Commutative Algebra Lecture Notes
Commutative Algebra Lecture Notes
Sommersemester 2017
Mohamed Barakat
Der Nachdruck dieses Textes, auch von einzelnen Teilen daraus, ist nicht gestattet.
Vorwort
Dies ist die geTEXte Version meiner Vorlesungsnotizen, die ich fortlaufend aktualisieren
werde. Der Stand der Notizen ist mindestens um eine Woche verzögert. Daher gilt: Kommt
zur Vorlesung und macht Eure eigenen Notizen. Die sind sowieso besser als jedes Skript. Die
Form eines Skriptes erreichen diese Notizen vermutlich erst gegen Ende der Vorlesung,
dies kann ich aber nicht garantieren. Die aktuelle Version ist unter der folgenden Adresse
zu finden:
http://www.mathematik.uni-kl.de/~barakat/Lehre/SS17/CA/Skript/CA.pdf
http://www.mathematik.uni-kl.de/~barakat/Lehre/WS16/Algebra/Skript/
Algebra.pdf
Als Vorlage für das Kapitel über Gröbnerbasen benutz(t)e ich das online-verfügbare Buch
von Prof. Wolfram Decker und Prof. Frank-Olaf Schreyer
http://www.mathematik.uni-kl.de/~decker/Lehre/WS12/
CommutativeAlgebra/skript/BookDeckerSchreyer.pdf.
Als Vorlage für das Kapitel über Lokalisierung benutz(t)e ich das online-verfügbare
Skript von Prof. Wolfram Decker
http://www.mathematik.uni-kl.de/fileadmin/AGs/agag/decker/WS1516/
CommAlg/CommutativeAlgebra.pdf.
iii
iv
Contents
1 Gröbner bases 1
1 Monomials and monomial ideals . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Monomial orderings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3 Division with remainder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
4 Gröbner bases and Buchberger’s algorithm . . . . . . . . . . . . . . . . . . . . 7
5 First applications: Submodule membership and syzygies . . . . . . . . . . . . 16
6 Iterated syzygies and Hilbert’s Syzygy Theorem . . . . . . . . . . . . . . . . . 17
7 Elimination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4 Primary decomposition 75
1 Definition and Existence in N OETHERian rings . . . . . . . . . . . . . . . . . . 75
2 Uniqueness results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
v
vi CONTENTS
Gröbner bases
This chapter is an excerpt from unpublished lecture notes of Wolfram Decker and Frank-
Olaf Schreyer.
Proof. Exercise.
Hint: Use an induction on the number of variables.
In fact, by removing those generators which are divisible by others, one obtains a uniquely
determined set of monomial generators for I. Its elements are called minimal generators
of I.
The ideal membership problem provides our first example of how monomial ideals are
easier to handle than arbitrary ideals: If I R is a monomial ideal, given by monomial
generators m1 , . . . , mr , a term is contained in I iff it is divisible by at least one of the mi ; an
arbitrary polynomial g ∈ R is contained in I iff all its terms are contained in I.
Most of the terminology used when working with polynomials extends to elements of
free modules overL R. In what follows let F be a free R-module with a fixed basis {e1 , . . . , es },
in particular F = si=1 Rei .
Definition 1.4. A monomial in F involving the basis element ei is a monomial in the
direct summand Rei ∼= R. A term in F is a monomial in F multiplied by a coefficient c ∈ k.
Every nonzero element f ∈ F can be uniquely expressed as the sum of finitely many
nonzero terms involving distinct monomials. These terms resp. monomials are called the
terms resp. monomials of f .
1 β
x | xα iff α − β ∈ Nn .
1
2 CHAPTER 1. GRÖBNER BASES
Example 1.5. Let F = k[x, y]3 . If e1 = 1 0 0 , e2 = 0 1 0 , e3 = 0 0 1 are the
canonical basis vectors, then
f := x2 y + x2 1 x = x2 y · e1 + x2 · e1 + 1 · e2 + x · e3 ∈ F.
2 Monomial orderings
Definition 1.6. A monomial order on R is a total order2 > on the set of monomials in R
such that
xα > xβ =⇒ xγ xα > xγ xβ for all α, β, γ ∈ Nn .
Definition 1.7 (Lexicographic order). Set
Remark 1.8. Note that we have defined >lex such that the variables are ordered accord-
ing to their appearance when writing R. For instance, in k[x, y, z],
(1) > is Artinian, i.e., each nonempty set of monomials has a smallest element with respect to >.
Proof. The only nontrivial part of the proof is to show that condition (3) implies condition
(1). If condition (3) holds, and X is a nonempty set of monomials, the monomial ideal
I = hXi R generated by X is, in fact, generated by a finite subset Y of X due to Gordan’s
lemma. Hence, every monomial in X is divisible by a monomial in Y , and the smallest
element of Y is the smallest element of X.
We use the word global to distinguish the monomial orders considered in this chap-
ter from so-called local orders, used to compute in local rings. The lexicographic order is
global.
Buchberger’s algorithm requires the choice of a global monomial order. The perfor-
mance of the algorithm and the resulting Gröbner basis depend in a crucial way on this
choice. For most applications, in principle, any Gröbner basis and, thus, any order will do.
With regard to efficiency, however, the global monomial order which we will now define
appears to be best possible (see the discussion following Remark 1.15 below and [BS87] for
some evidence):
2
We also write xβ < xα for xα > xβ .
2. MONOMIAL ORDERINGS 3
This order is extended to free R-modules as in Remark 1.17 (we suggest to give priority to
the monomials in R).
Note that as in the case of >lex , we have defined >drlex such that the variables are ordered
according to their appearance when writing R. In contrast to >lex , however, >drlex refines
the partial order by total degree:
For monomials of the same degree, the difference between >lex and >drlex is subtle but
crucial. The use made of these orders relies on their key properties (which, as we will see
in Exercise 1.1, characterize >lex and >drlex among all global monomial orders). End
Definition 1.12. Let > be a monomial order on R, and let f ∈ R be a nonzero polyno- lect. 1
mial. The leading term (or leading term) of f with respect to >, written 25.04
L> (f ) = L(f ),
is the largest term of f with repect to >. By convention, L> (0) = L(0) = 0. If L(f ) =
cxα , with c ∈ k, then c is called the leading coefficient of f and xα is called the leading
monomial of f .
Remark 1.13. Since monomial orders are compatible with multiplication,
The inequality is strict iff L(f ) and L(g) cancel each other in f + g.
This last equivalence shows that if L(h) is divisible by L(f ), and if we think of comput-
L(h)
ing h − L(f )
f as a single step of a division process, then the new dividend in such a step
will be zero, or its leading term will be smaller than that of the preceeding dividend h. This
does not imply, however, that the process terminates:
Example 1.14. In k[x], choose the terms of lowest degree as the leading terms. This
order is not global! Divide g = x by f = x − x2 using division steps as described above.
Then, the successive intermediate dividends are g = x, x2 , x3 , . . . .
Remark 1.15.
4 CHAPTER 1. GRÖBNER BASES
Theorem 1.18 (Division with Remainder, Determinate Version). Let > be a global mono-
mial order on R, and let f1 , . . . , fr ∈ R \ {0}. For every g ∈ R, there exists a uniquely determined
expression
g = g1 f1 + . . . + gr fr + h, with g1 , . . . , gr , h ∈ R,
such that
(DD1) no nonzero term of gi L(fi ) is divisible by L(fj ) for j < i;
is the resulting representation, then either g (1) := g − ri=1 gi fi − h is zero, and we are
P
done, or L(g) > L(g (1) ). By recursion, since > is Artinian, we may assume in the latter
(1)
case that g (1) has a representation g (1) = ri=1 gi fi + h(1) satisfying (DD1) and (DD2). Then
P
(1)
g = ri=1 (gi + gi )fi + (h + h(1) ) is a representation for g satisfying (DD1) and (DD2).
P
Conditions (DD1) and (DD2) are best understood by considering a partition of the
monomials in R as in the following example:
Example 1.19. Let f1 = x2 , f2 = xy + x ∈ k[x, y] with >lex . Then L(f1 ) = f1 = x2 and
L(f2 ) = xy. In the picture below, the monomials divisible by L(f1 ) correspond to the dots
in the region which is shaded in light grey:
− • • • •
− • • • •
− • • • •
− • • • •
(1, 1)
| •
(2, 0)
• •
g = (x + y 3 ) · L(f1 ) + y · L(f2 ) + 0,
6 CHAPTER 1. GRÖBNER BASES
g = (x + y 3 ) · f1 + (y − 1) · f2 + x.
End
lect. 2 It should be particularly clear from the picture in the example above that condition
02.05 (DD1) makes the order in which f1 , . . . , fr are listed play a crucial role in the determinate
division algorithm. We illustrate this by another example:
Example 1.20. Let f1 = x2 y − y 3 , f2 = x3 ∈ k[x, y] with >lex . Then L(f1 ) = x2 y. For
g = x3 y, the determinate division algorithm proceeds as follows:
x3 y = x · L(f1 ) + 0 · L(f2 ) + 0,
x3 y = x · (x2 y − y 3 ) + 0 · (x3 ) + xy 3 .
x3 y = y · (x3 ) + 0 · (x2 y − y 3 ) + 0.
(3) return(h).
With some extra bookkeeping as in Euclid’s division algorithm, the algorithm also returns
polynomials g1 , . . . , gr such that g = g1 f1 +. . .+gr fr +h. This representation of g satisfies the
conditions (ID1) and (ID2) below which are weaker than the conditions (DD1) and (DD2),
respectively:
Each such representation is called a standard expression for g with remainder h (in terms
of the fi , with respect to >).
In practical terms, it is often useful to give up uniqueness and allow choices to be made
since some choices are more efficient than others. In fact, there are various possible selection
strategies for the division process. It is not clear to us whether there is a “generally best”
strategy. Typically, the selection of the strategies depends on the particular application one
has in mind.
A version of the division algorithm which is even more indeterminate is discussed in
the exercise below.
Exercise 1.2. Show that we still get a division process which terminates if, at each stage,
we remove some term of the current dividend with the help of some L(fi ) by which it is
divisible, and if we stop as soon as this is no longer possible. Show that the resulting
representation g = g1 f1 + . . . + gr fr + h satisfies the conditions (ID1) and (DD2).
Remark 1.22 (Leading Terms in Standard Expressions). If g is a nonzero polynomial in
R and g = g1 f1 + . . . + gr fr + h is a standard expression, then L(g) is the maximum nonzero
term among the L(gi fi ) = L(gi )L(fi ) and L(h). Indeed, this follows from condition (ID1) in
conjunction with Remark 1.13. In particular, if the remainder h is zero, then L(g) is divisible
by one of L(f1 ), . . . , L(fr ). We, then, write
x3 y = x · f1 + 0 · f2 + xy 3
and
x3 y = y · f2 + 0 · f1 + 0,
which, in particular, have two different remainders. The problem with the first standard
expression is that x3 y and, thus, xy 3 are contained in the ideal hf1 , f2 i = hx2 y − y 3 , x3 i, but
xy 3 cannot be removed in the division process since it is not divisible by any of the leading
terms x2 y and x3 of the divisors. To decide ideal membership, we need to be able to cancel
any leading term of any element of I, using the leading terms of the divisors.
Based on this consideration, we make the following definition:
Definition 1.23. Let F be a free R-module with a fixed finite basis, > a global monomial
order on F , and A ⊂ F a subset.
(1) The leading submodule (or initial submodule) of A is the monomial submodule
That is, L(A) is generated by the leading terms of the elements of A. In the special
case where F = R, we refer to L(A) as the leading ideal (or initial ideal) of A.
For simplicity, we will say that a finite subset G of F is a Gröbner basis if it is a Gröbner
basis for the submodule it generates.
The terminology in the definition above is somewhat inaccurate in that we should have
written leading module with respect to > and Gröbner basis with respect to >. Indeed,
leading modules depend on the choice of the monomial order. Furthermore, if G is a Gröb-
ner basis with respect to >, and >0 is another monomial order, then G may fail to be a
Gröbner basis with respect to >0 .
Notation. For the rest of this section, > will be a fixed global monomial order on a free
R-module F with a fixed finite basis.
In contrast to the polynomials f1 , f2 in Example 1.20, the elements of a Gröbner basis
behave well under division with remainder and can, thus, be used to decide ideal and
submodule membership. In particular, we will now see that a Gröbner basis of I generates
I, although this was not part of the definition.
Proposition 1.24. Let P{f
r
1 , . . . , fr } ⊂ F \ {0} be a Gröbner basis for the submodule I :=
hf1 , . . . , fr i ≤ F . If g = i=1 gi fi + h is a standard expression for an element g ∈ F , then g ∈ I
iff the remainder h is zero. In particular, f1 , . . . , fr generate I.
Proof. If h is zero, then clearly g ∈ I. Conversely, if g ∈ I, then h ∈ I, which implies that
L(h) ∈ L(I) = hL(f1 ), . . . , L(fr )i. So L(h) and, thus, h are zero by condition (ID2) on the
remainder of a standard expression.
Corollary 1.25 (Hilbert’s Basis Theorem). Every submodule I ≤ F has a Gröbner basis and
is thus finitely generated. In particular, R is Noetherian.
Gordan’s Proof. As remarked earlier, Gordan’s lemma yields that each monomial submod-
ule of F is generated by finitely many monomials. In particular, there are finitely many
elements f1 , . . . , fr ∈ I such that L(I) = hL(f1 ), . . . , L(fr )i. That is, f1 , . . . , fr form a Gröb-
ner basis for I. The statement now follows from Proposition 1.24.
Remark 1.26. Gordan’s proof has, as every other proof of Hilbert’s basis theorem, two
ingredients, namely induction on the number of variables (here used to verify Gordan’s
Lemma 1.3) and division with remainder. The advantage of Gordan’s proof is that it sepa-
rates these ingredients.
End
lect. 3 Macaulay [1927] used the idea of obtaining information on an ideal from information on
04.05 its leading ideal to classify Hilbert functions. On his way, he proved the following crucial
result:
Theorem and Definition 1.1 (Macaulay). If I ≤ F is a submodule, the monomials not con-
tained in L> (I) represent a k-vector space basis for F/I. We refer to these monomials as standard
monomials (for I, with respect to >).
Proof. Let
B := {m + I | m ∈ F a standard monomial} ⊂ F/I.
To show that the elements of B are k-linearly independent, consider a k-linear combination
g of standard monomials such that the residue class g + I is zero. Then g ∈ I, so that
L(g) ∈ L(I). Since L(g) is a scalar times a standard monomial, this implies 0 = L(g) = g
by the very definition of the standard monomials.
To show that the elements of B generate F/I as a k-vector space, consider any element
g ∈ F . Choose elements f1 , . . . , fr ∈ F \ {0} which form a Gröbner basis for I, and let g =
P r
i=1 gi fi + h be a standard expression satisfying condition (DD2) of determinate division
with remainder. Then no term of h is in hL(f1 ), . . . , L(fr )i = L(I). Hence, the residue class
g + I = h + I is a k-linear combination of the elements of B, as desired.
4. GRÖBNER BASES AND BUCHBERGER’S ALGORITHM 9
I : m := {r ∈ R | mr ∈ I}
lcm(mi , m) mi
= , 1 ≤ i ≤ r.
m gcd(mi , m)
where
lcm(L(fj ), L(fi ))
mji = ∈ R.
L(fi )
and note that the terms mji for j < i form the set of generators for Mi mentioned in Exercise
1.4. As it turns out, there is no need to consider all the mji with j < i (that is, all the
corresponding S-vectors S(fi , fj )) in our test: For each i, and for each minimal monomial
generator xα of Mi , choose exactly one index j = j(i, α) < i such that mji involves xα , and
compute a standard expression for S(fi , fj ) in terms of the f` ’s with remainder hi,α (we
suppress the index j in our notation).
Theorem 1.30 (Buchberger’s Criterion). Let f1 , . . . , fr ∈ F be nonzero polynomial vectors.
With notation as above, f1 , . . . , fr form a Gröbner basis iff all remainders hi,α are zero.
Before proving the criterion, we illustrate it by an example, and formulate Buchberger’s
algorithm for computing Gröbner bases. For the example, recall that a k × k minor of a
matrix is the determinant of a k × k submatrix.
Example 1.31. Consider the ideal generated by the maximal minors, i.e., the ten3 3 × 3
minors of the matrix
x1 x2 x3 x4 x5
y1 y2 y3 y4 y5
z1 z2 z3 z4 z5
and the lexicographic order on k[x1 , . . . , z5 ]. The leading terms of the minors and the mini-
3 5 5
3 = 2 = 10.
4. GRÖBNER BASES AND BUCHBERGER’S ALGORITHM 11
x1 y2 z3
x1 y2 z4 M2 = hz3 i
x1 y3 z4 M3 = hy2 i
x2 y3 z4 M4 = hx1 i
x1 y2 z5 M5 = hz3 , z4 i
x1 y3 z5 M6 = hy2 , z4 i
x2 y3 z5 M7 = hx1 , z4 i
x1 y4 z5 M8 = hy2 , y3 i
x2 y4 z5 M9 = hx1 , y3 i
x3 y4 z5 M10 = hx1 , x2 i
So only 15 out of 10
2
= 45 S-vectors are needed in Buchberger’s criterion. In fact, the
criterion shows that the minors form a Gröbner basis.
In the situation of the criterion, we refer to the selection of the indices j = j(i, α) to-
gether with the computation of the remainders hi,α as Buchberger’s test. It is clear from the
criterion that the amount of computation needed for the test depends in a crucial way on
the order in which we list f1 , . . . , fr .
Algorithm 1.32 (Buchberger’s algorithm). Given I = hf1 , . . . , fr i ≤ F , compute a Gröb-
ner basis for I.
(1) Set ` = r.
(3) If some hi,α is nonzero, set ` = ` + 1, f` = hi,α , and go back to Step (2).
(4) Return f1 , . . . , f` .
End
Correctness and termination follow from Buchberger’s criterion and the discussion pre- lect. 4
ceeding it. 09.05
Corollary 1.33. Given polynomial vectors f1 , . . . , fr ∈ F \ {0}, a Gröbner basis for I :=
hf1 , . . . , fr i ≤ F can be computed in finitely many steps.
Example 1.34. Let f1 = x2 , f2 = xy − y 2 ∈ k[x, y] with >lex . Then L(f2 ) = xy and
M2 = hxi. We compute the standard expression
S(f2 , f1 ) = x · f2 − y · f1 = −xy 2 = 0 · f1 − y · f2 − y 3 ,
and add the nonzero remainder f3 := −y 3 to the set of generators. Then M3 = hx2 , xi = hxi,
and it suffices to consider the S-polynomial S(f3 , f2 ). Computing the standard expression
S(f3 , f2 ) = x · f3 + y 2 · f2 = −y 4 = 0 · f1 + 0 · f2 + y · f3
with remainder zero, we find that f1 , f2 , f3 form a Gröbner basis for the ideal I = hf1 , f2 i.
We visualize, once more, the monomials in k[x, y]:
12 CHAPTER 1. GRÖBNER BASES
• • • • •
(0, 3) • • • • •
− • • • •
− •
(1, 1)
• • •
| •
(2, 0)
• •
The dots in the shaded region correspond to the monomials in the ideal L(I) which is
minimally generated by x2 , xy, and y 3 . The monomials 1, x, y, y 2 respresented outside the
shaded region are the standard monomials. Due to Macaulay’s Theorem 1.1, their residue
classes form a k-vector space basis for k[x, y]/I. Hence, every class g + I ∈ k[x, y]/I is
canonically represented by a uniquely determined k-linear combination a + bx + cy + dy 2
(see Remark 1.27). To add and multiply residue classes, we add and multiply the canonical
representatives according to the rules in Exercise 1.3. The multiplication in k[x, y]/I is, thus,
determined by the following table (we write f = f + I):
· 1 x y y2
1 1 x y y2
x x 0 y2 0
y y y2 y2 0
y2 y2 0 0 0
(2) I has a binomial Gröbner basis, that is, a Gröbner basis consisting of binomials.
Now, we come to the proof of Buchberger’s criterion. For this, recall that the S-vectors
are designed to cancel leading terms: The L(f` )’s fit into the relations
Buchberger’s criterion then reads that f1 , . . . , fr form a Gröbner basis iff every relation
of type (4.1) considered in Buchberger’s test “lifts" to a relation of type (4.2) such that
(ij)
L(S(fi , fj )) ≥ L(g` f` ) whenever both sides are nonzero. In general, we think of a relation
g1 f1 + · · · + gr fr = 0 ∈ F
ϕ : Rr → M, εi 7→ fi ,
where {ε1 , . . . , εr } is the canonical basis of Rr . We call ker ϕ the (first) syzygy module of
f1 , . . . , fr , written
Syz(f1 , . . . , fr ) = ker ϕ.
Syz(f1 , . . . , fr ) is as a module over a noetherian ring R finitely generated and we regard the
elements of a given finite set of generators for it as the rows of a matrix which we call a
syzygy matrix of f1 , . . . , fr .
Exercise 1.7. Determine a syzygy matrix of x, y, z ∈ k[x, y, z].
To handle the syzygies on the elements f1 , . . . , fr of a Gröbner basis, we consider the
free module F1 = Rr with its canonical basis {ε1 , . . . , εr } and the induced monomial order
>1 on F1 defined by setting
Note that >1 is global if this is true for > (what we suppose, here).
Proof of Buchberger’s criterion (Theorem 1.30). Write I = hf1 , . . . , fr i ≤ F . If f1 , . . . , fr form
a Gröbner basis for I, all remainders hi,α are zero by Proposition 1.24. Indeed, the S-vectors
are contained in I.
Conversely, suppose that all the hi,α are zero. Then, for every pair (i, α), we have a
standard expression of type
(ij)
S(fi , fj ) = g1 f1 + . . . + gr(ij) fr + 0,
be the corresponding syzygy on f1 , . . . , fr (we suppress the index j in our notation on the
left hand side), and S matrix of syzygies4 with rows S (i,α) in some order. Hence
f1
..
S . = 0.
fr
On F1 we consider the induced monomial order. The leading term of S (i,α) with respect to
this order is
L(S (i,α) ) = mji εi .
4
We will prove in Corollary 1.36 that S is even a syzygy matrix.
14 CHAPTER 1. GRÖBNER BASES
Indeed,
mji L(fi ) = mij L(fj ), but i > j,
and
(ij)
mji L(fi ) > L(S(fi , fj )) ≥ L(g` )L(f` )
whenever these leading terms are nonzero.
To prove that the f` ’s form a Gröbner basis for I, let g be any nonzero element of I,
say g = a1 f1 + . . . + ar fr , where a1 , . . . , ar ∈ R. The key point of the proof
Pr is to replace
this representation of g in terms of the f` ’s by a standard expression g = k=1 g` f` (with
remainder zero). The result, then, follows by applying Remark 1.22 on leading terms in
standard expressions:
L(g) = max{L(g1 )L(f1 ), . . . , L(gr )L(fr )} ∈ hL(f1 ), . . . , L(fr )i.
To find the desired standard expression, we go back and forth between F1 = Rr and F :
Consider the polynomial row vector A := (a1 , . . . , ar ) ∈ Rr , and let G = (g1 , . . . , gr ) ∈ F1
be the remainder of A under determinate division by the syzygies S (i,α) ’s (listed in some
order). Then
g = a1 f1 + . . . + ar fr = g1 f1 + . . . + gr fr (4.3)
since the S (i,α) are syzygies5 on f1 , . . . , fr . We show that the right hand side of (4.3) satisfies
condition (DD1) of determinate division by the f` ’s (in particular, it is a standard expres-
sion). Suppose the contrary. Then there is a pair k < i such that one of the terms of gi L(fi )
is divisible by L(f` ). In turn, one of the terms of gi is contained in the monomial ideal
Mi = hL(f1 ), . . . , L(fi−1 )i : L(fi ) R
which is generated by the mji ’s selected in Buchberger’s test. This means that one of the
terms of G is divisible by some mji εi = L(S (i,α) ), contradicting the fact that according to
how we found G, the terms of G satisfy condition (DD2) of determinate division by the
End S (i,α) in F1 .
lect. 5 Corollary 1.36. If f1 , . . . , fr ∈ F \{0} form a Gröbner basis with respect to >, the S (i,α) consid-
11.05 ered in the proof of Buchberger’s criterion form a Gröbner basis for the syzygy module Syz(f1 , . . . , fr )
with respect to the induced monomial order. In particular, the S (i,α) ’s generate the syzygies on
f1 , . . . , f r .
Proof. Let A ∈ Rr be an arbitrary syzygy on f1 , . . . , fr , and let G = (g1 , . . . , gr ) ∈ Rr be the
remainder of A under determinate division by the S (i,α) ’s (listed in some order). Then, since
A and the S (i,α) ’s are syzygies on f1 , . . . , fr , the same must be true for G:
0 = g1 f1 + . . . + gr fr .
Furthermore, as shown in the proof of Buchberger’s criterion, the gi ’s satisfy condition
(DD1) of determinate division by f1 , . . . , fr . Since standard expressions under determinate
division are uniquely determined, the gi and, thus, G must be zero. Taking, once more,
Remark 1.22 on leading terms in standard expressions into account, we find that L(A) is
divisible by some L(S (i,α) ). The result follows.
Remark 1.37. The S in S-vector stands for syzygy. In fact, as is now clear, the relations
mji L(fi ) − mij L(fj ) = 0 corresponding to the S(fi , fj ) generate Syz(L(f1 ), . . . , L(fr )). In
our version of Buchberger’s test, we focus on the minimal monomial generators of Mi .
This means that we select a subset X ⊂ {S(fi , fj ) | j < i} such that the relations mji L(fi ) −
mij L(fj ) = 0 corresponding to the S(fi , fj ) in X still generate Syz(L(f1 ), . . . , L(fr )). It is this
property of X which makes our proof of Buchberger’s criterion work. Hence, in stating the
criterion, we may choose any set of S-vectors satisfying this property.
5
g = Af = (BS + G)f = Gf since Sf = 0.
4. GRÖBNER BASES AND BUCHBERGER’S ALGORITHM 15
where m1 , . . . , mr are the minimal generators of L(I). Explain how to compute the reduced
Gröbner basis from any given Gröbner basis.
Exercise 1.9. Design an algorithm to decide whether an ideal hf1 , . . . , fr i ⊂ k[x1 , . . . , xn ]
is principal or not.6
Remark 1.40. In the situation of Corollary 1.36, the syzygies S (i,α) form a minimal Gröb-
ner basis.
Proof. Indeed, with respect to the induced order, L(S (i,α) ) = c(i,α) xα εi , where xα is a minimal
generator of the monomial ideal Mi . If L(S (i,α) ) is divisible by L(S (j,β) ), then i = j and xα
is disible by xβ . Hence, xα = xβ and, thus, L(S (i,α) ) = L(S (j,β) ).
Remark 1.41. Buchberger’s algorithm generalizes both Gaussian elimination and Eu-
clid’s GCD algorithm:
(1) Given homogeneous degree-1 polynomials
let > be a global monomial order on k[x1 , . . . , xn ] such that x1 > · · · > xn . Computing
a minimal Gröbner basis for hf1 , . . . , fr i amounts, then, to transforming the coefficient
matrix A = (aij ) into a matrix in row echelon form (REF) with pivots 1. And com-
puting a reduced Gröbner basis for hf1 , . . . , fr i amounts to computing a row-reduced
echelon form (RREF) with pivots 1.
6
The general problem of computing the minimal generating number of a polynomial ideal is still an open.
16 CHAPTER 1. GRÖBNER BASES
(2) In the case of one variable x, there is precisely one global monomial order: 1 < x <
x2 < · · · . Given f1 , f2 ∈ k[x], the reduced Gröbner basis for hf1 , f2 i with respect to this
order consists of exactly one element, namely the greatest common divisor gcd(f1 , f2 ),
and Buchberger’s algorithm takes precisely the same steps as Euclid’s algorithm for
computing the gcd.
g ∈ I := hf1 , . . . , fr i ⊂ F.
g = g1 f1 + . . . + gr fr .]
(1) Compute a Gröbner basis f1 , . . . , fr , fr+1 , . . . , fr0 for I using Buchberger’s algorithm.
[Store each syzygy arising from a division which leads to a new generator f` in Buch-
berger’s test.]
(2) Compute a standard expression for g in terms of f1 , . . . , fr0 with remainder h (use the
same global monomial order on F as in Step 1).
(3) If h = 0, then g ∈ I. [In this case, for ` = r0 , . . . , r + 1, successively do the following: in
the standard expression computed in Step 2, replace f` by the expression in terms of
f1 , . . . , f`−1 given by the syzygy leading to f` in Step 1.]
Example 1.43. Let
g = x3 − x2 + xy 2 , f1 = x2 , f2 = xy − y 2 ∈ k[x, y].
x · f2 − y · f1 = −xy 2 = 0 · f1 − y · f2 − y 3 .
g = (x − 1) · f1 + y · f2 − f3
I = J ⇐⇒ I = I + J and J = I + J.
6. ITERATED SYZYGIES AND HILBERT’S SYZYGY THEOREM 17
We already know that Buchberger’s algorithm computes the syzygies on the elements
of a Gröbner basis (see Corollary 1.36). Based on this, we can compute the syzygies on any
given set of generators:
Algorithm 1.44 (Syzygy Modules). Given a free R-module F with a fixed finite basis
and polynomial vectors f1 , . . . , fr ∈ F \ {0}, compute a syzygy matrix of f1 , . . . , fr .
(1) Compute a Gröbner basis f1 , . . . , fr , fr+1 ,. . . , fr0 for hf1 , . . . , fr i ⊂ F using Buchberger’s
algorithm. On your way, store each syzygy on f1 , . . . , fr0 obtained in Buchberger’s
test. Let t be the number of these syzygies.
(2) Arrange the syzygies such that those obtained from a division leading to a new gen-
erator f` are first (and those arising from a division with remainder zero
aresecond).
A B
Then the syzygies fit as rows into a t×r0 matrix which has block form , where
C D
B is a lower triangular square matrix of size r0 − r with diagonal entries 1 (if signs are
adjusted appropriately).
Computing the syzygies on the S (i,α) and so forth, we successively get minimal Gröbner
bases which generate the successive syzygy modules of f1 , . . . , fs1 of higher order. At each
stage, the new Gröbner basis depends, in particular, on how we arrange the elements of
the Gröbner basis computed in the previous step. We show that if this arrangement is done
properly, then the process just described will terminate with a zero syzygies matrix after at
most n steps.
To begin, fix an integer 1 ≤ ` ≤ n such that none of the leading terms L(fi ) involves the
variables x`+1 , . . . , xn (choose ` = n if one of the L(fi )’s involves xn ). In Buchberger’s test,
let the fi ’s be arranged such that, for j < i, the exponent of x` in L(fj ) is smaller than or
equal to that of x` in L(fi ) whenever these leading terms involve the same basis element of
Rs0 . Then none of the resulting leading terms L(S (i,α) ) = c(i,α) xα εi involves x` , . . . , xn .
Arranging the Gröbner basis elements at each stage of our process accordingly, we ob-
tain after, say, ` ≤ k steps a Gröbner basis G` of higher syzygies for which none of the
leading terms L(g), g ∈ G` , involves x1 , . . . , xn . Having (chosen) a minimal Gröbner basis
in the previous step implies that G` = {0}.
Example 1.47. Consider the ideal
I = hf1 , . . . , f5 i ⊂ R = k[w, x, y, z]
generated by the polynomials
f1 = w2 − xz, f2 = wx − yz, f3 = x2 − wy, f4 = xy − z 2 , f5 = y 2 − wz.
As Buchberger’s test shows, these generators already form a Gröbner basis for I with re-
spect to the degree reverse lexicographic order on k[w, x, y, z], which is obviously minimal.
The left most column of the table below is the matrix ϕ1 consisting of the minimal gener-
ators f1 , . . . , f5 . All leading terms are printed in bold. We successively obtain two syzygy
matrices ϕtr 2 and ϕ3 which we display in a compact way as follows:
w2 − xz −x y 0 −z 0 −y 2 + wz
wx − yz w −x −y 0 z z2
x2 − wy −z w 0 −y 0 0
xy − z 2 0 0 w x −y yz
y2 − wz 0 0 −z −w x w2
0 y −x w −z 1
−y 2 + wz z 2 −wy yz −w2 x
The transposed syzygy matrix ϕtr 2 resulting from the test is the 5 × 6 matrix in the middle
of our table. Note that, for instance, M4 = hw, xi can be read from the 4th row of ϕtr 2 . At this
point, we already know that the rows of ϕ2 form a Gröbner basis for Syz(f1 , . . . , f5 ) with
respect to the induced monomial order on R5 . Buchberger’s test applied to these Gröbner
basis elements yields a 2 × 6 syzygy matrix ϕ3 which is printed in the two bottom rows
of our table. There are no syzygies on the two rows of ϕ3 since the leading terms involve
different basis vectors.
Observe that once we have the leading terms of the Gröbner basis for I, we can easily
compute the leading terms of the Gröbner bases for all syzygy modules, that is, all bold
face entries of our table. This gives us an early idea on the amount of computation lying
ahead.
With respect to our proof of the syzygy theorem, note that the variable z is missing from
the L(f` )’s, and the f` ’s are ordered such that the exponent of y in L(fj ) is smaller than or
equal to that in L(fi ) whenever j < i. As a consequence, none of the leading terms of the
columns of ϕtr 2 involves y. In the next step, however, the leading term of the second row of
ϕ3 involves x. What is wrong with how we arrange the columns of ϕtr 2?
7. ELIMINATION 19
Exercise 1.11. Following the matrix version of Hilbert’s Syzygy Theorem 1.46 compute
all successive syzygy matrices of the 2 × 2 minors of the matrix
x0 x1 x2 x3
.
x1 x2 x3 x4
End
lect. 7
7 Elimination 23.05
In particular, I0 = I.
Algorithm 1.49 (Elimination using >lex ). Given I = hf1 , . . . , fr i ⊂ R, compute all elimi-
nation ideals I` using the following steps:
(1) Compute a Gröbner basis G for I with respect to >lex on k[x1 , . . . , xn ].
(2) For any k, the elements g ∈ G with L(g) ∈ k[x`+1 , . . . , xn ] form a Gröbner basis for I`
with respect to >lex on k[x`+1 , . . . , xn ].
Proof of correctness. If f ∈ I ∩ k[x`+1 , . . . , xn ], then L(f ) is divisible by L(g) for some g ∈
G. Since f does not involve x1 , . . . , x` , also L(g) does not involve x1 , . . . , x` . By the key
property of >lex (Remark 1.15), we must have g ∈ k[x`+1 , . . . , xn ].
Example 1.50. Let I = hf1 , f2 i ⊂ k[x, y, z], with f1 = x2 − y, f2 = xy − z. Then I
is the vanishing ideal of the twisted cubic curve t 7→ (t, t2 , t3 ) = (x, y, z). We compute a
lexicographic Gröbner basis for I. To begin with, M2 = hx2 i : xy = hxi, and we have the
standard expression
S(f2 , f1 ) = x(xy − z) − y(x2 − y) = −xz + y 2 =: f3 .
We add f3 to the set of generators. Then M3 = hx2 , xyi : xz = hx, yi, and we have the
standard expressions
S(f3 , f1 ) = x(−xz + y 2 ) + z(x2 − y) = xy 2 − yz = y(xy − z)
and
S(f3 , f2 ) = y(−xz + y 2 ) + z(xy − z) = y 3 − z 2 =: f4 .
In the next step, M4 = hx2 , xy, xzi : y 3 = hxi, and
S(f4 , f2 ) = x(y 3 − z 2 ) − y 2 (xy − z) = −xz 2 + y 2 z = z(−xz + y 2 )
is a standard expression with remainder zero. Hence, f1 , f2 , f3 , f4 form a Gröbner basis for
I.
We visualize the monomials in L(I) via their exponent vectors:
5
5 5
y
x
0 0
20 CHAPTER 1. GRÖBNER BASES
The computaion shows that the elimination ideal I1 ⊂ k[y, z] is generated by the polynomial
f4 = y 3 − z 2 . The geometric interpretation of this is projection:
Proposition 1.55 (Algebra relations in affine rings). Let I be an ideal of R, and let f 1 =
f1 + I, . . . , f m = fm + I ∈ S := R/I. Consider the homomorphism
ϕ : k[y1 , . . . , ym ] → S = R/I, yi 7→ f i .
If J is the ideal7
J = I k[x, y] + hf1 − y1 , . . . , fm − ym i ⊂ k[x, y]
then
ker ϕ = J ∩ k[y].
Proof. Let g ∈ k[y] ⊂ k[x, y]. To prove the assertion, we have to show:
g(f1 , . . . , fm ) ∈ I ⇐⇒ g ∈ J.
P
If g = h+ gj (fj −yj ) ∈ J, with h ∈ I k[x, y], then g(f1 , . . . , fm ) = h(x1 , . . . , xn , f1 , . . . , fm ) ∈
I k[x, y] ∩ k[x] = I.
For the converse, observe that substituting the fj −(fj −yj ) for the yj in g and expanding
gives an expression of type
X
g(y) = g(f1 , . . . , fm ) + gj (fj − yj ).
Since we already know how to compute in affine rings (i.e., residue class rings of poly-
nomial rings), a particular application of the proposition is a method for computing in the
algebra k[f 1 , . . . , f m ] ∼
= k[y1 , . . . , ym ]/ ker ϕ. Once we have the required Gröbner basis for J,
we know, in particular, whether ker ϕ = 0. That is, we can decide whether f 1 , . . . , f m are
algebraically independent over k. In addition, we can check whether ϕ is surjective:
Exercise 1.13 ((*) Subalgebra Membership). With notation as above, let g, f 1 , . . . , f m be
elements R/I, and let > be a global elimination order on k[x, y] with respect to x1 , . . . , xn .
Show:
(1) We have g ∈ k[f 1 , . . . , f m ] iff the normal form h = NF(g, J) ∈ k[x, y] is contained in
k[y]. In this case, g = h(f 1 , . . . , f m ) is a polynomial expression for g in terms of the f ` .
(2) The homomorphism ϕ : k[y1 , . . . , ym ] → R/I is surjective iff NF(xi , J) ∈ k[y] for
i = 1, . . . , n.
Exercise 1.14.
(1) Compute the algebra relations on the polynomials
f1 = x2 + y 2 , f2 = x2 y 2 , f3 = x3 y − xy 3 ∈ k[x, y].
7
the vanishing ideal of the graph of ϕ
22 CHAPTER 1. GRÖBNER BASES
Chapter 2
1 Categories
The notion of a category is a special case of what is nowadays called a horizontal cat-
egorification or oidification of the notion of a monoid. Another funny name would be
“monoidoid”. This will become clear later.
Definition 2.1. A quiver (or multi-digraph) A consists of a class of objects A0 = Obj A
and a class of morphisms (or arrows) A1 = Mor A together with two defining maps
s, t : A1 ⇒ A0 ,
called identity and composition, respectively, subject to a list of defining properties. To ex-
press these properties in a suggestive way we first introduce some notational conventions:
• The notation used in µ : A1 ×A0 A1 → A1 means that µ is only defined for pairs of
morphisms (ϕ, ψ) such that t(ϕ) = s(ψ). We call such a pair composable. If not stated
otherwise, we write ϕψ instead of µ(ϕ, ψ). We call this the left convention, where
the right convention1 would be to write ψϕ instead of ϕψ. In any case we call ϕ the
pre-morphism and ψ the post-morphism.
• We use latin letters for objects and greek letters for morphisms, so M ∈ A and ϕ ∈ A
means M ∈ A0 and ϕ ∈ A1 . We also refer to them as A-objects and A-morphisms.
23
24 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
(2) s(ϕψ) = s(ϕ) and t(ϕψ) = t(ψ) for all composable morphism ϕ, ψ ∈ A.
In words: the composition respects source and target.
We express this as follows:
(4) 1M ϕ = ϕ and ψ1M = ψ for all ϕ with s(ϕ) = M and all ψ with t(ψ) = M .
In words: The identity is a left and right unit of the composition.
Remark 2.3.
• The objects are in bijection to the identities via M 7→ 1M . So for many further con-
structions we only need to refer to morphisms.
• It follows from the axioms that the set End(M ) is a monoid for all M ∈ A. In partic-
ular, a category with one object is nothing but a monoid. This is what we mean by
saying that “a category is a horizontally categorified monoid”, or a “monoidoid”.
• Isomorphic objects in a category need not be equal. In fact, this is one of the reasons
for the flexibility of the notion of a category. A category in which isomorphic objects
in A are equal is called skeletal.
Example 2.4.
(1) The category (Sets) with sets as objects and maps as morphisms.
(2) The category (Sets0 ) with pointed sets as objects and maps preserving the distin-
guished points as morphisms.
(3) The category (Grps) with groups as objects and group homomorphisms as morphisms.
(4) The category (Ab) = Z−Mod with A BELian groups as objects and group homomor-
phisms as morphisms.
(5) The category (Rngs) with (not necessarily unital) rings as objects and ring homomor-
phisms (not necessarily respecting one) as morphisms.
(6) The category (uRngs) with unital rings as objects and ring homomorphisms (not nec-
essarily respecting one) as morphisms.
(7) The category (URngs) with unital rings as objects and unital ring homomorphisms as
morphisms.
(8) The category (Top) with topological spaces as objects and continuous maps as mor-
phisms.
(9) The category (Top0 ) with pointed topological spaces as objects and continuous maps
preserving the distinguished points as morphisms.
2. THE CATEGORY OF MODULES 25
(10) The category of truth values with objects {>, ⊥} and a single non-identity morphism
⊥ → > saying that false implies true.
(11) A proset (=preordered set) is a thin small category, i.e., with at most one morphism
between two objects.
(13) The simplicial category ∆ with totally ordered sets n = {0 < 1 < · · · < n} for
n ∈ N = {0, 1, ...} as objects and order-preserving maps (i ≤ j =⇒ f (i) ≤ f (j)) as
morphisms.
Non of the above category names in round brackets is standard.
2.1 Objects
Because of the possible noncommutativity of rings we have to distinguish2 between left
and right modules.
Definition 2.5. A left R-module M is an A BELian group on which R acts by endomor-
phisms from the left. This action is given by a ring homomorphism ρ : R → (End(M ), ◦),
called a left representation of R. A right representation ρ : R → (End(M ), ·) defines a
right R-module. We occasionally denote the left (resp. right) R-module M by R M (resp.
MR ).
As usual, the left representation ρ can be interpreted as the C URRYing3 of a left action
map
αρ : R × M → M, (r, m) 7→ rm,
which is bilinear, associative, and satisfies 1m = m for all m ∈ M . In other words,
αρ (r, m) = ρ(r)(m). Likewise for right representations and right action maps αρ : M × R →
M, (m, r) 7→ mr.
Example 2.6. Checking the following examples is an easy exercise:
(1) Each A BELian group is a Z-module in a unique way.
(3) A trivial group {0} is an R-module for any ring R. We call it a zero module.
(4) The multiplication map of the ring R endows R with a left and with a right module
structure, which we denote by R R and RR , respectively.
(5) Any A BELian group with endomorphisms applied from the left is a left End(M )-
module and a right End(M )op -module.
(7) If we identify ring elements with 1×1-matrices then the matrix multiplications R1×1 ×
R1×c → R1×c and Rr×1 × R1×1 → Rr×1 define a left R-module structure on the row
space R1×c and a right R-module structure on the column space4 Rr×1 . This general-
izes 4.
2.2 Morphisms
Immediately after defining a mathematical structure and providing some examples one
should define the structure preserving maps:
Definition 2.7. A group homomorphism ϕ : M → N, m 7→ mϕ of two left R-modules
M, N is called an R-module homomorphism or an R-linear map, if it is compatible5 with
the action of R, i.e., (rm)ϕ = r(mϕ) for all r ∈ R, m ∈ M . For right modules we write
ϕ : M → N, m 7→ ϕm and require that ϕ(mr) = (ϕm)r.
The set HomR (M, N ) of all R-linear maps from M to N with pointwise addition is an
A BELian group6 , called the homomorphism group of M and N . We occasionally drop the
index R when it is clear from the context. The zero map 0M N : M → N is the neutral
element of this group.
The A BELian group EndR (M ) := HomR (M, M ) of all R-self-maps with composition is a
ring7 , called the endomorphism ring of M . The identity map 1M is the unit of this ring.
Exercise 2.1. Let M be a left R-module. Show that (EndR (M ), ·) is the subring of the
opposite ring (End(M ), ◦)op which centralizes the action of R, i.e., which commutes with
the image of the left representation of R.
Remark 2.8. Here we use the so-called associative convention for both right and left
modules. It differs from the usual “mixed” convention, where the so-called commutative
convention is used for left modules and the associative for right ones. The mixed conven-
tion is unavoidable if one insists on applying R-linear maps from the left, regardless of the
parity of the modules. There one would write for left R-linear maps ϕ : M → N, m 7→ ϕ(m)
with ϕ(rm) = rϕ(m) and say that ϕ commutes with the action of R.
The associativity convention has at least two advantages:
and the associativity rules (rm)ϕ = r(mϕ) and ϕ(mr) = (ϕm)r follow from the asso-
ciativity of matrix multiplication.
The mixed convention breaks the above symmetry: EndR (RR ) ∼ = R but EndR (R R) ∼
= Rop . It
0
also makes working out the relation between applying R-linear maps in HomR (R R1×c , R R1×c )
and matrix multiplication a nontrivial exercise.
Convention. From now on we won’t specify the parity unless necessary. If the formula-
tion of some statement does depend on the parity then we will formulate it for left modules.
4
Like S INGULAR, the subsystem P LURAL only uses column spaces as a data structure for modules but
treats them as left modules (maybe because algorithms in the noncommutative G RÖBNER basis literature are
usually developed for left ideals). This results in a mess when the ring is noncommutative.
5
We deliberately drop the brackets (m)ϕ for the associativity appearance of the law.
6
If R is commutative then HomR (M, N ) is a again an R-module.
7
If R is commutative then EndR (M ) is a again an R-algebra.
2. THE CATEGORY OF MODULES 27
The case of right modules is then analogous. As a consequence we will silently apply R-
linear maps from the right. In particular, in the composition ϕψ we first apply the left map
ϕ, then ψ.
By replacing definitions which refer to elements by ones which exclusively refer to maps
we will gradually replace the set-theoretic language by the categorical one. The categori-
cal approach, although more cumbersome at the beginning, will help us to reduce many
known constructions to few basic ones. This will be very rewarding when we start dealing
with computability as it will streamline all our algorithms. To emphasize this “element-
less” approach we will replace the words “module” by “object” and “map” by “morphism”.
An R-linear map is called injective, surjective, or bijective, if the underlying set-theoretic
map is so. We will make an attempt to describe these properties using maps (or, better, mor-
phisms) exclusively.
Definition 2.9 ([HS97, §II.3, p. 48]).
Although the previous definition is the correct categorical one it may look awkward
at the beginning. Motivated by the underlying set-theoretic maps and using the axiom of
choice one might find the following definition more “natural”.
Definition 2.10.
ιχ = 1M .
(3) A morphism α : M → N is an isomorphism (or iso) if it is a split epi and a split mono.
∼ ∼
→ N , or α : M ∼
= N , or M ∼
=
We write α : M −
→ N , or α : M − =α N . We call two objects
∼
M, N isomorphic if there exists an isomorphism α : M − → N.
(3) The pre- and post-inverse of an isomorphism coincide, and are hence unique. We call
it the inverse.
K κ
Definition 2.16. We say that a morphism κ : K → M dominates τ : L → τ /κ τ M
M if κ is a post-factor of τ , i.e., if there exists a morphism τ /κ : L → K L
such that (τ /κ)κ = τ . Any such morphism τ /κ is called a lift of τ along κ.
It is an easy exercise to prove that mutual dominance is an equivalence
relation and that two R-linear maps which are mutually dominant have the same image
submodule. The converse of the last statement is true for monos. In other words, we can
identify a submodule K ≤ M with the class of all R-monos having K as their image.
This is how we can redefine submodules categorically:
Definition 2.17. A subobject K ≤ M is an equivalence class of monos with M as target
under the equivalence relation of mutual dominance.
End
One can also define the poset of submodules categorically: lect. 9
τ κ
Definition 2.18. For two subobjects L, K ≤ M represented by L ,→ M ←- K we say 01.06
that L is smaller than K and write8 L ≤ K (or K is larger than L and write K ≥ L) if κ
dominates τ . We denote by Sub M the poset of all subobjects of M .
Exercise 2.4. Show that the last definition is independent of the representatives of sub-
objects.
We could now ask if we can redefine the entire lattice of submodules, i.e., the sum
and intersection of two submodules, categorically. The answer is yes, but we first need
a categorical definition of kernels, cokernels, and direct sums, which we will treat in the
following subsections.
Before ending this subsection we dualize the above definition and get:
C ε
Definition 2.19. We say that a morphism ε : N → C codominates η :
N → L if ε is a pre-factor of η, i.e., if there exists a morphism ε\η : C → L ε\η η N
such that ε(ε\η) = η. Any such morphism ε\η is called a colift of η along ε. L
Dually, this leads to a categorical formulation of a factor object.
Remark 2.20. Lifts along monos and colifts along epis are unique.
(ker) “The” kernel of ϕ is an object K, usually denoted by ker ϕ, together with a morphism
κ : K → M satisfying the following universal property:
8
Exercise 2.5 will justify notation which was reserved for subobjects.
30 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
(i) κϕ = 0, and
(ii) κ dominates all such morphisms, i.e., for all objects L and all morphisms τ :
L → M with τ ϕ = 0 there exists a unique lift τ /κ : L → K (of τ along κ with
τ = (τ /κ)κ).
It follows from the uniqueness of the lift τ /κ that κ is a mono9 . We will usually call it
the kernel mono.
0
K κ
ϕ
τ /κ τ M N
L
0
K is called “the” kernel object of ϕ. Depending on the context ker ϕ sometimes stands
for the morphism κ and sometimes for the object K.
(coker) “The” cokernel of ϕ is an object C, usually denoted by coker ϕ, together with a mor-
phism ε : M → C satisfying the following universal property:
(i) ϕε = 0, and
(ii) ε codominates all such morphism, i.e., for all objects L and all morphisms η :
N → L with ϕη = 0 there exists a unique colift ε\η : C → L (of η along ε with
η = ε(ε\η)).
It follows from the uniqueness of the colift ε\η that ε is an epi. We will usually call it
the cokernel epi.
0
C ε
ϕ
ε\η η N M
L
0
Remark 2.20 implies that instead of insisting on the uniqueness of the lift and the colift
in the definition of kernels and cokernels we could have equivalently required that κ is
mono and ε is epi.
Luckily, this is already enough to define images and coimages without referring to ele-
ments.
Definition 2.22. Let ϕ : M → N be a morphism.
(coim) The co-image of ϕ, denoted by coim ϕ, is the cokernel of its kernel mono.
9
Since (ακ)ϕ = α(κϕ) = 0 we can reconstruct α as the unique lift of ακ along κ.
2. THE CATEGORY OF MODULES 31
0 0
ker ϕ coker ϕ
κ ε
ϕ
M N
0 0
εκ κε
coim ϕ im ϕ
We denote the cokernel epi of the kernel mono κ by εκ and call it the co-image epi. Dually,
we denote the kernel mono of the cokernel epi ε by κε and call it the image mono.
The homomorphism theorem for R-modules states that each R-linear map ϕ induces
an R-linear map ϕ : coim ϕ → im ϕ (between its coimage and its image) which is an iso-
morphism. Now we construct this induced morphism categorically by using the above
diagrams:
0 0
ker ϕ coker ϕ
κ ε
ϕ
M N
0 εκ \ϕ 0
εκ κε
coim ϕ im ϕ
0
ker ϕ coker ϕ
κ ε
ϕ
M N
εκ \ϕ 0
εκ κε
coim ϕ im ϕ
ϕ := (εκ \ϕ)/κε
Do we get the same morphism ϕ if we reverse the process, i.e., first lift then colift: εκ \(ϕ/κε )?
The answer is yes since for both of them pre-composing the epi εκ and post-composing the
mono κε yields ϕ. Hence, both are equal.
Summing up, in categories in which cokernels and kernels exist we have categorical
algorithms to compute coimages and images together with an induced morphism ϕ :
coim ϕ → im ϕ. This morphism is an isomorphism is the category of modules but not an
isomorphism in general categories with kernels and cokernels. However, being an isomor-
phism is one of the defining axioms of so-called A BELian categories. Hence, in A BELian
categories the homomorphism theorem, also called the first isomorphism theorem, will
hold true.
In order to talk about the second isomorphism theorem we still need to be able to per-
form intersections and sums of subobjects categorically. The last missing ingredient are
direct sums (more precisely, products and coproduct).
32 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
is an R-module,
Q called the (direct) product of {Ai }i∈I . For each i ∈ I we obtain an
R-epi πi : Aj → Ai , a 7→ ai , called the i-th projection.
For a finite index set I it is clear that products and coproducts of modules coincide
(this is false in general categories but still true in A BELian ones). We refer to them as “direct
sum” and write M ⊕N . Although we will only deal with finite index sets giving the general
definition was not more complicated. Now we give the categorical characterization:
Definition 2.24 ([HS97, §II.5, p. 58, p. 54]). Let I be an index set and {Ai }i∈I a family of
objects.
Q
(prod) “The”Qproduct of {Ai }i∈I is an object i∈I Ai , together with a family of morphisms
{πi : Aj Ai }i∈I , called projections10 , such that the following universal property
is satisfied:
For any object M and any family {ϕi : M → Ai }i∈I of morphisms there exists a
unique morphism
Y
ϕ = {ϕi } : M → Ai ,
called the product morphism or pairing, such that ϕπi = ϕi for all i ∈ I.
L
(sum) “The” coproduct of {A i }i∈I is an object i∈I Ai , together with a family of morphisms
Aj }i∈I , called embeddings11 , such that the following universal property
L
{ιi : Ai ,→
is satisfied:
For any object M and any family {ϕi : Ai → M }i∈I of morphisms there exists a
unique morphism
M
ϕ = hϕi i : Ai → M,
The universal properties imply that the product (resp. coproduct), in case it exists, is
unique up to a unique isomorphism.
Remark 2.25.
10
The projections are not necessarily epis unless the category has a zero object [Bru].
11
The embeddings are not necessarily monos unless the category has a zero object [Bru].
2. THE CATEGORY OF MODULES 33
(prod) The product morphism for a family {ϕi : M → Ai }i∈I of R-linear maps is given by
Y
ϕ = {ϕi } : M → Ai , m 7→ (i 7→ mϕi ) .
(sum) The coproduct morphism for a family {ϕi : Ai → M }i∈I of R-linear maps is given by
M X
ϕ = hϕi i : Ai → M, a 7→ ai ϕ i .
i∈I
Z ρ
ζ
λ
(pull-back diagram) X L
η
ϕ
J M
ψ
(push) Reversing all arrows in the above definition we obtain the dual definition of the push-
ϕ ψ
out X of two morphisms L ←
− M −
→ J, together with two push-out morphisms
λ
L→− X← − J.
Z ρ
ζ
λ
(push-out diagram) X L
η
ϕ
J M
ψ
Here we only considered finite pull-backs and push-outs. Next we show how a finite
pull-back can be reconstructed as a kernel of a coproduct morphism and, dually, a finite
push-out as a cokernel of a product morphism. We will make use of the fact that in our
context finite products and coproducts coincide.
Proposition 2.27. The following two dual statements hold.
34 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
ϕ ψ {λ,}
(pull) The pull-back of two morphisms L −
→M ←
− J is the kernel X ,→ L ⊕ J of the coproduct
hϕ,−ψi
morphism L ⊕ J −−−−→ M .
ϕ ψ hλ,i
(push) The push-out of two morphisms L ←
−M −
→ J is the cokernel X L ⊕ J of the product
{ϕ,−ψ}
morphism L ⊕ J ←−−−− M .
Proof.
λ
(pull) One can recover the pull-back morphisms L ←−X→ − J by post-composing the kernel
πL πJ
mono {λ, } with the projections L ←− L ⊕ J −→ J. This justifies the name {λ, }.
The unique morphism ζ in the pull-back diagram is the lift of the product morphism
{ρ, η} along the kernel mono {λ, }.
λ
(push) Dually, one can recover the push-out morphisms L → − X ← − J by pre-composing
ιL ιJ
the cokernel epi hλ, i with the embeddings L −→ L⊕J ← − J. This justifies the name
hλ, i. The unique morphism ζ in the push-out diagram is the colift of of the coproduct
morphism hρ, ηi along the cokernel epi hλ, i.
Now we are finally able to define the intersection of two submodules categorically. I
leave the sum as the dual exercise.
ιL ιJ
Definition 2.28. The intersection of two subobjects L ,→ M ←- J is their pull-back.
Remark 2.29. The module (or set) theoretic definition of intersection coincides with the
above categorical one.
Proof. The proof uses the reconstruction of the pull-back as the kernel of the coproduct
morphism. Indeed:
Before we see how to describe a constructive setup for all this and more it is time to
End define A BELian categories.
lect. 10
08.06
3 More categorical notions
Like with monoids, to each category A one can define the so-called opposite category Aop
with Aop op
0 := A0 , A1 := A1 , 1Aop = 1A , sAop = tA and tAop = sA (i.e., HomAop (M, N ) :=
HomA (N, M )) and composition µAop (ϕ, ψ) := µA (ψ, ϕ). Clearly, double dualizing recovers
A = (Aop )op .
Introducing the opposite category saves us half of our definitions:
• A (split) epi in A is (split) mono in Aop .
(4) ...
Definitions 2.9 and 2.10 of (split) epi, (split) mono, and isomorphism were already cate-
gorical.
Exercise 2.6. The set theoretic embedding Z ,→ Q is a mono and epi in (URngs), but
obviously not an isomorphism.
Definition 2.32 ([HS97, §II.1, p. 43]). Let A be a category.
• A terminal object T ∈ A is an object such that HomA (M, T ) is a singleton, i.e., consists
of a single morphism, for all M ∈ A.
• An initial object I ∈ A is an object such that HomA (I, M ) is a singleton, i.e., consists
of a single morphism, for all M ∈ A.
Both terminologies are dual: T is terminal A iff T is initial in Aop . Trivially, initial objects
and terminal objects are unique up to unique isomorphisms in A. A terminal object is the same
as an empty product and an initial the same as an empty coproduct.
Definition 2.33. A zero object12 0 ∈ A is an object which is both initial and terminal13 .
A category with zero is a category that has a zero object.
Define for each pair of objects M, N the unique zero morphism14
0M N := M → 0 → N.
(1) A singleton in (Sets) is terminal (and not initial) while the empty set is initial (and not
terminal).
(2) Z is an initial object in (URngs) which is not terminal. The zero ring is a terminal
object in (URngs) which is not initial.
(3) The trivial group is a zero object in (Grps) and (Ab). Both are categories with zero.
(pAdd1) the set HomA (M, N ) is an (additively written) A BELian group, in particular, there
exists an addition (and a subtraction) operation turning HomA (M, N ) into an A BELian
group;
(pAdd2) the composition HomA (M, N ) × HomA (N, L) → HomA (M, L) is bilinear.
(1) The definition of pre-additive is self-dual, i.e., the opposite category Aop is again pre-
additive.
(2) If A has a zero object then the unique zero morphism 0M N defined above is exactly
the zero element of the A BELian group HomA (M, N ), so there is no ambiguity talking
about the zero morphism from M to N .
(3) The A BELian group EndA (M ) is a unital ring for all M ∈ A. Conversely, a unital ring
is a pre-additive category with one object. Another name for a pre-additive category
would therefore be a “ringoid”. This is our third and last example of an oidification.
Remark 2.38. The following is true for additive categories (cf. [HS97, §II.9, p. 75ff]):
Then πM ιM + πN ιN = 1M ⊕N .
• It follows that finite coproducts also exist: (M ⊕ N ; ιM , ιN ) with ιM , ιN as above and the
coproduct morphism defined by
hϕ, ψi := πM ϕ + πN ψ : M ⊕ N → L,
• In particular, finite products and coproducts “coincide”. Such objects are called biprod-
ucts or direct sums.
We have just proved that the definition of an additive category is self-dual, i.e., the opposite
Aop of an additive category A is again additive.
{α,β} hϕ,ψi
• For K −−−→ L ⊕ M −−−→ N we have {α, β}hϕ, ψi = αϕ + βψ.
This means that if a category is additive, then its (in general non-unique) pre-additive struc-
ture is already determined by its structure as a category. In other words, there is then ex-
actly one possible definition of the A BELian group structure on the Hom-sets. We say the
addition for an additive category is internal to that category.
15
The use of the direct sum symbol ⊕ will be justified below.
4. ADDITIVE AND ABELIAN CATEGORIES 37
Definition 2.21 of kernel and cokernel is valid for any category with zero. Again, the
following definition is self-dual.
Definition 2.39. An additive category is called pre-A BELian if every morphism
Exercise 2.7. Let A be a pre-A BELian category. Prove that for an A-morphism ϕ:
Our definition of coimage and image and the construction of the induced morphism
ϕ
coim ϕ −
→ im ϕ in §2.4 are valid in any category with kernels and cokernels. Using this we
can finally state the definition of an A BELian category:
Definition 2.40 ([Rot09, §5.5, p. 307]). A pre-A BELian category is called A BELian if the
homomorphism theorem is valid, i.e., if for each morphism ϕ ∈ A
ϕ
(Ab) the induced morphism coim ϕ −
→ im ϕ is an isomorphism.
Note that this last axiom involves a single existential quantifier, namely the existence of
an inverse. This is the first reason why we prefer it over other equivalent formulations.
The second reason is original axiom G ROTHENDIECK’s gave in his seminal T ÔHOKU paper
[Gro57]. The third reason is that it emphasizes the homomorphism theorem, which is in my
opinion among the reasons behind various “tunnel effects” connecting apparently different
subfields of mathematics.
Again, these definitions are self-dual, i.e., the opposite Aop of a (pre-)A BELian category
A is again (pre-)A BELian.
ker ϕ coker ϕ
κ ε
ϕ
M N
(Diagϕ ) εκ εκ \ϕ ϕ/κε κε
coim ϕ ∼ im ϕ
εκ \ϕ/κε
| {z }
ϕ:=
From the identities εκ \ϕ = ϕκε and ϕ/κε = εκ ϕ we conclude that the colift εκ \ϕ is a
mono (as the composition of an iso and a mono) and that lift ϕ/κε is an epi (as the compo-
sition of an epi and an iso).
Definition 2.41. We call the colift εκ \ϕ the co-image mono16 and the lift ϕ/κε the image
epi.
Exercise 2.8. Show that the above diagram is functorial, i.e., a commutative diagram of
the form
16
It the categorical version of the co-restriction of ϕ to its image.
38 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
ϕ
M N
σ τ
ϕ0
M0 N0
K κ ε C
ϕ
(Sϕ ) M N
π ι
I
where
• πι = ϕ, i.e., ϕ is the composition of an epi and a mono,
• ι is the kernel of ε.
Proof. Set I := im ϕ, the image of ϕ, and define π and ι as the image epi and mono, respec-
tively17 .
Corollary 2.43. Let A be an A BELian category. Then
• every mono is a regular mono, i.e., is a kernel mono;
(3) More generally, the category R-Mod of left R-modules is an A BELian category. Like-
wise for the Mod-R, the category of right R-modules.
17
Alternatively, set I := coim ϕ and define π and ι as the co-image epi and mono, respectively. This is
equivalent since coim ϕ and im ϕ are isomorphic with the explicit isomorphism ϕ.
5. COMPUTABLE RINGS 39
(4) The full subcategory R-mod ⊂ R-Mod of finitely generated (f.g.) left R-modules over
a left N OETHERian ring R is an A BELian category. Likewise for mod-R ⊂ Mod-R, the
category of f.g. right R-modules over a right N OETHERian ring R.
(5) The dual of an A BELian category is again A BELian. An interesting discussion about
the opposite category of a module category can be found here [Ste].
Exercise 2.9. Show that the full subcategory (tfAb) ⊂ (Ab) of torsion-free A BELian
2
groups is pre-A BELian but not A BELian. For this consider the map Z ,→ Z and prove
that it is mono and epi (!) but not an isomorphism in (tfAb).
End
lect. 11
13.06
5 Computable rings
We could’ve started developing advanced abstract algorithms in A BELian categories but
it makes more sense to first describe a basic class of categories of finitely presented mod-
ules, which are not only A BELian but constructively A BELian [BLH11]. Once we have such a
constructively A BELian category A, i.e., once we have for A specified algorithms for all the
existential quantifiers and disjunctions indicated by boxes in the previous sections, then
all abstract algorithms become algorithms in the usual sense. In fact, the boxes describe all
the needed interfaces between the abstract and the concrete part of a well-structured com-
puter implementation. The above mentioned class of finitely presented module categories
has already a wide range of applications. And by modeling any other A BELian category A
(e.g., of coherent sheaves on certain varieties) on a constructively A BELian one (e.g., that of
finitely presented modules over a so-called computable ring), we get a proof that A itself is
constructively A BELian. Furthermore, all abstract algorithms which we will develop later
for A BELian categories will be automatically applicable for A.
In order to describe this basic class of categories of finitely presented modules which are
constructively A BELian we first need to specify the underlying rings. Recall that all rings
we consider are unital. Let R be such a ring. All matrices over R will have finite number of
rows and columns. Empty matrices with no rows or no columns are explicitly allowed.
Definition 2.45.
• A constructive set S consists of two algorithms, one to decide membership in S (for any
given input) and another to decide equality of elements in S.
Remark 2.46.
• We did not require to have an algorithm which decides if a constructive set is empty
or not.
– SA = 0;
0
– S row-dominates any other matrix S0 ∈ Ra ×r with S0 A = 0.
– AS = 0;
0
– S column-dominates any other matrix S0 ∈ Rc×a with AS0 = 0.
Definition 2.49. A left computable ring is a constructive ring R together with algo-
rithms for solving left sided inhomogeneous R-linear system of equations B = XA (for given
matrices A and B over R with equal number of columns), i.e., with three algorithms
X := −T =: RightDivide(B, A);
Analogously define a right computable ring with its three algorithms for solving right
sided equations AX = B
• DecideZeroColumns(B, A);
X := LeftDivide(A, B);
• SyzygiesOfColumns(A).
18
Both matrices generate the same row-space
19
So we do not require a “normal form”, but only a mechanism to decide if a row is zero modulo some
relations.
5. COMPUTABLE RINGS 41
(Obj) (R-fpres)0 := the set of all matrices M over R (including the empty ones). We call the
number of columns of a matrix M the number of generators21 and denote it by gM ;
(Mor) (R-fpres)1 := the set of left compatible triples (M, A, N) over R modulo a certain equiv-
alence relation ∼, where a triple of matrices (M, A, N) over R is called left compatible
if
– A is a gM × gN matrix22 ;
– N ≥row MA, i.e., the left sided equation XA N = MA is solvable 23 for some matrix XA .
– M = M0 ;
– N = N0 ;
– N ≥row A − A0 , i.e., there exists 24 a matrix Y = YA,A0 such that25 A − A0 = YN.
(µ) µ((M, A, N), (N, B, L)) = Compose((M, A, N), (N, B, L)) = (M, A, N)(N, B, L) = (M, AB, L), where
AB is the matrix multiplication.
A
We occasionally denote a morphism (M, A, N) by M →− N.
The category fpres-R of (finite) right presentations over R is defined analogously. Ev-
erything we do below also holds for fpres-R.
Indeed the composed tripple (M, AB, L) is again compatible: There exists an XA such that
MA = XA N and XB such that NB = XB L and therefore M(AB) = (MA)B = (XA N)B = XA (NB) =
XA (XB L) = (XA XB )L. Furthermore, the composition is well-defined for classes: if (M, A, N) ∼
(M, A0 , N) via YA,A0 and (N, B, L) ∼ (N, B0 , L) via YB,B0 then AB−A0 B0 = (A0 +YA,A0 N)(B0 +YB,B0 L)−A0 B0 =
NB0 +A0 YB,B0 L + YA,A0 NYB,B0 L = (...)L. The remaining defining properties are all obvious.
YA,A0 |{z}
=XB0 L
When R is left (resp. right) computable then R-fpres (resp. fpres-R) will turn out to be a
constructive category in the following sense (see Proposition 2.56):
Definition 2.54. We call a category A constructive if A0 , A1 are constructive sets and all
four structure maps s, t, 1, µ are realized by algorithms.
21
We will justify this name in the next chapter.
22
in particular, the matrix multiplication MA is defined.
23
This is part of the membership decision algorithm in (R-fpres)1 .
24
This is part of the membership equality algorithm in (R-fpres)1 .
25
We say that the matrix A is unique up to a multiple of N.
6. A CONSTRUCTIVELY ABELIAN CATEGORY OF MATRICES 43
Remark 2.55. More generally, we call a mathematical structure constructive if all ex-
istential quantifiers and all disjunctions appearing in its defining axioms are realized by
algorithms. All the above constructivity notions were in this sense special cases. However,
as this notion is usually not treated in classical math curricula we decided to define some
special cases explicitly to underline the various and maybe unexpected places where such
quantifiers occur.
Note that we do not require to have an algorithm which decides if two different objects
are isomorphic or not. However, in a constructively A BELian category A the discussion
below will provide an algorithm capable of deciding whether a given A-morphism α : M →
N is an isomorphism or not.
End
Proposition 2.56. If R is left computable then R-fpres is a constructive category. lect. 12
20.06
Proof.
• The membership and equality algorithms for (R-fpres)0 are derived from those of R.
• For the membership and equality algorithms for (R-fpres)1 we use the left computabil-
ity of R.
We call the corresponding matrix algorithms AddMat, SubMat, and a constructor ZeroMat
for (possibly empty) zero matrices, respectively. All of them rely on R merely being a
constructive ring.
This constructive pre-additive structure is, as expected, uniquely determined and hence
reproducible from the constructive additivity of the category discussed below.
Remark 2.58. If R is left computable then R-fpres is a constructive category with zero.
Proof. The 0 × 0 matrix 0 over R is a zero object in R-fpres. Indeed, there is a unique mor-
phism (M, 0, 0) and (0, 0, N) with middle matrices being of the shape ? × 0 and 0×?, re-
spectively. All involved empty matrices can be constructed using the matrix constructor
ZeroMat.
As expected, the zero morphism 0MN := M → 0 → N = (M, 0, 0)(0, 0, N) coincides with the
zero morphism (M, 0, N), where the zero matrix is the product of the two empty matrices.
44 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
together with the two morphisms (given by the two compatible triples)
πM := ProjectionOnLeftFactor(M, N) := (M ⊕ N, ( 10 ) , M) ,
πN := ProjectionOnRightFactor(M, N) := (M ⊕ N, ( 01 ) , N)
as a candidate for the product of M and N and its two projections. This relies on the matrix
constructor algorithms DiagMat and StackMat to stack matrices. Indeed for two arbitrary
A B
morphisms ϕ : L → − M and ψ : L →− N there is exactly one morphism (i.e., one equivalence
class of compatible triples)
such that {ϕ, ψ}πM = ϕ and {ϕ, ψ}πN = ψ, as can be verified by block matrix multiplication.
For this we use a constructor AugmentMat to augment matrices.
Remark 2.60. Constructing the embeddings in the biproduct and the coproduct mor-
phism now follows from Remark 2.38. The following algorithms are in this sense derived
algorithms:
A+B A
• As mentioned above, the sum ϕ + ψ : M −−→ N of two parallel morphisms ϕ : M →
− N,
B
ψ:M→− N can be reconstructed as
Before we discuss the pre-A BELian structure we need a relative version of the matrix
algorithm SyzygiesOfRows, which we can deduce as a derived algorithm.
Definition 2.61. Let A and N be two stackable matrices over R. We say that K is a matrix
of row syzygies of A modulo N if
• N ≥row KA;
Before we discuss the A BELian structure we need a relative version of the matrix algo-
rithm RightDivide, which we can deduce as a derived algorithm.
Definition 2.63. Let A, B, N be stackable30 matrices over R. The matrix A is said to row-
dominate B modulo N if there exists matrices X, Y such that B = XA + YN. The definition of
relative column-domination is analogous.
Since B = XA + YN = ( X Y ) ( AN ) we can derive X (and Y) form RightDivide and StackMat
( X Y ) := RightDivide(B, ( AN )).
X =: RightDivide(B, A, N).
Proof. First we will show that any mono is the kernel of its cokernel (cf. Corollary 2.43). Let
A
ϕ : M ,→ N be a mono. From the proof of Proposition 2.62 we know that ε = (N, 1, ( AN ) = coker ϕ)
is the cokernel epi of ϕ.
0
M A
1
γ/ϕ = X G N ( AN ) = coker ϕ
M0
0
G
Let γ : M0 → − N be another morphism with γε = 0, i.e., ( AN ) ≥row G = G · 1, or, equivalently,
A ≥row G modulo N, i.e., there exists X, Y such that G = XA + YN. A candidate for the unique
X
lift32 of γ along ϕ would be γ/ϕ : M0 →
− M with
X = RightDivide(G, A, N),
once we have shown that the triple (M0 , X, M) is compatible, i.e., that M ≥row M0 X: For this
note that N ≥row M0 G since γ is a morphism. Furthermore N ≥row M0 YN, tautologically. Hence
N ≥row M0 G − M0 YN = M0 XA = (M0 X)A. Now recall that ϕ is mono means that its kernel mono
vanishes by Exercise 2.2.7, which translates to M ≥row K := SyzygiesOfRows(A, N) (cf. the
proof of Proposition 2.62). But K row-dominates by definition any T with N ≥row TA, in
End particular M ≥row K ≥row M0 X.
lect. 13 A
To show that any epi ϕ : M N is a cokernel epi of its kernel mono κ := (ker ϕ, K, M)
22.06 (cf. Corollary 2.43) it suffices to show that the unique colift ε\ϕ = (coker κ = ( K ) , A, N)
M
of ϕ along the cokernel epi ε = (M, 1, ( KM ) = coker ϕ) of κ is an isomorphism, where K :=
SyzygiesOfRows(A, N) (cf. the proof of Proposition 2.62):
30
i.e., with the same number of columns.
31
In practice, one can derive more efficient algorithms to compute the relative version of RightDivide.
32
Lifts along monos are unique.
7. FUNCTORS 47
coker κ = ( KM )
1
K := SyzygiesOfRows(A, N)
ε\ϕ = A M ker ϕ
A
N
An inverse morphism α of the colift ε\ϕ must satisfy α(ε\ϕ) = 1N . This implies for α that
Y
α := (ε\ϕ)−1 : N →
− ( KM ) = coker κ with Y := RightDivide(1, A, N),
where 1 is the gN × gN identity matrix. The existence of the matrix Y is equivalent to our
assumption that ϕ is an epi. Now we show that the triple α = (N, Y, ( KM )) is compatible and
hence defines a morphism. By definition of Y there exists a matrix Z such that
YA + ZN = 1. (*)
Multiplying with N from the left we obtain NYA + NZN = N, or equivalently (NY)A = (1 − NZ)N,
i.e., N ≥row (NY)A. But K row-dominates by definition any T with N ≥row TA, hence K ≥row NY
and α is a compatible triple. Since α(ε\ϕ) = 1N by definition of Y it remains to show that
(ε\ϕ)α = 1coker κ . To show this we multiply (*) with A from the left and obtain AYA + AZN = A,
or equivalently (AY − 1)A = (AZ)N, hence N ≥row (AY − 1)A and as above K ≥row (AY − 1) and
we are done.
In the next sections we will develop enough language to show that R-fpres is equivalent
to the category R-fpmod of finitely presented modules over R. When R is left computable
this will prove that the category of finitely presented R-modules is constructively A BELian.
7 Functors
In this chapter we will study functors as “1-morphism” between categories. It turns out that
there are “2-morphisms” which we will call natural transformations; they are “morphisms
between the functors”.
Definition 2.65. Let A and B be categories. A (covariant) functor F : A → B consists of
two maps A0 → B0 (on objects) and A1 → B1 (on morphisms33 ) such that
• F (HomA (M, N )) ⊂ HomB (F (M ), F (N ));
• F (1M ) = 1F (M ) ;
• F (ϕψ) = F (ϕ)F (ψ)
for all M, N ∈ A and any pair of composable A-morphisms ϕ, ψ. We write F ∈ B A .
We will see later how to turn B A into a category of functors.
The composition of functors is again a functor and the composition is associative. The
notions of identity functor, invertible functor, and inverse functor of an invertible functor
are evident. However, invertible functors offer a far too restrictive notion of equivalence,
which does not take the relative nature of categories into account, i.e., the possibility of
many unequal but isomorphic objects to exist.
33
You cannot speak of a functor if you don’t know how to define it on morphisms!
48 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
F is called
F is called essentially surjective (or dense) if for each of L ∈ B there exists an object M ∈ A
such that F (M ) ∼
= L (in B).
We start with some trivial examples.
Example 2.67.
(6) The so-called “forgetful” (or “underlying”) functors which forget some structure of
the category
(7) The “free” functors which associate to a set the “free” object on this set, e.g.,
(8) ...
Main example 2.68. Let A be a category. Each object M ∈ A gives rise to a functor to
(Sets), namely the covariant Hom-functor
34
viewed as skeletally thin categories (cf. 2.4.(12))
8. NATURAL TRANSFORMATIONS 49
ψ N
M ϕ
ψϕ L
ψ M
N ϕ
ϕψ L
8 Natural transformations
Natural transformations will play the role of second level morphisms, i.e., the morphisms
between the functors.
Definition 2.71. A natural transformation η : F → G of the functors F, G : A → B
is a map η : A0 → B1 such for all A-morphisms ϕ : M → N that the following diagram
commutes
F (ϕ)
F (M ) F (N )
ηM ηN
G(ϕ)
G(M ) G(N )
F
A η
B
G
50 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
(1) We now show that B A with functors as objects and natural transformations as mor-
phisms is a category:
η
• The source of a natural transformation F =⇒ G is F and the target is G.
• The map 1F : A0 → B1 , M 7→ 1F (M ) is a natural automorphism of F , called the
identity transformation of F .
η τ
• The (vertical) composition of two natural transformations F =⇒ G =⇒ H is
defined in the obvious way (ητ )M = ηM τM , i.e., by composing the two commu-
tative diagrams
F (ϕ)
F (M ) F (N )
ηM ηN
G(ϕ)
G(M ) G(N )
τM τN
G(ϕ)
H(M ) H(N )
ητ
and is again a natural transformation F =⇒ H. The composition of natu-
ral transformations is associative since it is defined by the composition of mor-
phisms.
End
lect. 14
(2) The evaluation map εV : V → V ∗∗ , v 7→ (ϕ 7→ ϕ(v)) between a k-vector space and
27.06
its double dual is a natural mono between the identity functor Idk-Vect and double-
dualizing functor D2 : k-Vect → k-Vect. The natural transformation ε becomes a
natural isomorphism if we restrict ourselves to the category of finite dimensional k-
vector spaces. The natural transformation ε is also definable for R-Mod and R-mod,
where it is in general neither mono nor epi.
(3) Using Exercise 2.14 we see that the kernel mono, cokernel epi, the coimage mono and
epi, the image mono and epi, and the induced morphism ϕ : coim ϕ → im ϕ are all
natural transformations. In A BELian categories the latter is a natural isomorphism
between the coimage and the image functors.
Remark 2.73. If we view categories as objects and functors as morphisms we get a cate-
gory of categories, with the restrictive notion of equality of functors. If we want to consider
functors up to equivalence we get what is called a 2-category with the natural transforma-
tions as 2-morphisms. This process can be iterated indefinitely with marvelous applications
in advanced mathematics.
Exercise 2.15. The category of functors AB from a small category B to an A BELian cate-
gory A is again A BELian.
8. NATURAL TRANSFORMATIONS 51
F :AB:G
with two natural isomorphisms G ◦ F ' IdA and F ◦ G ' IdB . The two categories A, B are
then called equivalent and we then write F : A ' B : G, or simply A ' B.
F G
If F : A ' B : G then the composed map HomA (M, N ) → HomB (F (M ), F (N )) →
HomA ((G ◦ F )(M ), (G ◦ F )(N )), ϕ 7→ (G ◦ F )(ϕ) is a bijection, given by conjugating with
(the components of) the natural isomorphism. Hence the maps FM,N : HomA (M, N ) →
HomB (F (M ), F (N )) are bijections for all M, N ∈ A0 . In other words, F is fully faithful.
Now, for B ∈ B the B-component B ∼ = F (G(B)) of the natural isomorphism F ◦ G ' IdB
states that F is essentially surjective. Thus, we have proved one direction of the following
criterion.
Proposition 2.75. A functor F : A → B is part of an equivalence iff it is fully faithful and
essentially surjective.
Proof. The other direction is immediate with a sufficiently strong axiom of choice.
Remark 2.76. Equivalence of categories preserves all categorical properties, i.e.,
• having initial, terminal, or zero objects;
• ...
In fact, one can always improve the above notion of equivalence to an “adjoint equiva-
lence”, which depends on the notion of adjoint functors.
Now we prove that R-fpres is indeed equivalent to the category of finitely presented
R-modules. We already know that HomR (R1×r , R1×c ) ∼ = Rr×c , i.e., that R-matrices are in
bijection with R-linear maps between free modules of finite rank.
Definition 2.77. Let R be a ring. A left R-module N is called finitely presented if there
exists a matrix N ∈ RrN ×gN such that N ∼
N
= coker(R1×rN →
− R1×gN ). We call such an N a pre-
sentation matrix of N . We denote the category of finitely presented modules by R-fpmod.
It is a full subcategory of R-mod.
The images (f1 , . . . , fgN ) of the standard free generators (f1 , . . . , fgN ) of the free module
1×gN
R under the cokernel epi εN : R1×gN N is a set of generators of N . By definition,
N εN = 0N iff N ≥row N0 , i.e., there exists a matrix X such that N0 = XN.
0
The rows of the matrix N are the images of the standard free generators (f01 , . . . , f0rN ) of
the free module R1×rN under the R-linear map N : R1×rN → R1×gN . They are called a set of
relations of N .
We now want to describe R-linear maps between M = coker M and N = coker N. Let
A ∈ RgM ×gN be a matrix for which N ≥row MA, i.e., for which there exists an X satisfying
MA = XN.
M εM
R1×rM R1×gM M
X A ϕA
N εN
R1×rN R1×gN N
52 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
Such an A induces an R-linear map ϕ = ϕA : M → N, mεM 7→ mAεN forcing the right square
to commute: εM ϕ = AεN . We only need to see that ϕ is well-defined. For this we have to
check that mAεN = 0N whenever mεM vanishes. But mεM = 0M means that there exists a matrix
x ∈ R1×rM such that m = xM. Hence, mAεN = xMAεN = xXNεN = 0N . Two matrices A, A0 give
rise to same R-linear map ϕ iff AεN = A0 εN , or equivalently, iff (A − A0 )εN = 0N and hence iff
N ≥row A − A0 .
PgNConversely, an R-linear map ϕ : M → N defines a coefficients matrix 1×gM
A by ei ϕ = ei εM ϕ =
j=1 Aij fj , where (e1 , . . . , egM ) are the standard free generators of R . So, by definition,
εM ϕ = AεN and the right square commutes.
Summing up, we have just proved the following:
Theorem 2.78. The essentially surjective functor
R-fpres → R-fpmod,
coker : M 7→ M = coker M,
(M, A, N) 7→ ϕA : M → N = coker N
is fully faithful and is hence an equivalence of categories R-fpres ' R-fpmod. If R is left computable
then R-fpmod is A BELian and R-fpres is constructively A BELian. In particular, R-fpres is a
constructive model for R-fpmod.
π
represented by theπ cokernel
epi of ι. And to a factor object M C we can associate the
subobject ker M C represented by the kernel mono of π. Corollary 2.43 now says that
in A BELian categories this G ALOIS connection is a bijection. Hence, we can draw subobjects
and factor objects using just one H ASSE diagram. The second isomorphism of N OETHER
implies the modularity of these lattices.
τ κ
Let L ,→ M ←- K be two subobjects of M with L ≤ K in some A BELian category. In
τ /κ
Exercise 2.5 we saw how the unique lift L ,→ K of τ along the dominant κ realizes L as a
subobject of K.
Definition 2.79. We then call the factor object
τ /κ
K/L := coker L ,→ K
The two morphisms K/L ,→ M/L M is an example of what we will later call a
generalized mono from K/L to M .
The definition of subfactors is all what we need to define defects of exactness, one of the
central notions in homological algebra.
Definition 2.80. Let A be a category, with zero when necessary. Two composable A-
ϕ ψ
morphisms M − → N −
→ L are called a short sequence. We call it a differential short se-
quence if ϕψ = 0.
ϕ ψ
Let M − → N − → L be a differential sequence in an A BELian category. The defect of
exactness (at N ) is defined as the subfactor object of N
Def(ϕ, ψ) = ker ψ/ im ϕ,
where the two subobjects im ϕ ≤ ker ψ of N are represented by the image mono and the
kernel mono im ϕ ,→ N ←- ker ψ, respectively36 . The sequence is called exact (at N ) if
Def(ϕ, ψ) = 0.
ϕ ψ
Exercise 2.16. Let M −
→N − → L be differential sequence in an A BELian category. The
following statements are equivalent:
• The sequence is exact at N .
Proposition 2.87 (Bivariate left exactness of Hom). Let A be an A BELian category and N ∈
A.
Proof. Exercise.
Exercise 2.19. Show that the Hom-functor is generally not exact in any of its arguments.
10 Free resolutions
Let M be an R-module. We called an exact sequence of the form
ϕ
0 ← M ← F0 ←
− F1
with free R-modules F1 , F2 a free presentation of M . If F1 and F2 are of finite ranks, then
we can identify ϕ with the presentation matrix we encountered Definition 2.77. Iterating
this process for the kernel of the presentation morphism ϕ yields:
Definition 2.88 (Free resolution). A (possibly infinite) exact sequence of the form
ϕ1 ϕ2
0 ← M ← F0 ←− F1 ←− F2 ← · · ·
with free R-modules Fi is called a free resolution of M . The submodule ker ϕi = im ϕi+1
is called the i-th syzygy module of M . We say that a resolution is finite of length c ∈ N if
Fi = 0 for all i ≥ c + 1.
The isomorphism type of the i-th syzygy module of M depends on the choice of the free
resolution of M .
The matrix version of Hilbert’s Syzygy Theorem Theorem 1.46 implies:
Corollary 2.89 (Hilbert’s syzygy theorem (module version)). Let K be a field and R =
K[x1 , . . . , xn ]. Then each finitely generated R-module M has a finite free resolution of length ≤ n
consisting of free modules of finite rank.
Example 2.90. Let R := Z/4Z. Using Ext’s it can easily be shown that the R-module
R/h2i does not admit a finite free resolution.
56 CHAPTER 2. CONSTRUCTIVE ABELIAN CATEGORIES
Chapter 3
1 Preliminaries
We start by recalling some operations on ideals:
T
• the (set theoretic) intersection i∈I Ii ;
Qk DQ E
k
• the product i=1 Ii := i=1 ai | ai ∈ Ii ;
R
P S
• the sum i∈I Ii := i∈I Ii R
;
Example 3.1. Consider R := Z, I := hmi = mZ, and J := hni = nZ for m, n ∈ N≥1 . Then
• Ann(I) = {0}.
√ DQ
k
E
, where m = ki=1 pαi i is the prime factorization of m with αi > 0.
Q
• I= i=1 p i
Exercise 3.1. Write down categorical algorithms for the intersection, the sum, and the
ideal quotient.
Definition 3.2. A ring is called reduced if 0 is the only nilpotent element of R.
√
Exercise 3.2. R/I is reduced iff I is a radical ideal (i.e., I = I).
Remark 3.3.
1
for a ∈ R we set I : a := I : hai
57
58 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
(1) Clearly
k
Y k
\
Ii ⊂ Ii .
i=1 i=1
The example 4Z · 6Z = 24Z ( 12Z = 4Z ∩ 6Z shows that the inclusion could be strict.
(2) However, v v
u k uk
uY u\
t I =t I.
i i
i=1 i=1
Definition 3.9. A local ring R is a ring with exactly one maximal ideal m (i.e., Max R is
a singleton {m}). We occasionally write (R, m) and call R/m the residue field of the local
ring (R, m).
Proof.
√
“⊂” Let a ∈ I and p R be some prime ideal such that I ⊂ p. Then there exists an m ≥ 1
with am ∈ I ⊂ p by the definition of the radical. Hence a ∈ p, since p is prime.
√
“⊃” Let a ∈ R \ I. The inclusion relation on the set
Γ := {J R | I ⊂ J, am ∈
/ J ∀m ≥ 1}
“⊃” Let a ∈
/ J(R). Then there exists a maximal ideal m of R with a ∈/ m. Hence m+hai = R,
so one can find elements c ∈ m and b ∈ R such that c + ab = 1. But then 1 − ab = c ∈ m,
which cannot be a unit.
60 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
χϕ (ϕ) = 0 ∈ EndR M .
where A = (aij ) and In is the n × n identity matrix. Multiplying with adjoint matrix (x · In −
A)adj we get
m1
..
det(x · In − A) . = 0.
mn
So det(x · In − A)mi = 0 for all i, hence χϕ (x) := det(x · In − A) ∈ AnnR[x] M , or equivalently
χϕ (ϕ) = 0 by (**).
Corollary 3.14. Let M be a f.g. R-module.
(2) Let R 6= 0. If M ∼
= Rn , then any set of n elements generating M is a free basis. In particular,
rank M = n is well-defined.
Proof.
3. THE CAYLEY-HAMILTON THEOREM AND NAKAYAMA’S LEMMA 61
(1) We will construct the inverse of α. For this consider M as an R[t]-module via t · m :=
α(m) for all m ∈ M . Define I := hti R[t] and ϕ := idM ∈ EndR[t] M . Then
α epi
ϕ(M ) = M = α(M ) = t · M = I · M .
Now consider
p1 + · · · + pn
q(t) := ∈ R[t].
t
Then for all m ∈ M
n
! n
!
X X
m = idM (m) = − pi idM (m) = − pi m = t · (−q(t)) · m.
i=1 i=1
Example 3.20. Caution: The last statement may not hold for non-local rings: Consider
a field K and hx, 1 − xi = h1i = K[x]. Both sets of generators are minimal but their cardi-
nalities differ.
Corollary 3.21. Let (R, m) be a local ring, M a f.g. R-module, and ϕ : M → N a morphism of
R-modules. Then ϕ is an epi iff the induced map ϕ : M/mM → N/mN, m 7→ ϕ(m) is an epi.
Proof. We only proof the nontrivial direction. So suppose ϕ is epi. Then given n ∈ N , there
exists an m ∈ M such that ϕ(m) = n. That is ϕ(m) − n ∈ mN . Hence im ϕ + mN = N and
im ϕ = N by Corollary 3.18.
Φ:M ×N →L
t T
M ×N ϕ
0
t0 T
with mi ∈ M and ni ∈ N .
Remark 3.26. Let X and Y be sets of generators of M and N , respectively. Then {x ⊗ y |
x ∈ X, y ∈ Y } is a generating set of M ⊗ N . In particular, the tensor product preserves finite
generatedness.
Analogously, the general tensor product M1 ⊗ · · · ⊗ Mk can be defined as the universal
object of R-multilinear maps with source M1 × · · · × Mk .
Proposition 3.27. Let M, N, L be R-modules. Then there exist unique isomorphisms
(1) M ⊗ N ∼= N ⊗ M mapping m ⊗ n 7→ n ⊗ n.
(2) (M ⊗N )⊗L ∼
= M ⊗N ⊗L ∼
= M ⊗(N ⊗L) mapping (m⊗n)⊗` 7→ m⊗n⊗` 7→ m⊗(n⊗`).
(3) (M ⊕ N ) ⊗ L ∼
= (M ⊗ L) ⊕ (N ⊗ L) mapping (m ⊗ n, `) 7→ (m ⊗ `, n ⊗ `).
(4) R ⊗ M ∼
= M mapping r ⊗ m 7→ r · m.
Proof. These are all simple exercises using the universal property. We will prove the third
statement as an example:
(3) The map
(M ⊕ N ) × P → (M ⊗ L) ⊕ (N ⊗ L),
((m, n), `) 7→ (m ⊗ `, n ⊗ `)
is R-bilinear in (m, n) and `. Hence, there exists a unique morphisms of R-modules
ϕ : (M ⊕ N ) ⊗ L → (M ⊗ L) ⊕ (N ⊗ L) such that
(m, n) ⊗ ` 7→ (m ⊗ `, n ⊗ `).
For the inverse map first use the universal properties to construct the obvious morphisms
M ⊗ L → (M ⊕ N ) ⊗ L and N ⊗ L → L → (M ⊕ N ) and then take their product morphism
ψ : (M ⊗ L) ⊕ (N ⊗ L) → (M ⊕ N ) ⊗ L satisfying
(m ⊗ `, n ⊗ `) 7→ (m, 0) ⊗ ` + (0, n) ⊗ `.
The morphisms ϕ and ψ are mutually inverse on pure tensors and hence an all elements.
64 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
Example 3.28.
(1) The R-linear map
∼
=
Rm ⊗ Rn → Rm×n ,
(x1 , . . . , xm ) ⊗ (y1 , . . . , yn ) 7→ (xi · yj )
is an isomorphism. In particular, ei ⊗ ej form a free basis of Rm ⊗ Rn .
(2) Let M be an R-module and I R an ideal. Then the R-linear map
∼
=
M ⊗ R/I → M/IM,
m ⊗ (r + I) 7→ rm + IM
is an isomorphism.
(3) Let M, N, L be R-modules. Then the R-linear map
∼
=
HomR (M ⊗ N, L) → HomR (M, HomR (N, L))
(
M → HomR (N, L)
ϕ 7→ ϕc :
m 7→ (N → L, n 7→ ϕ(m ⊗ n))
is an isomorphism.
Remark 3.29. Let ϕ : M → N and ϕ0 : M 0 → N 0 be morphisms of R-modules. Then
M × M0 → N ⊗ N0
(m, m0 ) 7→ ϕ(m) ⊗ ϕ(m0 )
is R-bilinear. Hence, there exists a morphism
ϕ ⊗ ϕ0 : M ⊗ M 0 → N ⊗ N 0 such that
m ⊗ m0 7→ ϕ(m) ⊗ ϕ0 (m0 ).
End
lect. 17 This turns − ⊗R N and N ⊗R − into functors R-Mod → R-Mod.
06.07 Given a morphism ϕ : R → S of unital commutative rings, one can make S into an
R-module by setting
r · s =: ϕ(r) · s for all r ∈ R, s ∈ S.
Then the R-module structure is compatible with the ring structure of S, i.e., (r · s) · s0 =
r · (s · s0 ) for all r ∈ R, s, s0 ∈ S.
Definition 3.30. The ring S together with the above R-module structure is called an R-
algebra. An R-subalgebra of S is a subring S 0 ≤ S containing im ϕ. An R-algebra morphism
between two R-algebras S and T is a morphism α : S → T of rings which is an R-module
homomorphism. In terms of the ring homomorphisms ϕ : R → S and ψ : R → T defining
the R-algebra structures this means that α ◦ ϕ = ψ.
Example 3.31 (Tensor product of R-algebras). Let S, T be R-algebras defined by homo-
morphisms ϕ : R → S and ψ : R → T . Then the universal property of the tensor product
and Proposition 3.27 yield a multiplication on S ⊗R T such that
(s ⊗ t)(s0 ⊗ t0 ) = ss0 ⊗ tt0 .
This gives a unital ring with 1S⊗T = 1S ⊗ 1T . The ring homomorphism
R → S ⊗R T, r 7→ ϕ(r) ⊗ 1T = 1S ⊗ ψ(r)
turns S ⊗R T into an R-algebra.
5. LOCALIZATION OF RINGS 65
Jump
Ch2§9 Exercise 3.4. Prove the converse of Proposition 2.87.1 for the category of R-modules,
i.e., the sequence M 0 → M → M 00 → 0 is exact iff the induced sequence
0 → HomR (M 00 , HomR (N, L)) → HomR (M, HomR (N, L)) → HomR (M 0 , HomR (N, L))
is exact. Now use Exercise 3.4. The right exactness of the other functor now follows from
Proposition 3.27.1.
Definition 3.33. An R-module N is called flat if the tensor functor − ⊗R N is exact.
Exercise 3.5.
(1) Show that the tensor functor is generally not exact in any of its arguments.
5 Localization of rings
Let R be a commutative unital ring.
Definition 3.34. A subset U ⊂ R is called multiplicatively closed, if it is closed under
(finite) products, i.e., if 1 ∈ U (as the empty product) and the product of two elements of U
is again in U .
Definition 3.35. Let U ⊂ R be a multiplicatively closed subset. We define an equiva-
lence relation on R × U by2
r r0 ru0 + r0 u
+ := ,
u u0 uu0
r r0 rr0
· := .
u u0 uu0
We call this ring the localization of R at U .
2
The element v in the definition is essential to guarantee the transitivity of the relation.
66 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
Proof.
(2) We have J ce ⊂ J by Remark 3.5.2. For the reverse inclusion let r ∈ R, u ∈ U . Then
r r r
∈ J =⇒ ∈ J =⇒ r ∈ J c =⇒ ∈ J ce .
u 1 u
Example 3.39.
(1) Let U be the set of non-zero-divisors of R. Then Q(R) := R[U −1 ] is called the total
quotient ring of R. In the special case where R is an integral domain, Q(R) is a field,
the quotient field or field of fractions of R.
h i
(2) Let f ∈ R and U = {f m | m ≥ 0}. Then Rf := R f1 := R[U −1 ].
Remark 3.40. For p ∈ Spec R denote by πp : R R/p the natural epimorphism and by
ιR\p : R → Rp the localization morphism.
• The map πp−1 : Spec R/p → Spec R, q 7→ qc is an embedding. Its image is {p0 ∈ Spec R |
p0 ⊃ p}.
Geometrically speaking, the passage to the residue class ring R/p means to restriction to the
affine subscheme Spec R/p of Spec R, whereas the passage to Rp considers the infinitesimal
neighborhood Spec Rp of the subscheme Spec R/p.
68 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
6 Localization of modules
Definition 3.42. Let U be a multiplicatively closed subset of R and M an R-module. We
get an equivalence relation on M × U by setting
We write nm o
−1 −1
U M := M [U | m ∈ M, u ∈ U
] :=
u
for the set of equivalence classes and make M [U −1 ] into an R[U −1 ] module with addition as
for R[U −1 ] and scalar multiplication ur · um0 := u·u
r·m
0 . The R[U
−1
]-module M [U −1 ] is called the
localization of M at U .
Remark 3.43. A morphism ϕ : M → N of R-modules induces a morphism
ϕ[U −1 ] : M [U −1 ] → N [U −1 ]
m ϕ(m)
7→
u u
of R[U −1 ]-modules. Indeed, localization at U is a (covariant) functor R-Mod → R[U −1 ]-Mod:
(1) idM [U −1 ] = idM [U −1 ] ;
ϕ ψ
(2) If M 0 − → M 00 are morphisms of R-modules, then (ψ ◦ ϕ)[U −1 ] = ψ[U −1 ] ◦ ϕ[U −1 ].
→M −
From Remark 3.37.1 we deduce:
Remark 3.44. Let I R and ιU : R → R[U −1 ] the localization of R at the multiplicatively
closed subset U ⊂ R. Then
I e = I[U −1 ].
ϕ
Proposition 3.45. The localization functor R-Mod → R[U −1 ]-Mod is exact, i.e., if M 0 −
→
ψ ϕ[U −1 ψ[U −1 ]
M− → M 00 is an exact sequence of R-modules, then M 0 [U −1 ] −−−→ M [U −1 ] −−−−→ M 00 [U −1 ] is an
exact sequence of R[U −1 ]-modules.
Proof. From ψ[U −1 ] ◦ ϕ[U −1 ] = (ψ ◦ ϕ)[U −1 ] = 0 we know that im ϕ[U −1 ] ≤ ker ψ[U −1 ]. We
now show the reverse inclusion. Let m u
∈ ker ψ[U −1 ]. Then
m ψ(m)
−1
0 = ψ[U ] = .
u u
6. LOCALIZATION OF MODULES 69
Hence, there exists a v ∈ U such that 0 = v · ψ(m) = ψ(v · m). But then v · m ∈ ker ψ = im ϕ
and there exists an m0 ∈ M 0 with ϕ(m0 ) = v · m and we conclude that
ϕ(m0 )
0
m vm −1 m
= = = ϕ[U ] ∈ im ϕ[U −1 ].
u vu vu vu
In particular, the localization preserves monos: So if N is an R-submodule of M then
N [U −1 ] is an R[U −1 ]-submodule of M [U −1 ].
Corollary 3.46. Let N and L be R-submodules of M . Then
(1) (N + L)[U −1 ] = N [U −1 ] + L[U −1 ];
(2) (N ∩ L)[U −1 ] = N [U −1 ] ∩ L[U −1 ];
(3) (M/N )[U −1 ] ∼
= M [U −1 ]/N [U −1 ].
Proof. This follows by abstract nonsense about exact functors between A BELian categories.
End
Note that the localization morphism ιU : R → R[U −1 ] endows R[U −1 ] with an R-algebra lect. 20
structure, and in particular with an R-module structure. 13.07
Proposition 3.47. The localization functor at U is naturally isomorphic to the tensor functor
R[U −1 ] ⊗R −. More precisely, there exists a unique morphism
ϕM : R[U −1 ] ⊗R M → M [U −1 ] such that
r rm
⊗ m 7→ .
u u
Proof. Since the map R[U −1 ]×M → M [U −1 ], ( ur , m) 7→ rm u
is R-bilinear, there exists a unique
morphism ϕM sending ur ⊗ m to rm u
. This morphism
P ri is clearly surjective.
We show that ϕM is also injective. For this let i ui ⊗ mi ∈ R[U −1 ] ⊗R M be an element. Set
u := i ui ∈ U , vi := j6=i uj . Then i urii ⊗mi = i riuvi ⊗mi = i u1 ⊗ri vi mi = u1 ⊗ i ri vi mi .
Q Q P P P P
It follows that any element of R[U −1 ] ⊗ M is of the form u1 ⊗ m for some u ∈ U and m ∈ M .
Now suppose ϕM ( u1 ⊗ m) = 0, i.e., suppose m u
= 0. Then there exists a v ∈ U such that
vm = 0. Hence
1 v 1 1
⊗m= ⊗m= ⊗ vm = ⊗ 0 = 0.
u uv uv uv
Corollary 3.48. R[U −1 ] is a flat R-module.
Proof. R[U −1 ] ⊗R − is by Proposition 3.47 isomorphic to the localization functor, and the
latter is exact by Proposition 3.45.
Proposition 3.49. Let M and N be R-modules. Then there exists a unique isomorphism of
R[U −1 ]-modules
ϕM,N : M [U −1 ] ⊗R[U −1 ] N [U −1 ] → (M ⊗R N )[U −1 ] such that
m n m⊗n
⊗ 7→ .
u v u·v
Proof. This follows from Proposition 3.47 using the canonical isomorphism of Proposi-
tion 3.27.4.
Let M, N be R-modules, ϕ : M → N an R-linear map, p ∈ Spec R, and U = R \ p. Then
we write
Mp := M [U −1 ] and ϕp := ϕ[U −1 ].
If N is another R-module, then the previous proposition reads
Mp ⊗R Np ∼= (M ⊗R N )p
p
as Rp -modules.
70 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
7 Local properties
We now study local properties of R-modules, i.e., a property that an R-module M has iff all
its localizations at prime ideals has the same property. This is part of the so-called “local-
global-principle”.
The first local property is “being zero”:
Proposition 3.50. Let M be an R-module. Then the following are equivalent:
(1) M = 0;
(2) Mp for all p ∈ Spec R;
(3) Mm for all m ∈ Max R ⊂ Spec R.
Proof. The implications 1 =⇒ 2 =⇒ 3 are obvious. We prove the remaining direction by
contraposition:
Suppose M 6= 0 and m ∈ M \ {0}. Then the annihilator AnnR (m) is a proper ideal of R
(1 · m = m 6= 0). Thus, there is an m ∈ Max R containing Ann(m). But then the element
m
1
∈ Mm is nonzero, for otherwise there would exist some u ∈ R \ m such that um = 0. This
u would then be contained in AnnR (m) ⊂ m. A contradiction. Hence, Mm 6= 0 and we are
done.
The next local property is “being a mono”:
Proposition 3.51. Let M, N be R-modules and ϕ : M → N an R-linear map. Then the
following are equivalent:
(1) ϕ is a mono;
(2) ϕp is a mono for all p ∈ Spec R;
(3) ϕm is a mono for all m ∈ Max R.
Proof.
ϕ
1 ⇒ 2: That ϕ is mono means that 0 → M − → N is an exact sequence. By the exactness of the
ϕp
localization (Proposition 3.45) the sequence 0 → Mp −→ Np is exact for all p ∈ Spec R.
2 ⇒ 3: Trivial.
3 ⇒ 1: Apply the exactness of the localization to the exact sequence
ϕ
0 → ker ϕ → M −
→N
8 Chain conditions
We have already encountered an instance of Hilbert’s basis theorem for polynomial rings
over a field as Corollary 1.25.
Theorem 3.54 (Hilbert’s basis theorem). If R is a Noetherian ring, then the polynomial
ring R[x1 , . . . , xn ] is Noetherian as well. In particular, the polynomial rings Z[x1 , . . . , xn ] and
K[x1 , . . . , xn ] (where K is a field) are Noetherian rings.
Proof. Assume I R[x] is not finitely generated. Then, inductively, we can choose elements:
f1 ∈ I \ {0}, f2 ∈ I \ hf1 i, f3 ∈ I \ hf1 , f2 i, ... of minimal possible degree. For all i, set
di := deg fi and write fi = ai xdi + lower order terms in x. Then d1 ≤ d2 ≤ · · · and ha1 i ⊂
ha1 , a2 i ⊂ · · · is an ascending chain of ideals in R which eventually becomes stationary by
assumption on R, i.e., there exists a k ∈ N such that
Since fk+1 ∈ I \ hf1 , . . . , fk i, also g ∈ I \ hf1 , . . . , fk i. This contradicts the choice of fk+1 since
deg(g) < dk+1 := deg(fk+1 ).
Proposition 3.55. Let R be a Noetherian ring.
(1) If I R is an ideal, then R/I is Noetherian.
(2) The extension to R[U −1 ] along the localization morphism R → R[U −1 ] is an order
preserving embedding by Theorem 3.38.2.
Proof. Trivial.
p0 ( p1 ( p2 ( · · · pm .
(3) R is Artinian.
If these conditions are satisfied, then R has only finitely many maximal ideals.
End
lect. 21
18.07 Proof.
1 ⇒ 2:
Step 1: Suppose that R is Noetherian but not of finite length. Then the set
·f
0 → R/(p : f ) −
→ R/p → R/(p + hf i) → 0.
2 ⇒ 3: We already know that finite length modules are Noetherian and Artinian.
Step 1: We first show that dim R = 0: Consider prime ideals p1 ⊂ p2 ⊂ R and let f ∈
p2 /p1 ( R/p1 . Since the residue class ring R/p1 is also Artinian, there exists
m ≥ 1 such that hf m i = hf m+1 i. Then f m = gf m+1 for some g ∈ R/p1 . That
is, (1 − gf )f m = 0. Since R/p1 is an integral domain and f ∈ p2 /p1 ( R/p1 is
not a unit, we must have f = 0. Hence p1 = p2 . This proves that dim R = 0, or
equivalently Spec R = Max R.
Step 2: Since R is Artinian, it has only finitely many maximal ideals, as any infinite col-
lection m1 , m2 , . . . of pairwise distinct maximal ideals would yield an infinite de-
scending chain of ideals m1 ⊃ m1 ∩ m2 ⊃ · · · ⊃ m1 ∩ . . . ∩ mk ⊃ · · · , with strict
inclusions. Writing m1 , . . . , ms for the distinct ideals of Max R we have
s
\
p
h0i = N (R) = mi (*)
i=1
Then I 2 = I. We will use this to show that I = h0i. Suppose the contrary. Then
the set
Γ := {J R | J · I 6= h0i}
contains I since I · I = I 6= h0i by the assumption. In particular, Γ is nonempty.
Hence, since R is Artinian, Γ contains a minimal element J0 . Now pick f ∈ J0
such that f · I 6= h0i. Then hf i = J0 by the minimality of J0 . The same argument
gives f I = J0 = hf i since (f I) · I = f I 2 = f I 6= h0i. Now pick g ∈ I such that
f · g = f . Then
f = f g = f g2 = · · · = f gm = 0
for some m ≥ 0 since each element of I is nilpotent by (*). This contradicts our
choice of f . Hence, I = h0i.
Step 4: Each factor in the chain
s
Y
R ⊃ m1 ⊃ · · · ⊃ mN
1 ⊃ mN
1 m2 ⊃ ··· ⊃ mN
i = h0i (**)
i=1
Theorem 3.62 (Structure Theorem for Artinian rings). Every Artinian ring R is a finite
product ofQ
local Artinian rings. More precisely, if m1 , . . . , ms are the distinct maximal ideals of R,
then R ∼ s
= i=1 Rmi .
74 CHAPTER 3. LOCALIZATION AND TENSOR PRODUCTS
is an isomorphism. Note that R/mN i is a local Artinian ring (whose maximal ideal is the
image of mi in R/mN
i ). Moreover, by localizing ϕ we get
Rmi ∼
= (R/mN ∼ N
i )mi = R/mi
since (R/mN
j )mi = 0 for j 6= i.
End
lect. 22
20.07
Chapter 4
Primary decomposition
with a unit u ∈ Z× = {±1}, pi prime numbers, and mi > 0 for i = 1, . . . , k. This implies
mk
hzi = hpm
1 i ∩ . . . ∩ hpk i.
1
We call such an intersection the primary decomposition of the ideal hzi. Since Z is a princi-
pal ideal domain any primary decomposition is already of this form.
In the attempt to prove Fermat’s Last Theorem mathematicians discovered subrings R
of the complex numbers which are, unlike Z, not unique factorization domains. We call
them nowadays Dedekind domains. In these rings the notion of a primary decomposition
of a ring element fails. For example
√ √ √
Z[ −5] 3 6 = 2 · 3 = (1 + −5)(1 − −5).
√ √ √
The
√ numbers 2, 3, 1 + −5, 1 − −5 are irreducible in Z[ −5] but obviously not prime in
Z[ −5]. In order to rectify this failure Dedekind introduced ideals which for this class of
rings still admit a primary decomposition. This is even the historical reason for the name
“ideal”: unlike numbers in R, ideals behave as “ideal numbers” as they admit a primary
decomposition:
√ D √ E D √ E D √ E D √ E
Z[ −5] h6i = 2, 1 + 5 ∩ 3, 1 + 5 ∩ 2, 1 − 5 ∩ 3, 1 − 5 .
| {z } | {z }
√ √
=h1+ −5i =h1− −5i
75
76 CHAPTER 4. PRIMARY DECOMPOSITION
√ √
(3) If I R is any ideal such that I is a maximal ideal, then I is I-primary. In particular, the
powers of a maximal ideal m R are m-primary.
Proof.
√
(1) By Proposition 3.10 it suffices to show that p := q is prime. Let f, g ∈ R such that
√
f g ∈ q. Then there exists some m ≥ 1 with f m g m = (f g)m ∈ q. Hence, either f m ∈ q
√ √ √
or g mk ∈ q for some k ≥ 1. That is, either f ∈ q or g ∈ q, so q is prime.
v
u r r
√ u\ \ √
q= t qi = qi = p.
i=1 i=1
of that ring. In particular, R/I is a local ring with maximal ideal m/I. Then each
element of R/I is either a unit or nilpotent, i.e., every zero-divisor in R/I is nilpotent.
In other words, I is m-primary.
Example 4.3.
(1) The primary ideals in the ring Z of integers are h0i and ideals of type hpn i, where p is
a prime number and n ≥ 1. Indeed they are the only ideals of Z whose radical is a
prime ideal and they are clearly primary.
√
(2) Let R = k[x, y] where k is a field and q = hx, y 2 i. Then p := q = hx, yi is maximal and
R/q ∼= k[y]/hy 2 i. In R/q the zero-divisors are the multiples of y and hence nilpotent.
Thus q is p-primary. We have p2 q p, so a primary ideal is not necessarily a prime
power!
(3) In contrast to the case of maximal ideals, a prime power is not necessarily prime: Consider
R = k[x, y, z]/hxy − z 2 i, where k is a field. Write x, y, z for the residue classes of
x, y, z. Then p := hx, zi R is prime
√ since R/p ∼= k[y] is an integral domain. We have
2 2 2
x · y = z ∈ p , but x ∈ / p and y ∈/ p = p. Hence, p2 is not primary.
2
Suppose Γ 6= ∅. Then Γ has a maximal element I0 . In √particular, I0 is not primary. That is,
there exists f, g ∈ R with f g ∈ I0 but f ∈/ I0 and g ∈ / I0 . The ascending chain condition in
N OETHERian rings applied to
I0 : g ⊂ I0 : g 2 ⊂ . . .
shows that I0 : g m = I0 : g m+1 for some m ≥ 1. Then
I0 = (I0 : g m ) ∩ hI0 , g m i.
are minimal primary decompositions. In fact, by Proposition 4.2.(3) both ideals hx, y 2 i and
hx2 , xy, y 2 i are hx, yi-primary. Moreover, hx, yi ⊃ hyi.
End
Nevertheless, there are certain “parts” of PDs that are indeed uniquely determined by lect. 23
I. We will study them in the next section. 18.10
2 Uniqueness results
Lemma 4.7. Let R be a ring and f ∈ R. Then:
(1) If I R is an ideal with f ∈ I then I : f = R.
Let now q R be a primary ideal with radical p ∈ Spec R. Then:
(2) If f ∈
/ p then q : f = q.
(3) If f ∈
/ q then q : f is p-primary.
Proof.
(1) & (2) follow immediately from the definitions.
TheoremTt 4.8 (1st Uniqueness Theorem). Let R be a ring, I R an ideal admitting a PD,
√
and I = i=1 qi a minimal PD with pi := qi ∈ Spec √ R for all i = 1, . . . , t. Then the pi ’s are
precisely the prime ideals occurring in the set of ideals I : f where f ∈ R.
Hence, the pi ’s are independent of the choice of the PD.
Tt
Proof. Let f ∈ R. We already know that I : f = i=1 (qi : f ), hence
t
4.7
p \ p \
I:f = qi : f = pj .
i=1 f ∈q
/ j
√ √
√ I : f is prime, we conclude by Exercise 3.3 that pj ⊂ IT: f for some j and thus pj =
If
I : f . Conversely, for all i = 1, . . . , t there exists some fi ∈ j6=i qj such that fi ∈
/ qi since
√
the PD is minimal. Again by Lemma 4.7 we get I : fi = pi .
In short, the First Uniqueness Theorem states that even if the primary ideals of a PD are
not uniquely determined, their radicals are. Each ideal I thus has its “own” set of prime
ideals occurring as radicals in its primary decomposition. We should give them a name in
order to address them in the future.
Definition 4.9. Let R be a ring and I R an ideal. We call
p
Ass(I) := {p ∈ Spec R | ∃f ∈ R : I : f = p}
the set of minimal associated primes or isolated primes of I. Furthermore, each prime in
Ass(I) − minAss(I) is called an embedded prime of I.
The names “isolated” and “embedded” are of geometric origin.
Proposition 4.10. Let R be a ring and I R an ideal admitting a PD. Then each prime p ⊃ I
contains a minimal associated prime of I. Thus, the primes in minAss(I) are precisely the primes
containing I which are minimal w.r.t. the inclusion. In particular,
(2) and each N OETHERian ring has only finitely many minimal primes.
Tt
Proof. If p ⊃ I = i=1 qi then
t t
√ \√ \
p= p⊃ qi = pi .
i=1 i=1
Proof. We have
t s
−1 3.46 −1 4.12
\ \
I[U ] = qi [U ] = qi [U −1 ]
i=1 i=1
Definition 4.14. Let R be a ring, I R a proper ideal, and Σ ⊂ Ass(I). We call Σ isolated
if it is inclusion-closed, i.e., p0 ∈ Ass(I) and p0 ⊂ p for some p ∈ Σ implies p0 ∈ Σ.
Theorem 4.15 (2nd Uniqueness Tt Theorem). Let R be a ring and I R an ideal admitting √
a
primary decomposition. Let I = i=1 qi be a minimal primary decomposition T with pi := qi for
all i = 1, . . . , t. If Σ ⊂ Ass(I) is an isolated set of primes of I, then pi ∈Σ qi is independent of the
chosen primary decomposition. In particular, the primary components corresponding to primes in
minAss(I) are uniquely determined by I.
S T
Proof. Set UΣ := R − pi ∈Σ pi = pi ∈Σ (R − pi ). Then UΣ is a multiplicatively closed subset
of R. Furthermore, we have for all pj :
[
pj ∩ UΣ = ∅ ⇐⇒ pj ⊂ pi
pi ∈Σ
3.8
⇐⇒ pj ⊂ pi for some i.
Definition 4.16. Let I R be a proper ideal of the ring R. Then we call each primary
ideal in a primary decomposition of I corresponding to a prime in minAss(I) and isolated
primary component. Each other primary ideal occurring in any minimal primary decom-
position of I is called an embedded primary component.
Chapter 5
1 Basic definitions
We start by recalling basic results about integral ring extensions from the Algebra lecture.
Definition 5.1. Let S be a ring (not necessarily commutative), R a subring in the center
Z(S) and I R.
X d + r1 X d−1 + · · · + rd ∈ R[X],
More generally:
X d + r1 X d−1 + · · · + rd ∈ R[X]
• S is called integral over R or the ring extension R/S is integral, if each element of S
is integral over R.
ϕ : RhX1 , . . . , Xn i → S, Xi 7→ si
with ϕR = idR .
81
82 CHAPTER 5. INTEGRAL RING EXTENSIONS
p
(2) R[s] is finite over R (and s ∈ I · R[s]);
√
(3) R[s] is contained in a commutative subring S 0 ≤ S which is finite over R (and s ∈ I · S 0 );
is a subring of S.
and
R := k[y] ≤ k[x, y]/hxyi =: S 0
are not integral: In both cases S = R[x] is not finite over R, i.e., x is not integral over R.
End
lect. 25
08.11
2 Lying over and going up
From now on S we will be commutative ring and S/R a ring extension.
Notation. If P ∈ Spec S then p := P ∩ R = Pc ∈ Spec R. We say that P lies over p.
Define (
Spec S → Spec R,
ι# :
P 7→ p := Pc = ι−1 (P) = P ∩ R,
associated to the ring embedding ι : R ,→ S.
Theorem 5.9 (Lying over). Let S/R be an integral (commutative) ring extension and let
p ∈ Spec R. Then:
(1) There exists a P ∈ Spec S lying over p (and containing pS), i.e., ι# is surjective.
2. LYING OVER AND GOING UP 83
(2) There are no strict inclusions between prime ideals of S lying over p.
(3) If P ∈ Spec S lies over p, then P is maximal iff p is maximal, i.e., ι# (Max(S)) = Max(R).
(4) If S is N OETHERian, then only finitely many prime ideals of S lie over p, i.e., the fibers of ι#
are finite.
Γ := {J R | J ⊃ I, J ∩ U = ∅}.
Then Γ is partially orderedS by inclusion, Γ 6= ∅ since, e.g., I ∈ Γ and if {Jλ }λ∈Λ is a totally
ordered subset of Γ then λ∈Λ Iλ ∈ Γ is an upper bound. So we can apply Z ORN’s Lemma
and there exists a maximal element of p ∈ Γ. We shall now prove that p ∈ Spec R:
First, p is a proper ideal of R, since otherwise 1 ∈ p ∩ U = ∅ , a contradiction. Second, let
r1 , r2 ∈ R − p. Then for j = 1, 2 the ideal p + hri i ∈ / Γ by the maximality of p. Hence for
j = 1, 2 we have (p + hri i) ∩ U 6= ∅. That is, we can find elements pj ∈ p and aj ∈ R such
that pj + aj rj ∈ U . Then (p1 + a1 r1 )(p2 + a2 r2 ) ∈ U ⊂ R − p, so a1 a2 r1 r2 ∈
/ p. In particular,
r1 r2 ∈/ p, so p is prime.
We still need one more remark:
Remark 5.11. Let S/R be an integral ring extension.
(1) Let J S be an ideal. Regard R/(J ∩ R) as a subring of S/J in the obvious way. Then
the ring extension (S/J)/(R/(J ∩ R)) is also integral.
Proof.
(1) Let s = s + J ∈ S/J. An integral equation for s over R/(J ∩ R) is obtained from one
for s over R by taking residue classes of coefficients.
Pm
(2) An s ∈ IS can be written as s = i=1 ai si with ai ∈ I and si ∈ S for all i, i.e.,
s ∈ IR[s1 , . . . , sm ] and we are done by Theorem 5.3.
(1) Consider the extended ideal pe = pS and the multiplicatively closed subset U :=
R − p ⊂ S. We will show that (pS) ∩ U = ∅ using the integrality of S/R. Then we will
apply K RULL’s prime existence lemma.
If s ∈ pS, then s is integral over p by Remark 5.11.(2). Hence, we have an integral
equation sd + a1 sd−1 + · · · + ad = 0 with all ai ∈ p. Now we show that s ∈ / U . Suppose
d d−1
the contrary. Then, in particular, s ∈ R and s = −a1 s − · · · − ad ∈ p, so s ∈ p. This
contradicts s ∈ U = R − p. So (pS) ∩ U = ∅. The prime existence lemma now yields a
P ∈ Spec S such that p ⊂ pS ⊂ P and P ∩ R ⊂ R − U = p. Hence, P is a prime ideal
lying over p.
84 CHAPTER 5. INTEGRAL RING EXTENSIONS
(2) If P1 ⊂ P2 are two prime ideals in S lying over p, then R := R/p ⊂ S/P1 =: S is
an integral extension of integrals domains by Remark 5.11.(1) such that (P2 /P1 ) ∩
R = h0i. Suppose P1 ( P2 . Then there exists a nonzero element s ∈ P2 /P1 and
we obtain a contradiction considering an integral equation for s over R of smallest
possible degree d:
sd + a1 sd−1 + · · · + ad = 0.
Indeed, since ad ∈ (P2 /P1 ) ∩ R is zero, and S is an integral domain, we may divide
the equation by s to get an integral equation of smaller degree.
(3) If p is maximal, then P is maximal by (2). For the converse we consider the integral
extension R/p ⊂ S/P. If S/P is a field, its only maximal ideal is h0i. Then, in turn, h0i
is the only maximal ideal of R/p by (1), i.e., R/p is also a field.
(4) If P ∈ Spec S lying over p, then pS ⊂ P and by (2) the ideal P is a minimal prime
containing pS. Since S is N OETHERian by assumption, Proposition 4.10 says that P is
a minimal associated prime of pS and there exists only finitely many such primes.
Example 5.12. The ring extension
√
R := Z ⊂ Z[ −5] ∼
= Z[x]/hx2 + 5i =: S
3 Going down
“Going up” was a relatively easy result once we had “lying over”. To prove a dual “going
down” we need more assumptions, as we will show by examples.
We recall the following definition from Algebra.
Definition 5.14. Let R be an integral domain. The integral closure
R := IntR (Frac R) = {s ∈ Frac R | s integral over R}
of R in its field of fractions Frac R is called the normalization of R. We call R normal if
R = R.
3. GOING DOWN 85
Example 5.15.
(1) If R is a unique factorization domain, then R is normal.
(2) In particular, k[x1 , . . . , xn ] is normal, where k is a field or k = Z.
(3) Q[x, y, z]/hx2 − y 2 zi is not normal. Geometrically, this is the so-called W HITNEY um-
brella:
Proof. Exercise.
End
Theorem 5.16 (Going-Down-Theorem of C OHEN and S EIDENBERG). Let S/R be an inte- lect. 26
gral ring extension and assume furthermore that R and S are both integral domains with R normal. 15.11
If p1 ⊂ p2 are prime ideals in R and P2 ∈ Spec S lying over p2 , then there exists P1 ∈ Spec S lying
over p1 , such that P1 ⊂ P2 .
For the proof we need a lemma:
Lemma 5.17. Let R be a normal ring, K := Frac R (its field of fractions), and K ⊂ L a field
extension. Let p ∈ Spec R and s ∈ L be integral over p. Then s is algebraic over K and all
coefficients of its minimal polynomial over K lie in p.
Proof. Since s is integral over p, it is algebraic over K. Let ps = xd +c1 xd−1 +· · ·+cd ∈ K[x] be
the minimal polynomial of s over K and s1 = s, s2 , . . . , sd be the roots of ps in the algebraic
closure K. Then, for each j there exists a ϕj ∈ Gal(K/K) with ϕj (s) = sj . Thus if f (s) = 0
is an integral equation for s with coefficients in p, then also f (sj ) = 0 for all j, so the sj are
all integral over p. By Corollary 5.6 the same is true for the coefficients ci of ps since these
are polynomial expressions in the roots sj (by V IETA’s formulas). Now since R = R and
√
p = p all ci lie in p.
Proof of Theorem 5.16. Consider three multiplicatively closed subsets of S:
U1 := R − p1 ,
U2 := S − P2 ,
U := U1 U2 = {r · s | r ∈ U1 , s ∈ U2 } ⊃ U1 , U2 .
In Step 1 below, we will show that pe1 ∩ U = (p1 S) ∩ U = ∅. Then we will apply K RULL’s
prime existence Lemma 5.10.
Step 1: Suppose there exists an s ∈ (p1 S) ∩ U . Then since s ∈ p1 S, it is integral over p1
by Remark 5.11.(2). Applying Lemma 5.17 we see that the minimal polynomial of
s ∈ L := Frac S over K := Frac R is of type ps = xd + c1 xd−1 + · · · + cd , with all ci ∈ p1 .
Since s ∈ U we can write s = r · se, where r ∈ U1 and se ∈ U2 . Then
c1 cd
pse = xd + xd−1 + · · · + d
r r
86 CHAPTER 5. INTEGRAL RING EXTENSIONS
is the minimal polynomial of se over K. Again applying Lemma 5.17 we see that the
coefficients ci /ri of pse lie in R, since se is integral over R by assumption. In fact, ci /ri
i
√p1 since ci ∈ p1 and r ∈
lie in / p1 for all i. It follows that se is even integral over p1 . So
se ∈ p1 S ⊂ P2 by K RONECKER’s Theorem 5.3. This contradicts se ∈ U2 .
Step 2: K RULL’s prime existence Lemma 5.10 yields P1 ∈ Spec S such that p1 S ⊂ P1 and
P1 ∩ U = ∅. In particular, P1 ∩ U1 = ∅, so that P1 ∩ R = p1 , i.e., P1 is lying over p1 .
Also P1 ∩ U2 = ∅, hence P1 ⊂ P2 .
To indicate that the additional assumptions in the Going-Down-Theorem are needed,
we will give an example. To prepare for it, we state a couple of remarks.
Remark 5.18. Let k be a field. If p = (a1 , . . . , an ) ∈ k n is a point, each polynomial
f ∈ k[x1 , . . . , xn ] can be written as a polynomial in the (xi − ai )’s:
f = f (p) + terms of degree at least 1 in (xi − ai ). (*)
Indeed, to obtain the Taylor expression of f at p, substitute the (xi − ai ) + ai for xi in f and
expand. It follows from (*) that mp := hx1 − a1 , . . . , xn − an i is the kernel of the evaluation
map at p
εp : k[x1 , . . . xn ] → k, f 7→ f (p).
Hence k[x1 , . . . , xn ]/mp ∼
= k is a field, so mp is a maximal ideal.
If k is not algebraically closed, then not every maximal ideal in k[x1 , . . . , xn ] is of this
form (for n ≥ 1). Consider for example the maximal ideal hx2 + 1i R[x].
We will come back to this later in the context of H ILBERT’s Nullstellensatz.
Examples 5.8 and 5.15.(3) involved rings of the following type:
Definition 5.19. Let k be a field. A finitely generated k-algebra is necessarily a residue
class rings of type k[x1 , . . . , xn ]/I (where I is an ideal). We call such an algebra an affine
ring or affine algebra. If I is a prime ideal then we talk about an affine domain.
The reason for this naming convention is that the canonical ring epimorphism
π : k[x1 , . . . , xn ] → k[x1 , . . . , xn ]/I
induces an embedding
π # : Spec(k[x1 , . . . , xn ]/I) → Spec k[x1 , . . . , xn ]
where we identify Spec k[x1 , . . . , xn ] with the n-dimensional affine space Ank over k.
Remark 5.20. Every extension R ⊂ S of affine rings is obviously of finite type. Hence, in
this case, S/R is integral iff S/R is finite. And finiteness can be computationally checked us-
ing Gröbner bases: Let R be the k-subalgebra generated by p1 , . . . , pm ∈ S := k[x1 , . . . , xn ]/I.
This defines a ring homomorphism
(
k[y1 , . . . , ym ] → k[x1 , . . . , xn ]/I,
ϕ:
yi 7→ pi .
Consider the ideal
J = hI, p1 − y1 , . . . , pm − ym i k[x1 , . . . , xn , y1 , . . . , ym ]
corresponding to the graph of the morphism ϕ. Then ker ϕ = J ∩ k[y1 , . . . , ym ] and R =
k[y1 , . . . , ym ]/ ker ϕ. Let {g1 , . . . , g` } be a Gröbner basis of J w.r.t. a block order to eliminate
x1 , . . . , xn (cf. Proposition 1.55). Recall that ker ϕ is generated by all gk ’s not involving any
xi . One can show that R ,→ S is finite iff for each i = 1, . . . , n there exists a Gröbner basis
element gk whose leading monomial is of type xαi i for some αi ≥ 1.
3. GOING DOWN 87
the converse inclusion let us ∈ S[U −1 ] be integral over R[U −1 ]. Then we have an integral
equation
s d r s d−1
rd
1
+ + ··· + =0 (*)
u u1 u ud
with all ri ∈ R and ui ∈ U . Set v = u1 · · · ud . Multiplying (*) by (u · v)d we get an integral
equation for sv over R. Hence sv ∈ Re and therefore s = sv ∈ R[U e −1 ].
u uv
88 CHAPTER 5. INTEGRAL RING EXTENSIONS
(1) R is normal.
89
90CHAPTER 6. KRULL DIMENSION, NOETHER NORMALIZATION, HILBERT’S NULLSTELLEN
2 Noether Normalization
Lemma 6.5. Let k be an infinite filed and f ∈ k[x1 , . . . , xn ] − {0}. Then there exists a point
p = (a1 , . . . , an ) ∈ k n such that f (p) 6= 0.
Proof. We use an induction over n. For n = 1 the statement is clear, since each non-zero
polynomial in one variable has at most finitely many roots. If n > 1 we write f in the form
f = c0 (x2 , . . . , xn )xe1 + · · · + ce (x2 , . . . , xn ). Then ci 6= 0 for at least one i and by induction
there exists a point p0 ∈ k n−1 such that ci (p0 ) 6= 0. Then 0 6= f (x1 , p0 ) ∈ k[x1 ], hence there
exists an element a such that f (a, p0 ) 6= 0.
For the main theorem we still need a lemma for which we need the following notation:
Remark 6.6. Let R be a ring. A polynomial in R[x1 , . . . , xn ] is called homogeneous (of
degree e) if all its monomials have degree e or if the polynomial is zero.
Every nonzero polynomial f ∈ R[x1 , . . . , xn ] of degree e can be uniquely written as a
sum f = fe + fe−1 + · · · + f0 , where fi are homogeneous of degree i, and where fe 6= 0. The
fi ’s are called the homogeneous components of f .
Given an extra x0 , the polynomial
h e x1 xn
f := x0 f ,..., ∈ R[x0 , x1 , . . . , xn ]
x0 x0
f h (1, x1 , . . . , xn ) = f and F = xm h
0 F (1, x1 , . . . , xn ) ,
k[x1 , x2 , . . . , xn ] → k[x1 , x
e2 , . . . , x
en ]
x1 7→ x1 , xi 7→ x ei + ai x1 i = 2, . . . , n
axe1 + c1 (e en )xe−1
x2 , . . . , x 1 + · · · + ce (e
x2 , . . . , x
en ),
k[x1 , x2 , . . . , xn ] → k[x1 , x
e2 , . . . , x
en ]
i−1
ei + x1r
x1 7→ x1 , xi 7→ x i = 2, . . . , n,
Proof.
2
See the proof for the construction of such elements.
3
See the proof for the construction of such a power.
2. NOETHER NORMALIZATION 91
f e = xm
1 fe (1, x2 , . . . , xn )
h
i−1
and let r ∈ N. After substituting xi 7→ x ei + xr1 for i = 2, . . . , n the terms depending
α1 +α2 r+···+αn rn−1
only on x1 are of type aα1 ,...,αn x1 . If r is strictly larger than all αi ’s ap-
pearing in a term of f , the numbers α1 + α2 r + · · · + αn rn−1 x are distinct for distinct
α1 , . . . , αn (the αi ’s are then the digits in the base-r expansion of the exponent). Hence,
the terms depending only on x1 cannot cancel with each other, and we are done.
k[x1 , . . . , xn ] → S, xi 7→ si
ϕ : k[x1 , . . . , xn ] → k[x1 , x
e2 , . . . , x
en ]
w.r.t. to this f as in Lemma 6.7 (we may also assume without loss of generality that ϕ(f )
is monic in x1 ). Setting Ie = ϕ(I) and Ie1 = Ie ∩ k[e en ], we then get an integral ring
x2 , . . . , x
extension
Se1 := k[e
x2 , . . . , x
en ]/Ie1 ,→ k[x1 , x en ]/Ie = Se1 [x1 ] =: Se ∼
e2 , . . . , x = S.
By induction, there exists a N OETHER normalization k[y1 , . . . , yd ] ⊂ Se1 with d ≤ n − 1. The
result follows from the transitivity of finite/integral ring extensions.
92CHAPTER 6. KRULL DIMENSION, NOETHER NORMALIZATION, HILBERT’S NULLSTELLEN
trdegK K(x1 , . . . , xn ) = n.
J ∩ k[y1 , . . . , yd ] = hy1 , . . . , yc i.
Step 1. We show that m ≤ n. Applying Theorem 6.12 and the transcendence degree argument
from its proof to P1 , we get a N OETHER normalization k[y1 , . . . , yn ] ⊂ k[x1 , . . . , xn ]
such that P1 ∩ k[y1 , . . . , yn ] = hy1 , . . . , yc i for some c ≤ n. Set pi := Pi ∩ k[y1 , . . . , yn ] for
all i. Then
p1 /p1 ( p2 /p1 ( . . . ( pm /p1
is a chain of prime ideals in the ring k[y1 , . . . , yn ]/p1 ∼ = k[yc+1 , . . . , yn ]. By the induction
hypothesis, m − 1 ≤ n − c. Hence m ≤ n since c ≥ 1 (otherwise, both P0 and P1
would lie over the zero ideal of k[y1 , . . . , yn ], a contradiction to Theorem 5.9.(2)). This
finishes the first step.
Step 2. Now assume that the chain (*) is maximal. Then the induced chain
h0i = p0 ( p1 ( . . . ( pm (**)
of prime ideals in k[y1 , . . . , yn ] is a maximal as well: Suppose not, i.e., suppose there
is a q ∈ Spec k[y1 , . . . , yn ] with strict inclusions pi ( q ( pi+1 for some i. Then 0 ≤
i ≤ m − 1 since pm is maximal by Theorem 5.9.(3). Applying Theorem 6.12 to pi we
get a N OETHER normalization k[z1 , . . . , zn ] ⊂ k[y1 , . . . , yn ] such that pi ∩ k[z1 , . . . , zn ] =
hz1 , . . . , zc0 i for some c0 ≤ n. Then the composition
Step 3. The maximal chain (**) corresponds to a maximal chain of prime ideals in
• If I = p is a prime ideal, take the supremum of the lengths of all chains of primes
ideals in R with largest prime p.
• If I is arbitrary then take the minimum of the codimension of prime ideals containing
I.
Corollary 6.16. Let R be an affine domain over a field k and I R be a proper ideal. Then
Proof. This follows from the Corollary 6.14: dim I can be expressed in terms of a maximal
chain in Spec R which includes a prime ideal p ⊃ I such that codim I = codim p.
Example 6.17. Consider the affine ring
and p = hx, y, z − 1i ∈ Spec R.5 The ring R is obviously not a domain and codim p + dim p =
1 + 0 6= 2 = dim R. Observe that Spec R contains maximal chains of different lengths.
Remark 6.18. If R is a ring and p ∈ Spec R, then
dim Rp = codim p.
Proof. Indeed, by Theorem 3.38.(3) chains in Spec Rp correspond to chains in Spec R with
largest ideal p.
4 H ILBERT’s Nullstellensatz
We begin by introducing some basic notions from algebraic geometry. Throughout this
section k will denote a field.
Definition 6.19.
V (T ) := Vk (T ) := {a ∈ k n | f (a) = 0 ∀f ∈ T }
is called the vanishing locus of T . Any such locus is called an (affine) algebraic set.
The basic idea of algebraic geometry is to use the constructions V and I to set up a cor-
respondence between ideals of k[x1 , . . . , xn ] and algebraic sets in k n , expressing geometric
properties in algebraic terms and vice versa.
Remark 6.20. Let T ⊂ k[x1 , . . . , xn ], J k[x1 , . . . , xn ], and A ⊂ k n . Then
(2) V (I(A)) = A;
5
Geometrically, the set of closed points of Spec R with residue field k is the union of the (x, y)-plane and
the z-axes, and p is corresponds to the point (0, 0, 1) on the z-axes.
4. HILBERT’S NULLSTELLENSATZ 95
Equality does not hold in general for two reasons. We demonstrate them by two examples:
• Consider J := hx2 i k[x]. Then V (J) = {0} ⊂ k and I(V (J)) = hxi ) hx2 i = J.
• Consider J := hx2 + 1i Q[x]. Then V (J) = ∅ ⊂ Q and I(V (J)) = h1i ) hx2 + 1i = J.
To get rid of problems of the first type, we may restrict ourselves to radical ideals. With
respect to the second problem, algebraic geometers prefer to work over an algebraically
closed field.
H ILBERT’s Nullstellensatz tells us, that over an algebraically closed field V and I define
a 1-1 correspondence between algebraic subsets of k n and radical ideals of k[x1 , . . . , xn ]. Un-
der this correspondence, irreducible algebraic sets correspond to primes ideals and points
to maximal ideals.
Theorem 6.21. Let k be an algebraically closed field. Then then Max k[x1 , . . . , xn ] consists of
the ideals of type
ma := hx1 − a1 , . . . , xn − an i,
where a = (a1 , . . . , an ) ∈ k n .
Proof. We already know from Remark 5.18 that any such ideal is maximal. Conversely, let
m ∈ Max k[x1 , . . . , xn ] be any maximal ideal. Then L = k[x1 , . . . , xn ]/m is a field. Hence, if
R = k[y1 , . . . , yd ] ⊂ L is a N OETHER normalization, we must have R = k since
Then k ⊂ L is a finite, hence algebraic field extension. But since k is algebraically closed,
the inclusion k ⊂ L is an isomorphism. Hence, there exist a1 , . . . , an ∈ k such that ai can be
identified with the residue classes xi , i.e., ai = xi for all i = 1, . . . , n. Thus hx1 − a1 , . . . , xn −
an i ⊂ m. Since the former ideal is maximal and m is proper both ideals coincide.
This theorem states that points a ∈ k n correspond to maximal ideals ma (these are the
closed points in Spec k[x1 , . . . , xn ]).
Theorem 6.22 (H ILBERT’s Nullstellensatz, weak version). Let k be an algebraically closed
field and I k[x1 , . . . , xn ] be an ideal. Then the following are equivalent:
(1) V (I) = ∅;
(2) 1 ∈ I.
Proof.
Proof.
96CHAPTER 6. KRULL DIMENSION, NOETHER NORMALIZATION, HILBERT’S NULLSTELLEN
√
"⊃": If f ∈ J, then f m ∈ J for some m ≥ 1. This implies f m and, thus, also f vanish on
V (J).
"⊂": Let now f ∈ I(VP (J)) and J = hf1 , . . . , fr i. Then f vanishes on V (J), and we have to
show that f = ri=1 gi fi for some m ≥ 1 and some g1 , . . . , gr ∈ k[x1 , . . . , xn ] =: R. For
m
J 0 := hf1 , . . . , fr , 1 − tf i R[t],
where t is a new indeterminate. We show that V (J 0 ) ⊂ k n+1 is empty and apply the
weak version of the Nullstellensatz. Suppose the contrary, i.e., suppose there exists an
a0 = (a1 , . . . , an , an+1 ) ∈ V (J 0 ). Set a = (a1 , . . . , an ) ∈ k n . Then f1 (a) = . . . = fr (a) = 0,
so that a ∈ V (J) and an+1 f (a) = 1. This contradicts the fact that P f vanishes on V (J).
Now applying the weak Nullstellensatz we get 1 ∈ J , that is 1 = ri=1 hi fi + h(1 − tf )
0
codim p ≤ 1.
codim q = dim Rq = 0.
q(n) := {a ∈ R | ∃v ∈
/ q : va ∈ qn }, n ≥ 1.
Then, by Theorem 3.38, q(n) = (qn )ec is the preimage of (qn )e = qn Rq under the localization
morphism R → Rq . Since the maximal6 ideal p/hf i of R := R/hf i is also a minimal prime
of R by the assumption on p, the ring R is zero-dimensional. Being also N OETHERian, it is
A RTINian by Theorem 3.61. Hence, the descending chain
q(1) + hf i ⊃ q(2) + hf i ⊃ . . .
q(n) + hf i = q(n+1) + hf i.
6
recall that we are assuming (R, p) a local ring
5. KRULL’S PRINCIPAL IDEAL THEOREM AND REGULAR LOCAL RINGS 97
We may then write any g ∈ q(n) as g = h + af , with h ∈ q(n+1) and a ∈ R. Then af ∈ q(n) .
Since p is a minimal prime of hf i, we have f ∈ / q ( p and hence a ∈ q(n) , by the definition
(n)
of q . This shows that
q(n+1) + f q(n) = q(n) .
Since hf i ⊂ p and p is a maximal ideal by our above assumption, N AKAYAMA’s lemma
(Corollary 3.18) applied to R yields q(n) = q(n+1) . Then qn Rq = qn+1 Rq by Theorem 3.38.
Applying N AKAYAMA’s lemma over Rq we get qn Rq = h0i, in words, the maximal ideal
qRq of Rq is nilpotent. We conclude from Proposition 3.10 and Definition 3.11 that dim Rq =
0.
The generalization of K RULL’s Principal Ideal Theorem provides us with a lower bound
for the number of equations of an algebraic set of a given codimension.
Theorem 6.25 (Generalization of K RULL’s Principal Ideal Theorem). Let R be a N OETHER-
ian ring and I = hf1 , . . . , fc i R an ideal generated by c elements. Then each minimal prime p of
I satisfies
codim p ≤ c.
Conversely, if p ∈ Spec R is of codimension c, then there exists y1 , . . . , yc ∈ R, such that p is a
minimal prime of hy1 , . . . , yc i.
Proof. For the first statement let p be a minimal prime of I. As in the preceding proof we
may assume that (R, p) is a local ring. We do induction on c. The case c = 1 is covered by
K RULL’s Principal Ideal Theorem 6.24.
Suppose now c > 1 and that q ( p is a prime ideal of R with no other prime ideal of R
between q and p (if no such q exists, then codim I = codim p = 0 and we are done). Since
p is a minimal prime of I = hf1 , . . . , fc i, at least one of the fi ’s is not contained in q, say
fc ∈
/ q. Then the maximal ideal p/(q + hfc i) of R/(q + hfc i) is also a minimal prime of this
ring. Arguing as in the previous proof we see that R/(q + hfc i) is a local A RTINian ring.
This implies that the maximal ideal p/(q + hfc i) is nilpotent (see the proof of Theorem 3.61).
In particular, all fi ’s are nilpotent modulo q + hfc i, say
Remark 6.26. If R is a N OETHERian ring with only finitely many maximal ideals, then
according to our definitions, the K RULL dimension of R is the maximum of the codimen-
sions of its maximal ideals (all of which are finitely generated). It then follows from the
generalization of K RULL’s Principal Ideal Theorem 6.25 that dim R is finite. This applies, in
particular, to every local Noetherian ring.
We show an important corollary, whose geometric meaning lies in the algebraic charac-
terization of non-singular (i.e., smooth) points of algebraic sets.
98CHAPTER 6. KRULL DIMENSION, NOETHER NORMALIZATION, HILBERT’S NULLSTELLEN
In particular,
dimR/m m/m2 ≥ dim R.
Proof. For the first statement let d = dim R = codim m. Further let d0 be the minimum of the
right hand side. Then d ≤ d0 and d0 ≤ d follow from the first and second of statement of the
generalization of K RULL’s Principal Ideal Theorem 6.25, respectively. The second statement
follows from the first one since m is generated by dimR/m m/m2 elements by Corollary 3.19.
The inequality in Corollary 3.19 may be strict. The special case of equality deserves an
own name.
Definition 6.28. A local N OETHERian ring (R, m) is called regular, if
r ∈ p and p1 ( p ( p3 .
Proof. Since h0i = p1 /p1 ( p2 /p1 ( p3 /p1 is a chain in Spec R/p1 , we have codim p3 /p1 ≥ 2.
Then, by K RULL’s Principal Ideal Theorem 6.24, p3 /p1 is not a minimal prime of of hri
R/p1 . Hence, there exists some p ∈ Spec R such that r ∈ p/p1 ( p3 /p1 . Now take preimages.
Proof.
(1) Choose a maximal chain p0 ( p1 ( . . . ( pd = m in Spec R with d = dim R. Then
r ∈ pi for some minimal i. Because of Corollary 6.29 we may assume that i ≤ 1. Then
p1 /hri ( . . . ( pd /hri is a chain in Spec R/hri. Hence dim R/hri ≥ d − 1 = dim R − 1.
End
lect. 31 Exercise 6.1. Prove the following general version of prime avoidance: Let p1 , . . . , pk−2 ∈
20.12 Spec R and pk−1 , pk , I R. Then
k
[
I⊂ pi =⇒ I ⊂ pi for some i.
i=1
dim R ≥ dim R − 1 = d − 1.
by Corollary 6.30. On the other hand, by Corollary 3.19 and since R is regular of dimension
d, we can find m2 , . . . , md ∈ m such that m1 := r, m2 , . . . , md generate m. Hence m can be
generated by d − 1 elements. Thus
These two classes of rings play an important role in algebraic geometry and number theory.
1 Valuation rings
Definition 7.1. An integral domain R is called a valuation ring (VR) if the following
condition holds:
1
0 6= a ∈ Frac R =⇒ a ∈ R ∨ ∈ R.
a
A valuation ring R is called a discrete valuation ring (DVR) if R is N OETHERian, but not a
field.
Note that each field is a valuation ring. We will justify the name “valuation ring” later
on.
Proposition 7.2 (First properties of valuation rings). Let R be a valuation ring. Then
(3) R is normal.
I, J ∈ Γ =⇒ I ⊂ J or J ⊂ I.
Proof.
101
102 CHAPTER 7. VALUATION RINGS AND DEDEKIND DOMAINS
d i−1
X 1
a=− ri · ∈ R.
i=1
a
(4) Exercise.
(5) By (4) there exists some i such that haj i ⊂ hai i for all j = 1, . . . , r. Hence I = hai i.
Moreover, if R is a DVR, then R is N OETHERian and, thus, a principal ideal domain.
Since R is not a field, we have dim R = 1 by Example 6.2.
Now we start justifying the names “valuation” and “discrete valuation” ring.
Definition 7.3. A totally ordered group is an additive A BELian group (G, +) together
with a total order ≤ on G, which is compatible with addition: For all g, g 0 , h ∈ G it holds: if
g ≤ g 0 , then g + h ≤ g 0 + h.
Note that no nonzero element of such a group can have finite order n, since otherwise
0G < g < 2g < · · · < n · g = 0G .
Example 7.4. (R, +, ≤) where ≤ is the usual order, is a totally ordered group and so is
every subgroup. In particular, (Z, +, ≤) is an ordered group.
Definition 7.5. Let K be a field and (G, +, ≤) is a totally ordered group.
1. VALUATION RINGS 103
ν : (K ∗ , ·) → (G, +),
such that ν(a + b) ≥ min(ν(a), ν(b)) for all a, b ∈ K ∗ where we set ν(0) := ∞.
• If (G, +, ≤) ∼
= (Z, +, ≤) and ν is an epimorphism, then ν is called a discrete valuation.
Proposition 7.6. In the situation above, the subset Rν := {a ∈ K | ν(a) ≥ 0} ⊂ K is a
subring of K, called the valuation ring of K w.r.t. ν. We have K = Frac Rν .
Proof.
(1) Rν is a subring: If a, b ∈ Rν , then ν(a + b) ≥ min(ν(a), ν(b)) ≥ 0, so a + b ∈ Rν , and
ν(a · b) = ν(a) + ν(b) ≥ 0, so a · b ∈ Rν . Furthermore, ν(1) = ν(1 · 1) = 2 · ν(1), hence
ν(1) = 0 and 1 ∈ Rν . Also 0 = ν(1) = ν(−1 · −1) = 2 · ν(−1) and −1 ∈ Rν . So also
−a = (−1) · a ∈ Rν .
(2) Clearly Frac Rν ⊂ K. For the other inclusion let a ∈ K − Rν . Then ν( a1 ) = −ν(a) > 0.
Hence, a1 ∈ Rν and a = 11 ∈ Frac Rν .
a
Rν := {a ∈ K | ν(a) ≥ 1}
= {a ∈ K | a ≥ 1}
n a o
= a ∈ K | a = ∈ R = R.
1
(2) (⇒) Since R is a DVR it is a local and a PID by Proposition 7.2. Hence R has a unique
maximal ideal m = hpi for some p ∈ R. Then R = Rhpi and it follows from
Lemma 7.8 below that R = Rν for some discrete valuation ν.
104 CHAPTER 7. VALUATION RINGS AND DEDEKIND DOMAINS
(⇐) By (1) we know that Rν is a valuation ring. We now show that Rν is a PID, since
then Rν is N OETHERian and, thus, a DVR:
Let h0i 6= I Rν . Pick 0 6= f ∈ I with ν(f ) minimal. Then I = hf i, since if 0 6=
g ∈ I, we have ν(g) ≥ ν(f ) =⇒ ν( fg ) ≥ 0 =⇒ fg ∈ Rν =⇒ g = fg f ∈ hf i.
(1) Let (G, +, ≤) be a totally ordered group and let ν : R − {0} → G be a map such that
Then
a
ν : K ∗ → G, 7→ ν(a) − ν(b)
b
is a valuation of K.
(2) Suppose R is a unique factorization domain and let p ∈ R be prime. Given a ∈ R − {0}, we
write a = a0 · pna with p - a0 . Consider the map ν : R − {0} → Z, a 7→ na . Then
a
ν : K ∗ → Z, 7→ na − nb
b
is a discrete valuation on K and
na o
Rν := | na ≥ nb = Rhpi
b
is its discrete valuation ring.
Proof.
a a0
= 0 ⇐⇒ ab0 = a0 b =⇒ ν(a) + ν(b0 ) = ν(a0 ) + ν(b).
b b
Example 7.9. Consider R = Z, K = Q, and p a prime number and obtain the DVR Zhpi .
Theorem 7.10. Let R be an integral domain and I R a proper ideal. Then there exists a
valuation ring R0 such that R ⊂ R0 ⊂ Frac(R) with I · R0 ⊂ mR0 (the latter is equivalent to
I · R0 6= R0 ).
Proof. Exercise.
Corollary 7.11. Let R be an integral domain. Then the integral closure of R in Frac(R) is
\
R= R0 .
R⊂R0 ⊂Frac R
R VR
2. DEDEKIND DOMAINS 105
Proof. Exercise.
Proposition 7.12. Let R be a local N OETHERian integral domain with maximal ideal m and
dim R = 1. Then the following are equivalent:
(1) R is a DVR;
(3) m is principal;
(6) There exists an element p ∈ R such that any ideal h0i 6= I R is of the form hpn i for some
n ∈ N0 ;
(7) R is normal;
Proof. Exercise.
(1) K[x]hxi .
(2) K[[x]].
(3) the local ring of a curve over an algebraically closed field at a smooth point.
2 D EDEKIND domains
Dedekind rings provide one possible answer to the search for unique prime factorization:
They allow for a unique factorization of each nonzero ideal into a product of prime ideals.
We start with one way of defining Dedekind rings:
Definition 7.14. An integral domain R is called a D EDEKIND domain if
(1) R is N OETHERian,
Theorem 7.16. Let R be an integral domain which is not a field. Then the following are equiv-
alent:
(2) Every nonzero proper ideal of R can be written as a product of finitely many prime ideals.
106 CHAPTER 7. VALUATION RINGS AND DEDEKIND DOMAINS
(1) K = Q and OK = Z.
The following are algebraic number fields of degree 2 (quadratic number fields):
1
i.e., a finite extension of Q
Bibliography
[BLH11] Mohamed Barakat and Markus Lange-Hegermann, An axiomatic setup for algorith-
mic homological algebra and an alternative approach to localization, J. Algebra Appl. 10
(2011), no. 2, 269–293, (arXiv:1003.1943). MR 2795737 (2012f:18022) 39
[Bru] Gregor Bruns, Are all projection maps in a categorical product epic?, StackExchange,
(accessed 2013-11-18): (http://math.stackexchange.com/questions/
238995). 32
[BS87] David Bayer and Michael Stillman, A theorem on refining division orders by the re-
verse lexicographic order, Duke Math. J. 55 (1987), no. 2, 321–328. MR 894583 2
[Gro57] Alexander Grothendieck, Sur quelques points d’algèbre homologique, Tôhoku Math.
J. (2) 9 (1957), 119–221, Translated by Marcia L. Barr and Michael Barr:
Some aspects of homological algebra, (ftp://ftp.math.mcgill.ca/barr/
pdffiles/gk.pdf). MR MR0102537 (21 #1328) 37
[HS97] P. J. Hilton and U. Stammbach, A course in homological algebra, second ed., Gradu-
ate Texts in Mathematics, vol. 4, Springer-Verlag, New York, 1997. MR MR1438546
(97k:18001) 23, 27, 29, 32, 35, 36, 38
[Ste] Greg Stevenson, What is the opposite category of the category of modules (or
hopf algebra representations)?, MathOverflow, (accessed 2013-11-18): (http://
mathoverflow.net/questions/29442). 39
107
Index
L(A), 7 CertainColumns, 44
R-algebra, 64 DecideZeroColumns, 40
Hom-set, 23 DecideZeroColumnsEffectively, 40
Mod-R, 38 DecideZeroRows, 40
R-Mod, 38 DecideZeroRowsEffectively, 40
R-mod, 39 DiagMat, 44
mod-R, 39 IdentityMat, 43
D EDEKIND domain, 105 MulMat, 43
K RULL dimension, 89 StackMat, 44
SubMat, 43
action map, 25
SyzygiesOfColumns, 40
algebra
SyzygiesOfRows, 40
relation, 20, 21
ZeroMat, 43
algebraic set, 94
G AUSSian normal form=RREF, 41
algebraically independent, 20
H ERMITE normal form algorithm=HNF,
algorithm
41
addition, 39
Gröbner basis, 41
additive
multiplication, 39
Add, 44
pre-A BELian
CoproductMorphism, 44
Cokernel, 45
EmbeddingOfLeftCofactor, 44
CokernelColift, 45
EmbeddingOfRightCofactor, 44
ProductMorphism, 44 CokernelEpi, 45
ProductObject, 44 Kernel, 45
ProjectionOnLeftFactor, 44 KernelLift, 45
ProjectionOnRightFactor, 44 KernelMono, 45
Buchberger, 11 submodule membership, 16
categorical, 31 subtraction, 39
category syzygy module, 17
Compose, 42 Artinian
IdentityMorphism, 42 monomial order, 2
Source, 42 associated primes, 78
Target, 42 minimal, 78
category with zero axiom of choice, 27, 51
ZeroMorphism, 43
decide equality, 39 binomial, 12
decide membership, 39 Gröbner basis, 12
determinate division, 5 ideal, 12
elimination, 19 biproduct, 36
matrix boundaries, 54
AddMat, 43 Buchberger
AugmentMat, 44 algorithm, 11
BasisOfRows, 41 criterion, 10, 13, 14
BasisOfRowsCoeff, 41 test, 11
108
INDEX 109
categorical complex, 53
object acyclic, 54
initial=empty coproduct, 35 left, 54
terminal=empty product, 35 right, 54
zero=empty biproduct, 35 exact, 54
properties, 51 factor, 54
categorification homology, 53
horizontal, 23 sub-, 54
category, 23 constructive
A BELian, 37 A BELian group, 39
additive, 36 category, 42
arrow, 49 mathematical structure, 43
enriched over, 35 ring, 39
equivalence of categories, 51 set, 39
identity, 23 convention
morphism, 23 associative, 26
object, 23 commutative, 26
of (finite) presentations left, 23
left, 42 mixed, 26
right, 42 right, 23
opposite, 34 coproduct, 32
pre-A BELian, 37 embedding, 32
pre-additive, 35 morphism, 32
simplicial, 25 criterion
skeletal, 24 Buchberger, 10, 13, 14
small, 23 cycles, 54
locally, 23
defect of exactness, 53
subcategory, 34
degree-compatible order, 3
full, 34
dehomogenization, 90
thin, 25
direct sum, 36
with zero, 35
divisible, 9
chain
Division
morphism, 54
with remainder
of degree n, 54
determinate version, 5
co-image, 28, 30
dominates
epi, 31
column-, 40
mono, 37
modulo, 46
co-sequence, 53 row-, 40
cocomplex, 53 modulo, 46
cohomology, 54
codimension, 93 elimination
coefficient, 1 ideal, 19
cohomology, 54 order, 20
cokernel, 28, 30 embedded prime, 78
epi, 30 epi
object, 30 regular, 38
colift, 29 epimorphism, 27
along an epi, 29 equivalence
cokernel, 30 of categories, 51
compatible triples of functors, 50
left, 42 exact, 53
110 INDEX
monomial, 2
subobject, 29
lattice, 29
syzygies
column, 40
row, 40
relative, 44
syzygy, 13
matrix, 13
module, 13
algorithm, 17
target, 23
tensor
product, 63
pure, 63
term, 1
initial, 3
leading, 3
theorem
division, 5
syzygy, 17
universal property, 29
cokernel, 30
coproduct, 32
kernel, 29
product, 32
pull-back, 33
valuation, 103
discrete, 103
vanishing ideal, 94
vanishing locus, 94