Ma4200-Commutative Algebra: Geirellingsrud
Ma4200-Commutative Algebra: Geirellingsrud
M A 4 2 0 0 — C O M M U TAT I V E
ALGEBRA
1
These are informal notes for the course MAT4200 autumn 2018. As you
will see they are "under construction", so they highly unfinished and errors
abound. Some chapters, the first two or three, are not that preliminary and
should be readable. Hopefully the rest will improve during the semester. Each
chapter has its own version number. If the number is positive, the chapter
should be readable(and as usual the higher version number the more ready
the chapter is), but if negative (like ´8) it is sketchy if existing at all. I shall
indicate major changes, so you can follow the development.
Geir
Contents
1 Rings 7
Rings 8
Polynomials 15
Direct products and idempotents 18
2 Ideals 23
Ideals 23
Quotient rings and kernels 28
Prime ideals and maximal ideals 31
Principal ideals 34
Existence theorems 39
Local rings 44
Direct products and the Chinese Remainder Theorem 46
Graded rings and homogenous ideals 48
The prime spectrum and the Zariski topology 52
4 Modules 65
The axioms 65
Direct sums and direct products 72
Finitely generated modules 77
Bases and free modules 79
Exact sequences 85
Snakes and alike 95
4
6 Localization 123
Localization of rings 124
Localization of modules 135
Nakayama’s lemma 139
The support of a module 143
Rings
Preliminary version 1.1 as of 2018-08-22 at 08:11 (typeset 3rd December 2018 at 10:03am)—Prone
to misprints and errors and will change.
2018-08-21 Added a few sentences in paragraph 1.2.
2018-08-22 Added some words about isomorphisms, paragraph 1.6.
2018-08-23 Added exersices 1.13 and 1.20. Section 1.3 rewritten in a simpler manner.
30/08/2018 Added a hint to exercise 1.15.
06/09/2018 Added exercise 1.16.
1.1 Rings
1.1 Recall that a ring is an algebraic structure consisting of a set endowed with
two binary operations; an addition which makes A an abelian group, and
a multiplication. The multiplication is assumed to be distributive over the
addition, and in this course it will always be associative and commutative
(or at least almost always). There are of course many extremely interesting
non-commutative rings and non-associative rings, but this course is dedicated
to rings that are associative and commutative.
The sum of two elements will of course be denoted as a ` b, and the product
will we indicated in the traditional way by a dot or simply by juxtaposition;
that is, as a ¨ b or just by ab. The left distributive law asserts that apb ` cq “ ab ` bc,
and since rings for us are commutative, it follows that the right distributive law
pb ` cqa “ ba ` ca holds as well.
We shall also assume that all rings have a unit element; that is, an element
1 A such that 1 A ¨ a “ a for all members a of the ring. At most occasions the
reference to A will be dropped, and we shall write 1 for the unit element
whatever the ring is.
Example 1.1 The simplest of all rings are the ring Z of integers and the
rings Z{nZ of residue classes of integers modulo n. The traditional numbers
systems of real numbers R and complex numbers C are well-known rings. K
1.2 Generally, ring-elements can behave quite differently than we are used to in
a classical setting. It might very well happen that ab “ 0 without neither a nor
b being zero. Such elements are called zero divisors. Be aware that the familiar Zero divisors (nulldivi-
cancellation law does not hold in a ring with zero divisors in that ab “ ac not sorer)
necessarily implies that b “ c. Rings without zero divisors are called integral
domains or, for short, domains. Integral domains (In-
Obviously, elements that are not zero divisors are called non-zero divisors, tegritetsområder)
Non-zero divisors (Ikke-
another name being regular elements. A regular element a has the virtue that nulldivisorer)
xa “ 0 implies that x “ 0 and can therefore be cancelled from equalities like Regular elements (Reg-
ab “ ac (the difference b ´ c being killed by a vanishes). ulære elementer)
For instance, in case n is a composite number, say n “ pq, the ring Z{nZ
has zero divisors; it holds true that pq “ 0, and p and q are both different from
zero (neither having n as a factor).
A more geometric example could be the ring of continuous functions on the
space X which is the union of the x-axis and the y-axis in the plane. On X the
function xy vanishes identically, but neither x nor y does (x does not vanish on
the y-axis and y not on the x-axis).
1.3 It might also happen that a powers of a non-zero elements vanish, i.e. one
has an “ 0 for some natural number n, but a ‰ 0. For instance, in the ring
rings 9
Z{p2 Z one has p2 “ 0, but p ‰ 0. Such elements are called nilpotent. Rings Nilpotent elements
deprived of nilpotent elements, are said to be reduced. (nilpotente elementer)
Reduced rings (reduserte
ringer)
Units and fields
Examples
1.3 We do not assume that 1 ‰ 0 although it holds in all but one ring. The
exception is the so-called null-ring. If in a ring it holds that 0 “ 1, one has The null-ring (nullrin-
a “ a ¨ 1 “ a ¨ 0 “ 0, so zero will be the sole element. The only role the gen)
null-ring plays and the only reason not to throw it over board, is that it allows
significantly simpler formulations of a few results. It does not merit a proper
notation (well. . . , one alway has the alternative 0).
1.5 The complex rational functions in a variable x form a field Cpxq. The
elements are meromorphic functions in C expressible as the quotient ppxq{qpxq
of two polynomials p and q and q not being identically zero.
1.6 For any set X Ď Cr one may consider the set of polynomial functions on X; Polynomial functions
that is, the functions on X that are restrictions of polynomials in r variables. (polynomiale funskjoner)
1.7 Associated with any topological space X are the sets CR pXq and CC pXq of
of continuous functions on X assuming respectively real or complex values.
Point-wise addition and multiplication make them (commutative) rings. When
X has more structure than just a topology, there are further possibilities. Two
10 ma4200—commutative algebra
instances being the ring of smooth functions on a smooth manifold, and the
ring O pΩq of holomorphic functions in an open domain Ω of the complex
plane.
complex numbers C and inherit their ring structure from C; they are closed
under addition (which is obvious) and multiplication:
? ? ?
pa ` b nqpa1 ` b1 nq “ paa1 ` nbb1 q ` pab1 ` a1 bq n.
Homomorphisms
When studying mathematical objects endowed with a certain structure—like
rings for instance—maps preserving the structure are fundamental tools.
Working with topological spaces one uses continuous maps all the time, and
linear algebra is really about linear maps between vector spaces. And of
course, the theory of groups is inconceivable without group homomorphisms;
that is, maps respecting the group laws. A new class of objects in mathemat-
ics is always accompanied by a new class of maps. This observation can be
formalized and leads to the definition of categories—see xxx. categories (kategorier)
1.5 In our present context the relevant maps are the so-called ring homomor-
phism, which also will be referred to as maps of rings or ring-maps. These are Ring homomorphisms
maps φ : A Ñ B between two rings A and B preserving all the structures (ringhomomorfier)
around; that is, the additive group structure, the multiplication and the unit
element 1. In other words, they comply to the rules
The sum of two maps of rings is in general not a map of rings (it is additive,
but does not respect the multiplication) neither is their product (it respects
multiplication, but not addition), but of course, the composition of two com-
posable ring-maps is a ring-map.
1.6 A homomorphism φ : A Ñ B is an isomorphism if there is a ring homomor-
phism ψ : B Ñ A such that the two relations ψ ˝ φ “ id A and φ ˝ ψ “ idB Isomorphisms of rings
(isomorfier av ringer)
hold true. One most often writes φ´1 for the inverse map, and it is common
usage to call isomorphisms invertible maps. For φ to be invertible it suffices it
be bijective. Multiplication will then automatically be respected since when φ is
injective, φ´1 pabq “ φ´1 paqφ´1 pbq is equivalent to ab “ φpφ´1 paqφ´1 paqq , and
rings 11
Examples
subset S of A, and then one speaks about polynomial expressions with coefficients
in S. They are thus elements of A shaped like
ÿ ÿ α
sα ¨ aα “ sα ¨ a1 1 ¨ . . . ¨ aαnn ,
α α
where the summation extends over all multi-indices, and where the non-zero
coefficients are finite in number and confined to lie in S.
A successive application of the distributive law gives the classical formula
for the product of two polynomial expressions:
ÿ ÿ ÿ ÿ
p sα ¨ aα q ¨ p t β ¨ a β q “ p sα t β q ¨ aγ . (1.1)
α β γ α` β“γ
12 ma4200—commutative algebra
1.8 A subring B of A is a ring contained in A whose ring operations are in- Subrings (Underringer)
duced from those of A. Phrased differently, it is an additive subgroup con-
taining the unit element which is closed under multiplication; to be specific,
it holds that 0 P B and 1 P B, and for any two elements a and b belonging to
B, both the sum a ` b and the product ab belong to B. The intersection of any
number of subrings of A is a clearly subring.
Example 1.11 The integers Z is a subring of the rationals Q. K
1.9 Given a ring A and a subring B and a set of elements a1 , . . . , ar from A.
One constructs a subring Bra1 , . . . , ar s of A as the set of all polynomial expres-
sions ÿ α
bα ¨ a1 1 ¨ . . . ¨ arαr
where α “ pα1 , . . . , αr q runs through the multi-indices and the bα ’s are elements
from B, only finitely many of which are different from zero. It is straightfor-
ward to check, using the classical formula (1.1) above, that this subset is closed
under multiplication and hence is a subring of A (it is obviously closed under
addition). It is called the subring generated by the ai ’s over B, and is the small- Subrings generated by
est subring of A containing the ring B and all the elements ai . Common usage elements (underringer
generert av elementer)
is also to say that Bra1 , . . . , ar s is obtained by adjoining the ai ’s to B.
This construction works fine even for infinitely many ai ’s since each polyno-
mial expression merely involves finitely many of them. Thus there is a subring
Brai |i P Is for any subset tai uiP I of A.
Examples
K
rings 13
1.10 Every ring has a canonical subring called the prime ring. The unit element
1 in A generates an additive cyclic subgroup of A whose elements are just
sums of 1 or ´1 a certain number of times; that is, they are shaped like n “
1 ` . . . ` 1 or n “ ´1 ´ . . . ´ 1. This subgroup is obviously closed under
multiplication and is hence a subring. It is called the prime ring of A. The prime ring (primrin-
As is well known, a cyclic group is either finite and isomorphic to Z{nZ for gen)
Algebras
Frequently when working in commutative algebra there are “coefficients”
around; that is, one is working over a “ground ring”. So the most natural
objects to work with are perhaps not rings, but the so-called algebras.
1.12 The notion of an algebra is a relative notion involving two rings A and B.
To give a B-algebra structure on A is just to give a map of rings φ : B Ñ A. One Algebras (Algebraer)
may then form products φpbq ¨ a of elements a from A with elements of the form
φpbq. The map φ (even though it is an essential part of the B-algebra structure
of A) is often tacitly understood and suppressed from the notation; one simply
writes b ¨ a for φpbq ¨ a. Later on, when we have introduced modules, a B-algebra
structure on a ring A will be the same as a B-module structure on A.
Example 1.15 Every ring has a canonical structure as a Z-algebra (defined as
in Paragraph 1.10 above). The class of algebras is therefore a strict extension of
the class of rings. Since a ring is an algebra over any subring, over-rings give a
large number of examples of algebras. K
1.13 Faithful to the principle that any new type of objects is accompanied by
a corresponding new type of maps; one says that a map of rings φ : A Ñ A1
between two B-algebras is an B-algebra homomorphism1 if it respects the action Algebra homomorphisms
of B; in other words, it holds true that φpb ¨ aq “ b ¨ φpaq for all elements a P A (algebrahomomorfismer)
1
Or any morphological
and b P B. Composition of two composable B-algebra homomorphisms is derivative thereof, like
a B-algebra homomorphism so that the B-algebras form a category denoted map of B-algebras or
Alg B . B-algebra-map etc.
1.14 One says that A is finitely generated over B, or is of finite type over B, if Finitely generated
A “ Bra1 , . . . , ar s for elements a1 , . . . , ar from A. algebras (endeliggenererte
algebraer)
Finite type algebras
(algebraer av endelig
type)
14 ma4200—commutative algebra
Problems
1.1 Find all nilpotent and all zero-divisors in Z{72Z. What are the units? ˇ
1.4 Show that prime ring is the smallest subring; i.e. it is contained in all other
subrings of the given ring.
1.5 Convince yourself that the binomial theorem persists being true in any
commutative ring; that is, check that your favourite proof still holds water.
1.6 Show that the sum of two nilpotent elements is nilpotent. Hint: You can
rely on the binomial theorem.
1.11 (The Eisenstein integers.) Let ω be the cube root ω “ e2πi{3 of unity. Show ˇ
that the subring Zrωs of C is given as
Zrωs “ t n ` mω | n, m P Z u.
1.2 Polynomials
We are well acquainted with polynomials with real or complex coefficients;
we met them already during the happy days at school. They were then intro-
duced as functions depending on a of real (or a complex one if you went to a
French school) variable whose values were given by a polynomial expressions.
We shall in this section introduce polynomials with coefficients in any (com-
mutative) ring A. The point of view will necessarily be formal and without
reference to functions, and there will be more than just one variable.
1.15 In an earlier paragraph we met polynomial expressions in a set of ring el-
ements. In the present situation where there is no surrounding ring, we must,
as signalled above, proceed in a formal way. A polynomial in the variables Polynomials (Polynomer)
x1 , . . . , xr is defined as a formal sum
ÿ α
f px1 , . . . , xn q “ aα x1 1 ¨ . . . xnαn , (1.2)
α
where the summation extends over all multi-indices α “ pα1 , . . . , αn q with the
αi ’s being non-negative integers, and where the coefficients aα are elements
from the ground ring A, only finitely many of which are non-zero. Do not
speculate much3 about what the term “formal sum” means, the essential 3
Or do Exercise 1.19
point is that two such “formal sums” are equal exactly when corresponding below where the more
general construction of
coefficients agree. the so-called monoidal
α algebras is described in
1.16 The “pure” terms aα x1 1 ¨ . . . xnαn occurring in (1.2) are called monomials. The
α a precise manner.
abbreviated notation x α “ x1 1 ¨ . . . xnαn is convenient and practical. The degree of
α
ř
a non-zero monomial aα ¨ x is the sum i αi of the exponents, and the highest Monomials (Monomer)
degree of a non-zero monomial term in a polynomial, is the degree of the The degree of a poly-
polynomial. Non-zero constants are of degree zero, but the zero polynomial nomial (Graden til et
polynom)
is not attributed a well-defined degree—it is rather considered to be of any
degree (it equals 0 ¨ x α for any α).
16 ma4200—commutative algebra
A polynomial is said to be homogenous if all its monomial terms are of the Homogenous polynomials
same degree. For example the polynomial x2 y ` z3 is homogeneous of degree (Homogene polynomer)
1.19 The polynomial ring Arx1 , . . . , xr s has a so-called universal mapping prop-
erty; one may freely assign values to the variables to obtain homomorphisms. A universal mapping
property (En universell
avbildningsegenskap)
ř
Proof: A polynomial p is given as p “ α aα x α . Since the coefficients aα are
ř α
unambiguously determined by p, setting φppq “ α aα β 11 ¨ . . . ¨ brαr gives a
well-defined map which easily is seen to be additive. Since a relation like the
one in (1.3) is universally valid in commutative rings, φ respects multiplication
as well, and we have an algebra homomorphism. o
Problem 1.14 Convince yourself that the universal mapping property holds ˇ
even for polynomial rings in infinitely many variables. M
1.22 The second type of ring we have in mind, are the rings of formal power
series. A formal power series is an expression like in (1.2) Rings of formal power
series (Ringen av formelle
ÿ potensrekker)
f px1 , . . . , xn q “ aα x α1 ¨ . . . xnαn ,
α
except that the sum is not require be to finite. Addition is done term by term,
and the multiplication is defined by formula (1.3), which is legitimate since the
expression for each coefficient involves only finitely many terms. The formal
power series ring is denoted AJx1 , . . . , xn K.
Problems
1.16 Let A be a reduced ring. Show that group of units in the polynomial ring
Arxs equals A˚ .
1.17 Assume that k is a field. Show that krts and krt, 1{ts are not isomorphic as ˇ
rings.
1.18 Assume that k is field. Show that a the number of zeros of a non-zero
polynomial in krts is less than the degree. Show that if two polynomials in
krts define the same function on k and k is infinite, then they coincide as
polynomials. If k is finite, exhibit a polynomial that vanish identically on k.
1.20 Let tAi u be a collection of subrings of the ring A. Prove that the intersec-
Ş
tion iP I Ai is a subring.
Give examples of two subrings A1 and A2 of A such that their union is not
a subring. Assume that the collection has the property that any two rings from
Ť
the collection are contained in a third. Prove that in that case the union iP I Ai
is a subring.
This indicates that in the interplay between geometry and rings, direct product
of rings correspond to disconnected spaces.
Below we define the direct product of a collection of rings regardless of the
cardinality of the collection and introduce the notion of idempotent elements
(elements e such that e2 “ e). Multiplication by idempotents are projection
operators (equal to their squares) and serve to decompose rings (and later on
modules) into direct products.
The archetype of an idempotent function is the characteristic function
eX of a connected component, say X, of a topological space Z; that is, the
function that assumes the value one on X and zero on the rest of Z. Since
X is a connected component of Z, this function is continuous and of course,
e2X “ eX . Moreover, the restriction f |X of any function f equals eX f , or put
more precisely, eX f is the restriction f |X extended by zero to the entire space
Z. Anyhow, in this way the set eX CR pZq is a ring naturally identified with
CR pXq with the idempotent eX corresponding to unit element in CR pXq. The
lesson learned is that idempotents are algebraic counterpart to the geometric
connected components.
The unit element is the pair p1, 1q, and the two projections πi : A Ñ Ai are ring
homomorphisms. Moreover, the direct product possesses two special elements
e1 “ p1, 0q and e2 “ p0, 1q, which satisfy ei2 “ ei and e1 e2 “ 0. The sets ei A
equal respectively A1 ˆ t0u or t0u ˆ A2 , and are, with a liberal interpretation5 , 5
Since the unit element
subrings of A isomorphic to respectively A1 and A2 . p1, 1q does not lie in
either Ai , they are
1.23 To generalize what we just did for a pair of rings, let tAi uiP I be any collec- properly speaking not
subrings even though
tion of rings, which can be of any cardinality. In our context it will mostly be
ś they are closed under
finite, but occasionally will be countable. The direct product iP I Ai has as un- both addition and
derlying additive group the direct product of the underlying additive groups multiplication.
of the Ai ’s. The elements are “tuples” or strings pai qiP I indexed6 by I whose 6
The reference to
i-th component ai belongs to Ai , and the addition of two such is performed the index set I will
frequently be dropped
component-wise. The same is true of the multiplication, also performed com- and the strings written
ponent for component; that is, it holds true that pai q ¨ pbi q “ pai ¨ bi q. The ring pai q.
axioms can be checked component-wise and thus come for free.
Ť
Interpreting tuples a “ pai q as maps a : I Ñ iP I Ai , the ring operations
of the direct product are just the point-wise operations. The unit element, for
instance, is the “constant7 ” function that sends each index i to 1. 7
Why the quotation
marks?
20 ma4200—commutative algebra
ś
1.24 The projections πi : iP I Ai Ñ Ai are ring homomorphisms (this is
just another way of saying that the ring operations are defined component-
wise) and enjoy the following universal property. Given any ring B and any B
φ
/ śi A i
collection φi : B Ñ Ai of ring homomorphisms, there is an unambiguously
ś πi
defined map of rings φ : B Ñ iP I Ai such that φi “ πi ˝ φ for all i P I: $
φi
Indeed, this amounts to the map given by φpxq “ pφi pxqqiP I being a ring Ai
homomorphism.
Idempotents
1.25 In any ring A an element e satisfying e2 “ e is said to be idempotent, and if Idempotent elements
f is another idempotent, one says that f and g are orthogonal when f g “ 0. The (idempotente elementer)
Orthogonal idempotents
element 1 ´ e is always idempotent when e is and is orthogonal to e as shown (ortogonale idempotenter)
by the little calculations
p1 ´ eq2 “ 1 ´ 2e ` e2 “ 1 ´ 2e ` e “ 1 ´ e,
ep1 ´ eq “ e ´ e2 “ e ´ e “ 0.
ea ¨ eb “ e2 ab “ eab,
A
» / ś ei A.
i
Proof: To begin with, we verify that the map in the proposition, call it φ, is a
ring homomorphisms. So let x and y be two elements from A. The ei ’s being
idempotents we find
Problems
1.22 (The p-adic integers.) Let p be a prime number and let ρn : Z{pn`1 Z Ñ
Z{pn Z be the reduction map that sends the residue class ras pn`1 of an integer
a modulo pn`1 to the residue class ras pn of a modulo pn . Let Z p be the subset
of the direct product nPN Z{pn Z whose members are strings pan qnPN such
ś
M
Lecture 2
Ideals
The ideals were first defined by Richard Dedekind in 1876, but the name comes
from the so called “ideal numbers” of Ernst Kummer which he introduced in a
series of papers around 1847.
Working with rings of integers in algebraic number fields, the algebraists of
that period realized that analogues of the Fundamental Theorem of Arithmetic
does not always hold in such rings. Recall that this theorem asserts that any
integer is a product n “ p1 ¨ . . . ¨ pr of signed primes and the factors are
Richard Dedekind
unique up to order and sign—changing the order of the factors does not (1831–1916)
affect the product, and changing the sign of one factor can be compensated by German mathematician
simultaneously changing the sign of another.
It is not too complicated to show that in a vast class of rings, including
the rings of algebraic numbers above, any element can be expressed as a
product of irreducible elements; that is, elements that can not be factored
further (they can of course always be altered by a unit, but that is not an
honest factorization). However the point is, that these factors are not unique
(up to order and units) in general.
The classical example omnipresent in text books, is the factorisazion 2 ¨ 3 “ Ernst Eduard Kummer
? ? ?
p1 ` i 5qp1 ´ i 5q in the ring Zri 5s. The four involved numbers are all (1810–1893)
irreducible and no two of them related by units. German mathematician
The ideals came about to remedy this fault and, in fact, in certain rings
called Dedekind rings, the situation can be salvaged; there is a factorization
theorem of ideals replacing the Fundamental Theorem of Arithmetic. Hence
the name ideals, they were “the ideal numbers”
Dedekind rings are however a very restricted class of rings, and today ide-
als play an infinitely wider role than just being “ideal numbers”. In algebraic
geometry for instance, they appear as the sets of polynomials in krx1 , . . . , xr s
vanishing along a subset of kr , and this is the clue to the coupling between
algebra and geometry.
2.1 Ideals
24 ma4200—commutative algebra
2.1 Let A be a ring. An additive subgroup a of A is called an ideal if it is closed Ideals (Idealer)
under multiplication by elements from A. That is, a satisfies the two following
requirements; the first merely being a rephrasing that a is a subgroup.
o If a P A and b P a, then ab P a.
Both the trivial additive subgroup 0 and the entire ring satisfy these require-
ments and are ideals, although special ideals. In many texts the ring when
considered an ideal, is denoted by p1q.
2.2 An ideal a is said to be a proper ideal if it is not equal to the entire ring. This Proper ideals (Ekte
is equivalent to no member of a being invertible. Indeed, if a P a is invertible idealer)
Proposition 2.3 A ring A is a field if and only if its only ideals are the zero ideal
and A itsef.
Examples
2.4 The set I pAq of ideals in the ring A has—in addition to being partially
ordered under inclusion—a lot of structure. One may form the intersection
Ş
iP I ai of any family tai uiP I of ideals. It is easily seen to be an ideal, and is the
largest ideal contained in all the ai ’s. Likewise, one has the notion of the sum of
a family of ideals. It is the ideal consisting of all finite sums of elements from
the ai ’s:
ÿ
ai “ t a1 ` . . . ` ar | ai P ai , r P N u,
iP I
ideals 25
and is the smallest ideal containing all the ai ’s. So I pAq is what one technically
calls a complete lattice; every subset of I pAq has a greatest lower bound (the
sum) and a smallest upper bound (the intsersection). It is the lattice of idealsin The lattice of ideals
A. (Ideallattiset(???))
2.5 A construct similar to the sum of a family of ideals is the ideal generated
by a set of elements tai uiP I from A. It will be denoted pai |i P Iq, or in case the
set S “ ta1 , . . . , ar u is finite, the alternative notation pa1 , . . . , ar q is common us-
age. Its members are all finite linear combinations of the ai ’s with coefficients
from the ring A; that is, it holds that
ř
pai |i P Iq “ t iP J ci ai | ci P A, J Ď I finite u.
The elements ai are called generators. Ideals which are generated by finitely Generators (Generatorer)
many elements are naturally called finitely generated. An ideal generated by Finitely generated ideals
a single element, is called a principal ideal; it consists of all multiples of the (Endeliggenererte idealer)
Principal ideals (Hoved-
generator. If a is the generator, it is denoted by paq or by aA, and it holds that idealer)
paq “ t c ¨ a | c P A u.
2.6 The product of two ideals a and b is the ideal generated by all products of The product of ideals
one element from a and one from b; that is, the product ab is formed of all (Produktet av idealer)
ab “ t a1 b1 ` . . . ` ar br | ai P a, bi P b, r P N u.
2.7 The last operation we offer is the formation of the transporter between two The transporter (Trans-
ideals. Some texts call it the quotient of the two deals—however, this term portøren)
should be reserved for another construction we shortly come to, and hence
should be avoided. So let a and b be two ideals in A. We define the transporter
pa : bq to be set of elements which on multiplication send b into a; that is
pa : bq “ t x P A | xb Ď a u.
Examples
2.3 In Z it holds that 100 : 10 “ p10q. More generally if a and b are elements
` ˘
from the ring A and b is a non-zero divisor, one has pab : bq “ paq. Indeed,
xb “ yab is equivalent to x “ ya since cancellation by b is allowed b being
a non-zero divisor. If b is a zero-divisor it anyhow holds that pab, bq “ paq `
Ann b.
26 ma4200—commutative algebra
2.4 In the polynomial ring Crx, ys it holds that pxy, y2 q : px, yq “ pyq.
` ˘
` ˘
Clearly pyq is contained in pxy, y2 q : px, yq . For the converse inclusion
assume that f x “ gxy ` hy2 where f , g and h are polynomials in Crx, ys. Since
x divides the terms f x and gxy, it divides hy2 as well, and by cancelling x, we
infere that f “ gy ` h1 y with h1 P Crx, ys; that is, f P pyq.
2.5 In Z{40Z one has Ann 2 “ p20q, that Ann 4 “ p10q and that Ann 20 “ p5q.
Functorially
A map of rings φ : A Ñ B induces two maps between the ideal-lattices I pAq
and I pBq, one in a covariant and one in a contravariant way. One can move
ideals forward with the help of φ and the the usual inverse image construct is
a way move ideals backwards along φ. The new ideals are in some texts either
called extensions or contractions in the respective case. Extensions of ideals
(Utvidlese av ideal)
2.8 We start with the contravariant one. The inverse image φ´1 pbq
of an ideal Contractions of ideals
b in B is an ideal in A; indeed, φpabq “ φpaqφpbq belongs to b whenever φpbq (Tilbaketrekning)
does, and this gives rise to a map φ´1 : I pBq Ñ I pAq. Obviously it preserves
inclusions and takes intersections to intersections since the pullback of sets
respects intersections in general. Sums and products of ideals however, are not
generally preserved. One has
or simply by aB. This induces a map I pAq Ñ I pBq. Inclusions are preserved,
and one easily verifies the following relations
o φpa ¨ bqB “ pφpaqBq ¨ pφpbqBq,
The inclusion in the last line may be strict, see Example 2.9 on page 30 below.
Problems
2.1 Let a and b be ideals in a ring A. Show that the relations a ¨ b Ď a X b and ˇ
pa X bq2 Ď a ¨ b hold. Show by giving examples that there might be a strict
inclusion in both cases.
2.2 Assume that a and b are ideals in a ring A satisfying a ` b “ p1q. Show that ˇ
a ¨ b “ a X b.
2.3 Let a,b and c be ideals in the ring A. Show that apb ` cq “ ab ` ac. Show
that a X pb ` cq Ď a X b ` a X c, and by exhibiting an example, show that the
inclusion can be strict.
2.5 Let tai u be a collection of ideals in the ring A. Show that for any ideal b it ˇ
Ş Ş ř Ş
holds true that p iP I ai : bq “ iP I pai : bq and that pb : iP I ai q “ iP I pb : ai q.
2.6 Show that any ideal in the ring Z of integers is generated by one element,
which is unique up to sign.
` ˘
2.7 Let m and n be two integers. Show that pnq ¨ pmq “ pnq X pmq ¨ pn, mq.
2.8 Consider the two ideals a “ p144q and b “ p24q in Z . Describe pa : bq.
In general, describe pn : mq in terms of the prime factorizations of the two
integers n and m.
2.9 Let krx, ys be the polynomial ring in the variables x and y over the field k, ˇ
and let m be the ideal generated by x and y; that is m “ px, yq. Let n denote a
natural number.
a) Exhibit a set of generators for the power mn .
b) Let µ and ν be two natural numbers. Show that mn Ď px µ , yν q for n suffi-
ciently large. What is the smallest n for which this holds?
? ?
2.10 Let A “ Zr 2, 3s. Show that as an abelian group A free of rank four ˇ
and exhibit a basis. Show that the underlying abelian groups of the principal
? ?
ideals p 2q and p 3q both are of rank four. Exhibit additive bases for both.
M
28 ma4200—commutative algebra
and we can conclude that ab P ker φ. Hence the kernel ker φ is an ideal.
2.11 To see that any ideal is a kernel, one introduces the concept of quotient
rings. An ideal a in A being an additive subgroup, there is a quotient group Quotient rings (Kvotient-
A{a. It consists of the residue classes ras “ a ` a of elements in A, and the sum ringer)
of two such, say ras and rbs, equals ra ` bs. To put a ring structure on A{a we
simply define the product of two classes ras and rbs as
ras ¨ rbs “ ra ¨ bs “ a ¨ b ` a.
Some checking is needed; the most urgent one being that the product only
depends on the residue classes ras and rbs and not on the representatives a and
b. This is encapsulated in the formula
pa ` aq ¨ pb ` aq “ a ¨ b ` a ¨ a ` b ¨ a ` a ¨ a Ď a ¨ b ` a.
It is left to the students to verify that this product comply with the associative
and the distributive laws. Finally, by definition of the ring operations in A{a,
the quotient map π : A Ñ A{a is a map of rings whose kernel equals the given
ideal a.
Example 2.7 It is appropriate to mention what quotients by the two “extreme”
ideals are. The quotient A{a equals A if and only if a is the zero-ideal, and it
equals1 the null-ring if and only if a “ A. K 1
This examplifies
what purpose the null-
2.12 The quotient ring A{a together with the quotient map π : A Ñ A{a enjoys ring serves; it allows
a so-called universal property—the rather pretentious notion “solves a universal a general existence
theorem (avoiding the
problem” is also common usage—which is convenient way of characterizing hypothesis a ‰ A).
many types of mathematical objects. The origin of the technique is found in A universal property (En
category theory where objects not always have “elements” and one relies on universell egenskap)
φpaq “ 0, the map φ is constant on every residue class ras “ a ` a, and we put
ψprasq equal to that constant value. This value is forced upon ψ, so ψ is unique,
and it is a ring-map since φ is. We have proven:
Ideals in quotients
Proof: We already saw that πpbq “ b{a is an ideal. That b is unique follows
from the second assertion since if πpbq “ πpb1 q it ensues that b ` a “ b1 ` a, and
hence b “ b1 when both contain a. The second assertion is clear since a “ ker π.
Likewise is the third for the same reason. o
2.16 Proposition 2.15 above may be rephrased in terms of the lattices of ideals
I pAq and I pA{aq. The map π ´1 : I pA{aq Ñ I pAq sending an ideal c to the
inverse image π ´1 pcq is injective and the image consists of the ideals in A con-
taining a. The map the other way round, which sends an ideal b in I pAq to its
image πpbq, serves as a section to π ´1 over the sublattice of ideals containing
a. The map π ´1 preserves inclusions and respects intersections, products and
sums.
2.17 The image in A{a of an ideal b Ď A, which not necessarily contains a,
is the ideal pb ` aq{a. This holds since obviously πpb ` aq “ πpbq. Now,
30 ma4200—commutative algebra
Finally, we mention that when a and b are two ideals with a Ď b, there is a
natural isomorphism
pA{aq{pb{aq » A{b. (2.2)
Indeed, in the diagramme in the margin, the composition πb{a ˝ πa has the A
πa
/ A{a
ideal b as kernel, and therefore factors through πb by say θ. The map θ is
πb πb{a
surjective since the composition is and injective since the composition has
b “ πa´1 pb{aq as kernel. These two formulas are often referred to as the A{b / pA{aq{pb{aq
θ
Isomorphisms theorems.
Theorem 2.18 (The Isomorphism Theorem) Let a and b be ideals in A.
Then the following two equalities hold, where in the second is assumed that a Ď b;
o b{b X a » pa ` bq{a;
o pA{aq{pb{aq » A{b.
Examples
2.8 We promised to give examples that strict inclusion can hold in the last
statement of in Paragraph 2.8 (on page 26). Here comes one. Let B “ krX, Y, Zs{pZ ´
XYq, and as usual the lower case version of a variable will stand for its class in
B. Then B “ krx, y, zs with z “ xy.
We let the map φ : A Ñ B be the natural inclusion of A “ krzs in B “
krx, y, zs; for any ideal c in B the inverse image φ´1 pcq is then nothing but the
intersection c X B.
Consider the two principal ideals a “ pxqB and b “ pyqB in B. Since z “ xy,
it holds that z P pxqB and z P pyqB, and since pzqA is a maximal ideal in
A “ krzs , we see that a X A “ b X A “ pzqA, and hence pa X Aq ¨ pb X Aq “
pz2 qA. On the other hand ab “ pxyqB “ pzqB so that a ¨ b X A “ pzqA; hence
φ´1 paq ¨ φ´1 pbq Ř φ´1 pa ¨ bq.
2.9 For an example of strict inclusion in the last statement Paragraph 2.9 (on
page 26) let A “ krX, Y, Zs and B “ krX, Y, Zs{pZX ´ ZYq and consider the
projection map φ : A Ñ B; then φpcq “ cB for any ideal c in A.
Let a “ pXq and b “ pYq. Then a X b “ pXYq. With the usual lower case
convention, since zx “ zy, it holds that and zx P aB X bB, but zx R pxyqB; and
hence φpaq X φpbq Ř φpa X bq.
K
ideals 31
An ideal a is said to be maximal if it is proper and satisfies the following Maximal ideals (Maksi-
requirement: male idealer)
In other words, a is maximal among the proper ideals. Notice that both prime
ideals and maximal ideals are proper by definition.
2.20 One has the following characterization of the two classes of ideals in
terms of properties of quotients.:
Proposition 2.21 An ideal a in A is a prime ideal if and only if the quotient A{a
is an integral domain. The ideal a is maximal if and only if A{a is a field.
Proof: It holds true that A{a is an integral domain if and only if rasrbs “ 0
implies that either ras “ 0 or rbs “ 0; that is, if and only if ab P a implies that
either a P a or b P a, which proves the first assertion.
Bearing in mind the relation between ideals in A{a and ideals in A con-
taining a (as in Proposition 2.15 on page 29), the second assertion is pretty
obvious. There is no ideal strictly between a and A if and only if A{a has no
non-trivial proper ideal; that is, if and only if A{a is a field (Proposition 2.3 on
page 24). o
Notice that the zero ideal p0q is a prime ideal if and only if A is an integral
domain, and it is maximal if and only if A is a field. When m is a maximal
ideal, the field A{m is called the residue class field of A at m and now and then Residue class fields
denoted by kpmq. Since fields are integral domains, we see immediately that (Restklassekropper)
maximal ideals are prime. The converse does not hold as we shortly shall see
examples of (Example 2.11 below).
2.23 Not only for elements is it true that a product lies in a prime ideal only
when one of the factors does, the same applies to products of ideals as well:
32 ma4200—commutative algebra
Proposition 2.24 Let a and b be two ideals in A such that ab is contained in the
prime ideal p. Then either a or b is contained in p.
Proof: Assume that neither a nor b lies in p and pick elements a P a and b P b
not being members of p. Since ab is contained in p, the product ab belongs to p,
and since p is prime, it either holds that a P p or that b P p. Contradiction. o
Proposition 2.26 Let A be a ring and a an ideal. The prime ideals in the quotient
A{a are precisely those of the form p{a with p a prime ideal in A containing a, and
the maximal ideals are those shaped like m{α with m a maximal ideal in A likewise
containing a.
Examples
2.10 The archetype of maximal ideals are the kernels of evaluation maps. For
instance, let a “ pa1 , . . . , ar q be a point in kr where k is any field, and consider
the map krx1 , . . . , xr s Ñ k sending a polynomial f to its value f paq at a. The
kernel m is a maximal ideal since krx1 , . . . , xr s{m is the field k. The kernel may
be described as m “ px1 ´ a1 , . . . , xr ´ ar q. This is obvious when a is the origin,
and introducing fresh coordinates xi1 “ xi ´ ai , one reduces the general case to
that case.
2.11 A simple example of a prime ideal not being maximal can be the princi-
pal ideal pyq in the the polynomial ring krx, ys in the two variables x and y over
a field k. One has krx, ys{pyq » krxs, for instance, since the partial evaluation
map krx, ys Ñ krxs sending f px, yq to f px, 0q has kernel pyq, and pyq is prime be-
cause krxs is an integral domain. Moreover pyq is not maximal being contained
in px, yq (or if you prefer, because krxs is not a field).
K
ideals 33
Proof: We can certainly assume that the union is irredundant; that is, none of
the prime ideals are contained in the union of the others. Then, for each j one
may pick an element y j lying in the prime ideal p j but not in any of the others.
Assume for a contrapositive argument that a is not contained in any of the
pi ’s, and chose elements x j from a not in p j . Consider for each j the element
ś
z j “ x j i‰ j yi . It lies in pi when i ‰ j but not in p j , and of course it belongs to
a since x j does. It ensues that the sum z1 ` . . . ` zr is a member of a, but not of
any of the pi ’s (each term lies in a and for a given j all but one lie in p j ). o
Notice that the proof merely requires a be closed under addition and multi-
plication, so the ideal a may be replaced with a “weak subring” (a subring
without a unit element).
2.29 At several later occasions we shall meet diverse unions and intersections
of prime ideals. In case there are no non-trivial inclusions among the involved
prime ideals, they enjoy strong uniqueness properties; in fact, the primes
involved are determined by their intersection or their union, as expressed in
the following twin lemmas.
Lemma 2.30 Let tp1 , . . . , pr u and tq1 , . . . , qs u be two families of prime ideals having
the same union; that is, p1 Y ¨ ¨ ¨ Y pr “ q1 Y ¨ ¨ ¨ Y qs . Assume that there are no
non-trivial inclusion relations in either family. Then the two families coincide.
Ť
Proof: For each index ν one has pν Ď j q j and the Prime avoidance lemma
gives that there is an index αpνq so that the relation pν Ď qαpνq holds. By sym-
metry, for each µ there is a βpµq such that qµ Ď pβpµq . Now
pν Ď qαpνq Ď pβpαpνqq
34 ma4200—commutative algebra
and there being no non-trivial inclusion relations among the pi ’s we infer that
βpαpνqq “ ν. In a symmetric manner one shows that αpβpµqq “ µ and we can
conclude that α is a bijection from t1, . . . , ru to t1, . . . , su with pν “ qα pνq, and
we are happy. o
Lemma 2.31 Let tp1 , . . . , pr u and tq1 , . . . , qs u be two families of prime ideals having
the same intersection; that is, p1 X ¨ ¨ ¨ X pr “ q1 X ¨ ¨ ¨ X qs . Assume that there are no
non-trivial inclusion relations in either family. Then the two families coincide.
Ş
Proof: For each index ν one has p1 . . . pr Ď q j Ď qν and therefore at least for
one index, say αpνq, the relation pαpνq Ď qv holds. By symmetry, for each µ there
is a βpµq such that qβpµq Ď pµ . Now
pαpβpµqq Ď qβpµq Ď pµ
and there being no non-trivial inclusion relations among the pi ’s we infer that
αpβpµqq “ µ. In a symmetric manner one shows that βpαpvqq “ ν and we can
conclude that α is a bijection from t1, . . . , ru to t1, . . . , su with with φαpvq “ qv o
Problems
Lemma 2.33 Two non-zero divisors a and b in a ring A generate the same principal
ideal precisely when they are related by a unit; that is, when a “ ub where u P A˚ .
Proof: When paq Ď pbq there is a ring element u so that a “ ub, and when
pbq Ď paq it holds that b “ va, Hence a “ vua. And a not being a zero-divisor,
we conclude that vu “ 1. o
In the case of the ring A possessing zero-divisors, things are more complicated.
For instance, in the ring krx, y, zs{pzp1 ´ xyqq it holds true that pzq “ pxzq, but
the two generators z and xz are not related by a unit (see Exercise 2.19 below).
ideals 35
Proposition 2.35 A principal ideal paq is a prime ideal exactly when a is a prime
element.
Proof: Recall what a being prime means: If a|bc then either a|b or a|c. Trans-
lated into a statement about ideals x|y means that y P pxq. Hence, bc P paq is
equivalent to a|bc, and b P paq or c P paq to a|b or a|c . o
2.36 An other aspect of prime numbers (which in fact is the usual definition) is
they can not be further factored; that is, their sole factors are 1 and the prime
itself. Irreducible polynomials in krxs share this quality except they can be
changed by non vanishing constant factors (of course f “ c´1 ¨ c f for any
non-zero constant c). Generalizing these two notions, one says that a non-
zero element a from the ring A is irreducible if it is not a unit, and if a relation irreducible elements
a “ bc implies that either b or c is a unit. (irreducible elementer)
Proof: Assume that a is prime element in A and that a “ bc. Since a is prime
it holds true that b “ xa or c “ xa for some x P A, say b “ xa. Substituting
back yields a “ xca and cancelling a, which is legal since A is supposed to be a
domain, we arrive at 1 “ xc which shows that c is a unit. o
36 ma4200—commutative algebra
The converse of this proposition is not generally valid, in fact one is tempted
to say that in most rings it does not hold, but we shall shortly meet classes
(see Proposition 2.41 on page 36) of rings where it is true. There are simple
examples of irreducibles not being prime in quadatic extensions of Z. We give
?
one in the ring Zr ´5s below ( Example 2.14 on page 37).
these rings together with three of their cousins, but because fundamental rings
like Z and the polynomial ring krxs are pid’s, the pid’s merit an entrance early
in the play.
2.39 One of the particular properties enjoyed by a principal ideal domain is
there is no distinction between maximal and non-zero prime ideals.
Proposition 2.40 In a principal ideal domain A, any non-zero prime ideal is
maximal.
Euclidean rings
The classical division algorithm for integers, which also goes under the name
of Euclid’s algorithm (you certainly learned it in school), ensures that the ring
of integers Z is a pid:
Proposition 2.42 The ring Z of integers is a principal ideal domain.
Proof: Let a be a non-zero ideal in Z, and let n be the smallest positive
number in a. Any other element a from a may be divided by n to give a
relation a “ q ¨ n ` r where the remainder r is of absolute value less than
n. Now, r “ a ´ q ¨ n is an element of a, and this contradicts the minimality of n
unless r “ 0; therefore a “ q ¨ n, and the ideal a equals the principal ideal pnq. o
ideals 37
2.43 The argument above is a classic but not limited to the ring of integers.
It goes through once there is a function on A for which there is a Euclidian
algorithm; that is, a function δ on A assuming values in N or N0 with the
following property2 : 2
One is not confined
to use functions with
o For any pair x and y of elements in A there are elements q and r in A so values in N; functions
with values in any
that x “ yq ` r and δprq ă δpyq. well-ordered set will
do.
Such a function is called a Euclidian function, and a domain A is said to be a Euclidian functions
Euclidean domain if it possesses one. (Euklidske funksjoner)
The absolute value used in the proof of Proposition 2.42 above is one exam- Euclidean domain
(Euklidske områder)
ple, but there are many others. For instance, on the polynomial ring krxs over
a fields, putting δp f q “ deg f ` 1 if f ‰ 0 and δp0q “ 0 we get one; indeed, the
Euclidean condition is fulfilled by the procedure of long division. The point
with these Euclidian functions is that they forces a domain to be a pid as in
the following proposition; whose proof is mutatis mutandis the same as that of
Proposition 2.42.
Corollary 2.45 Let k be a field. Then the polynomial ring krxs is a pid.
Examples
2.12 In the polynomial krxs over a field k principal ideals p f pxqq with f irre-
ducible, are maximal ideals. The quotient K “ krxs{p f pxqq is a field which is
obtained from k by adjoining a root of f . If you wonder what that root is, it is
just the residue class rxs of the variable x. This illustrates the devise that what
matters in modern mathematics is “what objects do, not what they are”.
2.13 The quotient Rrxs{px2 ` 1q is isomorphic to C as one sees mapping
x to i. In a similar vein, if p is a prime number, the polynomial Φ p pxq “
x p´1 ` x p´2 ` . . . ` x ` 1 is irreducible over Q, so that Qrxs{Φ p pxq is a field.
Sending x to a primitive p-root of unity ξ , gives an isomorphism with Qpξq.
2.14 Simple and concrete examples of irreducible elements that are not prime
?
are found in the ring Zr ´5s where among others the relation
? ?
2 ¨ 3 “ p1 ` i 5q ¨ p1 ´ i 5q (2.3)
?
holds. So, for instance, 2 is not a prime element since neither 1 ` i 5 nor
38 ma4200—commutative algebra
?
1 ´ i 5 is a unit (units have absolute value 1, cfr. Exercise 1.10 on page 14). The
?
element 2 is however irreducible in Zr ´5s, for if 2 “ zw, one has 2 “ }z}}w},
?
and as the square of the absolute value of members of Zr ´5s are natural
numbers, this entails that either }z} “ 1 or }w} “ 1; hence, again in view
of Exercise 1.10, either z or w is a unit. Of course, the three other numbers
appearing in (2.3) are irreducible as well, and Exercise 2.20 below asks you to
check this. For a generic example of irreducible elements not being prime, see
Exercises 2.43 and 2.44 on page 52.
2.15 (The Gaussian integers) The absolute value works as a Euclidean function
on the ring Zris. Indeed, geometrically the Gaussian integers form the integral
lattice in the complex plane; that is, the set of points both whose coordinates
are integers. Given two Gaussian integers a and q with q ‰ 0, the distance
from a to the nearest point in the integral lattice is obviously less then half
the diagonal of a lattice square; that is, there is an element c P Zris so that
?
}a{b ´ c} ď 2{2 ă 1. Putting r “ qpa{q ´ cq, we have a “ cq ` r with }r} ă }q}.
Problems
2.14 Given two ideals pnq and pmq in Z. Show that pnq Ď pmq if and only if
m|n. Conclude that the partially ordered set I pZqztp0qu of non-zero ideals in
Z is isomorphic to the the set of natural numbers ordered by reverse divisibil-
ity.
2.15 Let n and m be two natural numbers. Describe pn, mq and pnq X pmq.
2.17 Let p be a prime not congruent one modulo four. Show that the polyno- ˇ
mial x2 ` 1 is irreducible over the field F p , and that F p rxs{px2 ` 1q is a field
?
isomorphic to F p p ´1q. How many elements does it have?
2.18 Into which of the fields F3 , F5 and F7 is there a map of rings from Zris?
If there is one, describe the kernel.
2.19 Show that the units in krx, y, zs{pzp1 ´ xyqq are the constants k˚ . Show ˇ
that there is no unit u ‰ 1 so that uz “ xz. Conclude that z and xz are not
ideals 39
related by a unit even though pzq “ pxzq. Hint: Killing z gives a ring-map
krx, y, zs{pzp1 ´ xyqq Ñ krx, ys. Setting x “ 1, gives a ring map krx, y, zs{pzp1 ´
xyqq Ñ krz, ys{pzp1 ´ yqq.
2.20 Referring to Example 2.14 show that the three other involved numbers 3,
? ?
1 ` i 5 and 1 ´ i 5 are irreducible.
?
2.21 Show that Zr ´2s is principal ideal domain. Hint: Prove that A is
Euclidean with the absolute value as a Euclidean function.
?
2.22 Let A “ t n{2 ` im{2 3 | n, m P Z u. Prove that A is a subring of C. Prove
that A is Euclidean
Zorn’s lemma
Zorn’s lemma which is one of a few theorems that for some reason keep being
called lemmas, is usually attributed to Max Zorn, but as often happens, it can
be traced further back. At least Felix Hausdorff published versions of it some
ten years before Zorn. Any how, "Zorns lemma" is good name (so good that
an experimental and non-narrative film made by Hollis Frampton in 1970 was Max August Zorn
called "Zorns lemma"). (1906–1993)
German mathematician
2.46 A maximal element x in the partially ordered set Σ is one for which there
is no strictly larger element; that is, if y ě x then y “ x. One should not Maximal elements
(Maksimale elementer)
confuse "maximal elements" with "largest elements" the latter being elements
larger than all other elements in Σ. A partially ordered set can have several
maximal elements whereas a largest element, if there is one, is unique. There
is of course, analogous notions of minimal elements and least elements.
A partially ordered set is said to be linearly ordered or totally ordered if any Linearly ordered sets
two of its elements can be compared. Phrased differently, for any pair x, y of (Lineært ordnede
mengder)
elements either x ď y or y ď x should hold. A chain in Σ is a linearly ordered Totally ordered sets
subset of Σ. The chain is bounded above if for some element x P Σ it holds true (Totalt ordnede mengder)
that y ď x for all elements y in the chain, and then of course, x is called an Chains (Kjeder)
upper bound for the chain. Similarly, the chain is said to be bounded below when Bounded above (Opptil
begrenset)
Upper bounds (Øvre
skranker)
Bounded below (Nedtil
begrenset)
40 ma4200—commutative algebra
having a lower bound in Σ; that is, an element x P Σ satisfying x ď y for all Lower bound (Nedre
members y of the chain. skranke)
Theorem 2.47 (Zorn’s lemma) Let Σ be a partially ordered set in which every
chain is bounded above. Then Σ possesses a maximal element.
2.48 A chain C in Σ is called saturated or maximal if it is not properly contained Saturated or maximal
in any larger chain, or phrased differently, C is a maximal chain in Σ; that is, chains (Mettede eller
maksimale kjeder)
if C1 is another chain with C Ď C1 , then C “ C1 . A chain is saturated precisely
when it impossible to insert any new element in-between two members of
C. As an illustration of the mechanism of Zorn’s lemma, let us prove the
following
Proposition 2.49 Let C be a chain in the partially ordered set Σ. Then there is a
saturated chain containing C.
Theorem 2.50 (The Basic Existence Theorem for Ideals) Assume given
a ring A, an ideal a in A and a subset S not meeting a. Then there exists an ideal b
maximal subjected to the two following conditions
o S X b “ H;
o a Ď b.
ideals 41
Proof: We apply the proposition with S merely consisting of the unit element,
that is S “ t1u. The maximizing ideal is proper and not contained in any other
proper ideal. Hence it is maximal. To prove the second statement, apply the
first to the zero ideal. o
a “ t a P A | an P a for some n P N u.
‘
‘
The elements of a can also be characterized as the elements in A whose
residue class in A{a is nilpotent. Along the same line, taking a to be the zero
‘
ideal, we see that p0q is the set of nilpotent elements in A.
‘
2.53 The first thing to establish is that the radical a in fact is an ideal.
‘
Lemma 2.54 Let a be an ideal in the ring A. Then the radical a is an ideal.
42 ma4200—commutative algebra
2.56 An ideal a in A is a radical if it equals is own radical; i.e. it holds true that
‘
a “ a. One easily verifies that the radical of an ideal is a radical ideal: that Radical ideals (Radikale
‘‘ ‘ idealer)
is, one has a “ a. In the same manner with prime ideals and maximal
ideals, radical ideals can be characterized in terms of quotients:
Proposition 2.57 An ideal a in the ring A is a radical ideal if and only if the
quotient A{a is reduced.
by hypothesis. o
2.58 The radical of an ideal a must be contained in any prime ideal containing
it because if an P a and a Ď p with p prime, it holds that a P p. The converse is
also true and hinges on the basic existence theorem above.
‘
Proposition 2.59 Assume that a is a proper ideal in the ring A. The radical a
equals the intersection of the prime ideals containing a; that is,
‘ č
a“ p.
a Ď p, p prime
‘ Ş
Proof: We already saw that a Ď a Ď p p, so assume that a is an element
‘
not lying in the radical a. We shall apply Theorem 2.50 on page 40 with S
being the set t an | n P N u of powers of a, which obviously is closed under
‘
multiplication. Since a R a, it holds that S X a “ H, and by the theorem we
conclude that there is prime ideal p containing a disjoint from S; that is, a R p.
o
The special case that a “ p0q merits to be pointed out:
Corollary 2.60 The set of nilpotent elements in A equals the intersection of all
prime ideals in A; that is
‘ č
p0q “ p.
p prime
ideals 43
2.61 One might be tempted to discard the prime ideals not being minimal
among those containing a from the intersection in Proposition 2.59, and thus
write č
‘
a“ p. (2.4)
p minimal prime over a
Lemma 2.63 For every finite collection tai u of ideals in A it holds true that
Ş ? aŞ
i ai “ i ai .
‘
Proof: When an element a from A belongs to each of the radicals ai , there
are integers ni so that ani P ai . With n “ max ni , it then holds true that
an P ai for each i, and thus a P
‘Ş
Ş ‘ i ai . This shows that one has the inclusion
‘Ş
i a i Ď a
i i . The other inclusion is straightforward. o
Examples
‘ ‘
2.16 Even if a power of every element in a lies in a, no power of a will in
general be contained in a. A simple but typical example being the ideal a “
px1 , x22 , x33 , . . .q generated by the powers xii in the polynomial ring krx1 , x2 , x3 , . . .s
in countably many variables. The radical of a is equal to the maximal ideal
m “ pxi |i P N0 q generated by the variables, but no power of m is contained in a.
Indeed, the exponent needed to force a power of xi to lie in a, tends to infinity
with i.
2.17 The operation of taking radicals respects finite intersections, but this is no
more true when it comes to infinite intersections. For instance, if p is a prime
number, one has pr Z “ pZ and therefore r pr Z “ pZ, but evidently it
‘ Ş ‘
Problems
2.23 That minimal and maximal prime ideals exist ensues directly from ˇ
Zorn’s lemma as this exercise shows. Let tpi uiP I be a chain of prime ideals.
Ť Ş
Show that both the union iP I pi and the intersection iP I pi are prime ideals.
44 ma4200—commutative algebra
Show that every prime ideal containing a given ideal a contains a a prime
ideal minimal over a.
2.25 Let A be any ring. Show that the set of non-zero divisors in A form a ˇ
saturated multiplicative set. Conclude that the set of zero divisors is the union
of prime ideals.
Proposition 2.65 Let A be a ring and m a proper ideal in A. The following three
statements are equivalent.
o The group of units and the complement of m coincide; that is, A˚ “ Azm ;
o The ideal m is maximal and consists of the elements a such that 1 ` a is invertible.
Proof: To see the last statement ensues from the second, let a be a member of
m. Then of course 1 ` a is not in m and hence is invertible.
Finally, let m be maximal and assume that any element shaped like 1 ` a is
invertible. Let x be an element not in m. Since m is maximal, it holds true that
m ` pxq “ A; hence x “ 1 ` a for some a P m, and x is invertible. o
ideals 45
Proposition 2.67 Let A be a ring. The Jacobson radical of A consists of the ring
elements a so that 1 ` xa is invertible for all x P A.
2.68 Assume that A and B are two local rings whose maximal ideals are m A
and mB respectively. A map of rings φ : A Ñ B is said to be a local homomor-
phism, or a map of local rings, φpm A q Ď mB . Equivalently, one may request that Local homomorphism
(Lokal homomorfi)
φ´1 pmB q “ m A .
Examples
2.18 The set of rational functions over a field k that may be expressed as
Ppxq{Qpxq with Ppxq and Qpxq polynomials and Qp0q ‰ 0, is a local ring whose
maximal ideal equals the set of the functions vanishing at the origin. The
evaluation map Ppxq{Qpxq ÞÑ Pp0q{Qp0q identifies the residue field with the
field of complex numbers C.
2.19 Let p be a prime number and let Zp pq be the ring of rational numbers
expressible as n{m where the denominator m is relatively prime to p. Then
Zp pq is a local ring whose maximal ideal is generated by p.
More is true, the only ideals in Zp pq are the principal ideals ppv q; indeed,
every rational number lying in Zp pq may be written as pν n{m with ν ě 0 and
neither n nor m having p as factor. And among these ideals ppq contains all the
others.
2.20 In a polynomial ring Crx1 , . . . , xr s for all points a P Cr the ideal of polyno-
mials vanishing at a is a maximal ideal. It follows that the Jacobson radical f
Crx1 , . . . , xr s equals p0q; i
46 ma4200—commutative algebra
2.21 Assume that p and q are two prime numbers. Let A be the ring of ratio-
nal numbers with denominator relatively prime to pq. That is A “ t n{m |
n, m P Z, pm, pqq “ 1 u. The principal ideals by ppq and pqq are the only two
maximal ideals in A, and JpAq “ ppq X pqq “ ppqq.
Problems
2.26 Show that a ring has just one prime ideal if and only if its elements are ˇ
‘
either invertible or nilpotent. Prove that this is the case if and only if A{ p0q is
a field.
2.27 With reference to Exercise 1.22 on page 21, show that the ring Z p of
p-adic integers is a local ring whose sole maximal ideal is generated by p.
2.28 Let A be the subring of Q whose elements are the rational numbers a
expressible as a “ m{n where n does have neither 2 nor 3 as factor. Show that
A has two maximal ideals p2q and p3q whose intersection equals p6q. What are
the two residue fields? Show that 1 ` a is invertible in A for all members a P p6q.
2.31 Let f pxq be any polynomial in krxs where k is a field. Show that krxs{p f pxqq ˇ
is semi-local.
ś ś
Proposition 2.69 The ideals of A “ Ai are all of the form 1ďiďr ai
1ď i ď r ś
where each ai is ideal in Ai . It holds true that A{a » 1ďiďr Ai {ai . The ideal p is a
prime ideal if and only if pi “ Ai for all but one index i.
2.70 It is appropriate to give a comment about the zero ring at this stage. In
Proposition 1.27 the idempotents ei ’s are not required to be different from
zero, but if ei “ 0, of course ei A is the zero ring, and does not contribute in a
significant way to the product (it holds true that 0 ˆ A “ A). This is particularly
pertinent for the formulation of Proposition 2.69; it might happen that ai “ Ai
so that A{ai is the zero ring.
A classical result that at least goes back to third century AD is the so called
Chinese Remainder Theorem, a more informative name would be the Theorem
of Simultaneous Congruences: As long as the moduli n1 and n2 are relatively
prime, two congruences x ” y1 mod n1 and x ” y2 mod n2 have a common
solution. It seems that the first written account of this result is found in the
book Sunzi Suanjing by a Chinese mathematician “Master Sun”— hence the
Chinese theorem.
This can of course be generalized to any number of congruences as long
as the moduli are pairwise prime, and there is a formulation for general ring
Some old Chinese
with the moduli replaced by ideals. To formulate the appropriate condition on mathematics.
the ideals, the notion of comaximal ideals is introduced. Two ideals a and b
are said to be comaximal is a ` b “ A, equivalently, one may write 1 “ a ` b with Comaximal ideals
a P a and b P b. (Komaksimale idealer)
Given a finite collection tai u1ďiďr ideals in the ring A. There is an obvious
map
ź
AÑ A{ai
i
Theorem 2.71 (The Chinese Remainder Theorem) Let A be a ring and let
ta u be a finite collection of pairwise comaximal ideals. Then A{a1 X . . . X ar »
śi 1ďiďr
1ďiďr A{ai .
Problems
2.33 Show that 28y ´ 27z solves the simultaneous congruences x ” y mod 9
and x ” z mod 4.
2.35 Assume that a1 , . . . , ar are pair-wise comaximal ideals. Show that one
has a1 ¨ . . . ¨ ar “ a1 X . . . X ar .
Even more forceful techniques are available to handle graded rings which
satisfy appropriate finiteness conditions. For instance, when all the homoge-
nous components Rn are finite dimensional vector spaces over some field
k Ď R0 , the so called Hilbert function h R pνq “ dimk Rn is a very strong invari-
ant of R.
A the present stage of the course we merely scratch the surface of the
theory of the graded rings, but they will reappear at several later occasions.
2.72 A graded ring R is a ring together with a decomposition of the underlying Graded rings (graderte
abelian group as a direct sum ringer)
à
R“ Rν (2.5)
νPZ
one can not attribute a well-defined degree to it, but it will be considered to
be homogenous of any degree. From a decomposition as in (2.5) it ensues
ř
that elements a in R can be expressed as sums a “ ν aν whose terms av
are homogenous of degree ν with merely finitely many being different from
zero. The aν ’s are uniquely determined by a and go under the name of the
homogenous components of a. Homogenous components
Notice that R0 ¨ R0 Ď R0 , so R0 is a subring of R. Similarly, for any ν it holds (homogene komponenter)
Proposition 2.75 Let a be an ideal in the graded ring R. The following three
statements are equivalent.
Proof: That the first two statements are equivalent, is just a rephrasing of
how homogenous ideals were defined. So assume that a is a homogenous
50 ma4200—commutative algebra
2.76 A rich source of graded rings are the quotients of polynomial rings by
homogenous ideals, or more generally the quotient of any graded ring by a
homogenous ideal.
Proof: This follows immediately from the the direct sum decompositions R “
À À
ν Rν and aν “ ν aν . Notice first Ri X a “ ai and it holds
ř
that Ri {ai Ď R{a.
Hence any class ras P R{a with a decomposing as a “ i ai in homogeneous
ř
terms of degree i decomposes as ras “ i rai s, where we can consider the
rai s’s to be elements in R{a or in P Ri {ai . Moreover, the classes rai s are unique
ř ř
because is i ai and i bi were two such decompositions inducing the same
ř
element in R{a, it would hold true that i pai ´ bi q P a. The ideal a being
homogeneous and each term ai ´ bi being homogeneous of degree i, it would
follow that ai ´ bi P a, and hence rai s “ rbi s. o
Since already R is the direct sum R “ i,j k ¨ xi y j , one arrives at the direct sum
À
Problems
2.37 With reference to the Example 2.22 above , show that the subring R0
of elements of degree zero in the case α “ 1, β “ ´1 is isomorphic to the
polynomial ring over k in one variable. Describe Ri for all i.
2.40 Assume that k is an algebraically closed field and that f px, yq is a homo-
geneous polynomial in krx, ys. Show that f px, yq splits as a product of linear
factors. Hint: If f is of degree d, it holds that f px, yq “ yd f px{y, 1q. Consider
f px{y, 1q as a polynomial in t “ x{y.
À
2.43 Assume that A “ iě0 Ai is a graded integral domain with A0 being
a field. Show that any element homogenous of degree one is irreducible.
Hint: Work with components of highest degree.
2.44 Let A “ Zrx, y, z, ws{pxy ´ wzq show that the class of x is irreducible but
not prime.
the Zariski topology after Oscar Zariski, one of the great men of algebraic
geometry. This topological space depends functorially on the A, a ring map
φ : A Ñ B induces a map φ̃ : Spec B Ñ Spec A simply by sending a prime φ in
B to the inverse image φ´1 ppq (which is a prime ideal in A).
The spectra of rings enter as building blocks in Alexander Grothedieck’s
scheme theory , which is an infinitely larger ocean with as yet huge unex-
plored region, and the spectra form merely the shore.
In happy marriages the spouses exert a strong mutual influence, so also in
the relationship between algebra and algebraic geometry. Several geometric
features of Spec A are paramount to understanding algebraic properties of the
ring A, and, of course, vice versa. Even modern number theory and arithmetic
are inconceivable without the geometric language along with the geometric
intuition. However, in these matters we shall only superficially scratch the
surface; giveing the basic definitions and a few examples.
There is another geometric construct antecedent of the schemes by about a
century. Basically it goes back to René Descartes’s idea of using coordinates
and equations to describe geometric objects. We have all experienced parts of
the menagerie of plane curves and surfaces in the space. In general, subsets of
Cr being the common zeros of a set of polynomials, are called varieties and are Varieties (varieteter)
the subjects of interest for many algebraic geometers.
Prime spectra
2.78 Being a geometric gadget, the prime spectrum is endowed with a topol-
ogy which is called the Zariski topology. This topology is best defined by giving The Zariski topology
the closed subsets. These are denoted Vpaq where a is any ideal A, and given (Zariski topologien)
There are some axioms to be verified. First of all, Vp0q “ Spec A and VpAq “
H, so the empty set and the entire space are both closed. Secondly, we must
check that unions of finitely many closed subsets are closed (it suffices to
check it for unions of two) and that the intersection of any family of closed
sets is closed. To the former, observe that Vpaq Y Vpbq Ď Vpabq since both
ab Ď a and ab Ď b hold, and the other inclusion Vpabq Ď Vpaq Y Vpbq follows
since ab Ď p implies that either a Ď p or b Ď p according to Proposition 2.24 on
page 2.24 above. Hence Vpaq Y Vpbq “ Vpabq.
ř
To the latter, notice the trivial fact that i ai lies in p if and only if each of
the ai ’s lies in p. We have shown
Lemma 2.79 Let A be a ring.
o Vp0q “ Spec A and Vp1q “ H
o For any ideals a and b in A it holds true that Vpaq Y Vpbq “ Vpabq; Oscar Zariski
(1899–1986)
Russian-born American
ř Ş
o For any family tai uiP I of ideals one has Vp iP I ai q “ iP I Vpai q;
mathematician
o If a Ď b, then Vpbq Ď VpFaq
‘
o Vpaq “ Vp aq.
2.80 The Zariski topology has certain peculiar features never met when work-
ing with mundane topologies like the ones of manifolds. For instance, there
are lots of points in Spec A that are not closed; so in particular the prime
spectra tend to be seriously non-Hausdorff. One has:
Lemma 2.81 The closed points in Spec A are the maximal ideals.
Proof: We saw that every proper ideal is contained in a maximal ideal (The-
orem 2.51 on page 2.51); hence Vpaq will always have a maximal ideal as
member. So if tpu is closed; that is, equal to Vpaq for some a, the prime ideal p
must be maximal.
On the other hand whenever m is maximal, obviously Vpmq “ tmu since no
prime ideal strictly contains m. o
Examples
We do not intend to dive deeply into a study of prime spectra, but to give an
idea of what can happen, let us figure out which of the topological spaces with
only two points can be a prime spectrum.
There are three non-homeomorphic topologies on a two-point set; the
discrete topology with all points being close (and hence all being open as
well), the trivial topology whose sole closed sets are the empty set and the
entire space, and finally the so-called Sierpiński space, a two-point space with
just one of the points being closed (and consequently the other being open).
And two of these occur as Zariski topologies.
54 ma4200—commutative algebra
2.24 The direct product of two fields A “ k ˆ k1 has merely the two prime ideals
p0q ˆ k1 and k ˆ p0q which both are maximal. Hence Spec k ˆ k1 consists of two
points and is equipped with the discrete topology.
2.25 The ring Zp pq of rational numbers expressible as fractions with a de-
nominator prime to p has just two prime ideals, namely p0q and the principal
ideal ppq (Example 2.19 on page 45). Hence tp0qu is an open set being the
complement of the closed point ppq. Hence Spec A is the Sierpiński space.
2.26 Finally, the trivial topology having no closed point, can not be the Zariski
topology of any non-empty prime spectrum; in every ring different from the
null-ring there are maximal ideals, and the spectrum of the null-ring is empty.
K
2.82 The spectrum Spec A depends functorially on the ring A. If φ : A Ñ B
is a map of rings, pulling back ideals along φ takes prime ideals to prime
ideals; indeed, let p Ď B be prime and assume that ab P φ´1 ppq. This means
that φpaqφpbq P p, and so either φpaq P p or φpbq P p. In other words, either a lies
in P φ´1 ppq or b lies there. We thus obtain a map φ̃ : Spec B Ñ Spec A.
Proposition 2.83 The map φ̃ is continuos.
Proof: Let a Ď A be an ideal. The one has φ̃´1 Vpaq “ VpaBq; indeed, tautolog-
ically it holds true that a Ď φ´1 p if and only if φpaq Ď p. o
A byproduct of the proof is that the inverse image under φ̃ of the closed set
Vpaq is homeomorphic to Spec B{aB. Indeed, the prime ideals in B{aB are
in a one-to-one correspondence with the prime ideals in B containing aB
(Theorem 2.15 on page 29) , and these are, as we saw in the proof, precisely
the points in Spec B mapping to points in Vpaq. Moreover, the whole lattice of
ideals I pB{aBq is isomorphic to the lattice of ideals in B containing aB. That
takes care of the topology; closed sets correspond to closed sets.
Proposition 2.84 Let φ̃ : Spec B Ñ Spec A be induced by φ : A Ñ B. Then the
inverse image φ̃´1 Vpaq is homeomorphic to Spec B{aB. In particular, for any point
p P Spec A the fibre over p equals Spec B{pB.
2.85 The Zariski topology has a particular basis of open sets called the
distinguished open sets. For each element f P A there is one such open The distinguished open
set Dp f q whose members are the prime ideals not containing f ; that is, sets (Særskilte åpne
mengder)
Dp f q “ t p | f R p u. This is an open set since the complement is the closed set
Vp f q.
Lemma 2.86 The distinguished open sets form a basis for the topology on Spec A.
Proof: A typical open subset U is the complement of a set shaped like Vpaq.
Now, a prime p does not contain a if and only if there is an element f lying in
Ť
a but not in p. Hence U “ f Pa Dp f q, and we are through. o
ideals 55
Problem 2.45 Let A and B be two rings. Show that Spec pA ˆ Bq is the disjoint
union of Spec A and Spec B. M
‘
Problem 2.46 Show that Vp aq “ Vpaq. M
Complex Varieties
For any S Ď Crx1 , . . . , xr s of polynomials, the zero set VpSq is defined as the set
in Cr of simultaneous zeros of the polynomials in S; that is,
Curves in plane and surfaces in the space, as we know them fro earlier course
are most often of this type with S just containing one element, the equation of
the gadget under consideration. There is one difference however, we consider
points with complex coordinates not only reals.
Example 2.27 The good old parabola equals Vpy ´ x2 q in C2 , or more precisely
the old chap consists of the real points (those with both coordinates being real)
of Vpy ´ x2 q. K
2.87 One easily convinces oneself that the ideal a generated by S has the same
zero-set as S (sums of functions vanishing at point vanish there too, and even
more, multiples of one vanishing vanishes), so without loss one may confine
the study to sets shaped like Vpaq. One has the following formulas, where a
and b are ideals in Crx1 , . . . , xr s:
Preliminary version 0.1 as of 2018-09-13 at 14:13 (typeset 3rd December 2018 at 10:03am)—Prone
to misprints and errors and will change. 2018-09-13 Corrected proof of Prop 3.11.
13/09/2018 Added hypothesis that Rbe a doamain in Proposition 3.20.
3.1 To be precise a ufd is a domain where every non-zero element which is not
a unit, can be expressed in an essentially unique way as a product
a “ p1 ¨ . . . ¨ pr
One can always attempt a recursive attack. Any ring element a which is not
irreducible, is a product a “ a1 a2 of non-units. These being irreducible makes
us happy, but if they are not, they are in their turn products a1 “ a11 a12 and
a2 “ a21 a22 . If some of the fresh factors are not irreducible, we split them into
products of non-units. Continuing like this we establish a recursive process
which, if it terminates, yields a finite factorization of a into irreducibles.
However, the process may go onfor every—like it will for e.g. a “ sin z—but
in many cases there are limiting condition making it end. For instance, in the
case of Z, the number of steps is limited by the absolute value }a}, and in the
case of polynomials in krxs by the degree of a.
3.3 The second condition is the uniqueness requirement which is much more
of an algebraic nature, and hinges on the condition that irreducible elements
be prime. In fact, in any ring the uniqueness condition holds automatically for
finite factorizations into prime elements:
Lemma 3.4 Let A be a ring. Assume that tpi u1ďiďr and tqi u1ďiďs are two collec-
tion of prime elements from A whose products agree; that is it holds true that
p1 ¨ ¨ ¨ pr “ q1 ¨ ¨ ¨ q s .
Examples
The main examples of factorial rings are principal ideal domains and polyno-
mial rings over those, as we shortly shall see, and producing other examples
demands some technology not available to us for the moment. So we confine
ourselves to examples of non-factorial rings.
3.1 The classical example of a ring which is not factorial, which is ubiqui-
tous in texts, is the ring Zri 5s. In it 6 has two distinct factorizations in irre-
‘
3.2 The standard example from algebraic geometry is the quotient ring1 1
A geometer would
A “ krX, Y, Z, Ws{pXY ´ ZWq. Indicating the class of a variable by a lowercase call it "the cone over
a quadratic surface in
version of the name; the relation projective 3-space
xy “ zw (3.1)
unique factorization domains i 59
y2 ´ xpx ´ aqpx ´ bq
with a, b P C are called elliptic curves and have been at the centre-stage of
algebraic geometry since its beginning. They are closely related to the so-
called elliptic functions, and in fact, they were the very starting point for the
development of algebraic geometry.
3.4 The polynomial y2 ´ xpx ´ aqpx ´ bq is irreducible in krx, ys for any field k.
Indeed, if it were not, it would have a linear factor, and one could write
with α, β and γ constants from k and not both α and β being zero.
If β ‰ 0, one may substitute y “ ´β´1 pαx ` γq into this equation. The right
side then vanishes, and we obtain the polynomial identity
Irreducibles in a ufd
3.5 As we saw (Proposition 2.38 on page 35), in a domain prime elements are
always irreducible, but if the domain is factorial, the converse holds as well.
Among domains with an appropriate finiteness condition—like the Noetherian
domains we shall come to—this even characterises the factorial domains
(Lemma 3.4 above).
60 ma4200—commutative algebra
Proposition 3.6 For members of a ufd being prime is equivalent to being irre-
ducible.
xy “ p1 ¨ ¨ ¨ ps ¨ q1 ¨ ¨ ¨ qr
a ¨ r1 ¨ ¨ ¨ r m “ p1 ¨ ¨ ¨ p s ¨ q1 ¨ ¨ ¨ qr
3.7 In a ufd any two elements have a greatest common divisor; that is an element The greatest common
d so that the principal ideal pdq is minimal among the principal ideals con- divisor (Største felles
divisor)
taining both principal ideals paq and pbq where a and b are the two elements
under consideration. Expressed in terms of divisibility, there is a d so that d|a
and d|b and any other x dividing both a and b divides d; that is x|a and x|b
implies that x|d. The greatest common divisor of a and b is not unique, but
only determined up to an invertible factor; however the principal ideal pdq is
unambiguously defined.
3.8 The two elements also have a least common multiple. This is an element m The least common
in A so that the principal ideal pmq is maximal among the principal ideals multiple (Minste felles
multiplum)
contained in paq and pbq; or phrased in terms of divisibility, it holds that a|m
and b|m, and for any other member x of A it ensues from a|x and b|x that m|x.
Again, merely the principal ideal pmq is unambiguously defined.
Proposition 3.9 In a udf any two elements have a greatest common divisor and a
least common multiple.
Proof: Let a and b be the elements. Proceed to write down factorizations say
a “ p1 ¨ ¨ ¨ pr and b “ q1 ¨ ¨ ¨ qs into irreducibles, and pick up the “common
factors”: Reordering the factors, we find a non-negative integer t so that
ppi q “ pqi q for i ď t and ppi q ‰ pqi q for t ą i. Then d “ p1 ¨ ¨ ¨ pt is a greatest
common divisor. It might of course happen that no ppi q equals any pq j q, in
which case t “ 0, and the greatest common divisor equals one.
To lay hands on a least common multiple, make the product of all the
ideals ppi q and all the ideals pq j q not found among the ppi q’s. This product is a
principal ideal, and any generator serves as a least common multiple. o
unique factorization domains i 61
Kaplansky’s criterion
Proposition 3.11 A domain A is a ufd if and only if every non-zero prime ideal
contains a prime element.
Proof: The implication one way is clear. Let p be a non-zero prime ideal and
consider any of its non-zero elements. It factors as a product of primes, and
one of the factors must lie in p.
To prove the other implication it suffices, in view of Lemma 3.4 on page 58,
to show that any non-zero member of A is either a unit or a finite product of
prime elements, so let Σ be the set of elements in A that can be expressed as
a finite product of one or more prime elements. It is certainly multiplicative
closed, and A having at least one maximal ideal it is not empty (maximal
ideals are prime and contain prime elements by assumption). We contend that
Σ coincides with the set of non-zero non-units of A.
Assume this is not true; that is, there is a member x of A, neither zero nor a
unit, that does not lie in Σ. Then pxq X S “ H since if xy “ p1 ¨ ¨ ¨ pr with the pi ’s Irving Kaplansky
(19171–2006)
being prime elements and y not being a unit, it follows by an easy induction Canadian mathematician
on r that x belongs to S. By the Basic Existence Theorem (Theorem 2.50 on
page 40), there is a prime ideal in A maximal subjected to not meeting S
and containing x. That prime ideal is not the zero ideal, and by assumption
there is therefore a prime element lying in p, which is a contradiction since all
primes lie in Σ. Hence Σ fills up the entire set of non-zero non-units in A. o
An immediate corollary is the following (which also can be proved in several
other ways):
up to an invertible factor.
Lemma 3.14 (Gauss’ lemma) Assume that A is a ufd. Let f and g be primitive
polynomials in Arxs. Then the product f g is primitive.
It ensues immediately from the lemma that for any two polynomials f and g it
holds true that c f g “ uc f c g with u P A˚ ; in other words, the content depends
up to units on the polynomial in a multiplicative manner
Proof: Write f “ 0ďiďn ai xi and g “ 0ďiďm bi xi . Let d be a non-unit in A.
ř ř
The polynomial f being primitive, there is a least i0 so that d does not divide
ai0 and ditto a least j0 so that d does not divide b j0 . Consider the coefficient of
xi0 ` j0 in the product f g; that is, the sum
ÿ
ai b j .
i` j“i0 ` j0
Theorem 3.16 If A is a ufd, then the polynomial ring Arxs is a ufd. The ir-
reducible elements in Arxs are the irreducibles in A and the primitive, irreducible
polynomials.
Induction and the fact that krxs is a ufd immediately give the corollary that
polynomial rings over fields are ufd’s. Ot is also worth while mentioning that
the same applies to polynomials rings over the integers.
a ¨ p “ b ¨ g1 ¨ ¨ ¨ g r
a i f i “ bi g i
hence ai “ vi bi with vi P A˚ o
3.19 Factoring a polynomial into a product of irreducibles, or for that matter,
showing a polynomial is irreducible, is often an unpleasant task. That the
polynomial is homogenous might sometimes be helpful because then one
knows a priori that the irreducible components are homogeneous; indeed, one
has the following proposition.
Proof: Let ai be the irreducible factors of a and develop each ai as the sum
ř
ai “ ν ai, ν of the homogeneous components. Denote by ai, νi the non-zero
component of ai of lowest degree. Since the degree of every element is non-
negative, it holds true that
ź ź
a“ ai “ ai, νi ` terms of higher degree,
i i
ś
and i ai is non-zero as R is assumed to be a domain. But now, homogeneous
components are unambiguously defined, and a is homogenous. Hence the
sum of the high degree terms vanishes, and we have expressed a as a product
of homogeneous elements
ź
a“ ai, νi .
i
By induction on the degree, each ai, νi has only homogeneous irreducible
components, and the same applies therefore to a. o
64 ma4200—commutative algebra
Example 3.6 The polynomial y p ´ x q are irreducible when p and q are rel-
atively prime. To see this give krx, ys the grading for which deg x “ p and
deg y “ q. An irreducible polynomial has both non-zero terms of the form α ¨ yn
and of the form β ¨ x m , unless it equals x or y. If it additionally it is homoge-
neous, it holds true that nq “ mp and n “ ap and m “ bq for some natural
numbers a and b. It follows that y p ´ x q is irreducible since any irreducible
factor is not reduced to x or y. K
âĂć
Problem 3.1 Assume that ppxq is a monic polynomial in Zrxs wich factors
as ppxq “ rpxqspxq in Qrxs. Show that rpxq and spxq both lie in Zrxs and are
monic. Hint: multiply by the least common multiples of denominators of
coefficents and appeal to Gauss’s lemma. M
?
Problem 3.2 Let d “ 2k be an even integer and consider the ring Zri 2ks ˇ
?
with k an odd natural number. Show that p “ p2, i 2kq is not a principal ideal
?
but that p2 “ p2q. Prove that Zri 2ks is not a udf M
?
Problem 3.3 Consider the ring Zri 14s. Prove that ˇ
? ?
34 “ p5 ` 2i 14qp5 ´ 2i 14q
?
and show that the involved elements all are irreducible elements of Zri 14s.
M
Lecture 4
Modules
Preliminary version 1.05 as of 2018-10-02 at 14:37 (typeset 3rd December 2018 at 10:03am)—To be
completed. Prone to misprints and errors and will change.
2018-08-21: Rewritten paragraph 4.23 Proposition 4.26 corrected; added exercise 4.20
27/08/2018 Finished the paragraph 4.48 on page 87 about split exact sequences;
29/08/2018 Reworked section 4.5 about exact sequences.
15/09/2018 Reworked the snake section. Added exercise 4.15. Added examples 4.12, 4.13 and
4.14.
17/09/2018 Added example 4.18 exercises 4.25 and 4.26.
18/09/2018 Added example 4.24 on page 93; Added exercise 4.38 on page 95;
25/09/2018 Rewritten proof of Splitting Criterion, Prop 4.51; Corrected misprints; changed
exercise 4.38 slightly. Added a new exercise 4.16.
26/09/2018 Added an exercise about elliptic curves and free modules, ex 4.27 on page 84.
Added exercise 4.34 on page 89.
27/09/2018 Corrected several minors misprints.
Extended problem 4.34 about additive functors. Several minor improvements.
02/10/2018 Added exercise 4.5. Added example 4.17 Added a simple lemma snake-section
(lemma 4.65).
Along with every ring comes a swarm of objects called modules; they
are the additive groups on which the ring acts. The axioms for modules
resemble the axioms for a vector spaces, and modules over fields are in fact
just vector spaces. Over general rings however, they are much more diverse
and seriously more complicated than vector spaces. Ideals for instance, are
modules, and any over-ring is a module over any subring to mention two
instances. An abelian group is nothing but a Z-module and a module over the
polynomial ring krts over a field k is just a vector space over k endowed with
an endomorphism; so the module theory encompasses all abelian groups and
the entire linear algebra!
group, which will be written additively, on top of which lies a linear action of
the ring A. Such an action is specified by a map A ˆ M Ñ M, whose value at
pa, mq will be denoted by a ¨ m or simply by am. It is subjected to the following
66 ma4200—commutative algebra
four conditions:
o apm ` m1 q “ am ` am1 ;
o pa ` a1 qm “ am ` a1 m;
o 1 ¨ m “ m;
o a ¨ pa1 ¨ mq “ paa1 q ¨ m.
Examples
4.1 The primordial examples of modules over a ring A are the ideals a in A
and the quotients A{a. Already here, the difference from the case of vector
spaces surfaces; fields have no non-zero and proper ideals. There are also the
"subquotients" b{a of two nested ideals. Of course, these examples include the
ring itself; every ring is a module over itself.
4.2 Another examples more in the flavour of vector spaces are the direct
sums of copies of A. The underlying additive group is just the direct sum
A ‘ A ‘ . . . ‘ A of a certain finite number, say r, copies of A. The elements are
r-tuples pa1 , . . . , ar q, and addition is performed component-wise. The action of
A is also defined component-wise: a ¨ pa1 , . . . , ar q “ pa ¨ a1 , . . . , a ¨ ar q. We insist
on this being an additive1 construction and shall write rA for this module, not 1
The direct sum plays
Ar as is common usage in linear algebra. an additive role in the
category of A-modules,
the multiplicative role
4.3 Another familiar class of modules are the abelian groups. They are noth-
is taken by another
ing but modules over the ring Z of integers. An integer n acts on an element construct, the tensor
from the module by just adding up the appropriate number of copies of the product. But that’s for
a later chapter.
element and then correcting the sign.
instance, krx, ys is a krxs-module as is krx, x´1 s. And krxs will be a module over
the subring krx2 , x3 s.
? ?
If m and n are two natural numbers the ring Zr m{ns is a Zr ms-module
(and, of course, it is a Z-module, and also Zr1{ns-module for that matter).
More generally, any ring homomorphism φ : A Ñ B induces an A-module
structure on B through the action a ¨ b “ φpaqb of an element a P A on b P B.
One says that B is an A-algebra. Algebras (Algebraer)
Problem 4.1 Let A and B be two rings and assume that B has an A-module ˇ
structure compatible with the ring structure; i.e. a ¨ bb1 “ bpa ¨ b1 q. Show that
there is ring homomorphism A Ñ B inducing the module structure. M
underlying abelian groups that respects the action of A; that is, φpamq “ aφpmq
for all a’s in A and all m’s in M. Simply said, a module homomorphism is just
an A-linear map from M to N. With this notion of morphisms, the A-modules
form a category Mod A .
4.3 The set Hom A pM, Nq is naturally contained in the set HomZ pM, Nq of
group homomorphism from M to N (which are just the additive maps) and
consists of those commuting with the action of A on both M and N. It is well-
known that the sum of two additive maps is additive (left for the zealous
students to check ), and when both commute with the actions of A, the sum
does so as well. Therefore Hom A pM, Nq is an abelian group, and defining a ¨ φ
as the map sending m to aφpmq gives it an A-module structure. One must of
course verify that a ¨ φ is A-linear, but this is easy: If b P A is another element, 2
For non-commutative
one finds rings A the set
a ¨ φpbmq “ apbφpmqq “ bpaφpmqq, Hom A pM, Nq is merely
an abelian group in
where the first equality holds since φ is A-linear and the second because A is general. It does not
carry an A-module
commutative2 . structure unless fur-
ther hypotheses are
4.4 The module Hom A pM, Nq depends functorially on both the variables M
imposed on A.
and N, and historically, it was one of the very first functors to be studied. The
dependence on the first variable is contravariant—the direction of arrows are
reversed— whereas the dependence on the second is covariant—directions
68 ma4200—commutative algebra
are kept. The induced maps are just given by composition. Of course, such
constructions are feasible in all categories, what is special in Mod A is that
Hom A pM, Nq is an A-module and the induced maps are A-linear. The techni- M
φ
/N
cal name is that category Mod A has internal homs—the set of maps stay within
ψ
the family! ψ˚ pφq
To be presice let M, N and L be three A-modules and ψ : N Ñ L an A-linear L
where the maps are composable3 and a and a1 denote ring elements. 3
That is, φ and φ1 are
maps from M to N and
ψ and ψ1 from N to L.
Problems
4.4 Let p and q be two prime numbers. Show that HomZ pZ{pZ, Z{qZq “ 0 if ˇ
p ‰ q, and that HomZ pZ{pZ, Z{pZq » Z{pZ.
4.5 Let a and b be two ideals in the ring A. Show that there is a canonical
isomorphism Hom A pA{a, A{bq » pa : bq{b.
Submodules
4.6 A submodule N of an A-module M is a subgroup closed under the action Submodules (Undermod-
of A; in other words, for arbitrary elements a P A and n P N it holds true that uler)
an P N, and of course, N being a subgroup the sum and the difference of two
elements from N belong to N.
modules 69
Examples
4.6 Ideals in the ring A are good examples of submodules, and by definition
they constitute all submodules of A.
4.7 If a Ď A is an ideal and M an A-module, the subset aM of M formed by all
multiples am with a P a and m P M is a submodule.
4.8 Given an ideal a in A. The set p0 : aq M of elements in M annihilated by all
members of a form a submodule. It holds true that Hom A pA{a, Mq “ p0 : aq M .
4.7 Just like the ideals in A the submodules of a given A-module M form a
partially ordered set4 denoted I pMq under inclusion. 4
The set I pMq forms
Ş
The intersection iP I Ni of a collection tNi uiP I of submodules of M is a what is called a com-
plete lattice. A partially
submodule. It is the largest submodule of M contained in all the submodules ordered set is called a
from the collection. In the similar way, the smallest submodule containing lattice when every pair
ř of elements possesses
all the modules in the collection is the sum iP I Ni whose elements are finite a least upper bound
A-linear combinations of elements from the Ni ’s; that is, elements shaped like and a greatest lower
bound, and it is said
ÿ to be complete if the
ai mi , (4.1) same is true for any
subset. In our case the
greatest lower bound is
where mi P Ni , and the ai ’s are elements from A only finitely many of which the intersection and the
are non-zero. More generally, for any set S Ď M there is a smallest submodule least upper bound the
of M containing S; it is called the submodule generated by S and consists of sum.
Quotients
70 ma4200—commutative algebra
4.9 Just as with ideals in a ring, one can form quotient of a module by a
submodule.
Let M be the module and N the submodule. From the theory of abelian
groups we already know that the two underlying additive groups have a
quotient group M{N, which is formed by the cosets rms “ m ` N. Endowing
M{N with an A-module structure amounts to telling how elements a P A act
on M{N, and one does this simply by putting a ¨ rms “ rams. Of course, some
verifications are needed. Routinely one checks that the class rams does not
depend on the representative m, which is the case since apm ` Nq “ am ` aN “
am ` N. Secondly, the module axioms in paragraph 4.1 must be verified; this is
straightforward and left to the zealous students.
4.10 By definition the canonical map π : M Ñ M{N that sends m to the class
rms is A-linear, and characterized by the universal property that any A-linear
map vanishing on N factors through it:
Proof: The proof is mutatis mutandis the same as for (abelian) groups. The M
π
/ M{N
map φ is constant on the residue classes rms “ m ` N, and ψprmsq is defined
as (and compelled to be) that constant value. Since φ is A-linear, the constant φ
ψ
Proof: Routine. o
o pN ` N 1 q{N 1 » N{N X N 1 ;
Proof: Routine. o
L
Proof: As mentioned above, the quotient N{ im φ serves as the cokernel. The
kernel is the usual tangible subset of M consisting of the elements sent to zero.
o
Example 4.9 In Mod A the fabulous axiom cited above boils down to the
obvious: The kernel of the cokernel and the cokernel of the kernel both equal
the image. (If you find this rather cryptical than obvious, think twice!!) K
Direct products
4.18 In this section we work with a collection tMi uiP I of modules over the
ring A. From earlier courses direct product of a collection of groups is well
ś
known, and here we consider the direct product iP I Mi of the underlying Direct products of mod-
additive groups of the Mi ’s. The elements are strings or tuples pmi qiP I indexed ules (Direkte produkter
av moduler)
by the set I, and the addition is performed component-wise; i.e. pmi q ` pm1i q “
pmi ` m1i q. In case I is finite, say I “ t1, . . . , ru, an alternative notation for a
tuple is pm1 , . . . , mr q. The actions of A on the different Mi ’s induce an action
ś
on iP I Mi , likewise defined component for component; a ring element a acts
like a ¨ pmi q “ pa ¨ mi q. The module axioms in paragraph (4.1) are cheap to verify,
ś
and we have an A-module structure on iP I Mi .
ś
The projections πi : iP I Mi Ñ Mi are A-linear since the module operations
of the product are performed component-wise.
Direct sums
À
4.19 The direct sum of the module collection tMi u is denoted iP I Mi , and Direct sums of modules
is defined as the submodule of the direct product consisting of strings m “ (Direkte summer av
moduler)
pmi qiP I with all but a finite number of the mi ’s vanishing—or so to say, there
is an i0 (that depends on m) such that mi “ 0 for i ě i0 . Obviously this is a
submodule, and again, the projections are A-linear.
When the index set I is finite requiring strings to merely have finitely many
non-zero components imposes no constraint, so in that case the direct sum
and the direct product coincide. However, when the index set I is infinite, they
are certainly not isomorphic; they are not even of the same cardinality. For
instance, the direct sum of countably many copies of Z{2Z is countable (being
modules 73
the set of finite sequences of zeros and ones) whereas the direct product of
countably many copies of Z{2Z has the cardinality of the continuum (every
real number has a 2-adic expansion).
Universal properties
4.20 Both the product and the direct sum are characterised by universal prop-
erties. It is noticeable that these properties are dual to each other; reversing all
arrows in one, yields the other. For this reason the direct sum is frequently
called the co-product in the parlance of category theory.
We proceed to describe the two universal properties and begin with the N
φ
/ śiPI Mi
direct product. In that case, the set-up is formed by an A-module N and a
πi
collection of A-linear maps φi : N Ñ Mi . The conclusion is there exists a
&
φi
ś
unique A-linear map N : N Ñ iP I Mi such that πi ˝ φ “ φi . Indeed, this Mi
amounts to the map φpnq “ pφi pnqqiP I being A-linear.
In the case of the direct sum the universal property does not involve the
À
projections, but rather the natural inclusions ι j : M j Ñ iP I Mi that send an
m P M j to the string having all entries equal to zero but the one in slot j which
equals m. The given maps are maps φi : Mi Ñ N, and the conclusion is that
N of φ À
Mi
iPI
O
À
there exists a unique map iP I Mi Ñ N so that φ ˝ ι j “ φj . The map φ is
compelled to be defined as φi
ιi
` ˘ ÿ Mi
φ pmi qiP I “ φi pmi q,
iP I
and this is a legitimate definition since merely finitely many of the mi ’s are
non-zero.
Problem 4.7 Work out all the details in the above reasoning. M
4.21 With the stage rigged as in the previous paragraphs we round off the
discussion of the universal properties of direct products and direct sums by
offering equivalent formulations in terms of the hom-modules:
Notice that in the first isomorphism, which involves the contravariant slot, the
direct sum is transformed into a direct product. It further warrants a special
comment that when the index set is finite, the direct product coincides with
the direct sum, and the proposition may be summarized by saying that the
hom-functor commutes with finite direct sums. In the vernacular of category
theory one says that it is additive in both variables.
74 ma4200—commutative algebra
The slogan is: Killing one addend of a direct sum yields the sum of the others.
On the other hand, im e X ker e “ 0 since if x “ epyq lies in ker e, it holds that
epxq “ e2 pyq “ epyq “ x, and hence x “ 0. This takes care of one implication.
Suppose then that M “ N ‘ N 1 and let π : M Ñ N and ι : N Ñ M respec-
tively be the projection and the inclusion map, then ι ˝ π “ id N . Put e “ π ˝ ι.
Then
e2 “ pι ˝ πq ˝ pι ˝ πq “ ι ˝ pπ ˝ ιq ˝ π “ ι ˝ π “ e.
o
Problems
4.12 Let tMi uiP I be a family of submodules of the A-module N. Assume that
they comply to the following rule: For any index ν P I and any finite subset
ř
J Ď I not containing ν, the intersection of Mν and jP J M j vanishes; that is,
ř ř
M X iP J M j “ p0q. Prove that iP I Mi is isomorphic with the direct sum
Ài
i P I Mi .
Unions of submodules
There is no reason that the union of a collection of submodules in general
should be a submodule; no more than the union of two lines through the
origin in space is a plane! This an additive issue (unions of submodules are
obvious closed under multiplication by ring elements) and takes place in
most abelian groups: If H1 and H2 are two subgroups of an abelian group H,
neither containing the other, their union cannot be a subgroup. Indeed, if x1
lies in H1 but not in H2 and x2 in H2 but not in H1 , the sum x1 ` x2 cannot
belong to H1 Y H2 . Suppose for instance it belonged to H1 , then x2 would lie
there too which it was chosen not to.
4.29 However, there is one natural condition that ensures the union to be a
submodule. One says that the collection is directed if for any two modules Directed collection of sub-
modules (Rettet samling
in the collection there is a third containing both; that is—if tMi uiP I is the
av undermoduler)
collection—for any pair Mi and M j there should be an index k so that Mi Ď Mk
Ť
and M j Ď Mk . The union iP I Mi will then be closed under addition (and as
multiplication poses no problem, will thence be a submodule); indeed, let x
and y be two members of the union. This entails there are indices i and j so
that x P Mi and y P M j , and the collection being directed, one may find an
index k so that Mi Y M j Ď Mk . Both elements x and y then lies in Mk , and hence
their sum does as well. So the sum belongs to the union. We have proven:
many elements generate B it would hold true that BN “ B for some N. This is
obviously absurd, as p can appear in a denominator with any exponent. K
Cyclic modules
4.32 Modules requiring only a single generator are said to be cyclic or mono- Cyclic modules (Sykliske
genic, and they are omnipresent. Among the ideals, the principal ideals are moduler)
Monogenic modules
precisely the cyclic ones, and more generally, if M is any module and m P M (Monogene moduler)
an element5 , the submodule Am “ t a ¨ m | a P A u is cyclic. 5
The zero module is
Now, assume that M is a cyclic A-module and let m P M be a generator. counted among the
cyclic ones, so m “ 0 is
Multiplication induces an A-linear map φ : A Ñ M that sends a to am, and admitted.
this map is surjective since m was chosen as a generator. The kernel consists
by definition of those a’s that kill m, or which amounts to the same, that kill
M. Hence ker φ “ Ann M, and by Corollary 4.12 on page 70, we arrive at an
isomorphism M » A{ Ann M.
Simple modules
4.34 The simplest modules one can envisage are the ones without other sub-
modules than the two all modules have, and they are simply called simple: A
non-zero A-module M is said to be simple if it has no non-zero proper submod- Simple modules (Simple
ule. Simple modules are cyclic and any non-zero element generates; indeed, moduler)
Problems
4.15 Let N be a submodule of M. Show that if N and M{N both are finitely ˇ
generated, then M is finitely generated as well. Give an example of modules
M and N so that M and M{N are finitely generated but N is not.
4.17 Show that krx, x´1 s is not a finitely generated module over krxs. ˇ
4.18 Show that krxs{krx2 , x3 s is a cyclic module over krx2 , x3 s. What is the ˇ
annihilator? What can you say about krxs{krx2 , x p s where p is an odd prime?
4.19 Assume that k is a field. Consider the polynomial ring krxs as a module ˇ
over the subring krx3 , x7 s. Prove it is finitely generated by exhibiting a set of
generators. Determine the annihilator of the quotient krxs{krx3 , x5 s.
4.20 (Schur’s lemma.) Assume that if M and N are two simple A-module that ˇ
are not isomorphic. Prove that Hom A pM, Nq “ 0. Prove that Hom A pM, Mq “
A{ Ann M.
Problem 4.21 Show that the property, familiar from the theory of vector
ř
spaces, that i ai mi “ 0 implies that ai “ 0 is sufficient for a generating set
m1 , . . . , mr to be a basis. Hint: Consider the difference of two equal linear
combinations of the mi ’s. M
80 ma4200—commutative algebra
The gist of this example is that x and y commute, and this indicates that the
phenomenon is inherent in commutative rings. Any set of generators for an
ideal consisting of at least two elements can never be a basis simply because
the generators commute. K
Free modules
4.36 The lack of bases for most modules, leads to a special status of those
that have one. One says that an A-module F is free if it has a basis. The reason Free modules (Fri
behind the suggestive name “free” is that one may freely prescribe the values moduler)
a linear map takes on the basis elements—a principle that goes under the
name of the Universal Mapping Principle:
Proof: We saw in Example 4.16 above that when a requires at least two
generators, it has no basis and therefore is not free. Nor can principal ideals
generated by a zero divisor be free since if a f “ 0 with a ‰ 0, the relation
a f “ 0 f gives two representations of 0. The other way around, if the non-zero
divisor f is a generator for a, it is basis; indeed f being a non-zero divisor it
can be cancelled from an equality like a f “ b f . o
4.40 Archetypes of free modules are the direct sums nA “ A ‘ . . . ‘ A of n
copies of the ring A which we already met in Example 4.2 on page 4.2. They
come equipped with the so-called standard basis familiar from courses in linear
algebra. The basis elements ei are given as ei “ p0, . . . , 0, 1, 0, . . . 0q with the one
sitting in slot number i.
There is no reason to confine these considerations to direct sums of finitely
À
many copies of A. For any set I, the direct sum iP I A has a standard basis
modules 81
tei uiP I and is a free module; the basis element ei is the string with a one in slot
i and zeros everywhere else.
Proposition 4.41 Assume that F is a free A-module with basis t f i uiP I . Then there
À
is an isomorphism between F and the direct sum iP I A that sends each basis vector
f i to the standard basis vector ei .
À
Proof: By Proposition 4.37 above, we may define a map φ : F Ñ iP I A
À
by sending f i to the standard basis vector ei ; conversely, since iP I A is free,
À
sending ei to f i sets up a map ψ : iP I A Ñ F. These two maps are obviously
mutual inverses. o
Corollary 4.42 Any two bases of a free module have the same cardinality. Two
free modules are isomorphic if and only if they possess bases of the same cardinality.
The common cardinality of the bases for a free module is called the rank of the Rank of a free module
module, and the rank is the sole invariant of free modules; up to isomorphism (Rangen til en fri modul)
it determines the module. When the module is a vector space over a field and
the rank is finite, the rank is just the dimension of the vector space.
Proof: After Proposition 4.41 above we need merely to verify that when two
À À
direct sums iP I A and jP J A are isomorphic as A-module, the index sets I
and J are of the same cardinality. This is well known from the theory of vector
spaces when A is a field, so take any maximal ideal in A and consider the
À À
isomorphic vector spaces iP I A{m and jP J A{m over A{m (isomorphic in
view of exercise 4.10 on page 75). One has a basis of the same cardinality as I,
the other one of cardinality that of J; hence I and J are equipotent. o
Example 4.17 (Free modules with given basis) From time to time it will be
convenient to operate with free A-modules with a given set S as basis. There
is no constraint on the set S, it can be whatever one finds useful. The formal
way to construct AS is as the set of maps α : S Ñ A of finite support; that is the
maps such that αpsq ‰ 0 for at most finitely many members s of S; in symbols
AS “ t α : S Ñ A | α of finite support u.
A ¨ A: “ det A ¨ I (4.2)
and the determinant of a composition being equal to the product of the deter-
minants, we infer that NB{ A pbb1 q “ NB{ A pbqNB{ A pb1 q. The norm is therefore a
multiplicative functions on B with values in A. Moreover, for elements a P A
one has NB{ A paq “ an . This follows because the map ras is just a ¨ idB and
detpa ¨ idB q “ an (the map ras is represented as a scalar matrix in any basis). K
Problems
4.23 Prove that any finitely generated A-module is the quotient of a free ˇ
module of finite rank.
4.24 Show that any A-module is the quotient of a free module. Hint: let ˇ
À
FM “ mP M A be the direct sum of copies of A, one for each element m P M.
? ?
4.25 Let A “ Zr ds. Show that A is a free Z-module with basis 1, d. Use ˇ
?
this to show that NA{Z px ` y dq “ x2 ´ dy2 .
?
4.26 Assume that d is an innteger such d ” 1 mod 4. Let α “ p1 ` dq{2. ˇ
Show that α2 “ α ` pd ´ 1q{4. Prove that A “ Zrαs is free Z-module of rank 2
with 1, α as a basis. Determine the matrix of the map x ÞÑ αx in this basis and
compute the characteristic polynomial.
4.27 Let A “ krX, Ys{pY 2 ´ XpX ´ aqpX ´ bqq be the coordinate ring of an ˇ
elliptic curve and as usual let x be the class of X and y that of Y.
a) Show that krxs is isomorphic with the polynomial ring krXs and that A is a
free krxs-module of rank 2.
b) Show that krys is isomorphic to the polynomial ring krYs and that A is a
free krys module of rank 3.
c) Show that x and y are irreducible elements in A and conclude that A is not
a udf. M
The next series of exercises, which culiminate with problem 4.30, are aimed at
giving an example that countable products of free modules are not neceras-
saryli free
4.28 Show that in a free Z-module every element is divisible by at most
finitely many integers.
4.29 Show that the direct sum of countably many copies of Z is countable,
whereas the direct product of countably many copies is not (it has the cardinal-
ity of the continuum).
4.30 (Infinite products are not always free.) The task is to show that the direct
product P “ iPN Z of a countable number of copies of Z is not a free Z-
ś
module. Aiming for a contradiction, suppose that the product has a basis
iPN Z lies in P and has the standard basis
À
t f i uiP I . The direct sum Q “
modules 85
This sequence is said two be exact if ker ψ “ im ψ. It frequently happens that Exact sequences (Eksakte
such a sequence is part of a longer sequence of maps, extending to the left or følger)
to the right, and then the extended sequence is said to be exact at N as well. A
sequence exact at all places, is simply said to be exact.
4.47 Two special cases warrant mentioning; the first being when M “ 0:
/N / L.
ψ
0
The image of the zero map being the zero submodule p0q, exactness boils
down in this case to ψ being injective. Similarily, when L “ 0, the sequence is
shaped like
/N / 0,
φ ψ
M
0 / ker α / M1 /M / coker α / 0.
K
86 ma4200—commutative algebra
/ M1 /M / M2 /0
α β
0 (4.3)
is called a short exact sequence when it is exact. This means that α is injective,
that β surjective and that im α “ ker β. Of course, the term “short” in the name Short exact sequence
(Korteksakte følger)
implies there are long exact sequence as well, and indeed there are, as we shall
see later on.
It ensues from the First Isomorphism Theorem (Corollary 4.12 on
page 70) that there is a unique isomorphism θ : M2 » M{αpM1 q shaped in a
way that β corresponds to the quotient map. In other words, θ enters in the
following commutative diagram
/ M1 /M / M2 /0
α β
0 (4.4)
α| M1 θ
0 / αpM1 q /M / M{αpM1 q /0
where the maps in the bottom row are respectively the inclusion of im M1
into M and the quotient map. In short, up to isomorphisms all short exact
sequence appear as
0 /N /M / M{N / 0,
where N Ď M is a submodule, and the two maps are respectively the inclusion
and the quotient map.
Example 4.20 (Direct sums) The direct sum M ‘ N of two A-modules fits
naturally into the short exact sequence
0 /N / N‘M /M /0
where the left-hand map is the natural inclusion sending x to px, 0q and the
one to the right is the projection onto M, which maps px, yq to y. If N and N 1
are two submodules of the A-module M, then there is a short exact sequence
0 / N X N1 ι / N ‘ N1 σ / N ` N1 /0
rem 2.71 on page 48) for two ideals may be generalized by saying that the
sequence
where the two maps in the middle are given by the following two assignments
x ÞÑ prxsa , rxsb q and prxsa , rysb q ÞÑ rxsa`b ´ rysa`b . Having four non-zero terms
it is to long to be called short exact, but it may be obtained by splicing the two
short exact sequences
0 / aXb /A / A{a X b /0
and
0 / A{a X b / A{a ‘ A{b / A{a ` b /0
Problem 4.31 Verify that the two maps defined in the example above are
well defined and that the sequence is exact. Deduce the Chinese Remainder
Theorem from it. M
Problem 4.32 Write down a “Chinese sequence” involving three ideals that
generalizes the sequence (4.5) above. Prove it is exact and deduce the Chinese
Remainder Theorem for three ideals from it. Hint: It will have six non-zero
terms. M
/ M1 /M / M2 / 0.
α β
0 (4.6)
is split exact when being isomorphic to the standard sequence of Example 4.20
on page 86. This not only requires that M be isomorphic to the direct sum Split exact sequences
(Splitteksakte følger)
M1 ‘ M2 , but the somehow stronger requirement that there be an isomorphism
inducing the identity on M1 and pairing β with the projection, must be met:
that is, the isomorphism must fit into the following commutative diagram
/ M1 /M / M2 /0
α β
0
»
0 / M1 / M1 ‘ M2 / M2 / 0.
where the maps in the bottom sequence are the projection and the inclusion.
4.49 Of course all sequences are not split exact, and even if the two extreme
modules are the same the middle modules need not be isomorphic. The
88 ma4200—commutative algebra
easiest example is found among abelian group: Both Z{p2 Z and Z{pZ ‘ Z{pZ
appear in the midst of short exact sequences with both extreme modules being
Z{pZ. In general, it is an unsurmountable challenge to classify all possible
middle modules given the two extreme ones.
4.50 There is nice criterion for a short exact sequence to be split involving
only one of the maps α or β; to formulate it we need two new concepts. A
right section for β is an A-linear map σ : N Ñ M such that β ˝ σ “ id N , and a left Right sections (Høyresek-
section for α is one τ : M Ñ M1 with the property that τ ˝ α “ id M1 . sjoner)
Left sections (Venstresek-
Proposition 4.51 (Splitting criterion) Let the short exact sequence sjoner)
/ M1 /M / M2 /0
α β
0
Before starting with the proof we observe that a map β : M Ñ M2 that pos-
sesses a right section σ is automatically surjective since if x P M2 it holds true
that βpσpxqq “ x. Thus, once the map β has a right section its source M will be
isomorphic to M2 ‘ ker β. Moreover, the restriction of β to the image σpM2 q
is easily seen to be an isomorphism so that M is the direct sum of the two
submodules ker β and σpM2 q.
Proof: We start out by assuming that the sequence is split and that θ : M Ñ
M1 ‘ M2 is an isomorphism of the right type. Then one easily verifies that
the maps given by m2 ÞÑ θ ´1 p0, m2 q and m1 ÞÑ θ ´1 pm1 , 0q are sections for
respectively β and α of the right variance and hence the first assertion implies
each of the two others.
Assume then that the map β has a right section σ and define a map Φ : M1 ‘
M2 Ñ M simply by the assignment px, yq ÞÑ αpxq ` σpyq. This map is clearly
A-linear and it lives in the diagram
/ M1 /M / M2 /0
α β
0 O
Φ
0 / M1 / M1 ‘ M2 / M2 /0
ι π
whose rows are exact, and where ι and π are respectively the canonical inclu-
sion and projection. The left square commutes since Φpιpxqq “ αpxq ` σp0q “
αpxq, and that the right does, ensues from the following little calulation
where we use that σ is a right section for β. Now, there is a general fact (called
The Five Lemma, see Exercise ?? on page ?? below) that a map in the midst
of two isomorphisms as in the diagram above is an isomorphism; but it is
also easily checked ad hoc: If 0 “ Φpx, yq “ αpxq ` σpyq, it follows that y “ 0
since the right square above is commutative, hence αpxq “ 0, and consequently
x “ 0 because α is injective. To prove surjectivity of Φ, observe that for z P M,
the difference z ´ σpβpzqq is killed by β, and because ker β “ im α, one has
z ´ σpβpzqq “ αpyq for some y so that z “ αpyq ` σpφpzqq.
Finally, let us treat the case that α has a left section. The map Ψ : M Ñ
M1 ‘ M2 defined by Ψpzq “ pτpzq, βpzqq fits in the following diagram analogue
to the one above
/ M1 /M / M2 /0
α β
0
Ψ
0 / M1 / M1 ‘ M2 / M2 /0
ι π
The right square commutes trivially, and the left coomutes since one has
Problem 4.33 In this exercise the setting is as in the proposition, and the
point is to connect up with the splitting criterion formulated in terms of idem-
potents we gave when discussing direct sums and products (Proposition 4.26
on page 75).
a) Assume that σ is a right section for β. Prove that σ ˝ β is an idempotent
endomorphism of M, and that the corresponding decomposition in a direct
sum is the one in remark abovef; that is M “ ker β ‘ σpM2 q.
b) Assume that τ is a left section for α. Prove that α ˝ τ is idempotent, and that
the corresponding decomposition is M “ αpM2 q ‘ ker τ. M
/ M1 /M / M2 /0
α β
0
F pαq F p βq
0 / FpM1 q / FpMq / FpM2 q /0
is split exact as well. In particular, the functor F transforms finite direct sums
into finite direct sums. Hint: If σ and τ are sections of respectively β and α,
then Fpσq and Fpτq will be sections of respectively Fpβq and Fpαq. M
4.52 Suppose given a short exact sequence like (4.3) above and let N be an
A-module. Applying the covariant hom-functor Hom A pN, ´q to the sequence,
we obtain an induced sequence which is shaped like
β˚ α˚
0 / Hom A pM2 , Nq / Hom A pM, Nq / Hom A pM1 , Nq , (4.8)
and an argument like above shows that this also is an exact sequence.
Problem 4.35 Give the argument referred to in the previous paragraph in
detail. M
4.54 It is common usage to refer to the phenomena described above as saying
that Hom A pN, ´q and Hom A p´, Nq are left exact functors. There are crowds of Left exact functors
(Venstre-eksakte funk-
torer)
modules 91
examples that β ˚ and α˚ are not surjective, so Hom A pN, ´q and Hom A p´, Nq
are seldom exact functors in the sense that they take short exact sequences to Exact functors (Eksakte
short exact sequences. funktorer)
A large part of homological algebra was developed just to describe the
"missing cokernels" coker β ˚ and coker α˚ . In general the only answer to this
challenge is that the two sequences can be extended ad infinitum to the right to
yield long exact sequences that involve so-called Ext-modules. These modules
do not depend on the original short exact sequence only on the modules
involved and N of course (but the maps in the long exact sequence do), and in
some cases these long exact sequences can be controlled.
Two classes of modules stand out namely the ones with the property that
either the functor Hom A pN, ´q or the functor Hom A p´, Nq exact. The former
are called projective modules (they are ubiquitous in commutative algebra, and Projective modules
we shall come back to them) and the latter are the so-called injective modules. (Projektive moduler)
Injective modules
4.55 In paragraph 4.52 above we proved the "only-if-part" of the following (Injektive moduler)
proposition (although we worked with short exact sequences like in (4.3) we
never used that β was surjective).
Proposition 4.56 (Left exactness I) Let the sequence
/ M1 /M / M2
α β
0 (4.9)
be given and assume that β ˝ α “ 0. The sequence is exact if and only if for all
A-modules N the sequence
is exact.
Proof: To attack the "if-part" assume that (4.10) is exact for all A-modules N.
If α is not injective, take N “ ker α, which is non-zero, and let ι be the
inclusion of ker α in M1 . Then α ˝ ι “ 0, but ι is non-zero so α˚ is not injective.
In a similar vein, if the image im α is strictly smaller than the kernel ker β,
take N “ ker β and consider the inclusion map ι of N in M. By choice it holds
that β ˝ ι “ 0, but ι cannot factor though α since im α is strictly contained in im ι.
o
4.57 As alluded to in Paragraph 4.54 above, even if the map β is surjective, the
induced map β ˚ will in most cases not be surjective. An easy example is given
below. That β ˚ surjective means that any A-linear map φ : N Ñ M2 can be
lifted to an A-linear map into M, as illustrated in the following diagram.
/ M1 /M / M2 /0
α β
0 a O
φ
N
92 ma4200—commutative algebra
Proof: Let P be the module. And assume first that P is projective. Let tpi uiP I
be a generating set for P (finite or not, we do not care about the size. One
could e.g. use the canonical free module FP from Example 4.17 on page 81)
and consider the exact sequence
/K /À /P / 0,
φ
0 iP I A
ř
where a string pai qiP I is sent to i ai pi (the string lives in the direct sum so
merely finitely many of the ai ’s are non-zero). Since the pi generate P, this
map is surjective. Now, since P is projective, the identity map idP : P Ñ P
À
can be lifted to a σ : P Ñ iP I A. This means that φ ˝ σ “ idP , and the lifting P
ι
/ P ‘ P1
À
σ is a section of φ . Hence P lies split in iP I A by the Splitting Criterion π
(Proposition 4.51 on page 88). ψ
ψ˝ι P
To prove the other implication, let φ : P Ñ M2 be given assume that P ‘ P1
is free. Consider the map P ‘ P1 Ñ M2 sending px, yq to φpxq. This map can be φ
lifted as P ‘ P1 is free, and restricting the lifted map to P yields a lifting of φ. o M / M2 /0
β
4.60 There is also an assertion dual to the one of Proposition 4.56 above. The Diagram illustrating the last
part of the proof; π is the
proof of the two being quit similar, we leave all the checking to the zealous projection and ι the inclusion.
and not so zealous students; it is a good training for these diagram-arguments. Notice that π ˝ ι “ idP .
The assertion reads as follows:
/M / M2 /0
α β
M1 (4.11)
be given and assume that β ˝ α “ 0. The sequence is exact if and only if for all
A-modules N the sequence
is exact.
4.62 In most cases the map α˚ will not be surjective even if α is. Surjectivity
means that any map φ : M Ñ N can be extended to a map M Ñ N, like in the
modules 93
diagram below
/ M1 /M / M1 /0
α β
0
φ
}
N
Examples
p
0 /Z /Z / Z{pZ /0 (4.13)
0 /0 /0 / Z{pZ,
4.23 We observe that HomZ pZ, Z{pZq “ HomZ pZ{pZ, Z{pZq “ Z{pZ, so the
functor HomZ p´, Z{pZq when applied to the sequence (4.13) in example 4.22
above, yields the exact sequence
p˚
0 / Z{pZ / Z{pZ / Z{pZ.
The map p˚ is just multiplication by p which on Z{pZ is the zero map; hence
the rightmost arrow is zero. The left arrow is therefore an isomorphism; this
ensues from exactness of the sequence, but neither is it hard to verify ad hoc
that it equals the identity map.
Example 4.24 (A projective module that is not free) The prime ideals one meets
in number theory which are not principal are in most cases projective, and
later on we shall prove this for ideals in the so-called Dedekind rings. For
the moment we content ourself with giving just one example illustrating this
phenomenon.
? ?
The example will be the ideal a “ p2, 1 ` i 5q in the ring A “ Zri 5s, which
turns out to be a projective but not a free A-module. Indeed, we shall with an
explicit construction see that a is a direct summand in the free module A ‘ A
and hence it will be projective (Proposition 4.59 on page 92), and since it is not
principal, it is not free.
94 ma4200—commutative algebra
φ: A ‘ A Ñ aĎ A
?
defined by the assignments e1 ÞÑ 2 and e2 ÞÑ 1 ` i 5. We shall identify a
submodule M inside A ‘ A that φ maps isomorphically onto a and thereby
proving that a lies split in A ‘ A. The submodule M in question is generated by
? ?
the two elements a1 “ ´2e1 ` p1 ´ i 5qe2 and a2 “ ´p1 ` i 5qet1 ` 3e2
We begin with checking that the restriction φ| M is surjective. This ensues
from the very definition of φ as the following little calculations show:
? ?
φpa1 q “ ´2 ¨ 2 ` p1 ´ i 5q ¨ p1 ` i 5q “ ´4 ` 6 “ 2,
? ? ?
φpa2 q “ ´p1 ` i 5q ¨ 2 ` 3 ¨ p1 ` i 5q “ p1 ` i 5q.
and since
` ? ˘ ` ? ? ˘
2 p1 ´ i 5qx ` 3y “ ´ p1 ` i 5qp1 ´ i 5q ` 6 y “ 0,
?
it follows that p1 ´ i 5qx ` 3y “ 0 as well. Hence a “ 0, and we are through. K
Problems
of B-modules that regarded as A-modules are free of finite rank. Prove that
rk A F “ rkk E ` rk A G.
In the following exercise we let Pφ ptq “ detpt ¨ idE ´φq be the characteristic
polynomial of an endomorphism φ : E Ñ E of a free A-module of finite rank.
4.37 (Multiplicativity of the characteristic polynomial.) Assume given a commuta- ˇ
tive diagram
0 /E /F /G /0
ψ φ θ
0 /E /F /G /0
ψ
modules
‹‹
95
where the rows are exact and the involved modules are free of finite rank. 0 θ
Prove that det φ “ det ψ ¨ det θ. Show that Pφ ptq “ Pψ ptq ¨ Pθ ptq. Hint: Exhibit a
basis for F in which the matrix of φ has an appropriate block decomposition
(as in the margin). Conclude that det φ “ det θ ¨ det ψ and that tr φ “ tr θ ` tr ψ.
?
4.38 With reference to Example 4.24 above, let A “ Zri 5s and a “ p2, 1 ` i5q.
Let Ψ : 2A Ñ 2A be the A-linear map given by the matrix
˜ ? ¸
´3 1`i 5
Ψ“ ?
1´i 5 ´2
Ψ / 2A
φ
/a / 0,
2A
where φ is the map from Example 4.24, and hence that coker Ψ » a.
b) Show that Hom A pa, Aq » a.
c) Let Φ : 2A Ñ 2A be the map given the matrix
˜ ? ¸
2 1`i 5
Φ“ ?
1´i 5 3
Φ / 2A Ψ / 2A Φ / 2A Ψ / 2A Φ / ...
...
/ M2 / M3 /0
φ1 φ2
M1 (4.14)
α1 α2 α3
0 / N1 / N2 / N3
ψ1 ψ2
where the rows are exact and the squares are commutative. There exists a map
δ : ker α3 Ñ coker α1 rendering the following sequence exact
Why snake?
The reason for the name “Snake Lemma” is apparent when one considers the
diagram below.
There the map δ connecting ker α3 to coker α1 we constructed in the Snake
Lemma, zig-zags like a green snake through the diagram.
M1 M2 M3 0
α1 α2 α3
0 N1 N2 N3
In applications of the Snake Lemma one frequently happens that the maps
φ1 : M1 Ñ M2 is injective and one ψ2 : N2 Ñ N3 is surjective. The diagram we
start with thus is shaped like
/ M1 / M2 / M3 /0
φ1 φ2
0 (4.16)
α1 α2 α3
0 / N1 / N2 / N3 /0
ψ1 ψ2
Such a diagram induces two three term exact sequences, one formed by the
kernels of the αi ’s and one by their cokernels, and point is that the snake map
δ connects these two sequences. In other words, we have a six term exact
sequence
Lemma 4.65 (Snake lemma II) Assume given a commutative diagram with
exact row as in (4.16). Then the six term sequence (4.17) above is exact.
Proof: The sequence is trivially exact at the two extreme slots ker α1 and
coker α3 , and that the snake-part is exact, is just the Snake Lemma, so what
remains to be done is checking exactness at ker α2 and coker α2 , and this
98 ma4200—commutative algebra
Problems
4.39 (The five lemma I.) Use the Snake Lemma to prove the following abbrevi- ˇ
ated and preliminary version of the five lemma in the next exercise. Assume
given a commutative diagram
0 / M1 / M2 / M3 /0
α1 α2 α3
0 / N1 / N2 / N3 /0
with exact rows. If two of the αi ’s are isomorphisms, then the third one is as
well.
M0 / M1 / M2 / M3 / M4
α0 α1 α2 α3 α4
M0 / N1 / N2 / N3 / N4
ψ φ
0 /F /E /A /0
modules 99
4.41 Infer from the previous exercise that if A “ Z and φ has non-vanishing ˇ
determinant, it holds true that the cokernel coker φ is finite and that } det φ} “
# coker φ. Hint: Use the Snake Lemma.
4.42 In the same vein as in exercise 4.41 above, assume that A “ krts is ˇ
the polynomial ring over the field k and that φ has non-vanishing determi-
nant. Prove that coker φ is of finite dimension over k and that deg det φ “
dimk coker φ Hint: Again, the snake is the solution.
0 /N /M / M{N /0
O O O
0 / Ar / A r `s / As /0
with exact rows and all three vertical map being surjective. Then apply the
Snake Lemma and Proposition 4.15 on page 79.
M
Lecture 5
Tensor products
Version 2.1 as of 2018-10-18 at 09:53 (typeset 3rd December 2018 at 10:03am). Prone to misprints
and errors and will change.
The term “tensor” appeared for the first time with a meaning resembling
the current one in 1898. The German physicists Woldemar Voigt used the
word in a paper about crystals. Tensors are these days extensively used in
physics, and may be the most prominent example is the so-called “stress-
energy-tensor” of Einstein. It governs the general theory of relativity and
thereby our lives in the (very) large!
A slightly less influential occurrence took place in 1938 when the American
mathematician Hassler Whitney when working on the universal coefficient
theorem in algebraic topology introduced the tensor product of two abelian
groups. Certain isolated cases had been known prior to Whitney’s work, but
Whitneys construct was general, and it is the one we shall give (although sub-
sequently polished by several mathematicians, in particular Nicolas Bourbaki,
and generalized to modules).
How far apart stress in crystals and the universal coefficient theorem may
appear, the concept of tensors is basically the same—the key word being Woldemar Kummer
(1850–1919)
bilinearity. German physicist
A typical example from the world of vector spaces over a field k, would be
a scalar product on a vector space V, and within the realm of commutative
algebra, the products of an A-algebra B is a good example; the multiplication
map pa, bq ÞÑ ab is an A-bilinear map B ˆ B Ñ B.
5.1 There is naturally also the notion of multilinear maps, which involves more Multilinear maps (Multi-
ś
than two modules. In that case, the source of the map is a product i Mi lineære avbildninger)
5.2 The tensor product captures in some sense all possible bilinear maps
defined on the product of two A-modules M and N, or at least makes them
linear. This rather vague formulation becomes precise when phrased as a
universal property.
5.3 The tensor product is a pair consisting of an A-module Mb A N together
with an A-bilinear map τ : M ˆ N Ñ Mb A N that abide by the following rule: The tensor product
(Tensorproduktet)
!
L
Exsistence
The construction of the tensor product is rather abstract and serves the sole
purpose of establishing the existence. It will seldom be referred to in the
sequel, if at all. To ease getting a grasp on the tensor product remember the
mantra, so true in moderen mathematics: “Judge things by what they do, not
by what they are”.
5.4 The construction starts out with the free A module F “ A Mˆ N on the set
ř
M ˆ N. The elements of F are finite, formal linear combinations i ai ¨ pxi , yi q
with xi P M, yi P M and ai P A. In particular, every pair px, yq is an element of
F, and by definition these pairs form a basis for F. We proceede by letting G
be the submodule of F generated by all expressions either of the form
or of the form
px, ay ` a1 y1 q ´ apx, yq ´ a1 px, y1 q, (5.2)
tensor products 103
where a and a1 are elements from A while x and x1 lie in M and y and y1 in N.
The tensor product Nb A M is defined as the quotient F{G, and the residue
class of a pair px, yq will be denoted by xby. Having forced the two expres-
sions (5.1) and (5.2) above to be zero by factoring out the submodule G, we
have made xby a bilinear function of x and y; that is, the following two rela-
tions hold true in Mb A N:
Proposition 5.5 The pair τ and Mb A N as constructed above satisfy the universal
property in paragraph 5.3; in other words, they are the tensor product of M and N.
wanted map γ : Mb A N Ñ L. L
Elements shaped like xby generate the tensor product, and because the
value at xby of any factorization of β is compelled to be βpx, yq, the unique-
ness of γ comes for free. o
Decomposable tensors.
5.6 For several reasons, tensors of the form xby deserve a special name; they
are dubbed decomposable tensors. Only in a very few highly special cases all Decomposable tensors
elements in a tensor product will be decomposable; the usual situation is that (Dekomponerbare
tensorer)
most are not (A simple example is discussed in Problem 5.13 on page 113
below. See also Example 5.6 on page 120). A general element in Mb A N may
ř
however, be expressed as a finite linear combination i ai ¨ xi byi of decompos-
able tensors since this is already true in the free module F “ A Mˆ N .
104 ma4200—commutative algebra
paxqby “ xbpayq.
Functoriality
Linear maps between A-modules are fundamental tools in algebra, and and
it comes as no surprise that exploring how maps behave when exposed to
tensor products occupies a large of the theory. As a modest start we shall
observe that the tensor product construct is functorial, in the precise meaning
that when considered a function of either variable, it gives a functor Mod A Ñ
Mod A ; so we have to tell how to tensorize maps.
5.9 Any A-linear map φ : M Ñ M1 gives rise to an A-linear map Mb A N Ñ
M1 b A N that on decomposable tensors acts as xby ÞÑ φpxqby: The expression
φpxqby depending bilinearily on x and y, this is a viable definition, and the
resulting map is naturally babtized φbid N .
Obviously, it holds true that ψbid N ˝ φbid N “ pψ ˝ φqbid N when ψ and φ
are two composable maps (the identity holds for decomposable tensors), and
clearly id M bid N “ id Mb A N . This means that the pair of assignments
M ÞÑ Mb A N and φ ÞÑ φbid N
This follows easily from how ´b A N acts on maps together with the basic
bilinear relations in (5.3) on page 103. Indeed, one finds
and the two sides of (5.4) agree on decomposable tensors. Hence they are
equal since the decomposable tensors generate Mb A N.
5.11 The situation is completely symmetric in the two variables, so if ψ : N Ñ
N 1 is a map, there is a map id M bψ from MbN to MbN 1 that sends xby to
xbψpyq, and naturally, on sets φbψ “ pφbid N q ˝ pid M1 bψq.
Some formulas
5.12 When working with tensor products a series of formulas are invaluable.
Here we given the most basic ones revealing the multiplicative nature of the
tensor product; together with the direct sum it behaves in a way resembling
the product in a ring.
Proposition 5.13 Suppose that M, N and L are modules over the ring A. Then
we have the following four canonical isomorphisms.
o Neutrality: Mb A A » M;
o Symmetry: Mb A N » Nb A M;
Neutrality: For the first formula the actions on decomposables are xba Ñ xa
and x ÞÑ xb1. The product xa is bilinear in x and a and therefore the map
xba Ñ xa descends to an A-linear map Mb A A Ñ M which obviously have
x ÞÑ xb1 as inverse.
Symmetry: In this case the short hand assignments are xby Ø ybx. In both
directions the assignments are bilinear in view of the fundamental relations
(5.3), and obviously they define inverse maps.
Associativity: This case is more subtle than one should believe at first sight;
the (very short) shorthand definition would be pxbyqbz Ø xbpybzq, but
this is not viable since xby is not a general member of Mb A N. To salvage the
situation one introduces some auxiliary maps, one for each z P L.
So, fix an element z from L and define an A-linear map ηz : Mb A N Ñ
Mb A pNb A Lq by the assignment xby ÞÑ xbpybzq, which is legitimate since
the expression xbpybzq is bilinear in x and y (the third variable z is kept
fixed).
Obviously the map ηz ptq is linear in z and a priori being linear in t, it de-
pends bilinearily on t and z. We infer that sending tbz to ηt pzq induces a map
pMb A Nqb A L Ñ Mb A pNb A Lq. On decomposable tensors this map behaves as
wanted; that is, it sends pxbyqbz to xbpybzq.
A symmetric construction yields a map the other way which sends a de-
composable tensor xbpybzq to pxbyqbz. Finally, these two maps are mutually
inverses since they act as inverse maps on the decomposables, and the decom-
posables generate the tensor products.
Distributivity: Another way of phrasing this is to say at the tensor product
respects directs sums, and this we already established in Proposition 5.10
above. However to at least halfway fulfil our promise, we vaguely indicate the
beginning of an ad hoc proof in the flavour of the three preceding parts of the
proof. It is based on the short hand version px, yqbz Ø pxbz, ybzq, and the
salient point is to extend these to maps defined on the entire tensor products
using bilinearity; but, of course, details are left to the benefit of the zealous
students. o
5.14 It is worth while to dwell a little on the associativity. Since the paren-
theses are irrelevant, we may as well skip them and write Mb A Nb A L for
MbpNb A Lq (or for that matter for pMb A Nqb A L). According to the universal
property of the tensor product bilinearity is the clue for defining maps having
source Mb A N, and there is a similar trilinearity principle for defining maps
sourced in a triple tensor product Mb A Nb A L. It suffices to specify φ on de-
composable tensors xbybz as long as the specifying expression is trilinear in
x, y and z; the precise statement is a as follows:
Proof: The argument is mutatis mutandis the same as we gave in the proof
of Proposition 5.13 concerning the associative law: Fix an element z P K and
consider φpx, y, zq; it depends in a bilinear manner on x and y and hence gives
rise to a linear map ηz : Mb A N Ñ L. The dependence of ηz on z obviously
being linear, assigning ηz ptq to tbz for t P Mb A N and z P K is a bilinear in z
and t and therefore yields the desired map pMb A Nqb A K Ñ L.
And again, the map is unambiguously determined since its values are
prearranged on the decomposable tensors which generate pMb A Nqb A K. o
As a final comment, there is nothing special about the number three in this
context. A similar statement—that is, a principle of multi-linearity—holds
true for tensor products with any number of factors, but we leave that to the
imagination of the reader.
Problem 5.1 Consider the four isomorphisms in Proposition 5.13 on page 105.
Be explicit about what it means that they are functorial (in every variable in-
volved) and prove your assertions. M
Proposition 5.17 Let A be a ring. For any two ideals a and b in A it holds true
that A{a b A A{b » A{pa ` bq. In particular, one has A{a b A A{b “ 0 if and only if
the two ideals a and b are comaximal.
Proof: Sending a pair prxsa , rysb q from A{a ˆ A{b to the element rxysa`b in
A{a ` b is well defined and bilinear; indeed, changing x (resp. y) by a member
of a (resp. b) changes xy by a member of a (resp. b) as well, so the map is well
defined, and a product clearly depends bilinearily on the factors. This induces
a map A{ab A A{b Ñ A{a ` b.
One the other hand, the tensor product A{ab A A{b is a cyclic A-module
generated by 1b1 because rasa brbsb “ ab ¨ 1b1, and clearly elements from both
the ideals a and b kill it since
The proposition shows that the tensor product of two non-zero modules
very well may vanish, and for cyclic modules this happens precisely when
the respective annihilators are comaximal. We also observe that taking b “ a
yields an isomorphism A{ab A A{a » A{a.
5.18 A modest instance of the tensor product being zero, is found among
finite abelian groups. Powers of two relatively prime integers p and q are
comaximal—for natural numbers µ and ν it holds that ppµ , qν q “ Z—and we
infer that
Z{pµ ZbZ Z{qν Z “ 0.
We also infer that if the two natural nu,bers satify µ ď ν it holds that true
that
Z{pµ ZbZ Z{pν Z » Z{pµ Z
for any prime number p since ppµ , pν q “ ppminpν,µq q. Together with the formu-
las from Proposition 5.13 and the Fundamental Theorem for Finitely Abelian
Groups, these two formulas make it clear how to compute the tensor product
of any pair of finite abelian groups.
5.19 In the second example we show that the tensor product of two free
modules is free.
Proposition 5.20 Assume that E and F are free A-modules. Then the tensor
product Eb A F is free. More precisely, if tei uiP I and t f j u jP J are bases for respectively E
and F, the tensors ei b f j with pi, jq P I ˆ J form a basis for Eb A F.
This proposition holds true regardless of the cardinalities of I and J, but the
case when E and F are of finite rank, warrants to be mentioned specially. One
may deduce the finite tank case from Proposition 5.13 by a sta esforward
induction, however we offer another and simple proof that is generally valid.
Corollary 5.21 If E and F are free A-modules of finite ranks n and m respectively,
the tensor product Eb A F is free of rank nm. In particular, for vector spaces V and W
of finite dimension over a field k it holds true that dimk Vbk W “ dimk V ¨ dimk W.
of an element y P F. All the ai pxq’s and all the b j pyq’s depend linearly on their
arguments.
For each pair of indices µ P I and ν P J the expression aµ pxqbν pyq depends
bilinearly on x and y and therefore xby Ñ aµ pxqbν pyq gives a map δµν : EbF Ñ
A. This map vanishes on ei b f j unless i “ µ and j “ ν, and takes the value one
on eµ b f ν .
With the δµν ’s up our sleeve, the rest is a piece of cake: Just apply δµν to a
potential dependence relation
ÿ
cij ¨ ei b f j “ 0
The word “functorial” refers to all three variable. The dependence is covariant
in L and contravariant in M and N (A sanity check is that the variances are the
same on both sides!). We refrain from diving into the general details afraid of
laying a notational smokescreen over the matter which is quit simpe, but in
Example 5.1 below we discuss the situation with M being the variabale.
Proof: The salient point is that Hom A pM, Hom A pN, Lqq is canonically isomor-
phic to the space of bilinear maps M ˆ N Ñ L; and once one realizes that, the
proposition becomes just another reformulation of the Universal Property of
the tensor product.
One may think about the members of Hom A pM, Hom A pN, Lqq as being
maps Φpx, yq defined on M ˆ N: When x is a specified member of M, the
corresponding map Φpx, ´q from N to L is given as y ÞÑ Φpx, yq. That Φpx, ´q
is A-linear, means that Φpx, yq is linear in y, and that Φpx, ´q depends linerly
on x, means that Φpx, yq is linear in x: Hence at the end of the day Φpx, yq is
bilinear!
And that’s it: By the Universal Property of the tensor product any such
map can be written unambiguously as Φpx, yq “ φpxbyq with a linear map
φ : Mb A N Ñ L. o
θ M ˝ pβbid N q˚ “ β˚ ˝ θ M1 , (5.5)
or in other words verify the diagram below commutes, whose vertical maps
are the isomorphisms θ M and θ M1 furnished by Proposition 5.23:
p βb id N q˚
Hom A pM1 b A N, Lq / Hom A pMb A N, Lq
θ M1 θM
Hom A pM1 , Hom A pN, Lqq / Hom A pM, Hom A pN, Lqq
β˚
β˚ θ M1 pφqpx, yq “ φpβpxqbyq.
Problem 5.2 Convince yourself (along the lines of this example) that the
isomorphisms in Proposition (5.23) are functorial in N and L as well. M
Right exactness
In analogy with the notion of left exactness, which we discussed in connec-
tion with the hom-functors, a covariant and additive functor F between two
module-categories1 Mod A and ModB is said to be right exact if it transforms 1
Or more generally,
exact sequences shaped like between two abelian
categories
Our approach relies on Proposition 5.23 above and illustrates the general fact
that adjoint functors tend to share exactness properties; if one is exact in some
sense, the other tends to be exact in a related sense. It is possible to give a
proof of right exactness based on the construction of the tensor product. This
is however tedious, cumbersome and not very enlightening, and according to
our mantra should be avoided.
It is common usage to call N a flat A-module if the functor p´qb A N is exact; Flat modules (Flate
i.e. when it transforms injective maps into injective maps. moduler)
p βb id N q˚ pαb id N q˚
0 / Hom A pM2 b A N, Lq / Hom A pM1 b A N, Lq / Hom A pM0 b A N, Lq
»
»
θ M2
0 / Hom A pM2 , Hom A pN, Lqq / Hom A pM1 , Hom A pN, Lqq / Hom A pM0 , Hom A pN, Lqq
β˚ α˚
The next step is to evoke Proposition 5.23 above and replace the complex
Hom A pM‚ b A N, Lq with Hom A pM‚ , Hom A pN, Lqq; the latter is displayed
as the bottom line of the grand diagram. The crux of the proof is that this
latter sequence is exact, again by Left Exactness II, so once we know
that the two rows in the grand diagram are isomorphic (as sequences) we
are through. But indeed, they are since with the vertical maps being the
canonical isomorphisms from Proposition 5.23 (Adjointness), the two
squares commute according to Example 5.1 above. o
5.26 Proposition 5.17 on page 107 describes the tensor product of two cyclic
module. An analogous result holds true with just one of the modules being
cyclic while the other can be arbitrary:
Proposition 5.27 Let a Ď A be a an ideal and M and A-module. Then one has a
canonical isomorphism Mb A A{a » M{aM that sends mbras to rams.
0 /a /A / A{a / 0,
112 ma4200—commutative algebra
ab A M
α /M / Mb A A{a / 0,
because the tensor product is right exact. The map α sends abx to ax, hence
its image is equal to aM, and we are done. o
Example 5.2 Be aware that the tensor product can be a bloodthirsty killer.
Injective maps may cease being injective when tensorized, and they can even
become zero. The simplest example is multiplication by an integer n, that is;
the map Z Ñ Z that sends x to nx. It vanishes when tensorized by Z{nZ.
This also illustrates the fact that the functor ´b A N is not always exact, even
though always being right exact. In this example the exact sequence
0 /Z n /Z / Z{nZ / 0,
Proposition 5.29 Let a P A and assume that M and N are A-modules such that
M is divisible by a and a P Ann N, then Mb A N “ 0.
Proof: The short argument goes like this. Every x P M is of the form ax1 for
some x1 P M, so that xby “ ax1 by “ x1 bay “ 0, and as the decomposable
tensors generate Mb A N we are through. o
tensor products 113
Problems
5.5 Show that Q{ZbZ Q “ 0. Show that one has QbZ Q » Q, but that ˇ
Q{ZbZ Q{Z “ 0. Hint: Proposition 5.29.
5.6 Let a be a proper ideal in the ring A. Assume that M is an A-module for
which there is surjection M Ñ A{a. Show that aM ‰ M.
5.7 Let M be finitely generated module over the ring A: Show that Mb n ‰ 0 ˇ
for any natural number n. Hint: Exhibit a surjective map M Ñ A{p. Proceed
by induction on n and right exactness of the tensor product.
5.9 Assume that G is a finite abelian group. Show that GbZ Q “ 0. Leopold Kronecker
(1823–1891)
5.10 Let A be a domain contained in a field K and let M be an A-module. German mathematician
Assume that Ann M ‰ p0q. Prove that Kb A M “ 0. The Kronecker product
(Kronecker-produktet)
5.11 Let A Ñ B be a surjective ring homomorphism. Show that for any two ˇ
B-modules M and N (which automatically are A-modules) it holds true that ˇ
ˇ
Mb A N “ MbB N.
5.12 Show that ZrisbZ Zris is a free abelian group of rank 4 while ZrisbZri sZris “ ˇ
Zris and is of rank two as an abelian group.
5.13 Merely in a very few highly special cases will all elements in a tensor
product be decomposable; the commonplace situation is that most are not. A
simple example is W “ Vbk V where V is a two-dimensional vector space over
k. Let te1 , e2 u be a basis for V. The tensor product W is of dimension four with
a basis tei be j u where 1 ď i, j ď 2. Let xi be coordinates relative to this basis;
that is, any vector v is expressed as v “ x1 ¨ e1 be1 ` x2 ¨ e1 be2 ` x3 ¨ e2 be1 ` x4 ¨
e2 be2 ,
a) Establish that he decomposable tensors are shaped like
5.14 (The rank stratification.) Let V be a vector space over a field k of finite
dimension. The dual space V ˚ “ Homk pV, kq consists of linear functionals on
V and is a vector space of the same dimension as V. Chose a basis tei u for V
and let tφi u be the dual basis for V ˚ . That is, the functionals φi are defined as
φi pe j q “ 0 when i ‰ j and φi pei q “ 1.
a) Let W be a second vector space of finite dimension over k with basis t f j u.
Prove that the assignment φbw to v ÞÑ φpvqw induces an isomorphism
»
Γ : V ˚ bk W Ñ Homk pV, Wq.
b) Given a linear map θ : V Ñ W whose matrix relative to the bases tei u
and t f j u is paij q. Show that the element in V ˚ bW corresponding to θ equals
ř ř ř
i φi bθpei q and that i φi bθpei q “ ij aij φi b f j .
c) Show that the non-zero decomposable tensors in V ˚ bk W under the map Γ
correspond to the linera maps of rank one. Hint: Chose an appropriate basis
for V.
d) Show that a linear map in Homk pV, Wq is of at most rank r if and only if the
corresponding tensor in V ˚ bW is the sum of at most r decomposable tensors.
Hint: Chose an appropriate basis for V
5.31 Notice, that elements in B coming from A may be moved past the b-sign;
i.e. if the structure map is u : A Ñ B one has xbab “ axbb. Be aware however,
there is a hidden pitfall. For any b P B the notation a ¨ b is a sloppy version
of the correct notation upaq ¨ b, where the product is the product in the ring B.
Hence axbb “ xbupaqb would be the correct way of writing.
This leads to equalities like apxbbq “ upaqpxbbq where on the left side
Mb A B is considered an A module (through the left factor) and on the right a
B-module (through the right factor).
5.32 Sometimes one wants to perform consecutive base changes, and the
tensor product behaves well in such a situation. It is transitive in the following
sense.
Proof: The short description of the maps are mbxby ÞÑ mbxy and mbz ÞÑ
mb1bz. By the principles of bi- and tri-linearity both extend to maps between
the tensor products, and they act as mutually inverses on the decomposable
tensors. Hence the are mutually inverse maps. o
5.34 Changing the base ring preserves tensor products, but not hom-modules
in general. One has the following:
In other words, the functor p´qb A B takes tensor products into tensor products.
5.36 In a base change situation there are two natural functors between the
two module categories. In addition to the base change functor, given as the
116 ma4200—commutative algebra
tensor product p´qb A B : Mod A Ñ ModB , there is a functor going the other
way: p´q A : ModB Ñ Mod A . The action of this functor is kind of trivial. If N is
a B-module, NA is equal to N but regarded as an A-module—one forgets the
B-module structure. With maps the same happens; they are kept intact, but
merely regarded as being A-linear. Such functors that “throws away” part of
the structure, are called forgetful functors in the parlance of the category theory.
The point is that the tensor product p´qb A B and the forgetful functor p´q A
are adjoint functors:
Proposition 5.37 (Adjointness) Given an A-algebra B. Then there is a canon-
ical isomorphism
HomB pMb A B, Nq » Hom A pM, NA q.
Proof: The from left two right is simply a “restriction” map. It sends a given
B-linear map φ : Mb A B Ñ N to φpxb1q which is A-linear in x. To define
a map from the right side to the left side, let φ : M Ñ NA be A-linear. The
expression ψpxbbq “ bφpxq is A-bilinear and by the universal property enjoyed
by the tensor product it extends to an A-linear map ψ : Mb A B Ñ N which
turns out to be B-linear (remember, multiplication by elements from B is
performed in the right factor):
At last and as usual, the two maps are inverses to each other, since they are
when acting on decomposable tensors. o
of φpei q in terms of the basis elements f j . Applying φbidB to the basis element
ei b1, yields
ÿ ÿ
φbidB pei b1q “ φpei qb1 “ p a ji f j qb1 “ upa ji q ¨ f j b1
j
tensor products 117
Notice that the C-module structure on AbC B is in pace a priori being the in-
duced structure on the tensor product, so merely the ring structure is lacking.
118 ma4200—commutative algebra
Proof: To argue that the assignment of (5.6) can be extended to give a prod-
uct of arbitrary tensors, we once more we appeal to the principle of bilinearity
(at the end of paragraph 5.2 on page 103). In fact, we shall apply it twice—
once for each factor—the basic observation being that the right side of (5.6) is
C-bilinear both in pa, bq and px, yq
The first application of the principle shows that multiplication by a fixed
pair pa, bq extends to a C-linear map AbC B Ñ AbC B. This yields a map
ηpaa1 bbb1 q “ η A paa1 qηB pbb1 q “ η A paqη A pa1 qηB pbqηB pb1 q “ ηpabbqηpa1 bb1 q ηA
AbC B
ηB
? _
where the two extreme equalities hold true by the very definition of η, and the
ιA ιB
middle one because both ηa and ηB are ring maps.
A_ ?B
Next, one has η ˝ ι A “ η A and η ˝ ι B “ ηB since η A p1q “ ηB p1q “ 1, and finally,
that η is unique follows, since it is determined by the values on decomposable
tensors, and these satisfy
C
ηpabbq “ ηppab1qp1bbqq “ ηpab1qηp1bbq “ η A paqηB pbq
Polynomial rings
We continue with the stage set as above, with B being an A-algebra through
the structure map u : A Ñ B. In the name of the inherent human laziness, we
shall write Arx‚ s for a polynomial ring Arx1 , . . . , xr s.
5.43 A natural question is how polynomial rings behave under base change,
and the answer is they do in the obvious and simplest way.
There is a map of A-algebras Arx1 , . . . , xr s Ñ Brx1 , . . . , xr s sending xi to xi
ř ř
(hence a polynomial α aα x α maps to α upaα qx α ). Together with the inclu-
sion of B as the constants in Brx1 , . . . , xr s it induces, in view of the universal
property of the tensor product, a map of A-algebras Arx1 , . . . , xr sb A B Ñ
Recall the notation
Brx1 , . . . , xr s. It sends the variable to the variables, and gives an isomorphism: from Paragraph 1.16
on page 15 where the
Lemma 5.44 Let A be a ring and B be an A-algebra. Then Arx1 , . . . , xr sb A B “ α
monomial x1 1 ¨ . . . ¨ xrαr
was denoted by x α ,
Brx1 , . . . , xr s
with α being the multi-
index pα1 , . . . , αr q.
Proof: Considered as A-module, the polynomial ring Arx1 , . . . , xr s is free, and
the monomials x α , with α running through all multi-indices, form a basis. The
same holds for Brx1 , . . . , xr s; the monomials x α form a B-basis.
120 ma4200—commutative algebra
According to what we figured out in Paragraph 5.36 on page 116 the ele-
ments x α b1 form a B basis for Arx1 , . . . , xr sb A B; but our map sends these to
x α which form a basis for Brx1 , . . . , xr s. o
Problems
5.15 Let k be a field and let f pxq be an irreducible polynomial in krxs so that ˇ
K “ krxs{p f pxqq is field. Moreover let L be field extension of k over which f
splits as a product f pxq “ f 1 pxq . . . f r pxq of irreducible polynomials. Use the
ś
Chinese remainder theorem to prove that Kbk L » Li where Li is the field
Li “ Lrxs{p f i pxqq.
tensor products 121
5.16 Assume that K is an algebraic number field; i.e. a finite extension of the ˇ
field Q of rational numbers. Show that KbQ R is isomorphic to a product of
fields each being isomorphic either to R or C. Show that rK : Qs “ r1 ` 2r2
where r1 denotes the number of factors isomorphic to R and r2 the number of
factors isomorphic to C. Hint: By the Primitive Element Theorem one may
assume that K “ Qrxs{p f pxqq where f pxq P Qrxs is irreducible polynomial.
5.17 Show that A “ Rrx, ys{px2 ` y2 q is an integral domain, but that AbR C is ˇ
not.
M
Changes:
2018-09-23 Have reworked almost everything until section 5.4; not yet finished.
05/10/2018 Reworked the last part about tensor products of algebras.
07/10/2018 Added an exercise, corrected lots of misprints and minor errors.
08/10/2018 Have rewritten the proof of Proposition 5.23 about Adjointness on page 109 and
the subsequent example with substantial simplifications.
10/10/2018 Have added a prop about tensor prods and bases change; prop 5.35 on page 115.
11/10/2018 Added Exercise 5.6 on page 113
18/10/2018 Corrected misprints, minor changes in the text.
22/10/2018 Minor changes of the language.
Lecture 6
Localization
Preliminary version 2.0 as of 2018-10-22 at 09:43 (typeset 3rd December 2018 at 10:03am)—Prone
to misprints and errors and will change.
28/10/2018 Added new exercise 6.6
Very early in our mathematical carrier, if not in our lives, we were intro-
duced to fractions, and we should be familiar with their construction. Anyhow,
recall that to every pair of integers m and n with n ‰ 0, one forms the “frac-
tion” m{n with m being the enumerator and n the denominator. Two such
fractions m1 {n1 and m{n are considered equal—that is, have the same numeri-
cal value—precisely when nm1 “ n1 m. The fractions—or the rational numbers
as we call them—are entities per se and not only results of division. They obey
the rules for adding and multiplying we learned in school—that is, the field
axioms—and they thus form a field Q.
There is a simple and very general version of this construction. It gives us
the freedom to pass to rings were a priori specified elements become invertible.
Virtually any set of elements can be inverted; there is merely one natural
constraint. If s and t occur as denominators, their product st will as well;
indeed, one has s´1 t´1 “ pstq´1 . Hence the natural notion to work with is the
concept of multiplicatively closed sets.
The process is indeed very general. It even accepts zero-divisors as denom-
inators, but can then be "murderous". If a is a zero-divizor, say a ¨ b “ 0, and
a becomes inverted, b gets killed; indeed, it will follow that b “ a´1 ¨ a ¨ b “
a´1 ¨ 0 “ 0. In principle, one can even push this so far that 0 is inverted; but
this will be devastating and the resulting ring not very interesting, it will be
the null-ring.
There are several ways of defining the localized rings. We shall follow most
text books and mimic the way one constructed the rational numbers. This is a
direct and intuitive construction not requiring much machinery.
The name “localization” has its origin in geometry and rings of functions,
say on a topological space X. If U Ď X is an open set, every function whose
zeros all lie in the complement XzU of U, become invertible when restricted to
U; hence one obtains many functions on U by inverting certain functions on X.
In general far from all are shaped like that, but in special situations, important
124 ma4200—commutative algebra
o 1 P S;
o If s, t P S, then st P S.
In the sequel we shall mostly suppress the reference to the map ι S and just
write a for ι S paq. This calls however for a word of warning: The map ι S is not
always injective! Membres of A killed by elements from S become zero in AS ;
indeed, from as “ 0 follows ι S paqι S psq “ 0, and ι S psq being invertible forces
ι S paq “ 0.
localization 125
we infer that
However, some checking is necessary. First of all, the definitions are expressed
in terms of representatives of the equivalence classes, and it is paramount
they do not dependent on which representatives are used. Secondly, the
ring axioms must be verified; once the definitions are in place, this is just
straightforward high school algebra safely left to volunteering students.
6.8 As an illustration, let us check that the sum is well defined. Notice that
it suffices to vary the representatives of one of the addends at the time; so
assume that pa, sq „ pa1 , s1 q; i.e. upas1 ´ a1 sq “ 0 for some u P S. We find
which is killed by u. Therefore the sum does not depend on the representative
of the first addend used, and by symmetry, neither on the representative of the
second, and the sum is well defined.
Problem 6.1 Show that the product is well defined. On a rainy day when all
your friends are away, verify the ring axioms for AS . M
6.9 The localisation map ι S is nothing but the canonical map ι S : A Ñ AS that
sends an element a in A to the class of the pair pa, 1q; that is, a is mapped to
the fraction a{1. By the very definition in (6.1) of the sum and the product
in the localization AS this a ring homomorphism. (Seemingly, this map does
nothing to a, but kill it if necessary).
6.10 Our next proposition gives a more practical description of the ring AS :
Proposition 6.11 All elements in AS are of the form a{s. It holds true that
ι S paq “ 0 if and only if a is killed by some element from S; i.e. if and only if there
is an s P S such that sa “ 0.
and the salient point is that this is a legitimate definition, i.e. φpaq ¨ φpsq´1 is
independent on the chosen representative pa, sq. But from a{s “ a1 {s1 ensues
that t ¨ pas1 ´ sa1 q “ 0 for an element t P S, and hence, since φ is map of rings,
that
` ˘
φptq ¨ φpaq ¨ φps1 q ´ φpa1 q ¨ φpsq “ 0.
The element φptq is invertible by assumption, and we conclude that φpaq ¨
φpsq´1 “ φpa1 q ¨ φps1 q´1 .
This proves most of the following proposition:
localization 127
Examples
6.1 We have not excluded that 0 lies in S. In this case however, the localized
ring will be the null-ring since 0 becomes invertible. This situation occurs e.g.
when S has nilpotent members.
6.4 The ring of integers Z within the field rationals Q is a particular instance
of the situation in the previous example. When S is the set of all powers of a
given number p, that is, S “ tpn u, the ring ZS “ Zr1{ps “ t a{pn | a P Z, n P
N0 u will be the ring of rational numbers whose denominators are powers of p.
(see also Example 1.12 on page 12).
In a similar vein, when p is a prime and S is the complement of the prin-
cipal ideal pZ, the localization ZS will be the ring Zp pq “ t a{b | a, b P
128 ma4200—commutative algebra
Problems
6.4 Prove that the group of units A˚ in A is a multiplicative set. Show that the
localization maps ι S is an isomorphism if and only if S is a subset of A˚ .
6.6 Consider be the polynomial ring A “ krx1 , . . . , xn s over the field k. Let
S be the set polynomials in A that depend only on the first r variables; i.e.
those on the form ppx1 , . . . , xr q. Show that S is multiplicatively closed and that
AS “ Krxr`1 , . . . , xn s where K is the field kpx1 , . . . , xr q of rational functions.
6.15 Extension of ideals from A to AS yields a map from the set I pAq of ideals
in A to the set of ideals I pAS q in the fraction ring AS . The map operates by
sending an ideal a to the ideal aAS generated by the image ι S paq, and one
verifies that members of aAS are all shaped like as´1 for some a P a and some
s P S. The map is inclusion preserving and preserves products and sums of
ideals as extension maps always do (see Paragraph 2.9 on page 26).
localization 129
6.16 In general the extension map from I pAq to I pAS q is not injective. For
instance, it may happen that S X a ‰ H, in which case the extension aAS will
contain an element invertible in AS and consequently be equal to AS ; and
of course, this may be the case for several different ideals. In quite another
corner, ideals a contained in the kernel of ι S reduce to the zero ideal in AS . So,
some ideals are blown up to AS (those meeting S) and some collapsed to zero
(those contained in ker ι S ).
Example 6.6 A simple instance of the extension map not being injective is
the case when A “ Z and S “ t pn | n P Z u for some prime p. All the
ideals a “ ppm q extend to the entire ring ZS . This also illustrates that forming
extensions does not commute with forming infinite intersections; indeed, one
has m ppm q “ 0 whereas m pm ZS “ ZS .
Ş Ş
K
6.17 The extension map is however surjective. Any ideal b Ď AS equals
1
ι´
S pbqAS ; that is, when pulling an ideal back to A and subsequently extending
the result, one recovers the original ideal. To see this, notice that if b “ a{s
1
belongs to b, the element a belongs to ι´
S pbq as ι S paq “ b ¨ s, and therefore b lies
1
in the extension ι´ S pbqAS .
Proof: We have already proved most of the proposition, only the assertions
about sums, products and intersections remain unproven. It is a general
feature of extension of ideals that products and sums are preserved, and we
concentrate on the finite intersections; and of course, the case of two ideals
will suffice.
Clearly pa X a1 qAS Ď aAS X a1 AS . So assume that b P aAS X a1 AS . One may
then express b as b “ a{s “ a1 {s1 with elements a and a1 from respectively a
and a1 . This yields tpa ¨ s1 ´ a1 ¨ sq “ 0 for some t P S. But then tsa1 “ ts1 a P a X a1
and consequently b “ tsa1 {tss1 lies in pa X a1 qAS . o
6.19 Prime ideals behave more lucidly under localization than general ideals.
Either they blow up and become equal to the entire localized ring AS , or they
persist being prime. Moreover, every prime ideal in AS is of the shape pAS for
an unambiguous prime ideal p of A. One has:
bb1 P pAS and that b “ a{s and b1 “ a1 {s1 with a, a1 P A and s, s1 P S. We infer
that tss1 bb1 “ taa1 P p for some t P S, and hence either a or a1 lies in p since t
does not. Moreover, if ι S paq “ a1 s´1 for some a1 P p, if follows that sta “ ta1 P p,
hence a P p since p is a prime ideal. o
Proposition 6.21 The prime ideals of AS are precisely the ideals of the form pAS
for p a prime ideal in A not meeting S. The prime ideal p is uniquely defined.
Proof: By the previous proposition, the ideals pAS are all prime, so let q
1
be a prime ideal in AS . Then p “ ι´S q is prime, and by the last sentence in
proposition 6.18 above it holds that q “ pAS . o
6.22 The localization processe commutes with the formation of radicals:
Problems
6.7 Show that if aAS “ AS , then the same holds for all powers am of a.
6.8 Let p and q be different prime numbers and let S be the multiplicative set ˇ
S “ t pn | n P N0 u. Describe Z X ppqqZS .
6.10 Suppose that p is a prime ideal and that a an ideal contained in p. Show
1
that ι´
S paAs q Ď p.
6.11 Let j : Spec AS Ñ Spec A be the map induced from the localization map
ι S : A Ñ AS . Show that j is a homeomorphism onto its image (when the image
is endowed with induced topology). Show that the image jpSpec Aq equals the
intersection of all the open sets containing it.
6.13 Given an example of a ring A and a non-zero prime ideal p such that
Ap “ A{p. Hint: Let A be the product of two fields.
localization 131
Functoriality
Proposition 6.27 The localisation Ap is a local ring with maximal ideal pAp . The
assignment q ÞÑ qAp is a one-to-one correspondence bewteen prime ideals in Ap and
prime ideals q in A contained in p.
Proof: This is nothing but Proposition 6.21 above on page 6.21, according
to which the prime ideals in Ap are precisely the ideals in Ap of the form
qAp where q is a prime ideal in A contained in p, and it holds true that q “
ι´1 pqAp q. o
Notice that the kernel of the localization map ι : A Ñ Ap is contained in any
prime ideal q Ď p; indeed, if sa “ 0 with s R p, a fortiori s R q and hence a P q.
132 ma4200—commutative algebra
The residue field Ap {pAp is frequently written kppq. Since ι´1 ppAp q “ p
it holds true that A{p maps invectively into Ap {pAp . And in fact, Ap {pAp
equals the fraction field KpA{pq of A{p.
Example 6.7 If p is a prime number, the localized ring Zp pq at the maximal
ideal generated by p consists of rational numbers whose reduced form is a{b
where b is relatively prime to p. The maximal ideal is generated by p and the
residue field is the field F p with p elements. K
6.28 Clearly an arbitrary intersection of multiplicatively closed sets is mul-
tiplicatively closed, therefore complements of unions of prime ideals are
multiplicatively closed. If the prime ideals involved are tpi uiP I ; and hence
Ť
that S “ Az iP I pi , the maximal ideals of AS will be the extension pi AS . In
particular, if the pi ’s are merely finite in numbers, AS will be a semi-local ring.
invertible in KpAq. Consequently KpAq is a field; it is called the field of fractions The field of fractions
(Kvotientkroppen)
localization 133
of A, which we met already in Example 6.3 on page 6.3. The ring KpAq is in
general not a field, but by definition has the property that all non-zero divisors
are invertible.
In any case, the canonical map A Ñ KpAq is injective; indeed, if s is a
non-zero divisor, by definition sa “ 0 implies that a “ 0.
Proposition 6.29 The total ring of fractions KpAq of a ring A has the property
that every non-zero divisor is invertible. The natural map A to KpAq is injective.
Moreover, KpAq is a field if and only if A is an integral domain.
6.30 We give two examples. The first is a simple but typical situation in
algebraic geometry. The ring A is the ring of polynomial functions on a variety
X with two components; in our specific example the two components are the
coordinate axes in the plane. The total ring of fractions of A turns out to be
the product of two fields, the elements are pairs of rational functions, one on
each of the axes. Contrary to what is requested of polynomial functions on X,
there is no continuity condition at the origin to be fulfilled; the two rational
functions may take different values, or for that matter, may have poles at the
origin.
This behavior is fairly general; for any ring A without nilpotent elements
and which merely has finitely many minimal prime ideals (that is, there is a
representation p0q “ p1 X . . . X pr of the zero ideal as a finite intersection of
prime ideals) it holds that KpAq is a product of fields. The only moral of the
second example, is that if A has nilpotents, the ring of fractions KpAq can be
rather involved; and far from being a product of fields.
Example 6.9 Let A “ krX, Ys{pXYq and let x and y be the classes of X and
Y respectively. Since pXYq X krXs “ p0q it holds that krXs » krxs and ditto
krYs » krys, thus we may talk about the rational function fields kpXq and KpYq
as kpxq and kpyq.
We shall see the total quotient ring KpAq is isomorphic to the product
kpxq ˆ kpyq of the rational functions fields in kpxq and kpyq. In geometric terms,
A is the ring of polynomial functions on the union Z “ Vpxq Y Vpyq of the
y-axis and the x-axis in the plane, and the elements of KpAq are just a pair of a
rational function, one on the x-axis and one on the y-axis.
All cross-terms are killed in A, and the elements are all shaped like a `
ppxq ` qpyq where ppxq and qpyq are polynomials vanishing at zero and a is
scalar in k. The zero-divisors in A is the union1 pxq Y pyq, so that the non-zero- 1
If Ppx, yq is a zero di-
divisors of A are of the form a ` ppxq ` qpyq with either a ‰ 0 or neither ppxq visor, there is a relation
like PpX, YqQpY, Yq “
nor qpyq identically equal to zero. ApX, YqXY, so either X
There is a map of rings or Y divides PpX, Yq
A Ñ kpxq ˆ kpyq
sending the class of a polynomial a ` ppxq ` qpyq to the pair pa ` ppxq, a ` ppyqq,
134 ma4200—commutative algebra
these are invertibel in kpxq ˆ kpyq and the there is induced a ring map
which clearly is injective; but it is slightly more subtle that it is surjective. Let
ξ “ pQpxq{Rpxq, Spyq{Tpyqq be an element in kpxq ˆ kpyq; then Rpxq and Tpyq are
non-vanishing polynomial. Then
ξ “ pQpxqx{Rpxqx, Spyqy{Tpyqyq,
Example 6.10 Let B “ krX, Ys{pX 2 , XYq and let as usual x and y be the classes
of X and Y in B. Then m “ px, yq is a maximal ideal (one has B{m “ k). Let
A “ Bm be the ring B localized at m. As in every local ring the units of A are
precisely the members not lying in the maximal ideal m. We contend that all
elements in m are zero-divisors, and hence A is its own total ring of fraction.
Indeed, x kills both x and y, hence the whole ideal m “ px, yq.
It worth while noticing that pxq is a prime ideal, since it holds true that
B{pxq “ krX, Ys{pX 2 , XY, Xq “ krYs is an integral domain, and pxq is the only
a
other prime ideal in B and is equal to the radical p0q of B; indeed, if p is
prime one has x2 P p, hence x P p.
Thus B has two prime ideals, the maximal ideal px, yq whose elements
constitute all zero-divisors and a sole minimal ideal pxq whose elements are all
the nilpotents of B. K
Problem 6.16 Show that Apxq equals the rational function field kpYq. M
Problems
6.17 Let n be a natural number. Determine the total quotient ring of Z{nZ. ˇ
6.18 Let A be any ring. Show that the nil-radical of KpAq is equal to the ˇ
extension of the nil-radical of A.
a) Show that if the elements of A are either zero divisors or invertible, then
A “ KpAq.
b) If A has only one prime ideal, prove that KpAq “ A.
c) Let A be a direct product (of any cardinality) of any number of rings each
having only one prime ideals. Prove that KpAq “ A.
localization 135
A last example
6.31 There is multiplicatively closed set associated with any ideal a in A which
is not that frequently met. It equals the set S “ 1 ` a; that is, the set consisting
of elements shaped like 1 ` a with a P a. one has the subset S of elements of
the form 1 ` a with a P a which is multiplicatively closed. In this case the the
maximal ideals in AS are those of the form mAS ; hence aAS is contained in the
Jacobson radical of AS .
tpms1 ´ m1 sq “ 0 (6.2)
6.34 The same applies to the maps ι S when it comes to modules as with rings.
To simplify the notion, one soon drops the reference to the map and writes x
or x{1 for ι S pxq, but with some cautiousness since the image very well can be
zero.
When the module M is finitely generated, say by members m1 , . . . , mr , the
images ι S pmi q of the mi ’s will obviously generate MS . Indeed, pick a member
xs´1 from MS and write x “ ai mi then of course xs´1 “ ai s´1 mi .
ř ř
Functoriality
where φ and ψ are A-linear maps from M to N and a and b ring elements.
And it is equally clear that the association is functorial; it holds true that
pψ ˝ φqS “ ψS ˝ φS
whenever φ and ψ are composable since it already holds at the level of the
Cartesian products—and of course, pid M qS “ id MS .
Proof: The only subtle point is that the functor is exact. In other words that it
brings an exact sequence
/M /L
ψ φ
N (6.3)
/ MS / LS
ψS φS
NS
Submodules
Given a submodule N Ď M. The localized module NS can be considered to be
submodule of MS . The inclusion map localizes to an injection whose image
consists of elements shaped like fractions ns´1 with n P N and s P S, and thus
it can naturally be identified with NS .
6.38 Localization behaves nicely with respect to sums and finite intersections
of submodules; the following two assertions hold:
ř ř
o p i Ni qS “ i pNi qS
o pN X N 1 qS “ NS X NS1
To establish the first equality observe that it ensues from the inclusion
ř ř ř ř
Ni Ď i Ni that pNi qS Ď p i Ni qS , hence that i pNi qS Ď p i Ni qS . Any element
in p i Ni qS is of the form p i xi qs´1 with merely finitely many of the xi ’s being
ř ř
ř
non-zero, which obviously lies in i pNi qS .
The second follows because y “ n{s “ n1 {s1 means ts1 n “ tsn1 for some t P S,
and putting x “ ts1 n we then infer that x P N X N 1 and y “ x{tss1 .
Problem 6.22 Locization does not commute with infinite direct products. Let
p P Z be a number and denote by S the multiplicative set S “ tpn u in Z. Show
that there is a natural inclusion
ź ź
p ZqS Ď ZS ,
iPN i PN
pMb A NqS » MS b AS NS .
6.43 When it comes to hom-sets, the behaviour is rather nice, at least for
modules of finite presentation. In general, sending an A-linear map φ be-
tween two A-modules M and N to the localized map φS is an A-linear map
Hom A pM, Nq Ñ Hom AS pMS , NS q. By the universal property of localization it
extends to a map Hom A pM, NqS Ñ Hom AS pMS , NS q; and in case M is of finite
presentation, this map is an isomorphism:
Proof: Recall that both localization and the hom-functors are additive func-
tors, hence the proposition holds true whenever M is a free A module of finite
rank n; indeed, one finds
where the isomorphisms are the natural ones (the one in the middle is an
isomorphism since localization is additive, and the two others are because
hom-functors are additive). Since M is assumed to be of finite presentation, it
lives in an exact sequence
/ nA /M / 0,
φ π
mA (6.4)
with m, n P N and where φ and π are A-linear maps. Consider the diagram
0 / Hom A pMS , NS q / Hom A pnAS , NS q / Hom A pmAS , NS q
S S S
The upper sequence is obtain from (6.4) by applying Hom A p´, NqS to it and is
therefore exact by left exactness of hom-functors. The bottom sequence is the
localization in S of the upper one, and since localization is an exact functor, it
is exact. The vertical maps are the canonical maps induced by sending φ to φs ,
and it is a matter of simple verification to check that the squares commute.
Now, the final point is that the two rightmost maps are isomorphisms by
the beginning of the proof, and then the Five Lemma tells us that the third
map is an isomorphisms as well, which is precisely what we aim at proving! o
6.46 It is not true that every A-module has a simple quotient, an elementary
example being the Z-module Q; every ideal a in Z satisfies aQ “ Q and hence
Q has no simple quotient. However finitely generated modules always have,
and this is our first version of Nakayama’s lemma:
Proof: If M is cyclic, it is of the form A{a for some proper ideal and has A{m
as a quotient for any maximal ideal m containing a. Assume next that M is not
cyclic, and let n be the the least number such that M can be generated by n
elements. Then n ě 2. Let x1 , . . . , xn generate M. The submodule N generated
by x2 , . . . , xn is a proper submodule and M{N is cyclic (generated by the class
of x1 ), and has a simple quotient by the first part of the proof. o
Nakayama classic
6.49 To assure anyone (hopefully there are none) that finds our approach a
blasphemous assault on their most cherished tradition, we surely shall include
Nakayama classic; here it comes:
Proof: Recall that the Jacobson radical of A equals the intersection of all the
maximal ideals in A. Assume M ‰ 0. By Nakayama I (Proposition 6.47 above)
there is a maximal ideal m such that mM is a proper submodule, which is
impossible since a Ď m and aM “ M by assumption. o
localization 141
6.52 The by far most common situation when Nakayama’s lemma is applied,
is when A is a local ring and a “ m is the maximal ideal, and this situation
merits to be mentioned specially:
Other formulations
6.54 There are several other reformulations of Nakayama’s lemma, and here
we offer a few of the most frequently applied version.
Proof: The tensor product is right exact so cokerpφbid A{a q “ pcoker φqb A{a.
The cokernel coker φ is finitely generated since M is, and cokerpφbid A{a q “
pcoker φqb A{a “ 0 by hypothesis. Hence coker φ “ 0 by Nakayama III
(Proposition 6.51 above). o
Proof: The quotient M{N is finitely generated since M is, and it holds true
that aM{N “ M{N because any m from M lies in aM modulo elements in N;
or if you prefer, use the previous proposition with φ being the inclusion map
N ãÑ M. o
6.58 There is a version of Nakayama valid for all ideals not only those lying in
the Jacobson radical; but of course, when weakening the hypothesis you get a
weaker conclusion. The proof relies on a localization technique.
Problems
6.24 (Demystifying Nakayama’s lemma.) Let A be a local ring with residue class ˇ
field k. Assume that φ : E Ñ F is an A-linear map between free module of
finite rank, and let Φ be the matrix of φ in some bases.
a) Show that if one of the maximal minors of Φ
s does not vanish, one of the
maximal minors of Φ is invertible in A. Conclude φ is surjective when φbidk
is.
b) Show the classical Nakayama’s lemma for finitely presented modules over a
local ring by using the previous subproblem.
c) (Mystifying the demystification.) Show Nakayma’s lemma for finitely gener-
ated modules over a local ring by using subproblem a). Hint: The key word
is "right sections" of linear maps, if you don’, prefer coping with maximal
minors of n ˆ 8-matrices!!
6.26 Let M an A-module such that mM “ M for any maximal ideal m. Show
that M has the property that if one discards any finite part from a generating
set one still has a generating set.
6.30 Let A be ring and P a finitely generated projective module. Show that
there is a set of elements t f i u in A such that the distinguished open subsets
Dp f i q cover Spec A, and such that each localized module Pf i is a free module
over A f i .
6.31 Let A a ring and let e be a non-trivial idempotent element. Show that
À
the principal ideal I “ peqA is projective, and that a direct sum i I of a any
number, finite or not, of copies of I never can be free. Hint: Such sums are
killed by 1 ´ e.
prime ideals p so that Mp ‰ 0. This subset is called the support of M and is a The support of a module
geometric structure associated with M. In many case (e.g. when M is finitely (støtten til en modul)
That the ideals be maximal in the second condition can (of course) be replaced
by they being prime.
Proof: One way is obvious. So assume that M is non-zero and let x be a non-
zero element in M. Then the annihilator Ann x of x is a proper ideal as x ‰ 0
and hence contained in a maximal ideal m. By the simple lemma above, the
image of x in Mm is non-zero and a fortiori Mm is non-zero. o
is,
Supp M “ t p P Spec A | Mp ‰ 0 u.
Remember that elements of Mp are all shaped like xs´1 , and such an element
is zero precisely when tx “ 0 for some t R p.
6.67 The hypothesis that M be finitely generated was used only in the last
part of the proof, and it holds true for a general module M that Supp M Ď VpAnn Mq.
The other inclusion may however fail when M is not finitely generated.
Examples
6.11 Clearly the support of a cyclic module A{a equals the closed set Vpaq. In-
deed, an element s lying in a, but not in p would kill A{a and hnece pA{aqp “
0.
6.12 One has Supp Q “ Spec Z since QS “ Q for any multiplicative set S
in Z. More generally, for the fraction field K of any domain A it holds that
Supp K “ Spec A.
6.13 An example of this failure for “large modules”, can be the Z-module
Z p8 “ Zrp´1 s{Z where p is a prime.
Every element of Z p8 is killed by a power of p; indeed, an element x is the
class of a rational number shaped Every element of Z p8 is of the form x “
a{pr with a prime to p. Since ax P Z is and only if a is divisible by pr , one has
Ann x “ ppr q and from Lemma 6.61 above it follows that Supp Z p8 “ ppqZ.
146 ma4200—commutative algebra
Proof: This follows immediately from localization functor being exact. For
each prime p the localized sequence
0 / Np / Mp / Lp /0
is exact, and the middle module vanishes if and only if the two extreme do. o
6.70 The aim of this paragraph is the prove that the support of a tensor prod-
uct is the intersection of the supports of the two factors, at least when the
involved modules are finitely generated. The result hinges on Nakayama’s
lemma.
Lemma 6.72 Let A be a local ring with maximal idea m. Let M and N be two
finitely generated A-modules. Then Mb A N “ 0 if and only if either N “ 0 or
M “ 0.
localization 147
The tensor product of two non-zero vector spaces being non-zero (e.g. Proposi-
tion 5.20 on page 108), we infer that pMb A Nqb A k ‰ 0, and hence Nb A M ‰ 0
a fortiori. o
Proof of Proposition 6.71: Since localization is an exact operation, the
localized modules Np and Mp are finitely generated over Ap whenever M and
N are finitely generated over A, and in view of the isomorphism
pMb A Nqp » Mp b Ap Np ,
Problem 6.32 Let p be a prime number and let M be the abelian group M “
iPN0 Z{p Z. Determine the annihilator Ann M and the support Supp M.
i
À
M
09/10/2018 Still Much work remaining; but the first section should be readable now. More
follows!
10/10/2018 Added a exercise 6.8; Minor changes in the first section; brushed up the second
section.
11/10/2018 Added prop 6.23 on page 130 and a new exercise 6.14 on page 131.
15/10/2018 Have moved the section about Nakayama’s lemma to this chapter.
Lecture 7
Chain conditions
Preliminary version 1.2 as of 2018-11-01 at 10:16 (typeset 3rd December 2018 at 10:03am)—Much
work remaining. Prone to misprints and errors and will change.
20/10/2018 Almost every thing has been redone!!
29/10/2018 Corrected a few misprints in lemma ?? in section about Krull’s intersection
theorem.
29/10/2018 Have added exercise 7.13 in connection with Krulls’ intersection theorem. Several
minor cahnges
01/11/2018 Added two exercises on page 172. Rewritten prop 7.50 on page 169. Added
theorem 7.58 on page 172.
M0 Ď M1 Ď . . . Ď Mi Ď Mi`1 Ď . . . .
Similarly, a descending chain is a sequence tMi uiPN0 of submodules fitting into a Descending chain
chain of inclusions shaped like (nedstigende kjeder)
. . . Ď Mi`1 Ď Mi Ď . . . Ď M1 Ď M0 .
152 ma4200—commutative algebra
Such chains are said to be eventually constant or eventually terminating if the Eventually constant
submodules become equal from a certain point on; that is, for some index i0 it chains (terminerende
kjeder)
holds that Mi “ M j whenever i, j ě i0 . Common usage is also to say the chain
stabilizes at i0 .
7.2 An A-module M is said to be Noetherian if every ascending chain in M is Noetherian modules
eventually constant. This condition is frequently referred to as the Ascending (noetherske moduler)
Chain Condition abbreviated to acc . The module is Artinian if every descending The Ascending Chain
chain terminates, a condition also called the Descending Chain Condition with Condition (Den oppsti-
gende kjedebetingelsen)
the acronym dcc . Artinian modules
A ring A is called Noetherian if it is Noetherian as module over itself, and (artinske moduler)
of course, it is Artinian if it is Artinian as module over itself. The submodules The Descending Chain
Condition (Den nedsti-
of A are precisely the ideals, so A being Noetherian amounts to ideals of A gende kjedebetingelser)
satisfying the acc, and similarly, A is Artinian precisely when the ideals Noetherian and Artinian
comply with the dcc. rings (noetherske og
artinske ringer)
7.3 The two conditions, being Noetherian and Artinian, might look similar,
but in fact there is a huge difference between the two. Noetherian and Ar-
tinian modules belong in some sense to opposite corners of the category Mod A .
In what follows we shall treat Noetherian modules and Noetherian rings and
establish their basic properties, but will lack time to discuss the Artinian ones
in any depth, although Artinian rings will be discussed (in section 7.5 below).
In fact, according to a result of Akizuki, they turn out to be Noetherian as well.
M0 Ď M1 Ď . . . . Ď Mi Ď Mi`1 Ď . . .
Ť
be given. The union N “ i Mi is by assumption finitely generated and have
say x1 , . . . , xr as generators. Each x j lies in some Mνj , and the chain being
ascending, they all lie in Mν with v “ max j νj . Therefore N “ Mν , and the
chain stabilizes at ν. o
7.6 It warrants a comment that Zorn’s lemma is not used in the proof of
families of submodules having maximal elements. As chains are eventually
constant, we avoid problems with limit ordinals, and ordinary inductions
works perfectly. As a consequence a large portion of the theory of Noethe-
rian modules does not depend on the Axiom of Choice. Notably existence
theorems like Theorem 2.50 on page 40 come for free for Noetherian rings.
7.7 The Noetherian modules, as do the Artinian modules, form a subcategory
of Mod A which enjoys a strong closedness property. They are what in category
theory are called thick subcategories. Submodules and quotients of Noetherian
modules are Noetherian as is an extension of two, and the same is true for
Artinian modules.
Proposition 7.8 Let M1 , M and M2 be three A-modules fitting in the short exact
sequence
0 / M1 /M / M2 / 0.
Then the middle module M is Noetherian (respectively Artinian) if and only if the two
extremal modules M1 and M2 are.
and hence Ni “ Nj . o
Proof: The proof is based on the simple observation that for any submodule
N Ď MS one has pι´1 NqS “ N, which is clear since elements in N are of the
form ιpyqs´1 . Now, any chain tNi u in MS , whether ascending or descending,
induces a chain tι´1 Ni u in M, and if this chain stabilizes, say ι´1 Ni “ ι´1 Nj
for i, j ě i0 , it holds true that Ni “ pι´1 Ni qS “ pι´1 Nj qS “ Nj , and the original
chain stabilizes at i0 as well. o
Example 7.2 (Finite product of fields) The conclusions of the preceding example
ś
extend to rings that are finite products of fields; say A “ 1ďiďr k i . Modules
À
over such rings are direct sums V “ 1ďiďr Vi where each Vi is a vector
space over k i with the A-module structure induced by the projection A Ñ k i .
From Proposition 7.8, or rather the succeeding comment, ensues that V is
Noetherian (or Artinian) if and only if each Vi is of finite dimension over k i . K
Problems
7.1 Prove the assertion just after Proposition 7.8 that if a direct sum of Noethe-
rian modules is Noetherian, the sum is finite and all the summands are
Noetherian.
7.2 Show that Z is a Noetherian Z-module, but that Z p8 “ Zrp´1 s{Z is not.
Show that Z p8 is an Artinian Z-module, but that Z is not.
chain conditions 155
7.4 Show that a direct sum of finitely many simple modules is both Noethe-
rian and Artinian.
o A is Noetherian, that is, the ideals in A comply to the ascending chain condition;
It is trivial that fields are Noetherian and we shall shortly see that polynomial
rings over fields are Noetherian; this is the celebrated Hilbert’s Basis Theorem.
Other examples are the principal ideal domains (see Exercise 7.6 below).
7.13 Quotients of Noetherian rings are Noetherian (Proposition 7.12), but
subrings are not necessarily Noetherian. A stupid example being the fraction
field of a non-noetherian domain, a more subtle example will be given below
(Example 7.4).
156 ma4200—commutative algebra
Examples
which does not stabilize. Indeed, if p´pi`1q x P pp´i xq, one would have
p´pi`1q x “ Ppxqp´i x for some polynomial Ppxq P A. Cancelling p´i x would
give p´1 “ Ppxq, which contradicts that Pp0q P Z.
7.5 A large class of important non-Noetherian rings are formed by the rings
HpΩq of holomorphic functions in an open domain Ω in the complex plane.
Chains that do not terminate arise from sequences of distinct points in Ω
that do not accumulate in Ω. If tzi u is such a sequence, let an be the ideal of
functions in HpΩq vanishing in the set Zn “ tzn`1 , zn`2 , . . .u. These ideals
clearly form an ascending chain, and from Weierstrass’ Existence Theorem
ensues that there are functions f n holomorphic in Ω whose zeros are exactly
the points in Zn . Then f n P an , but f n R an´1 , and the chain can not stabilize at
any stage.
A structure theorem
As an illustration of the strength and elegance the Noetherian method can
show, we offer a structure theorem for finitely generated modules over Noethe-
rian rings- it does not reveal the fine features of a module , but rather de-
scribes the overall structure. Every such module is obtained by successive
extensions of cyclic modules shaped like A{p with p a prime ideal.
7.17 The structure theorem builds on the following result which is of indepen-
dent importance and will be use later.
Corollary 7.19 Any non-zero module over a Noetherian ring contains a module
isomorphic to A{p for a prime ideal p.
Proof: The set of annihilators of non-zero elements is non empty and has a
maximal element since A is Noetherian. Then we cite Proposition 7.18 above.
o
7.20 The ground is now prepared for the announced structure theorem; here it
comes:
158 ma4200—commutative algebra
0 / M i ´1 / Mi / A{pi /0
for 1 ď i ď ν.
Problems
7.5 Let A Ď Q be any proper subring. Show that the polynomials in Qrts ˇ
assuming values in A at the origin, is not Noetherian.
7.7 Let tA1 , . . . , Ar u be a finite family of Noetherian rings. Show that the ˇ
ś
product i Ai is Noetherian.
ś
7.8 Let k be a field. Show that the product i PN k of a countable numbers of ˇ
copies of k is not Noetherian.
7.9 Show that the ring of numerical polynomials in Qrxs is not Noetherian.
7.10 Let A be the subring of the complex numbers C whose elements are
algebraic integers; that is, they are solutions of equations of the type
zn ` an´1 zn´1 ` . . . ` a1 z ` a0 “ 0
where the coefficients ai are integers. Show that A is not Noetherian. Hint:
?
For instance, the principal ideals p 2n 2q form an ascending sequence that does
not terminate.
M
chain conditions 159
like the integers, but the spirit was entirely the same. The abstract and non-
constructive proof was revolting at a time when that part of mathematics was
ruled by long and soporific computations, making it extremely difficult to
obtain general results, and it opened up the path to modern algebra. Part of
the mythology surrounding the theorem is the exclamation by the “König der
Invariant Theorie” Paul Gordan: “Das ist nicht Mathematik, das ist Theolo-
gie!”. The truth is that Hilbert had proved in a few pages what Gordan and Paul Albert Gordan
his school had not proved in twenty years. (1837–1912)
German mathematician
7.22 There are several different proofs in circulation, and we shall give one
of the shortest. These days many constructive proofs are known and good
algorithms exist for exhibiting explicit generators for ideals in polynomial
rings; however, we shall present a non-constructive proof in the spirit of
Hilbert’s.
Before giving the proof of Hilbert’s basis theorem we state three important
corollaries. A straightforward induction on the number variables immediately
yields the following:
Corollary 7.24 Assume that A is a Noetherian ring. Then the polynomial ring
Arx1 , . . . , xn s is Noetherian.
160 ma4200—commutative algebra
An important special case is when the ground ring A is a field. Since fields are
Noetherian, the Basis Theorem tells us that polynomial rings over fields are
Noetherian. Moreover, quotients of Noetherian rings are Noetherian, and we
obtain directly the next corollary. In particular it says that algebras of finite
type over a field, a class of rings that include the coordinate rings of affine
varieties, are Noetherian.
Corollary 7.25 Any algebra finitely generated over a Noetherian ring is Noethe-
rian.
Finally, the last corollary we offer before giving the proof of Hilbert’s Basis
Theorem, combines that theorem with Proposition 7.10 on pag 154 which
states that localization preserves Noetherianess. Recall that and A-algebra is
said to be essentially of finite type if it is the localization of a finitely generated
A-algebra.
Corollary 7.26 Any ring essential of finite type over a Noetherian ring is Noethe-
rian.
Proof of Hilbert’s basis theorem: Let a be an ideal in Arxs and for each n
let an be the set of leading coefficients of elements from a of degree at most n.
Each an is an ideal in A, and they form an ascending chain which eventually
stabilizes, say for n “ N. Each an is finitely generated, so for each n ď N
we may chose a finite set of polynomials of degree at most n whose leading
coefficients generate an . Let f 1 , . . . , f r be all these polynomials in some order
and let a1 , . . . , ar be their leading coefficients.
We contend that the f i ’s generate a. So assume not, and let f be of minimal
degree n among those polynomials in a that do not belong to the ideal gen-
erated by the f i ’s. If the leading coefficient of f is a, it holds that a P an and
ř
we may write a “ j b j ai j with the f i j ’s corresponding to ai j of degree at most
ř pdeg f ´deg f i j q
the degree of f . Then f ´ j b j x f i j is of degree less than deg f and
does not lie in the ideal generated by the f i ’s since f does not contradicting the
minimality of deg f . o
Cohen’s criterion
One may wonder if there are conditions only involving prime ideals that en-
sure a ring be Noetherian. An acc-condition on prime ideals is far from
sufficient; there are non-Noetherian rings with merely one prime ideal.
For instance, let m be the ideal generated by all the variables xi in the ring
krx1 , x2 , . . .s of polynomials in infinitely many variables. The quotient krx1 , x2 , . . .s{m2
has only one prime ideal, namely the one generated by the images of all the
variables, but is not Noetherian since that prime ideal is not finitely generated.
However, a result of Irvin Cohen’s tells us that for a ring to be Noetherian it
suffices that the prime ideals are finitely generated.
chain conditions 161
7.27 We begin with stating a lemma about maximal ideals that are not finitely
generated; it joins the line of results of type ideals maximal subjected to some
condition being prime:
Lemma 7.28 Let a be maximal among the ideals in A that are not finitely generated.
Then a is a prime ideal.
Proposition 7.29 Assume that all prime ideals in the ring A are finitely gener-
ated. Then A is Noetherian.
Proof: Assume that A is not Noetherian. The set of ideals that are not finitely
generated is then non-empty and according to Zorn’s lemma has a maximal
Ť
element, say a; indeed, if the union i ai of an ascending chain of ideals were
finitely generated, the chain would stabilize (argue as in the last part of the
proof of Proposition 7.12 on page 155) and a member of the chain would
be finitely generated. From the lemma we infer that a is prime, which is a
flagrant contradiction. o
Proof of Lemma 7.28: The ring A{a is Noetherian since all its ideals are
finitely generated. Let a and a1 be two members of A and assume that the
product aa1 lies in a, but that neither a nor a1 lies there. Let c “ a ` paq and
c1 “ a ` pa1 q. These ideals both contain a properly and are therefore finitely
generated by the maximality of a, moreover it holds that cc1 Ď a because aa1 P c.
The quotient c{cc1 is a finitely generated module over A{a, and contains a{cc1 .
Hence a{cc1 is finitely generated, and by consequence a since cc1 is finitely
generated. o
vanish to the ν-th order. In this simple situation Krull’s theorem expresses
the well-known and obvious fact that no non-zero polynomial vanishes to
all orders. Of course, Krull’s result is a vast generalization of this prosaic
example; ideals in local Noetherian rings are infinitely more intricate than
maximal ideals in a ring of complex polynomials. K
7.30 The show begins with a technical lemma, and again submodules maxi-
mal subjected to a specific condition enter the scene:
Corollary 7.37 Assume that a is a proper ideal in the Noetherian integral domain
ai “ 0.
Ş
A, then i
Problems
7.11 The aim in this exercise is to exhibits a domain A with a maximal ideal
m, which is principal, such the intersection i mi is non-zero. It is in some a
Ş
7.12 Let A be a local ring with maximal ideal m. Assume that m is a principal
ideal. Prove that if i mi “ p0q, then the powers mi are the only ideals in A.
Ş
7.38 An ascending chain tMi u in an A-module M is called a composition series Composition series
if all of its subquotients Mi`1 {Mi are simple modules. Even though most (komposisjonsserier)
composition series we shall meet are finite, we do not exclude they being
infinite. By convention simple modules are non-zero, so in particular the
inclusions Mi Ĺ Mi`1 are strict. A finite composition series when displayed
appears like
0 “ M0 Ď M1 Ď . . . Ď Mn´1 Ď Mn “ M,
where each subquotient Mi`1 {Mi is shaped like A{mi for some maximal ideal
mi in A. The number n is called the the length of the series; it is the number of The length of a composi-
non-zero constituencies of the series. In case the series is infinite, the length is tion series (lengden av en
komposisjonskjede)
of course said to be infinite. More generally, the length of any finite chain will The length of a chain
be the number of inclusions; that is, one less than the number of modules. (lengden av en kjede)
Being a composition series is equivalent to being a maximal chain as de-
scribed in the introduction to this section; that is, no module is lying strictly
chain conditions 165
7.39 The main result of this section is that once a module has one finite
composition series, they are all finite and have the same length. This is a result
of Jordan-Hölder type, but one on the weak side—the true Jordan-Hölder
theorem states that the subquotients of any two composition series are the
same up to a permutation. The original Jodan-Hölder theorem is about (finite)
groups, but one finds analogues in many categories, so also in the subcategory
of Mod A of finite lengths modules.
The common length of the composition series is called the length of the mod- The length of modules
ule and denoted ` A pMq; for modules not being of finite length this means that (lengden til moduler)
0 “ M0 Ĺ M1 Ĺ . . . Ĺ Mn “ M.
The image of M in the quotient M{M1 is one shorter than M, hence all
composition series in M{M1 are of length n ´ 1 by induction. Denote by
β : M Ñ M{M1 the quotient map.
Given another another composition series N “ tNj u in M. Its length r is at
least n, and by induction its image in M{M1 is of length n ´ 1. Consequently
at least one of the inclusions in N becomes an equality in M{M1 ; that is, for
some index ν it holds that βpNµ q “ βpNµ`1 q, and we pick ν to be the least such
index. Then βpN q displays as
We contend that there is just one index µ for which equality occurs— this is
the fulcrum of the proof from which it clearly ensues that r “ n; indeed, on
the one hand, the length of βpN q is one less than that of N and on the other, it
equals n ´ 1 by induction .
166 ma4200—commutative algebra
Proof: We continue with the above proof and go on with induction on the
length; hence the two series βpMq and βpN q have the same subquotients up to
order. Now, the subquotients of M and βpMq differ only at the bottom stage
M1 , so M has the subquotient M1 in addition to those shared with βpMq. On
the other hand, the subquotients of N and βpN q coincide except at stage ν,
but in the proof above, we showed that Nv ` M1 “ Nv`1 , and the additional
subquotient of N is M1 as well. o
7.43 Just like the dimension of vector spaces the length is additive along
short exact sequence, and this is an indispensable property making it possible
to compute the length in many cases. Observe also that a submodule (or a
quotient) of M having the same length as M must be equal to M.
Examples
7.7 (Vector spaces) Over fields, of course the length of modules coincide with
their vector space dimensions. In a similar fashion, if the A-module M is killed
by a maximal ideal m in A, and therefore is a vector space over the field A{m,
one has ` A pMq “ dim A{m M.
7.8 (Finite groups) The only abelian groups that are of finite length are the
finite groups. They are all direct sums of cyclic groups of shape Z{pν Z where
p is a prime and ν a natural number; that is such a group M enjoys a direct
sum decomposition
Z{pi i Z.
à ν
M“
i
ř
We contend that `Z pMq “ i νi . By additivity of the length it suffices to show
that for a each prime p the length of Z{pν Z is given as ` A pZ{pv Zq “ v, and
this one does by an induction argument based on the standard4 mod pν . short 4
The map p is close to
exact sequences being a “multiplication-
by- p-map”; it sends a
p class rxs pν´1 mod pν´1
0 / Z{pν´1 Z / Z{pv Z / Z{pZ / 0. to the class rpxs pν
168 ma4200—commutative algebra
0 / mn´1 {mn / Mn / M n ´1 /0
M “ F3 piq ‘ F5 ‘ F5 ‘ F7 piq,
and we conclude that `Z pMq “ 6 but `Zris pMq “ 4. (The primes 3 and 7 persist
being primes in Zris, but 5 splits up in the product 5 “ p2 ` iqp2 ´ iq.)
Problems
7.17 Show that the the length `Z pZ{nZq equals the number of prime factors ˇ
of n (counted with multiplicity).
chain conditions 169
7.18 (Modules of finite length over pid’s.) Let A be pid and let f P A be an ˇ
element. Show that ` A pA{p f qAq is the number of prime factors in f (counted
with multiplicity).
7.49 We finish of the story about modules of finite length with a criterion for
a module being of finite length in terms of the support of the module, and a
structure theorem, essentially saying that a module is of finite length is just a
finite direct sum of "local contributions"—but of course, it says nothing about
how the local contributions are shaped.
/ ker Φ /M / / coker Φ / 0.
À
0 mPSupp M Mm
Given a prime ideal p there are two outcomes when the sequence is local-
ized at p. Either p does not contain any of the mi ’s, and then all terms of the
sequence become zero when localized, or p is one of the maximal ideals in
Supp M; say p “ m. If m1 ‰ m is another member of Supp M, there are ele-
ments in m1 not belonging to m, and hence pMm1 qm “ 0. Therefore the only
term in the sum that survives being localized at m is Mm , and the sum lo-
calises to Mm . But of course M localises to Mm as well, and the localization
map localizes to the identity! It follows that ker Φ and coker Φ localizes to
zero everywhere, and hence they are zero by the"The Localness of Being Zero"
principle (Proposition 6.62 on page 144). o
7.55 Recall that a module which is both Noetherian and Artinian is of finite
length. Hence Artinian rings are of finite length (regarded as modules over
themselves) and hence come with a natural numerical invariant, the length
lpAq; that is, the number of (simple) subquotients in a composition series.
chain conditions 171
This lemma concludes the proof of Akizuki’s theorem. Since A{J, being the
product of finite number of fields is Noetherian, we infer that A is Noetherian.
Proof: Consider the set of powers J r that are not finitely generated. If J is
not finitely generated, that set is non-empty and since A is Artinian it has
a a minimal element, say J m . In any case, there is an m so that J v is finitely
generated for v ě m.
The descending chain of powers J r becomes stationary at a certain stage as
well, that is, there is an m such that J ν`1 “ J ν whenever ν ě m.
Altogether, we infer there is an r so that J r is finitely generated and satisfies
J 1 “ J r and we are thence in the position to apply Nakayama’s lemma to the
r `
0 / J v `1 / Jν / J ν {J ν`1 / 0.
172 ma4200—commutative algebra
Theorem 7.58 Let A be an Artinian ring. Then Spec A is finite and discrete, and
the localisation maps A Ñ Am induce an isomorphism
ź
A» Am .
mPSpec A
Saying Spec A is finite and discrete is just another way of saying that all prime
ideals in A are maximal and finite in number. Anticipating the notion of Krull
dimension, a ring all whose prime ideals are maximal is said to be of Krull
dimension zero. Hence a Noetherien ring A is Atinian if and only if its Krull
dimension equals zero.
The theorem says nothing about local Artinian rings, even if they might
appear small and innocuous, they can be extremely intricate creatures.
Problems
7.20 Let n and m be two natural numbers and let a be the ideal a “ px m , yn q in ˇ
krx, ys. Show that A “ krx, ys{a is Artinian and compute its length.
7.21 Let n, m and r be three natural numbers and let A “ krx, y, zs{px n , ym , zr q.
Prove that A is Artinian and compute its length.
M
Lecture 8
Primary decomposition
Preliminary version 1.0 as of 2018-11-15 at 11:29:54 (typeset 3rd December 2018 at 10:03am)—
Much work remaining. Prone to misprints and errors and will change.
2018-11-15 Added exercise 6.20, corrected exercises 8.9 and ??
new important players, the primary ideals, and establish their basic properties.
Next follows the announcement of the main theorem and a discussion around
it, and finally the main theorem is proven through a series of lemmas.
‘
It is customary to say that a primary ideal q is p-primary when p “ q, or one p-primary ideals (p-
also says that p belongs to q. The converse of Proposition 8.3 does not hold in primære idealer)
general; the radical being prime is not sufficient for an ideal to be primary.
Example 8.1 below is an easy and concrete instance of this, and more elaborate
example in a polynomial ring is found in Exercise 8.4 below. However, if the
radical of q is maximal, q is primary:
Examples
8.1 The ideal a “ px2 , xyq in the polynomial ring krx, ys has a radical that is
prime, but a is not primary. Clearly the radical of px2 , xyq equals pxq which is
prime, but in the quotient krx, ys{px2 , xyq multiplication by y is neither injective
nor nilpotent (y kills the class of x, but no power of y lies in px2 , xyq). One
decomposition of px2 , xyq as an intersection of primary ideals is
indeed, the polynomials in px2 , xyq are those with x as factor that vanish at
least to the second order at the origin. This gives an example of the primary
decomposition not being unique.
Obviously the ideal px, y, zq being maximal is primary. The ideal pyq is more
interesting. Killing y, we obtain the ring A{pyq “ krX, Zs{pX 2 q, whose elements
are either non-zero divisors or nilpotent4 and pyq is a primary ideal. It’s 4
The elements are of
radical equals px, yq. the form apzq ` bpzqx
with a, b P krzs, and
one easily sees that this
is a non-zero divisor
K unless a “ 0, but then
the square is zero.
8.6 The intersection of finitely many p-primary ideals persist being p-primary.
In the analogy with the integers, this reflects the simple fact that the greatest
common divisor of some powers of the same prime number is a power of that
prime.
Proposition 8.7 If tqi u is a finite collection of p-primary ideals, then the intersec-
Ş
tion i qi is p-primary.
Proof: Taking radicals commutes with taking finite intersection (Lemma 2.63
‘Ş Ş ‘
on page 43), and therefore one has i qi “ i qi “ p. Assume next that
Ş Ş
xy P i qi , but y R i qi ; that is, xy P qi for each i, but y R qν for some ν. Since
‘
qν is p-primary x lies in the radical qν of qν , which equals p, but as we just
Ş
checked, p is as well the radical of the intersection i qi . o
The hypothesis that the intersection be finite cannot be discarded. Powers
mi of a maximal ideal are all primary and have the same radical, namley m,
but but at least when A is Noetherian, their intersection equals the zero ideal
p0q by Krull’s intersection theorem— which might be primary, but certainly
not m-primary (in most cases). There are however instances when infinite
intersection of primary ideals are primary.
Problems
8.1 With notation as in Example 8.2 on the previous page, show that pn is not ˇ
primary for any n ě 2. Hint: Show that zyn´1 P pn but yn´1 R pn .
8.2 Let tqi uiP I be a collection of primary ideals (of any cardinality) all with the ˇ
same radical p. Assume that there is a natural number n so that pn Ď qi for all i.
Ş
Prove that the intersection iP I qi is primary with p as radical.
8.3 Let the group G act5 on the Noetherian ring A and let q be a p-primary ˇ
ideal. Assume that p is invariant under G. Prove that gPG qg is p-primary and The action induces
Ş 5
an action on the
invariant under G. ideals by setting
ag “ t ga | g P G, a P a u
Proposition 8.9 Let S a multiplicative set in the ring A and let q be a p-primary
ideal. Assume that S X p “ H. Then q AS is p AS -primary and it holds true that
1
ι´
S pq AS q “ q.
8.12 Given a collection tSi u of set. It might very well happen that the inter-
Ş
section i Si does not change if one throws away one or more of the Si ’s (for
instance, if S1 Ď S2 , one stupidly has S1 X S2 “ S1 ) and in that case one says that
intersection is redundant. In the opposite case, that all the Si contribute to the Redundant intersections
Ş Ş
intersection, or in other words, when i Si Ř i‰ j Si for all j, the intersection (redundant snitt)
a “ q1 X . . . q r (8.1)
where qi ’s are primary ideals. We have already seen several examples (exam-
ples 8.1 and 8.2 above).
Without further constraints there are several trivial6 ways such a decom- 6
And it can be in non-
trivial ways too; we
position can be ambiguous. First of all, it can be an irredundant intersection.
shortly return to those.
Secondly a p-primary ideal can be the intersections of other p-primary ideals
in infinitely many different ways (see the upcoming example 8.3). The first
type of ambiguity is coped with by just discarding superfluous ideals, and
Proposition 8.7 above helps us coping with the second. We just group those
qi ’s with the same radical together, and replace them by their intersection,
which will be primary with the same radical.
The primary decomposition (8.1) is called minimal or reduced if all the Minimal or reduced
‘
radicals qi are different and the intersection is irredundant. primary decompositions
(minimale eller reduserte
primærdekomposisjoner)
Lemma 8.14 Any primary decomposition a “ q1 X . . . X qr can be rendered a
‘
minimal one; that is, an irredundant intersection with the radicals qi being distinct.
Example 8.3 Consider m2 “ px2 , xy, y2 q in krx, ys (where m “ px, yq). For all
scalars α and β with α ‰ 0 one has the equality
Indeed, this amounts to the two lines generated by the class of y and the class
of αx ` βy in the two dimensional vector space m{m2 intersecting in 0. K
Example 8.4 If the ring A is a pid, there is nothing much new. The prime ide-
als are the principal ideals ppq generated by an irreducible p. The ppq-primary
ideals are those generated by powers of p; that is, those on the form ppv q. In
primary decomposition 179
ν
general, if f “ p11 . . . prν is a factorisation of f into a product of irreducible
elements, the primary decomposition of p f q is unambiguous and it is given as
The same applies to principal ideals in any ufd; but be warned that not all ideals
are principal! K
Finally in this paragraph, we notice that by Proposition 8.9 a primary
decomposition localises well:
Proposition 8.15 Assume that S is a multiplicatively closed subset of A and that
a “ q1 X . . . X qr is a primary decomposition of a with radicals pi . Then a AS “
q1 AS X . . . X qr As , and either qi AS is primary with radical pi AS or qi AS “ AS
Proof: Since the ring A is assumed to be Noetherian, the set of ideals for
which the conclusion fails, if non-empty, has a maximal element a. Replacing
A by A{a we may assume that the zero ideal is not the intersection of finitely
many primary ideals (in particular it is not primary), but that all non-zero
ideals in A are.
In A there will be two elements x and y such that xy “ 0, but with x ‰ 0
and y not nilpotent. The different annihilators Ann yi form an ascending
chain of ideals, hence Ann yν “ Ann yν`1 for some ν. We contend that p0q “
Ann y X pyν q. Indeed, if a “ byν is an element in pyν q that lies in Ann y, one
has ay “ byν`1 “ 0 and therefore b P Ann yν`1 “ Ann yν , and it follows that
a “ byν “ 0. Now, x P Ann y is a non-zero element, and since y is not nilpotent,
both ideals pyν q and Ann y are non-zero and are therefore finite intersections
of primary ideals; the same is thus true for p0q. o
180 ma4200—commutative algebra
Proof: Start with any decomposition of an ideal a into primary ideals (there
is at least one according to the proposition above). By Lemma 8.14 on page 178
it can be made minimal by regrouping ideals with the same radical and
discarding redundant ones. o
Passing to the quotient A{a and observing that pa : xq{a equals the annihilator
p0 : rxsq of the class rxs in A{a, the theorem has the equivalent formulation
(remember Proposition 8.11 on page 177) which is the one we shall prove:
8.22 The radicals of the primary components are of course tightly related to
the ideal, vaguely analogous to the prime factors of an integer, and they merit
a proper name. They are called the associated prime ideals of a, and the set they Associated prime ideals
constitute is denoted by Ass A{a. In particular, Ass A will be the set of prime (assosierte primidealer)
relations between members of either family, the families coincide (Lemma 2.31
on page 34). Hence the sets Spec A and Ass A have the same minimal primes.
We have proved:
Proposition 8.25 In a Noetherian ring A the sets Spec A and Ass A have the
same minimal elements; in other words, the minimal primes of A are precisely the
isolated associated primes. In particular, there are finitely many minimal primes.
8.26 We have seen the intersection of the associated primes of A is the set of
nilpotent elements, and their union turns out to be the set of zero divisors:
Proposition 8.27 The set of zero-divisors in a Noetherian ring A equals the union
Ť
pPAss A p of the associated primes.
a “ px2 y, y2 z, z2 xq.
This is however not the whole story. The element xyz lies in the intersection
to the right, but not in a. Now, clearly px, y, zq Ď pa : xyzq and px, y, zq being
maximal, it holds that pa : xyzq “ px, y, zq, so there must be an px, y, zq-primary
component. After a few tries (and a more failures) one finds the equality
The associated primes are pz, yq, px, yq, pz, xq and px, y, zq. Once one has
guessed correctly, it is relatively easy to check the answer. All involved ideals
are generated by monomials, and such ideals have the nice property that they
primary decomposition 183
have a polynomial as member if and only if all the monomial terms of the
polynomial are members. Hence it suffices to check that every monomial lying
in the ideal to the right lies in the one to the left as well. But monomials in
px2 , y2 , z2 q have either x2 , y2 or z2 as a factor, and by symmetry we may as-
sume it is y2 . Lying in pz, z2 x, x2 q too, our monomial must have either z or x2
as a factor and thereby also zy2 or x2 y2 ; but these two both lie in a, and we are
done. K
and both px2 , yq and px, yq2 are primary components. Notice that they have the
same radical (they must!) and that they are embedded components. Now, the
bad news are that px2 , xyq, have infinitely many different primary decompo-
sition (Example 8.6 below), but the good news are that merely the embedded
components differ. This is generally true. The Second Uniqueness Theorem
we are about to prove, states that isolated components are unique. The point
is that the isolated components may be expressed in terms of the correspond-
ing associated primes which are independent of the decomposition. We shall
see that if q is the component and q “ p it holds that q “ ι´1 pa Ap q where
‘
q “ ι´1 pa Ap q, (8.2)
from which the theorem ensues since the isolated components are invariants
of a.
Ş
To establish (8.2) express the decomposition of a as a “ q X i qi where the
intersection extends over the primary components different from q. Localizing
at p one finds
č
a A p “ q A p X qi A p “ q A q (8.3)
i
since the qi ’s blow up when localized; that is, qi Ap “ Ap . Indeed, since p is
184 ma4200—commutative algebra
holds true in the polynomial ring krx, ys for any natural number n, and is an
example of infinitely many different minimal primary decompositions. Indeed,
the Ď -inclusion is obvious, and to check the other, assume that a belongs to
the right side. Then
a “ p ¨ x “ q ¨ x2 ` r ¨ yn ` sxy
with p, q r and s polynomials in krx, ys. It follows that x divides r, and hence
that f P px2 , xyq. K
Example 8.7 So far all our examples have merely involved monomial ideals,
but of course most idelas are not shaped like that. Primary decompositions are
notoriously strenuous to lay hands on, and the monomial ideals are among
the easiest to handle, hence their tendency to appear in texts. However, we
are obliged to give at least one example of a more mainstream situation. It
illustrates as well that the decomposition is largely of a geometric nature; that
is, at least the isolated associated prime ideals are; the primary components
may conceal subtler structures.
We shall analyse the familiar case of the intersection of two quadratic
curves; the unit circle centred at p0, 1q and a standard parabola. So let a “
px2 ` py ´ 1q2 ´ 1, y ´ x2 q in krx, ys, where k is any field of characteristic
different from 2. A standard manipulation shows that the common zeros
of the two polynomials are the points p1, 1q, p´1, 1q and p0, 0q, and the same
manipulations give
a A “ px2 ´ 1, y ´ x2 q “ px ` 1, y ´ x2 q “ px ` 1, y ´ 1q.
primary decomposition 185
Finally, in C “ krx, ysp x, yq both x ` 1 and x ´ 1 have inverses, and we see that
aC “ px2 , yq.
a “ px ´ 1, y ´ 1q X px ` 1, y ´ 1q X px2 , yq.
Problems
and that this is a minimal primary decomposition. Show that different scalars
a give different decompositions.
8.7 Let a be the ideal in the polynomial ring krx, y, zs given as a “ pyz, xz, xyq. ˇ
Show that the minimal primary decomposition of a is shaped like
8.8 Let p be the ideal in the polynomial ring krx1 , . . . , xn s over a field k gener- ˇ
ated by the r first variables; that is, p “ px1 , . . . , xr q. Show that every power pm
is p-primary. Hint: Consider pm Krx1 , . . . , xr s where K “ kpxr`1 , . . . , xn q and
show that pm Krx1 , . . . , xr s X krx1 , . . . , xn s “ pm .
a “ p1 X p22 X . . . X prr´ 1 r
´1 X p r
8.10 (Symbolic powers.) We have seen that the powers pn of a prime ideal p
in A are not necessarily p-primary (unless p is maximal). But there is an in
some sense canonical primary ideal associated to the powers pn ; the so-called
symbolic power ppnq . They arise in the following way. The ideal p Ap is maximal Symbolic powers
in the local ring Ap and its powers are therefore primary by Proposition 8.4 (symbolske potenser)
on page 175. Pulling primary ideals back along the localization map ι results
in primary ideals (Proposition 8.9 on page 177), the ideal ppnq “ ι´1 pn Ap
(or when ι is injective ppnq “ A X pn Ap ) will be primary, and this is the n-th
symbolic power of p.
n`m
a) Show that if n and m are natural numbers it holds that ppnq ¨ ppmq Ď p
b) Show that pn “ ppnq if and only if pn has no embedded component.
c) With the notation as in Example 8.2 on page 175, let p “ px, yq and
determine the symbolic square pp2q .
Preparitions
8.29 We start with a little lemma (in fact, we already met a version of it in
Exercise 1.15 as early as on page 17).
Lemma 8.30 Let R be a graded ring and assume that x and y are two elements such
that x ¨ y “ 0. Let xe be the lowest homogeneous term of x. Then xeν y “ 0 for some ν.
In particular xeν yi “ 0 for each graded part yi of y.
primary decomposition 187
Proposition 8.32 Let R be a graded ring and let p be a prime ideal associated to
the homogeneous ideal a. Then p is homogeneous.
Lemma 8.34 Let q be primary ideal in the graded ring R whose radical is homoge-
neous. Then q7 is primary and
‘ 7 ‘
q “ q.
finally end up with an element whose term of lowest degree does not belong
to q7 . With y f in place outside q7 , it follows that y f does not belong to q, hence
x P q “ q7 .
‘ ‘
o
Proof: Observe first that all prime ideals associated to a homogeneous ideals
are homogeneous. (Proposition 8.32 above).
It is fairly clear that pa X bq7 “ a7 X b7 (the homogeneous elements in a X b
are the honogenous elements the lie in both a and b!!) so starting out with a
Ş
minimal primary decomposition a “ i qi and applying the 7-construction to
it, one arrives at a decomposition
č 7
a “ a7 “ qi , (8.4)
i
and according to Lemma 8.34 on the preceding page, this is a primary decom-
position. Moreover, the radicals of the sharpened ideals q7i are the same as
the radicals of the qi ’s, and we can conclude that (8.4) is a minimal primary
decomposition. o
component of a.
primary decomposition 189
Primary modules
Given a submodule N of the module M, one has the transporter ideal pN : Mq
consisting of ring elements that multiply M into N; that is
pN : Mq “ t x P A | xM Ď N u,
‘
and the radical pN : Mq is called the radical of N relative to M. The elements
are the ring elements such a power multiplies M into M. If N “ 0, they consist
of the ring elements inducing a nilpotent homothety on M.
and one extends the notion of primary ideals to modules in the following
fashion:
190 ma4200—commutative algebra
Or in other words
Integral extensions
Preliminary version 1.1 as of 2018-11-19 at 16:45 (typeset 3rd December 2018 at 10:03am)—For the
moment nothing here!! Prone to misprints and errors and will change.
Completely rewritten and extended a lot. For the moment not many exercises or examples—
hopefully this will follow soon. Also probably peppered with misprints and impressions (sorry,
improvements will shortly follow) and organisation will also be better soon.
19/11/2018 Corrected som mistakes and imoroved the text many places.
From an algebraic point of view there is a huge difference between the ring
of integers Z and the field of rationals Q; we need only mention the primes.
They are visible in Z as generators of the prime ideals, but in Q they are, at
least from an algebraic point of view, on an equal footing with the other units.
When the exploration of number fields1 begun, an immediate want arose 1
That is, finite field
to have subrings playing the role of the integers and where the deep secrets extensions of the
rationals
of the field could be revealed. These rings was made up of the “integral
elements” in the field, or more precisely those being integral over Z.
In algebraic geometry these rings give rise to what is called normal varieties
where the geometry of the codimension one subvarieties strongly influence the
geometry of the entire space.
9.1 If all the elements in B are integral over A one says that B is integral over
A or that B is an integral extension of A. The subset of B consisting of the Integral extension (hel
elements integral over A is called the integral closure of A in B and is denotes utvidlese)
Integral closure (helavs-
by A.
s Of course it depends on B, but to keep notation simple, we do not
lutningen)
include a reference to B in the notation—the context will make it clear where
the integral closure is taken.
It is a basic, but subtle fact that the integral closure is ring, which we shortly
shall prove. Finally, one says that A is integrally closed in B if A “ A; s that is, Integrally closed (helavs-
luttet)
every element which is integral over A belongs to A.
9.2 Both in algebraic geometry and algebraic number theory the integral clo-
sure of a domain A in its field of fractions is a frequently used construction
and we shall use the notation A r for it to distinguish it from all the crowd. Do-
mains being integrally closed in their field of fractions; that is, those satisfying
r are called normal, and for general domians A
A “ A, r is sometimes called the Normal rings (normale
normalization of A. ringer)
Normalization (normalis-
9.3 To illustrate the difference between algebraic an integral dependence ering)
relation consider the two simple equations
y2 ´ z “ 0
pz ´ 1qy2 ´ z “ 0
‘
over the complex numbers. The first one “has z as a solution”, but due to the
ambiguity of the square root, it is impossible to find a continuous (yet alone
analytic) solution in the entire plane Only in simply connected domain not
containing the origin can a continuous solution be found. The solutions of the
second equation suffer the same defect, but additionally they acquire a pole at
z “ 1. The difference between solutions of algebraic and integral relations is
precisely the occurrence of poles in the former.
Example 9.1 The golden section p1 ` 5q{2 is integral over Z as it satisfies the
‘
equation
x2 ´ x ´ 1 “ 0.
The integral closure of Z in Qp 5q equals Zrp1 ` 5q{2s. Indeed, a number
‘ ‘
‘
a ` b 5 has the minimal equation
and if it is integral over Z, the coefficients are integral by Gauss lemma (Exer-
cise 3.1 on page 64). Then n “ 2a P Z and 20b2 P Z, and hence m “ 2b P Z as
well. Substituting back, gives 0 ” pn2 ´ 5m2 q ” pn2 ´ m2 q mod 4, which holds
if and only if m and n have the same parity. K
Example 9.2 The ring of integers in the number field Qpi 5q equals Zri 5s.
‘ ‘
x2 ´ 2ax ` a2 ` 5b2 “ 0.
integral extensions 193
x n “ ´pan´1 x n´1 ` . . . ` a1 x ` a0 q,
Proof: The only implication that shows any substantial resistance is that the
last statement entails the first. So let M be a module as in the last assertion
and let m1 , . . . , mn be generators for M over A. One may express each element
x ¨ mi in terms of the m j ’s, and this gives relations
x ¨ mi “ f ij m1 ` . . . f in mn
Corollary 9.6 If x is integral over A, all elements in Arxs are integral over A.
9.7 There is a close relationship between integral and finite extensions as was
unveiled in the previous proposition. Finite extensions are integral, but in gen-
eral the converse is not true. There are even examples of Noetherian domains
whose normalizasion A r is not a finite module over A; they are however, rather
exotic creatures, and the lions share of the rings appearing in mainstream
algebraic geometry—that is, domains finitely generated over field and their
localizations—have normalizations which are finitely generated as modules.
9.8 The first conclusion to be drawn from the Basic Characterization is that
finitely generated algebras which are integral, are finitely generated modules;
an important observations since the integral closure being finite over A or not,
is an issue. One has
Proof: This is just a combination of Corollary 9.6 and the transitivity property
(Corollary 9.10). Indeed, let x and y be integral over A. The ring A[x] is an
integral extension of A and y being integral over A, the extension Arx, ys
is integral over Arxs. Hence Arx, ys is integral over A by the transitivity. In
particular, this applies to both the product x ¨ y and the sum x ` y.
The last statement of the proposition might appear as a tautology, but an
argument is in fact needed. We have to see that elements integral over A s are
integral over A, which is exactly what Corollary 9.10 above tells us since A s is
integral over A. o
x n ` a n ´1 x n ´1 ` . . . ` a i x i ` . . . ` a 0 “ 0
and it follows that tx is integral over A. We conclude that xs´1 “ ptxqs´1 t´1 P
p Aq
s S. o
9.15 Next comes the quotients, and in addition to the usual the staging of this
chapter with an integral extension A Ď B, an ideal b in B is given. We let a be
the ideal b induces in A; that is, a “ A X b. Then A{a Ď B{b is an extension, and
it persits being integral.
yn ` an´1 yn´1 ` . . . ` a1 y ` a0 “ 0
with the ai ’s from A. Reducing that relation modulo b gives the relation
where rai s as usual denotes the classes of ai in A{a. Hence x is integral over
A{a. o
Problems
9.1 Let B be a ring and tBi uiP I a family of subrings of B. If each Bi is inte-
Ş
grally closed in B, then the intersection iP I Bi is integrally closed as well.
9.2 Assume that tBi uiP I is a family of subrings of a field K. Prove that if each
Ş
Bi is integrally closed in K, then the intersection iP I Bi is as well.
9.3 Show that x is integral over A if and only if there is a square matrix with
coefficients in A having x as an eigenvalue.
9.4 Show that if x and y are eigenvalues for Φ and Ψ then x ¨ y is and eigen-
value for the Kronecker product ΦbΨ and that x ` y is one for the matrix
Φb Im ` In bΨ where the In are Im are identity matrices of appropriate size.
Conclude that the integral closure is a ring.
M
integral extensions 197
Proof:
Notice first, that all the localization Am have K as fraction field as well.
Consider the inclusion A ãÑ A, r which fits into the short exact sequence
0 /A /A
r / A{A
r /0
0 / Am / p Aq
rm / p A{Aq
r m
/0
9.2 Examples
We indulge ourselves in two examples of rather large classes of rings that
are normal. Firstly, the unique factorization domains are always normal, and
secondly, there is a principle that rings of invariant of the actions of a finite
groups are normal, at least when the ring upon which the group act is normal.
This class includes the so-call "quotient singularities". We shall treate a few
simple examples in detail but leave the general case to the zealous students in
the form of a guided exercise.
198 ma4200—commutative algebra
z n ` a n ´1 z n ´1 ` . . . ` a i z i ` . . . ` a 0 “ 0
Every irreducible factor of y divides the right side, hence it divides the left
side and consequently also x. Contradiction. o
and t2 “ pt2 ´ 1q ` 1 P A.
Moreover, the ratio between the two generators of A equals t, so that its
fraction field is Cptq. Now, the polynomial ring Crts being a ufd is normal,
and we can conclude that Crts equals the normalization of A. It is typical for
curves that their normalisation “resolves the singularities”; that is, it separates
the different branches of the curve passing through the double points (or
points of of higher multiplicity). K
Abraham Seidenberg
(1916–1988)
American mathematician
integral extensions 199
Rings of invariants
9.22 Now comes the promised result on rings of invariants, and as promised,
we shall proceed rather in relaxed way merely treating the simples possible
case. That is, the case of a cyclic group of order two acting on a normal do-
main B. Such an action is just given by an involution on B; in other words, a Involutions (invo-
ring map σ : B Ñ B with σ2 “ idB . The map σ extends to an involution of lusjoner)
Kσ “ L Ď K
x2 ´ pσpxq ` xqx ` σpxqx “ 0. (9.6)
Ď
Bσ “ A Ď B
Both σpxq ` x and xσpxq are invariant under σ and belong therefore to A, hence
(9.6) is an integral dependence relation for x over A.
Finally, as B is integral over A, the integral closure of A in K equals B
r by
transitivity, and hence A “ B X L. From this ensues that A “ B X L “ A in case
r r r
B “ B.r o
which shows that y is invertible in A. Next assume that A is a field, and let
x P B be given. It satisfies a relation
x n ` a1 x n´1 ` . . . ` a1 x ` a0 “ 0,
1 n´1
x ¨ a´
0 px ` a1 x n´2 ` . . . ` a1 q ` 1 “ 0.
is integral by Proposition 9.16 on page 196, and the corollary ensues since the
quotient by an ideal is a field if and only if the ideal is maximal. o
prime ideals in B intersecting A in a given fixed prime ideal. The fibres are not
empty, or in other words, all prime ideals p in A are of the form p “ q X A, and
moreover, there are no inclusion relations between the members of the fibres.
In topological terms this means that the fibres of π are discrete topological
spaces.
9.27 Here it comes:
Proposition 9.28 (Lying-over) The map π is surjective, and the fibres are
discrete. In other words, for each p Ď A there is at least one q Ď B such that q X A “ p.
Moreover, if q and q1 are prime ideals in B with q X A “ q1 X A and q Ď q1 , then
q “ q1 .
integral extensions 201
Proof: We begin with treating the local case; the rest of the proof is a reduc-
tion to that case.
Assume then that A is local with maximal ideal m. Let n be any maxi-
mal ideal in B which exists according to the Basic Existence Theorem (Theo-
rem 2.50 on page 40). By Corollary 9.26 n X A is maximal, hence equal to m
since m is the only maximal ideal in A.
To see that no ideal in the fibre is contained in another, assume that q Ď q1
and that both intersect A in m (we are still in the local situation). Again by
Corollary 9.26 both q and q1 are maximal and must be equal since q Ď q1 .
Let p be a prime ideal in A. In order to reduce the general case to a local
situation we pass to the localized Ap Ď Bp extension, which persists being
integral in view of Proposition 9.14. According to what we established in the
local case, there is an ideal in Bp , which as all ideals in Bp is of the form qBp ,
such that qBp X Ap “ pAp . But then q X A “ p, this follows e. g. as
pAp X qBq q X A “ A X qAp “ a X pB X qBp q “ A X q
.
If q X A “ q1 X A “ p for two prime ideals in B, one included in the other, it
holds by the local case that qBp “ q1 Bp , and hence q and q1 are equal. o
Example 9.4 It might well happen that π is surjective and has finite fibres y
without B being integral over A. An example can be the extension
A “ krx2 s Ď krx, px ´ 1q´1 s “ B
where we assume that k is algebraically closed and not of characteristic two.
The geometric interpretation of the example is the parabola X given as y “ x2
with a hole punched in it; that is, the point p1, 1q is removed. The map π is just
projection Xztp1, 1qu to the y-axis.
The ring krx2 s is isomorphic to the polynomial ring krys (rebaptize x2 to y). x
Every maximal ideal m in A is of therefore the form x2 ´ a2 (all elements in k
have a square-root) with a ‰ 0, and it holds true that px ´ aqB X A “ px2 ´ a2 qA
since x2 ´ a2 “ px ` aqpx ´ aq. However px ´ 1q´1 is not integral over A; indeed,
if it were, multiplying an integral dependence relation of degree n by px ´ 1qn ,
would yield a relation
1 ` pn´1 px ´ 1q ` . . . ` p1 px ´ 1qn´1 ` p0 px ´ 1qn “ 0,
where the coefficients pi ’s are elements on krx2 s; that is, they are polynomials
pi px2 q in x2 . Putting x “ 1 gives an obvious contradiction.
As of the fibres, it is a nice exercise to check that if a ‰ 1, the two prime
ideals (x-a)B and px ` aqB are the ones lying over px2 ´ a2 qA, but the sole prime
ideal lying over px ´ 1qA is px ` 1qB. K
Problem 9.5 Let A Ď B be an integral extension. Show that the map π : Spec B Ñ
Spec A is a closed map, by showing that πpVpaqq “ Vpa X Aq for any ideal a in
A. Hint: Apply Lying–Over to the extension A{a X A Ď B{a. M
202 ma4200—commutative algebra
Going–Up
The Going–Up Theorem is about extending, or lifting as one also says, ascend- The Going–Up Theorem
ing chains of prime ideals in A to chains in B by climbing them; contrary to (Going–Up teoremet)
Ď
9.29 The one step case is what is usually called the Going–Up Theorem p0 Ď p1 Ď A
Corollary 9.31 (Going–Up II) Any finite chain tpi u of prime ideals in A has a
chain tqi u of prime ideals in B lying over it.
Proof: The proof goes by induction on the number of prime ideals in the
chain A, and one should find it completely transparent pondering the follow-
ing display:
q 0 Ă q 1 Ă . . . Ă q n´ 1
Ď
p0 Ă p1 Ă ... Ă pn´1 Ă pn
The upper chain exists by induction, an one just fils in the upper right corner
citing the Going–Up Theorem. o
9.32 A chain tqi u in B that lifts the chain tpi u, will be saturated whenever tpi u
is; indeed, any prime strictly in between qi and qi`1 would either meet A in
pi or pi`1 since tpi u is saturated, but this can not happen since Lying–Over
guarantees there are no inclusions among primes in the fibres. We conclude
that the suprema of the lengths of chains in the two rings coincide, and hence
we have
Going–Down
The Going–Down Theorem the version for descending along chains. One says
that Going–Down property holds for an extension A Ď B if for any pair p1 Ď p0 The Going–Down
of prime ideals in A and any prime ideal q0 in B lying over p0 there is a prime property (Going–Down-
egenskapen)
integral extensions 203
ideal q1 contained in q0 and lying over p1 ; in other words lifted chains may be
continued downwards. When A and B domains and A integrally closed in its
fraction field the Going-Down property holds for te extension A Ď B, but it is
significantly more subtle to establish than Going-Up.
9.34 As indicated we content ourself by formulating the Going–Down theo-
rem.
a function with a pole, namely the extension krx, 1{xs of krxs. The geometric
counterpart is the projection of the classical hyperbola, xy “ 1 onto the x-axis,
the hyperbola just being the graph of the function 1{x.
The ring krx, 1{xs is not finite over krxs, but perturbing x slightly, we obtain
a subring over which krx, 1{xs is finite. The subring krx ` 1{xs will do the job;
indeed, krx, 1{xs “ krx, x ` 1{xs is generated by x over krx ` 1{xs and one has
the integral dependence relation
x2 ´ xpx ` 1{xq ` 1 “ 0.
It is remarkable that almost any perturbation of x will work; that is, krx, 1{xs
is finite over krax ` b{xs as long as both the scalars a and b are non-zero. K
Problem 9.6 Show that krx, 1{xs is a finite module over krax ` b{xs for any
scalars a b both being different from zero. M
9.37 The proof of Noether’s Normalization Lemma goes by induction of the
number of generators A requires as an algebra over k, and the basic ingredient
in the induction step is the following lemma:
Lemma 9.38 Let k be an infinite field and let A “ krX1 , . . . , Xm s{a. Assume that
A is a domain whose fraction field K has transcendence degree at most m ´ 1 over
k. Then there are elements y1 , . . . , ym´1 in A such that A is a finite module over
kry1 , . . . , ym´1 s.
Proof: Following our usual convention, we let xi denote the image of the
variable Xi in A. Since the transcendence degree of K over k is less than m, the
m elements x1 , . . . , xm can not be algebraically independent and must satisfy
and equation
ppx1 , . . . , xm q “ 0,
where p is a non-zero polynomial with coefficients in k. Let d be the degree of
p and let pd be the homogenous component of degree d. Now, perturbe the
xi ’s by putting yi “ xi ´ αi xm for i ď m ´ 1 where the αi ’s are scalars to be
chosen. This gives3
3
Recall that for any
0 “ ppx1 , . . . , xm q “ ppy1 ` α1 xm , . . . , ym´1 ` αm´1 xm , xm q “ polynomial ppxq
it holds true that
“ pd pα1 xm , α2 xm , . . . , αm´1 xm´1 , xm q ` qpxm , y1 , . . . , . . . , ym´1 q “ ppx ` yq “ ppxq `
d yqpx, yq where q is a
“ pd pα1 , . . . , αm´1 , 1qxm ` qpxm , y1 , . . . , . . . , ym´1 q polynomial of total
degree less than the
where q is a polynomial of degree less that d in the last variable xm . Now, degree of p. Apply this
since the ground field is infinite, for a generic choice of the scalars αi it holds to pd .
true that pd pα1 , . . . , αm´1 , 1q ‰ 0 (see Exercise 9.9 below). Hence the element xm
is integral over kry1 , . . . , ym´1 s and by consequence, A is a finite module over
the algebra kry1 , . . . , ym´1 s. o
9.39 By induction on m one obtains the full version of the normalization
lemma:
integral extensions 205
Problems
9.8 Prove the formula in Footnote number 3. Hint: Use the binomiall theo-
rem to check it for the polynomials px ` yqν .
9.9 In stead of the linear change of variables yi “ xi ´ αi x1 , show that the proof
n
of Lemma 9.38 goes through over any field with the change yi “ xi ` x1 i when
ni are chosen sufficiently large.
M
206 ma4200—commutative algebra
Corollaries
The Nullstellensatz
Hilbert’s Nullstellensatz is about the composition of I and V the other way
?
around, namely about IpVpaqq. Polynomials in the radical a vanish along
?
Vpaq and therefore a Ď IpVpaqq, and the Nullstellensatz tells us that this
inclusion is an equality. We formulate the Nullstellensatz here, together with
two of its weak avatars, but shall come back with a thorough discussion of the
proof(s) a little later.
Notice that the ground field must be algebraically closed. Without this as-
sumption the result is not true. The simplest example of an ideal in a polyno-
mial ring with empty zero locus is the ideal px2 ` 1q in Rrxs.
Proof: By version I, the field krx1 , . . . , xr s{m is algebraic over k hence equa
to since k is assumed to algebraically closed. The ensuing homomorphism
krx1 , . . . , xn s Ñ k has xi ´ ai in its kernel m, and since we already know that
integral extensions 207
9.45 We proceed to present the J.L. Rabinowitsch trick proving that the weak
version of the Nullstellensatz (Theorem ?? on page ??) implies the strong.
?
That is, we need to demonstrate that IpZpaqq Ď a for any proper ideal a in
krx1 , . . . , xn s.
The crux of the trick is to introduce a new auxiliary variable xn`1 and for
each g P IpZpaqq to consider the ideal b in the polynomial ring krx1 , . . . , xn`1 s
given by
b “ a ¨ krx1 , . . . , xn`1 s ` p1 ´ xn`1 ¨ gq.
9.6 Appendix—skirmishes
We have relegated a few simple results of preparatory character to this ap-
pendix. They do not take part in the main battle, but are merely skirmishes
on the flanks, though of significant importance for the progress. At least the
two firsts are easy and elementary. The concept of transcendence degree might
be a little more involved, but should be known to mosts students from earlier
courses.
det φ ¨ ι “ Φ ˝ Φ ˝ ι “ 0.
:
This means that pdet Φ ¨ m1 , . . . , det Φ ¨ mn q “ det Φ ¨ ιp1q “ 0; hence det Φ kills
M as the mi ’s generate M. o
Lemma 9.47 Let A Ď B Ď C be a tower of rings and assume that C i finite over B
and B is finite over A, then C is a finite over A.
Transcendence degree
a n x n ` a n ´1 x n ´1 ` . . . ` a 1 x ` a 0 “ 0 (9.7)
integral extensions 209
where the ai ’s are elements form k and an ‰ 05 , and the transcendental ones
are the rest; that is, those for which ppxq ‰ 0 for any non-zero-polynomial in 5
Since K is field we
krXs. may as well assume
More generally a collection x1 , . . . , xr of element from the bigger field K is that an “ 1
said to be algebraically dependent over k if for some non-zero polynomial p in Algebraically dependent
r variables and coefficients from k it holds true that ppx1 , . . . , xr q “ 0, and of elements (algebraisk
avhengige elementer)
course, if no such polynomial can be found, the collection is called algebraically
independent over k. Algebraically independent
elements (algebraisk
A collection of algebraically independent elements x1 , . . . , xr is a tran-
uavhengige elementer)
scendence basis of K over k if it is maximal; that is firstly, it is algebraically transcendence basis
independent and secondly, adding any new element to it will make it alge- (transcendence basis)
braically dependent. In terms of the field K that x1 , . . . , xr is a transcendence
basis means that the field K is an algebraic extension of kpx1 , . . . , xr q, and that
kpx1 , . . . , xr q is isomorphic to the rational function field over k in r variables..
We cite the following result but refrain from giving a proof:
Proposition 9.49 Every field extension K of k that is not algebraic, has a transcen-
dence basis. All transcendent bases for K over k have the same number elements.
That common number is called the transcendence degree of K over k and de- Transcendence degree
noted by trdegk pKq. (transcendensgrad)
9.50 In linear algebra The Tietze’s Extension Lemma is used to show that
bases of vector spaces have the same number of elements. An analoguous
result will give that different transcendence bases of fiels extension have same
cardinality.
Lemma 9.51 (Exchange emma) Let k Ď L Ď K be a tower of fields and let a and
b be two elements of K. Assume that b is transcendent over L but algebraic over Lpaq,
then a is algebraic over Lpbq.
Proof: Since b is algebraic over Lpaq, there is a polynomial f px, yq with coeffi-
cients from L such that f pa, yq ‰ 0 but f pa, bq “ 0. Expandingh f in powers of
x yields
ÿ
0 “ f pa, bq “ qi pbqai ,
wehere qi pyq P Lrys. This looks very much like a dependence relation for a over
Lpbq, it only remains to see that f px, bq is not identically zero, but since b is
trancsendent over L and at least one of the polynomials qi pyq’s is non-zero, this
holds true. o
Proof of Trancendence bases: We only treate the case of finite bases. Let
A “ ta1 , . . . , an u and B “ tb1 , . . . , bm u be two transcendence bases for K over k.
o
Lecture 10
Krull dimension
Very preliminary version 0.01 as of 2018-12-03 at 10:01 (typeset 3rd December 2018 at 10:03am)—
Prone to misprints and errors and will change.
03/12/2018 Several changes, still unfinished.
The definiton
212 ma4200—commutative algebra
10.1 Let A be a ring. The point is to consider strictly ascending and finite
chains tpi u of prime ideals
p0 Ă p1 Ă . . . Ă pν .
Recall that ν is called the length of the chain; it is one less than the number of
prime ideals, or if you want, the number of inclusions. The Krull dimension Krull dimension (Krull
of A be the supremum of the length of all such chains in A. It is denoted by dimensjon)
dim A. We shall say that a chain is saturated if there are no prime ideals of saturated chains (mettede
A lying strictly between two of the terms, and it is maximal if additionally it kjeder)
Maximal chains (maksi-
neither can be lengthened upwards nor downwards. male kjeder)
10.2 Even if each chain is finite, there might be arbitrary long chain, and the
Krull dimension will in that case be infinite. It is easy to find examples among
non-noetherian rings whose Krull dimension is infinite; Example10.1 below is
an obvious example of one with an infinite ascending chain. Even Noetherian
rings might have infinite Krull dimension. However, these examples live on
the fringe of the Noetherian society, and rings met in mainstream algebraic
geometry will all have finite dimension. When dim A ă 8, there are saturated
chains of maximal length, that is of length dim A. We shall also see that local
Noetherian rings have finite Krull dimension.
Example 10.1 The polynomial ring in A “ krx1 , . . . , xr , . . .s in infinite many
variables is of inifinite Krulll dimension. Each of the ideals pr “ px1 , . . . , xr q is
a prime ideal, and they form an infinite ascending chain.
There is also an infinite descending chain of prime ideals in A whose
members are the ideals qr “ pxr , xr`1 , . . .q. K
10.3 If A has several minimal prime ideals, the space Spec A will have several
irreducible components, as in the example above with p0, 1q and the x-axis be-
ing the components. The dimension of A will we the largest of the dimensions
of the components, or if there are infinitely many, the supremum. In algebraic
terms this translates as:
Proposition 10.4 If tpi u are the minimal primes of A, then dim A “ supi dim A{pi .
Proof: The intersection of the prime ideals of a chain being prime, any max-
imal chain starts at minimal prime (see also Paragraph 2.61 on page 43 and
Exercise 2.23 on page 43). o
10.5 It is common usage to call dim Ap the the height of p, or more generally Height of ideals (høyden
til idealer)
for any ideal a in A the heigh is the least height of any prime ideal containing
a; that is
ht a “ min ht p.
aĎp
A useful inequality
10.6 Any chain tpi ui of prime ideals in A may be broken in two at any stage,
say at term p “ pν . It is the concatenation of a lower chain, formed by the
members of the chain contained in p, and an upper chain, the one formed by
those containing p. And of course, one may as well splice two chains provided
one ends at the prime where the other one begins.
p0 Ă p1 Ă . . . Ă pν´1 Ă pν Ă pν`1 Ă . . . Ă pn
Notice that the proposition is still valid if one or more of the dimensions are
infinite, with the usual intepretation that n ` 8 equals 8.
In some rings there are maximal chains—that is, saturated chains which
cannot be lengthened—of different lengths, and the Krull dimension is of
course the length of the longest. For any prime ideal in a shorter chain, the
inequality as in the proposition will be strict. It is easy to give such examples
when the ring A is not a domain, one simply takes an A with components
214 ma4200—commutative algebra
of different dimension (see Example 10.4) below). However, there are also
examples when A is a Noetherian domain, but these are again rather exotic
constructs you don’t tumble over in algebraic geometry.
Example 10.4 Let A “ krX, Ys{a where a “ pXY, YpY ´ 1qq, and as usual, we y
let lower case letters denote the classes of their upper case versions. Geomet- p0,1q
rically Vpaq is the subset of k2 being the union of X-axis and the point p0, 1q.
x
The primary decomposition of p0q in A is p0q “ pyq X px, y ´ 1q. Hence pyq and
px, y ´ 1q are the minimal prime ideals in A. Now, px ´ a, yq is a maximal ideal
for any a P k, so A possesses saturated chains
pyq Ă px ´ αq,
Example 10.5 If you want an example that is a local ring consider the ideal
a “ pZq X pX, Yq “ pZX, ZYq in krX, Y, Zs and let A be the localization at
px, y, zq of the quotient krX, Y, Zs{a. Then p0q “ pzq X px, yq is the primary
decomposition of p0q and pzq and py, xq are minimal prime ideals in A; but
there maximal chains pzq Ă px, zq Ă px, y, zq and px, yq Ă px, y, zq in A. K
Going–Up again
that dim B ď dim A. On the other hand by Going–Up II, every chain in A has
a chain in B lyinig over it which means that dim B ď dim A. o
10.9 For example if Q Ď K is any field extension (finite or not) the ring of
integers in K, that is the integral closure A of Z in K, is of course integral
over Z and consequently is of dimension one. In particular this applies to the
quadratic extensions K “ Qp dq we have seen, but also to the more impressive
‘
extension K “ Q,s the field of algebraic numbers. The ring of algebraic integers
Z
s is therefore of dimension one, but recall, it is not a Noetherian ring.
krull dimension 215
Corollary 10.12 Let A be a domain finitely generated over the field k whose field
of fractions is K. Then dim A “ trdegk K
Proof: Let n “ trdegk K. By Noteher’s Normalization LemmaV there is V 9.40 on page 205
“may be the most important single theorem in the theory of Noetherian rings”.
There is however a rather simple underlying intuition that the dimension of a
solution space goes down by at most one when an additional equation is intro-
duced. We recongnize this from theory of linear equations, but the principle
has a much wider scope (as shows e. g. the Hauptidealsatz).
10.13 The scene is a ring A with the two main players, a prime ideal p and
an element x over which p is minimal; in other words, there is no prime ideal
properly lying between pxq and p. And the conclusion is that ht p is at most
one. Of course, the prime ideal p might be a minimal one and then the height
would be equal to zero, but if A is a domain and x is non-zero, the height will
be one.
Proof: We are to show that there are no chain of prime ideals of length two pxq Ď p
Ă
q1
our task is to prove that if q Ă p, then q “ 0.
The first observation is that, since p is minimal over pxq, the ring A{xA has
only one prime ideal and being Noetherian, it is Artinian, and we shall have
the opportunity to activate the descending chain condition. The chain we shall
exploit, is the descending chain tpxq ` qpnq un , where qpnq is the n-th symbolic2 2
Don’t panic! We
power of q, that is, qpnq “ A X qn Aq . The chain tpxq ` qpnq u corresponds to the would rather have used
the powers qn . But
powers of prime ideals
can be unruly, and at a
later point in the proof
they will not serve our
purpose.
krull dimension 217
descending chain t pxq ` qpnq {pxqu in A{xA and must eventually be stable, as
` ˘
This entails that if a P qpnq , one may write a “ b ` cx with b P qpn`1q , so that cx P
qpnq . This entails3 that c P qpnq , and consequently that qpnq Ď qpn`1q ` xqpnq Ď qpnq 3
This is where we
Nakayama’s lemma yields that qpnq “ qpn`1q ; this gives qn Aq “ qn`1 Aq , and use the symbolic
power; since x R q, it
appealing once more to Nakayama’s lemma, we may conclude that qAq “ 0; is invertible in Aq , and
that is, q “ 0. o c P A X qn A q
Theorem 10.15 (The Height Theorem) Let A be a Noetherian ring and let p
be a prime ideal minial over an ideal a generated by r elements. Then ht p ď r.
ait “ ci a1 ` bi p
pa1 q pd´1
with bi P pd´1 , and let b “ pb2 , . . . , br q. Then b is contained in pd´1 . We contend
that there is prime ideal q lying between b and pd´1 , properly contained in
q
pd´1 ; indeed, if pd´1 were minimal over b, the height of pd´1 would be at most pd´2
r ´ 1 by induction, but being next to the top in a chain of length d, the ideal
pd´1 is of height at least d ´ 1, and r ´ 1 ă d ´ 1. b
Now, the idea is to pass to the ring A{q. The ideal q ` pa1 q contains a power
of a, hence there is no prime ideal between q ` pa1 q and p, which means that the
p{q is minimal over the principal ideal q ` pa1 q{q, and therefore of height one
after the Principal Ideal Theorem, but there is also the chain 0 Ă pd´1 {q Ă p{q.
Contradiction. o
10.16 It ensues from the Height Theorem that any local Noetherian ring A
has a finite Krull dimension. Indeed, the maximal ideal m is finitely generated,
218 ma4200—commutative algebra
and by the Height Theorem the height of m, which is the same as dim A, is
bounded by the number of generators. Similarily, Noetherian rings enjoy a
descending chain condition for the prime ideals. Any term in a chain is finitely
generated and hence is of finite height.
10.18 Among the different numbers associated with a ring having some
flavour of a dimension, is the so-called embedding dimension of a local ring A. Embedding dimension
If m denotes maximal ideal of A, embedding dimension of A is defined as the (embeddingsdimensjon)
vector space dimension dim A{m m{m2 (the module m{m2 is killed by m and
therefore is a vector space over A{m). Any vector space basis of m{m2 is of the
form rx1 s, . . . , rxr s for members xi of m. Nakayama’s lemma implies that the
xi ’s generate m, and in its turn The Height Theorem yields that dim A ď r. We
have thus proved
Theorem 10.20 A Noetherian domain A is a ufd if and only if every prime ideal of
height one is principal.
Proof: Let A be the ring, let m the maximal ideal, and let n “ dim A. We
shall, by a recursive construction, exhibit a sequence x1 , . . . , xn of elements in
m so that the ideal ai “ px1 , . . . , xi q generated by the i first has all its minimal
primes of height i. Assume that aν has been constructed and consider the
prime ideals tp j u minimal over aν . They all have height ν so if ν ă n, none
of them equals m. Hence their union is not equal to m by Prime Avoidance
(Lemma 2.28 on page 33), and we may pick an element xν`1 from A so that
Ť
xν`1 P m, but xν`1 R i pi . Then any prime ideal minimal over aν`1 “
px1 , . . . , xν`1 q has height ν ` 1; indeed, let p be one of them. It is not among
the minimal prime ideals pi of aν , and therefore must contain one of pi ’s
properly, say p j , and we infer that ht p ą ht p j “ ν. The other inequality; that is
ht p ď ν ` 1, ensues from the Height Theorem. o
Problem 10.1 With notations as above, show that ` Ap pAp {pX ´ αYqAp q “ 2
when α is a scalar and that ` Ap pAp {pYqAp q “ 3. The excess length in the latter
case is explain by the x-axis, that is the line Y “ 0, being tangent C at the
origin in some sense. Show that ` A pA{pβX ´ αYqq “ 3 for all linear forms
βX ´ αY. M
10.25 From linear algebra we know that for a given linear map φ : V Ñ W one
has the formula
or using that dim im φ ď dim V, it yields the inequality dim V ď dim ` dim φ´1 p0q.
For a smooth map φ : X Ñ Y between manifolds there is a similar inequality
for y P φpXq, which in fact is just the inequality from linear algebra applied
to the derivative of φ. If now, φ : X Ñ Y is a map of varieties, or between
spectra of rings, there is a similar formula, however we confine ourselves to an
algebraic version valid for maps of local rings.
10.26 Recall that map of local rings is map of rings between two local rings Maps of local rings
which sends the maximal ideal into the maximal ideal. (lokale ringavbildning)
Proposition 10.27 Let A and B be the two local rings having maximal ideals m
and n respectively, and assume that φ : A Ñ B is a map of local rings. Then it holds
true that
dim B ď dim A ` dim B{mB.
Example 10.6 (Strict inequality may occur) Consider the map ψ : C2 Ñ C2 send-
ing a point pu, vq to pu, uvq. The fibre over p0, 0q is the entire line u “ 0, and
thus of dimension strictly larger than the difference between the dimensions of
the source and the target.
Transcribing this into local algebra we consider the map of rings φ : kru, vs Ñ
krx, ys that sends u ÞÑ x and v ÞÑ xy . The appropriates localizations are
A “ kru, vspu,vq with m “ pu, vqA and B “ krx, yspx,yq with n “ px, yqB. Then
mB “ px, xyq and the “fibre” B{px, yqB “ pkrx, ys{px, xyqqpx,yq “ kryspyq is of
dimension one, whereas of course, dim B ´ dim A “ 0. K
10.28
10.29 As an example connecting up with the geometric situation, assume that
A Ñ B is a mp of rings, not necessarily local, and that m is a maximal ideal in
A. Furthermore, assume that n is maximal ideal so that φ´1 n “ m. Then the
proposition yields
dim Bn ď dim Am ` dim B{mB.
Examples
that #Ii tends to infinity when i does. It can as simples as I1 “ t1u, I2 “ t2, 3u,
I3 “ t4, 5, 6u an so forth with Ii consisting of i consecutive numbers.
Let pi be the prime ideal generated by the variables x j for which the index
j belongs to Ii , moreover S will be the set of elements in krx1 , . . . , xr . . .s not
belonging to any of the prime ideals pi . Then S is multiplicatively closed and
we let A “ krx1 , x2 , . . .sS and qi “ pi A.
The main observation is that local rings Api are equal to polynomial rings
over a certain field in the variables x j with index j P Ii localized at the prime
ideal those variables generate; that is, Aqi “ Ki rx j |j P Ii spx j | jP Ij q , and where
222 ma4200—commutative algebra
Lemma 10.30 Assume that A is a ring such that all localisations Am at maximal
ideals are Noetherian and that any element i A is contained in finitely many maximal
ideals. Then A is Noetherian.
Proof: Let a be an ideal. Each ideal aAm is generated by finitely many ele-
ments, as Am is Noetherian), and they may be chosen to lie in a. Recollecting
all these generators for all the finitely many maximal ideals containing a, one
obtains a finite set of generators for a. o
Problem 10.2 Let A be a Noetherian ring and let p Ă q be two prime ideals.
Prove that if there is a prime ideal properly contained between p and q then
there are infinitely many. Hint: Assume that r1 , . . . , rr are the primes lying
Ť
properly between p and q. Then i ri is not equal to q according to the princi-
Ť
ple of Prime Avoidance. Pick an element x P pz i ri and apply the Principal
Ideal Theorem. M
Problem 10.3 Show that among the Noetherian rings only the Artinian ones
and the semi local one-dimensional ones have finitely many prime ideals. M
Lecture 11
A class of rings which are omnipresent in mathematics are the principal ideal
domains. As the name indicates, they are integral domains all whose ideals Principal ideal domains
are principal; that is, every ideal is generated by a single element. They are (Hovedidealområder)
simple kind of rings with a lot of nice properties. As we shortly shall see they
are automatically Noetherian and unique factorization domains.
The most prominent members of this club are the ring of integers Z and the
ring of polynomials krts over a field k. Modules over Z are just abelian groups,
incontestably met everywhere in mathematics. Modules over krts appear, if
possible, even more frequently but often disguished. Such a module is just a
vector space V together with a k-linear the action of t; that is, together with a
linear operator.
The of club of principal ideal domains is a subclub of the even more fab-
shionalble club of Dedekind domains, and these are as frequently met as the
pid. The ring of integers in an algebraic function field is a Dedekind domain,
so they are everywhere in algebraiv number theory. An in algebraic geomtery
they are the coordianet sings of non-singualr affine curves. So teh outspring of
both number theory and algebaric geoemtry is found in that club.
They are integrally closed domains of Krull dimesion one: So the ring of
regular affine curves!
The principal ideal domains live in a family of four. The closest relativ be-
ing the big brothers, rthe Bezout domains. They are not necessarly Noetehrein,
but have proerty that any fintely generated ideal is principal.
The two other family members are the so called Dede kind rings and the
Prufer rings. They There is a genaralization of pid’s called Bezout domains.
These rings are not requiered to be Noetherian, but they have the property
that any finitely generated ideal is principal. A theorem in the theory of
functions states that the ring of holomorphic functions i a domain Ω of C are
Bezout domians which makes the Bezout rings important. The Bezout rings
share several nice properties with the pid’s and we shall prove some whose
proof does not have addirional difficulties.
224 ma4200—commutative algebra
The finitely generated modules over a pid are one of the very few classes of
modules which are completely classified. The well-known “Main theorem for
finitely generated abelian groups” states that such a group M decomposes as a
direct sum of cyclic groups; that is, one has
M » Zν ‘ Z{pi i Z,
à ν
where ν is non-negative integer, the νi ’s are positive integers and the pi are
prime numbers; of course the sum is finite. Over any pid finitely generated
modules decomposes in total analogy to this; it holds true that
à ν
M » Aν ‘ A{pi i A (11.1)
i
associate is equivalent to generating the same principal ideal. For integers this
principal ideal domains 225
is just the good old ˘-thing, and polynomials over a field are associate when
they differ by a constant factor; but of course, units in most rings are not that
simple to describe.
Being associate is evidently an equivalence relation, and the set of principal
ideals in A incarnates the set of equivalence classes. Two elements a|b if and
only if pbq Ď paq; so the divisibility lattice mod assocition equals the opposit of
the lattice of principal ideals.
11.2 Recall that a greates common divisor of a finite collection a1 , . . . , ar of Greates common divisor
elements from a ring A is an element d so that d|ai for all i, and if b is another (Største felles divisor)
ring element dividing all the ai ’s, then b|d. If a greatest common divisor exists,
it is unique up to multiplication by a unit.
Greatest common divisors do not necessarily exist in general, but in a
Bezout domain any finite set of elements has one. The ideal pa1 , . . . , ar q is in
that case principal, and any generator serves as a greatest common divisor of
the ai ’s. Indeed, if d is generator, it holds that d|ai since ai P pdq, and to see that
ř ř
b divides d when b|ai for all i, just write d “ i ci ai . Then d “ p i ci di qb where
the di ’s are elements with ai “ di b.
Ş
In a similar vein, a generator d of the intersection i pai q is a least common
multiple of the ai ’s. It is clear that ai |d. Moreover, if ai |b for all indices i, the Least common multiple
Ş
element b lies in the intersection i pai q, and is therefore divisible by d. In a (Minste felles multiplum)
Lemma 11.3 Let a and b be elements from a domain A. If pa, bq is principal, then
the intersection paq X pbq is principal.
Proof: Assume that pa, bq “ pdq and let x and y be ring elements such that
a “ xd and b “ yd. Moreover there are ring elements u and v such that
ux ` vy “ 1. An element s from the intersection paq X pbq is shaped like
s “ za “ wb, and then zx “ wy. Multiplying the relation ux ` vy “ 1 by z,
one obtains z “ zux ` zvy “ ypwu ` zyq, or z “ γy with γ “ wu ` zy. Hence
za “ γya and ya (which equals xb) generates the intersection. o
The end of this story is that we have established the following:
Proposition 11.4 In a Bezout domain every finite set of elements has as greatest
common divisor and a least common multiple.
11.5 As signalled in the introduction, the principal ideal domains are the
Noetherian members of the Bezout club.
Example 11.2 The simplest example of irreducibles not being prime is found
?
in the ring Zr ´5s. There the relation
? ?
2 ¨ 3 “ p1 ` i 5q ¨ p1 ´ i 5q
Proof: We consider the set Σ of “bad guys”: The set of principal ideals pxq so
that x is a counterexample to the assertion; or in other words, so that x is not
a product of finitely many irreducibles. If non-empty, the set Σ has a maximal
member, say paq. Then a is not irreducible, and we may write a “ bc with
neither factor being a unit. This means that paq Ĺ pbq and paq Ĺ pcq, and by
consequence, neither pbq nor pcq belongs to Σ as paq was maximal there. Hence
both b and c can be expressed as finite products of irreducibles, and the same
is true for a. Contradiction. o
any domain Ω Ď C there are functions with infinitely many zeros which can
not be products of finitely many irreducibles. In fact, a famous theorem of
Weierstrass’ asserts that for any sequence tan u of distinct points in Ω not accu-
mulating in Ω, there is a function vanishing at the an ’s and nowhere else; one
may even prescribe the order of vanishing at each an . K
Problems
11.1 Referring to Example 11.2 show that the three other involved numbers 3,
? ?
1 ` i 5 and 1 ´ i 5 are irreducible.
11.2 This exercise gives a general version of Example 11.1. Assume that
À
A “ iě0 Ai is a graded integral domain with A0 being a field. Assume
there are four homogeneous elements of degree one x, y, z and w that satisfy
xy “ zw. Show that x is irreducible but not prime. Hint: Assume that
there is a factorization x “ pq. Work with the non-vanishing homogeneous
components of p and q of highest degree.
11.3 Let A “ Zrx, y, z, ws{pxy ´ wzq show that the class of x is irreducible but
not prime.
11.4 (Euclidean domains.) A domain A is said to be Euclidean if there is func- Euclidean domains
tion δ on A assuming values in the set N0 of non-negative integers that has (Euklidske områder)
o For any pair x and y with y ‰ 0 there are elements q and r in A so that
x “ yq ` r and either δprq ă δpyq.
11.5 Show that the conclusion that A is a pid in the previous exercise persist
when in the definiton of δ the set N0 is replaced with any well-ordered set W .
11.6 Let n P Z and put δpnq to be the number binary digit of }n}. Show that
δ is the smallest Eucliden function on Z. The article1 is a nice paper about 1
Eucliden algorithms.
11.9 Let r ą 0 be a real number. Show that set of power series that converge
for }z} ą r is an Euclidean ring. Hint: The number of zeros in the disk
t z | }z} ď r u is a Euclidean function.
M
principal ideal domains 229
The torsion module depends functorially on M, and it sits in the short exact
sequence
0 / TpMq /M / M{TpMq / 0.
pν x “ 0 for every x P M. The actual power need for the killing depends on x,
and need not be bounded. For intstance when p is a prime number, the group
Z p8 has elements of every odre a power of p. The annihilator of such modules
is zero eventhough the annihilator over single element has ppqA as radical.
1 “ c1 q1 ` . . . ` cr qr .
x “ c1 q1 x ` . . . ` cr qr x.
vj νj
Now a “ q j p j , so that each summand c j q j x is annihilated by p j and therefore
belongs to Tp j . o
Example 11.4 The abelian group Q{Z is a typical infinite torsion group, and
the p-torsion part equals the group Z p8 “ Zrp´1 s{Z. Indeed, that a rational
number x satisfies pν x P Z, means that x “ a{pv for some a P Z. It follows
230 ma4200—commutative algebra
that there is a decomposition Q{Z “ p Z p8 where the sum extends over all
À
primes.
Verbatim, this reasoning works for any principal ideal domain A, and one
has a decomposition
à
K{A “ A p8
p
where K is the fraction field of A and A p8 “ Arp´1 s, and where the sum is
taken over all prime ideals ppq. K
0 / Am / An /M / 0,
Am / An /M / 0.
In other words it is finitely generated and one may chose the generators in
way that the relations between them are finitely generated as well. Over a
Noetherian ring every submodule of An is finitely generated, hence being
finitely generated is equivalent to being finitely presented.
Lemma 11.16 Assume that M is a finitely generated torsion free module over the
domain A. Then there is non-zero A-linear map φ : M Ñ A.
Theorem 11.17 Assume that A is a Bezout domain. Then every finitely generated
submodule of a finitely generated torsion free A-module E is free. In particular, every
finitely generated torsion free A-module is free.
Torsion free modules over a pid which are not finitely generated is a devilish
class, and even for abelian groups there are large white areas on the map.
For instance, torsion free subgroup contained in Q2 seem to be impossible to
classify. However those contained in Q are rather easy to lay hands on (see
exercise xxx).
The case of principal ideal domains merits is own announcement. The
finitely general case i cover by the theorem above, but it holds without hy-
potheses on the projective modules—in fact almost Noetherian domains
projective modules that are not finitely generated are free.
Corollary 11.18 Assume that A is a pid. Then every projective module is free and
every submodule of a free module is free.
11.19 Just being slightly more careful in the proof the lemma can be general-
ized to
Lemma 11.20 Let A be Prufer domain, then every finitely generated torsion free
A-module is isomorphic to a direct sum of projective modules of rank one.
Recall that a Prufer domain is an integral domain with all finitely generated
ideals being projective. Dedekind domains are of this sort, and of course
principal ideal domains as well, their ideals are even free modules. In the
dictionary between algebra and geometry, projective modules correspond to
vector bundles. Hence on the affine curve X “ Spec A, any vector bundle
decomposes as the sum of line bundles. This is far from being the case on
projective curves, the projective line being the sole exception.
Proof: Induction on the rank will do. The rank one case being the hypothesis
that A be Prufer. Once we have non-zero map
EÑA
the kernel decomposes as a direct sum of rank one projectives, and since E is
finitely generated the image is a projective ideal. Hence E » I ‘ ker π, and
ker π is finitely generated projective. The ranks has dropped by one, and the
induction hypothesis applies to ker π.
So have to infer that E˚ has a non zero element, but this is just xxxx, any
element E that maps to Eb A K is torsion. o
Lemma 11.22 Assume that A is a Bezout domain. Let E be a free A-module of finite
rank and let e P E be an element. Assume there is an A-linear map π : E Ñ A with
πpeq “ 1. Then e is part of a basis for E.
Proof: The kernel of π is free by Proposition 11.17, so pick a basis tei u1ďin´1
for it. We contend that joining e to that basis we obtain a basis for E. Applying
ř
π to a linear dependence relation i ai ei ` ae “ 0 immediately gives a “ 0, so e
together with the ei ’s constitute a linearly independent set. They also generate
E since for any x P E the element x ´ πpxqe lies in ker π, and hence is a linear
combination of the ei ’s. o
Proposition 11.23 Let E and F be free A-modules of the same rank. Let φ : F Ñ
E be an injective linear map. Then there are bases t f i u for F and tei u for E and ring
elements λi so that φp f i q “ λi ei .
In the case A is a field and E and F are vector spaces of the same dimension,
the proposition just says that the image of basis under an injection is a basis;
in fact, in that case, the matrix may be taken to be the identity matrix. The
result is far less subtle than any existence theporem for eigenvalues.
Proof: Pick an element f P F that is part of a basis t fˆi u for F, and chose a
basis têi u for E. These two bases are auxiliary and not the ones we shall end
up with—hence the hat. Anyhow, φp f q may be expressed in terms of the êi ’s;
ř
that is, it can be written as a linear combination φp f q “ j ai êi with the ai ’s
from A. The ideal pa1 , . . . , an q generated by the ai ’s is principal since A is a
Bezout domain, and let us say it is generated by the element d.
/ ker π
φ|ker ρ
o The case d “ 1. ker ρ
ř
This means that there are ring elements ci so that i ci ai “ 1. Define an A-
φ
/E
ř ř
linear map π : E Ñ A by sending an element i xi êi to i ci xi . Put ρ “ π ˝ φ; F
then by the choice of the ci ’s, it holds true that ρp f q “ 1, and the maps ρ and ρ π
π are both split surjections. Their kernels are therefore free (Proposition 11.17)
A
id A
/A
and ranks have dropped by one. By induction the two kernels have bases
t f i u1ďiďn´1 and tei u1ďiďn´1 respectively so that
φp f i q “ λi ei (11.2)
The idea is to factor φ as a map falling under the first case and one which
apriori has diagonal martxi. Assume then that d ‰ 1 then there are ring
234 ma4200—commutative algebra
case, and there are bases tei u and t f i u like in the first case. Now define a map
σ : F Ñ F by putting σp f q “ d f and σp f i q “ f i for the rest of the basis f i . Then
it holds true that φ “ ψ ¨ σ and we are trhough. o
Theorem 11.24 Let E and F be two free A-module of finite rank, and φ : F Ñ F an
A-linear map. The there are bases for F and E so that the matrix is semi-diagonal.
Semi-diagonal means that the only non-zero elements of the matrix are situ-
ated on the diagonal that emanates from the upper left corner. Of course when
matrix is squar, this is just being diagonal.
Proof: The image of φ is a free module after Ppropostion ?? as is the kernel,
and hence F »P φ ‘ ker φ. The map φ factors trhroug an injective φ̃ : im φ Ñ E,
and by Proposition xxx there are bases for P φ and E relative to which φ̃ has a
diagonal matrix. Lift these basis elements to F and complement them with a
basis for ker φ to obtain a basis for F of the right sort. o
à ν
M » Aν ‘ A{pi i A
i
0 0 0
φ1
0 / E1 / F1 / M2 /0
/E /F /M /0
φ
0
ρ
π0
0 /A /A / M1 /0
a
0 0 0
By lemma xxx on page xxx we infer that the middle column is split so that
E » F1 ‘ A, and it follows that F1 is projective. By lemma xxx from this ensues
that F1 is free, and obviously, its rank is one less than F. Hence the induction
hypothesis applies, and we can conclude that E1 is free. Now, the leftmost
column is also split and E “ F1 ‘ A. Thus E is free and the first assertion is
established.
Then we attack the assertion about the bases, and we denote the rank of E
by M and that of F by n. By induction, there are bases e1 , . . . , em´1 for E1 and
f 1 , . . . , f n´1 for F1 such that φ|E1 sends ei to ai f i with ai P A We extend these
bases to bases for E and F by respectively adding an element em of E such that
ρpem q “ 1 and and element f n of F with π0 p f n q “ 1.
The crucial observation is that for any projection π : F Ñ A, it holds true
that πpρpem qq P paq. Indeed, assume not, and let b “ πpρpem qq. Then pa, bq “ 1,
and one may write αa ` βb “ 1 for appropriate elements α and β in A. Then
απ0 ` βπ sends em to 1, contradicting the maximality of paq.
Applying this observation to the projections onto the different basis ele-
ř
ments f i , leads us to an expression φpem q “ a f m ` i ci a f i for ci ’s in A, and
ř
simply replacing f m by f m ` i ci f i , we have the requested basis element. o
In fact observing that if E “ A ‘ M1 and M is finitely generated, if follows
that M1 is finitely generated. This follows for instance, by decomposing id M “
η1 ` η2 into a sum of two oprthogonal idempotents, such that η2 pMq “ M1 .
Å lese: Lam, T. Y. (2006), Serre’s Problem on Projective Modules, Springer Monographs in
Mathematics, Berlin, Heidelber
Ât’