R>j+1 then Rj+1 = 0, which implies that Rn = 0 for n ≥ j + 1.
So this means that Rn = 0 for n >> 0, so PR (t) = 0 which has degree −1
by convention.
We now assume that dim R > 0 and that the theorem is true for smaller
dimension. We first reduce to the case that m = R>0 is not an associated
prime. Since the zero divisors on R are the union of the associated primes, this
will let us find a homogeneous nonzerodivisor (of degree 1)

41
Let J = (0 :R m∞ ) = {f ∈ R|∃k with f mk = 0} = f ∈ R|∃k with f g = 0
for all g ∈ mk }.
Let R0 = R/J. First note that m is not associated to R0 , since if (0 : x) = m,
then x ∈ J. Also note that PR = PR0 because Jt = 0 for t >> 0, because
J = (f1 , . . . , fs ) for some s, where we can take fi homogeneous.
Then there is a k such that fi mk = 0 for all i, then Jt = 0 for t >>
k + max dim fi . So Rt = Rt0 for t >> 0 so PR = PR0 .
Also, dim R = dim R0 , so since if P0 ⊂ . . . ⊂ Pd is a chain of length d = dim R
in R, then since dim R > 0, P0 6= m, so we can find x ∈ m \ P0 homogeneous,
and then if f ∈ J, f xk = 0 for k >> 0 so f xk ∈ P0 , so f ∈ P0 . Thus J ⊆ P0 ,
so dim R0 = dim R.
Thus, we assume that m is not an associated prime of R. So we can find
x ∈ R1 a nonzerodivisor on R. Then dim R/x = deg PR/x +1, by induction. And
so, dim R = dim R/x+ 1 and deg PR = deg PR/x + 1, so dim R = deg PR + 1.
Fact: R graded, etc, then dim R = max codim of homogeneous Q. Why?
Follows from dim k[x1 , . . . , xn ]/I = tr degk R and all maximal ideals have the
same codimension. This itself follows from Nöther Normalization.
Theorem 8.27 (Nöther Normalization). If R = k[x1 , . . . , xn ]/J then there exist
y1 , . . . , yd ∈ R such that k[y1 , . . . , yd ] ⊆ R and R is finite over S. (can take yi
to be homogeneous).

42

You might also like