Field Theory
Field Theory
Pete L. Clark
Thanks to Asvin Gothandaraman and David Krumm for pointing out errors in
these notes.
Contents
The purpose of these notes is to give a treatment of the theory of fields. Some
aspects of field theory are popular in algebra courses at the undergraduate or grad-
uate levels, especially the theory of finite degree field extensions and Galois theory.
However, a student of algebra (and many other branches of mathematics which use
algebra in a nontrivial way, e.g. algebraic topology or complex manifold theory)
inevitably finds that there is more to field theory than one learns in one’s stan-
dard “survey” algebra courses.1 When teaching graduate courses in algebra and
arithmetic/algebraic geometry, I often find myself “reminding” students of field-
theoretic facts that they have not seen before, or at any rate not in the form I wish
to use them.
I also wish to fill in some gaps in my own knowledge. Especially, I have long
wished to gain a deeper understanding of positive characteristic algebraic geometry,
and it has become clear that the place to begin understanding the “pathologies”2
of algebraic geometry in characteristic p is the study of finitely generated field ex-
tensions in positive characteristic.
These notes are meant to be comprehensible to students who have taken a basic
graduate course in algebra. In theory one could get away with less – the exposi-
tion is mostly self-contained. As algebraic prerequisites we require a good working
knowledge of linear algebra, including tensor products. The reader should also be
comfortable with – and fond of – groups and rings. Such a benevolent familiarity
is used much more than any specific results of group or ring theory. Our approach
is sufficiently abstract and streamlined that it is probably inappropriate for most
undergraduates. In particular, more often than not our approach proceeds from
the general to the specific, and we make no apologies for this.
7
Some Conventions
By convention, all of our rings are associative and have a multiplicative unity,
called 1. Again by convention, a homomorphism of rings necessarily carries 1 to 1.
These notes contain many exercises, including some which ask for proofs of stated
results. In general I am not at all opposed to the idea of a text giving complete
details for all of its arguments.3 However, it is in the nature of this particular
subject that there are many more results than proof techniques, to the extent that
giving complete proofs of all results would create a lengthy repetitiveness that may
discourage the reader to read the proofs that we do give.
As a rule, exercises that ask for proofs of stated results are meant to require no
new ideas beyond what was (even recently) exposed in the text. A reader who feels
otherwise should contact me: there may be an unintended gap in the exposition.
On the other hand, if exercises are given at all, it certainly spruces things up to
have some more challenging and interesting exercises. I have also not hesitated to
give exercises which can in principle be solved using the material up to that point
but become much easier after later techniques are learned.
At some point I fell victim to the disease of not liking the look of a paragraph
in which only a few words appear on the last line. Because of this, in the exercises I
have sometimes omitted the words “Show that”. I hope the meaning remains clear.
3In fact I agree with Robert Ash that the prevailing negative reputation of such texts is
undeserved: the royal road to a particular destination may or may not exist, but it seems perverse
to claim that it ought not to exist.
9
CHAPTER 1
Introduction to Fields
The two branches of mathematics in which general fields play a principal role are
field theory (of course) and linear algebra. Most of linear algebra could be devel-
oped over a general division algebra rather than over a general field. In fact for the
most part the theory is so similar that it is not really necessary to consider division
algebras from the outset: one can just check, if necessary, that a certain result which
is true for vectors spaces over a field is also true for left modules over a division
algebra. On the other hand, when one studies things like roots of polynomials and
lattices of finite degree extensions, one immediately finds that non-commutative
division algebras behave in quite different and apparently more complicated ways.
Example 1.1. There are exactly two complex numbers z such that z 2 = −1:
z = i and z = −i. In general, any nonzero polynomial P (t) with coefficients in
a field can have no more solutions than its degree. But in Hamilton’s quaternion
algebra H there are clearly at least three solutions: i2 = j 2 = k 2 = −1, and in
fact there are uncountably many: a quaternion squares to −1 iff it is of the form
xi + yj + zk with x2 + y 2 + z 2 = 1.
Example 1.2. Let K/Q be a quartic field (i.e., a field extension of Q which has
dimension 4 as a Q-vector space). Then there are at most three intermediate sub-
fields Q ( F ( K. (More precisely there is either zero, one or three such fields, and
the first case happens “most of the time.”) However, any noncommutative division
algebra B/Q of degree 4 as a Q-vector space has infinitely many nonisomorphic
quadratic subfields.
11
12 1. INTRODUCTION TO FIELDS
The study of division algebras is closely related to field theory – via Brauer groups
and Galois cohomology – so that one can put one’s understanding of a field F
and its finite degree extensions to excellent use in studying noncommutative divi-
sion algebras over F . In fact, notwithstanding the above two examples, the finite
dimensional, central division algebra over a field F are significantly easier to under-
stand than finite dimensional extension fields of F : e.g. we understand quaternion
algebras over Q far better than quartic number fields.
CHAPTER 2
2. Fields of Fractions
If R is a domain, then one can define a field F whose elements are viewed as
fractions ab with a, b ∈ R, b 6= 0. Formally speaking one considers ordered pairs
(a, b) ∈ R2 , b 6= 0 and introduces the equivalence relation (a, b) ∼ (c, d) ⇐⇒
ad = bc, i.e., exactly the same construction that one uses to define the rational
numbers in terms of the integers. The field F is called the field of fractions, (or,
sometimes, “quotient field”) of the domain R.
Exercise 2.5. (Functoriality of the field of fractions) Let ϕ : R → S be an
injective homomorphism of domains. Show that ϕ extends uniquely to a homomor-
phism from the fraction field F (R) of R to the fraction field F (S) of S.
Exercise 2.6. (Universal property of the field of fractions) Let R be a domain
with fraction field F and let K be a field. For any injective homomorphism ϕ : R →
K, there exists a unique extension to a homomorphism F → K.
Exercise 2.7. Let R be a domain with field of fractions F (R). Show: #R =
#F (R).
2. FIELDS OF FRACTIONS 15
Thus any method which produces a supply of domains will also produce a supply of
fields (of course distinct domains may have the same fraction field, a rather trivial
example being Z and Q itself; there are in fact uncountably many isomorphism
classes of domains with fraction field Q).
Proposition 2.6. If R is a domain, then the univariate polynomial ring R[t]
is also a domain. Moreover, if F is the fraction field of R, then the fraction field
of R[t] is F (t), the field of all quotients of polynomials with F -coefficients.
Exercise 2.8. Prove Proposition 2.6.
Example 2.7. Applying the Proposition with R = F a field, we get a field F (t)
of rational functions in F . E.g., the field C(t) is the field of meromorphic functions
on the Riemann sphere (see the next section). Moreover, for any field F , F [t] is a
domain, so F [t1 , t2 ] := F [t1 ][t2 ] is also a domain. The fraction field is easily seen
to be F (t1 , t2 ), i.e., the fraction field of F [t1 , . . . , tn ] is F (t1 , . . . , tn ) the field of
rational functions in n indeterminates.
Although successive applications of Proposition 2.6 will yield polynomial rings in
only finitely many indeterminates, nothing stops us from considering larger poly-
nomial rings: let T = {ti } be any set of indeterminates, and R any commutative
ring. One can consider the polynomial ring R[T], defined as the union (or, if you
like, direct limit) of polynomial rings R[S] where S ⊂ T is a finite subset. In other
words, we consider the ring of polynomials in an arbitrary infinite set S of inde-
terminates, but any given polynomial involves only finitely many indeterminates.
One can again show that if R is a domain, so is R[T]. The corresponding fraction
field R(T) is the field of all quotients of polynomials in all these indeterminates.
Exercise 2.9. Let F be a field and T a nonempty set of indeterminates. Show
that the cardinality of the rational function field F (T) is max(ℵ0 , #F, #T).
Another way of manufacturing domains is to start with a commutative ring R and
take the quotient by a prime ideal p. Then we can get a field by (if necessary, i.e.,
if p is not maximal) taking the field of fractions of R/p. For example with R = Z
we get the finite fields Fp .
Example 2.8. Let R = F [T ] and p a nonzero prime ideal. Then, since R is
a PID, p = (f (t)), where f (t) is an irreducible polynomial. Moreover, assuming
f (t) 6= 0, p is maximal, so without having to take quotients we get a field
K = F [t]/(f (t)),
whose dimension as an F -algebra is the degree of f .
A domain R is finitely generated (over Z) if there exist n ∈ Z+ and elements
α1 , . . . , αr ∈ R such that the least subring of R containing all the αi ’s is R itself.
Another way of saying this is that the natural map
Z[T1 , . . . , Tn ] → R, Ti 7→ αi
is surjective. In other words, a domain is finitely generated iff it is, for some n, the
quotient of the ring Z[T1 , . . . , Tn ] by some prime ideal p.
Proposition 2.9. For a field F , the following are equivalent:
a) There are α1 , . . . , αn ∈ F so that the only subfield of F containing all the
αi ’s is F itself.
16 2. SOME EXAMPLES OF FIELDS
b) The field F is the fraction field of Z[x1 , . . . , xn ]/p for some prime ideal p.
Exercise 2.10. Prove Proposition 2.9.
A field satisfying the equivalent conditions of Proposition 2.9 is said to be finitely
generated. Applying part b) and Exercise 2.7 we see that any finitely generated
field is finite or countably infinite. In particular the fields R, C are not finitely
generated. Conversely, a countable field need not be finitely generated: if T is
a countably infinite set of indeterminates, then by Exercise 2.9 the field Q(T) is
countable. Moreover it is both plausible and true that Q(T) is not finitely gener-
ated, but we lack the tools to prove this at the moment: we will return to this later
on in the context of the concept of transcendence degree.
So for instance one can speak of fields which are finitely generated over the complex
numbers C, and such fields are especially important in algebraic geometry.
Proposition 2.11. Let F be a field.
a) If F has characteristic 0, there is a unique homomorphism ι : Q → F .
b) If F has characteristic p, there is a unique homomorphism ι : Fp → F .
Proof. For any ring R, there exists a unique ring homomorphism ι : Z → R,
which takes the integer n to n times the multiplicative identity in R. For R = F a
field, the map ι is an injection iff F has characteristic 0. So if F has characteristic 0,
ι is injective, and by Exercise 2.5 it extends uniquely to a homomorphism ι : Q → F .
Any homomorphism from Q to F must restrict to the canonical injection on Z and
therefore be ι. If F has characteristic p > 0, then ι factors through to give a map
ι : Fp → F . The uniqueness of ι can be seen in any number of ways: we leave it to
the reader to find an explanation that she finds simple and convincing.
It follows that Q (resp. Fp ) is the unique minimal subfield of any field F of charac-
teristic 0 (resp. p > 0). We refer to Q (resp. Fp ) as the prime subfield of F . Note
that since there are no nontrivial automorphisms of either Q or Fp (this follows by
applying the proposition with F = Q or F = Fp ), the prime subfield sits inside F
in an especially canonical way.
Exercise 2.12. Let K be a field and k its prime subfield. Show that the concepts
“K is finitely generated” and “K is finitely generated over k” coincide.
3. FIELDS OF FUNCTIONS 17
Exercise 2.13. For any field F , there exists a set of indeterminates T and a
prime ideal p of Z[T] such that F is isomorphic to the fraction field of Z[T]/p.
If F is infinitely generated (i.e., not finitely generated over its prime subfield) then
the set T in Exercise 2.13 will of course have to be infinite. In such a case this
“presentation” of F is not, in truth, so useful: e.g., with certain limited exceptions
(to be discussed!) this is not a very insightful way of viewing the complex field C.
Exercise 2.14. Let R be a commutative ring, ι : R → S an injective ring
homomorphism, and α ∈ S. Show that there is a unique minimal subring of S
containing R and α, namely the set of all polynomials P (α), P ∈ R[t]. This subring
is accordingly denoted R[α].
3. Fields of Functions
Let U be a domain – i.e., a nonempty connected open subset – of the complex plane.
In complex analysis one studies the set Hol(U ) of all functions holomorphic (a.k.a.
analytic) on all of U and also the larger set Mer(U ) of all meromorphic functions
on U , i.e., functions which are holomorphic on the complement of a discrete set
X = {xi } and such that for each xi there exists a positive integer ni such that
z ni f is holomorphic at xi . Under the usual pointwise addition and multiplication
of functions, Hol(U ) is a ring (a subring of the ring of all continuous C-valued
functions on U ). Similarly, one can view Mer(U ) as a ring in a natural way.
Theorem 2.12. Let U be a domain in the complex plane. Then:
a) The ring Hol(U ) is a domain.
b) The ring Mer(U ) is a field.
c) The field Mer(U ) is the field of fractions of Hol(U ).
Proof. a) A consequence of the principle of analytic continuation is that the
zero set of a not-identically-zero holomorphic function is discrete in U . For 0 6=
f, g ∈ Hol(U ), the zero set of f g is the union of the zero sets of f and g so is again
discrete and thus certainly a proper subset of U .
b) Because 0 6= f ∈ Hol(U ) has a discrete zero set {xi } and for each xi , there exists
a positive integer ni such that zfni extends to a continuous nonzero function at xi ,
it follows that f1i is meromorphic.
c) This lies deeper: it is a consequence of Weierstrass’ factorization theory, in
particular of the fact that for any discrete subset X = {xi } of U and any sequence
of positive integers {ni } there exists a holomorphic function on U with zero set X
and order of vanishing ni at xi .
Exercise 2.15. Show: the field Mer(C) is not finitely generated over C.
More generally, if M is a connected complex manifold, there is a ring Hol(M ) of
“global” holomorphic functions on M and a field Mer(M ) of meromorphic functions.
It need not be the case that Mer(M ) is the fraction field of Hol(M ).
Example 2.13. Take M = C ∪ {∞} to be the Riemann sphere. Then the only
holomorphic functions on M are the constant functions, whereas Mer(M ) = C(z),
the rational functions in z.
In various branches of geometry one meets many such “fields of functions”: a very
general example, for the highly trained reader, is that if X is an integral (reduced
18 2. SOME EXAMPLES OF FIELDS
and irreducible) scheme, then the ring of all functions regular at the generic point
η is a field. If X itself is a scheme over a field k, then this field is written k(X) and
called the field of rational functions on X. For example, the field of rational
functions on the complex projective line P1 /C is the rational function field C(t).
This is essentially the same example as the Riemann sphere above, but couched in
more algebraic language.
In general, one must restrict to functions of a rather special kind in order to get a
field of functions. Using the ideas of the previous subsection, it seems fruitful to
first consider rings R of functions on a topological space X. Then we want R to be
a domain in order to speak of fraction field F (R) of “meromorphic functions” on X.
Suppose X is a topological space and consider the ring R = R(X, C) of all continu-
ous functions f : X → C. A moment’s thought indicates that for the “reasonable”
topological spaces one considers in geometry, R will not be a domain. The question
comes down to: do there exist functions f1 , f2 : X → C neither of which is zero on
all of X but such that the product f1 · f2 is identically zero?
Here are some easy observations. First, if X is not connected, the answer is cer-
tainly yes: write X = Y1 ∪ Y2 where Yi are disjoint open sets. Take f1 to be the
characteristic function of Y1 an f2 = 1 − f1 to be the characteristic function of Y2 .
On the other hand, suppose the space X is irreducible: that is, if Y1 , Y2 are two
proper closed subsets of X then Y1 ∪ Y2 6= X. Then, applying this to Yi = fi−1 (0),
we get that the zero set of f1 f2 is Y1 ∪ Y2 6= X, so R(X, C) is a domain, and one
can take its fraction field, which consists of functions which are defined on some
dense (equivalently, nonempty!) open subset of X. If you have never studied al-
gebraic geometry, you will doubtless be thinking, “What kind of crazy topological
space would be irreducible?” However, the Zariski topology on a smooth, connected
algebraic variety over (say) the complex field C is irreducible.
4. Completion
None of the constructions of fields we have discussed so far give rise to either R or
C in a reasonable way. These fields are uncountable, and from a purely algebraic
perspective their structure is quite complicated. The right way to think about them
is via a mixture of algebra and topology, e.g. one thinks of R as the completion of
the field of rational numbers with respect to the standard absolute value.
We denote C/c by K̂ and call it the completion of K with respect to || ||. There
exists a natural embedding K ,→ K̂ – namely we map each element of K to the
corresponding constant sequence – and a natural extension of the norm on K to
a norm on K̂, namely ||x• || = limn→∞ ||xn ||, with respect to which ι : K ,→ K̂ is
an isometry of normed spaces in which the image of K in K̂ is dense. For more
details, the reader is invited to consult [Cl-NTII, Chapter 2].
Example 2.14. The completion of Q with the standard Archimedean absolute
value || pq || = | pq | is the real field R.
Remark 2.1. It is sometimes suggested that there is a circularity in this con-
struction, in that the definition of completion refers to a metric and the definition
of a metric refers to the real numbers.1 But one should not worry about this. On the
one hand, from our present point of view we can consider the reals as being already
constructed and then it is a true, non-tautologous statement that the metric com-
pletion of the rationals is the reals. But moreover, a careful look at the construction
in terms of equivalence classes of Cauchy sequences shows that one absolutely can
construct the real numbers in this way, just by being careful to avoid referring to
the real numbers in the course of the completion process. In other words, the real
numbers can be defined as the quotient of the ring of Cauchy sequences of rational
numbers (where the definition of Cauchy sequence uses only the metric as defined
on rational numbers) by the maximal ideal of sequences converging to zero. After
one constructs the real numbers in this way, one notes that the Q-valued metric on
Q extends to an R-valued metric on R: no problem.
1In particular, Bourbaki’s General Topology refrains from making any reference to real num-
bers or metric spaces for many hundreds of pages until the reals can be rigorously constructed.
20 2. SOME EXAMPLES OF FIELDS
Example 2.15. If k is any field, then defining ||0|| = 0 and ||x|| = 1 for all
x 6= 0 gives an absolute value on k. The induced metric is the discrete metric
and therefore k is, in a trivial way, complete and locally compact. This absolute
value (and any other absolute value inducing the discrete topology) is called triv-
ial; such absolute values are usually either explicitly or implicitly excluded from
consideration.
Example 2.16. || ab || = pordp (b)−ordp (a) , where for an integer a, ordp (a) denotes
the largest power of p dividing a. (To get the degenerate cases to work out correctly,
we set ordp (0) = ∞ and p−∞ = 0.) The induced metric on Q is called the p-adic
metric: in this metric, a number is close to zero if, after cancelling common factors,
its numerator is divisible by a high power of p. Since the induced topology has no
isolated points, the completeness of the metric would contradict the Baire category
theorem, hence the completion is an uncountable field, called Qp , the field of p-adic
numbers.
Example 2.17. Let k be any field and K = k(t). Any element r(t) ∈ K can be
P (t)
written as ta Q(t) where P (0)Q(0) 6= 0 for a uniquely determined integer a. Define
||r(t)||∞ := e−a . (There is no particular reason to use the number e = 2.718 . . .;
any real number greater than 1 would serve as well.)
Exercise 2.16. Show: || ||∞ gives an absolute value on K(t).
An element r(t) ∈ K(t) is close to 0 iff it is divisible by a high power of t.
Exercise 2.17. Show: the completion of K(t) with respect to || ||∞ is isomor-
phic
P∞ to thenLaurent series field K((t)), whose elements are formal power series
n=n0 an t with n0 ∈ Z, an ∈ f . (Hint: It is enough to show that the norm || ||∞
extends to all of K((t)) and that K(t) is dense in K((t)) in the induced topology.)
Exercise 2.18. Show: the fields Qp are locally compact in their natural topol-
ogy. Show: K((t)) is locally compact iff K is finite.
Remark 2.2. If k is a field complete with respect to an absolute value | | and
V is a finite-dimensional vector space over k, then viewing V ∼ = k dim V gives V the
canonical structure of a topological space – i.e., we can endow it with the product
topology, and this topology is independent of the choice of basis. In particular, if k
is locally compact, so is V . Moreover it has the canonical structure of a uniform
space, and if k is complete then so is V . In particular, if k ,→ l is a field embedding
such that l is finite-dimensional as a k-vector space, then l is a complete uniform
space and is locally compact iff k is. This implies that any finite degree extension
of the fields R, Qp or Fp ((t)) have a canonical locally compact topology.
Theorem 2.18. (Classification of locally compact valued fields) Let || || be a
nontrivial valuation on a field K. The following are equivalent:
(i) The metric topology on K is locally compact.
(ii) Either (K, || ||) = R or C; or the induced metric is complete and non-
Archimedean and the residue field is finite.
(iii) K is a finite degree extension of R, of Qp or of Fp ((t)).
Proof. See [Cl-NTII, Theorem 5.1].
There are more elaborate ways to construct complete fields. For instance, suppose
R is a domain and p is a prime ideal of R. Then in commutative algebra one learns
4. COMPLETION 21
Such fields arise in algebraic and analytic geometry: C((x1 , . . . , xn )) is the field
of germs of meromorphic functions at a nonsingular point P on an n-dimensional
analytic or algebraic variety.
Exercise 2.19. Show: the field k((x1 , x2 )) is properly contained in k((x1 ))((x2 )).
CHAPTER 3
Field Extensions
1. Introduction
Let K be a field. If ι : K → L is a homomorphism of fields, one says that L is an
extension field of K. As a matter of psychology, it often seems more convenient
to think of L as “lying above K” rather than K as embedding into L. We often
write L/K instead of ι : K → L, notwithstanding the fact that the latter notation
hides important information, namely the map ι.1
As an immediate application we can rederive the fact that the order of a finite
field is necessarily a prime power. Namely, let F be a finite field, and let Fp be
its prime subfield. Since F is finite, it is certainly finite-dimensional over Fp (any
infinite dimensional vector space over any field is infinite), say of dimension d. Then
F as an Fp -vector space is isomorphic to Fdp , so its cardinality is pd .
Theorem 3.1. (Degree multiplicativity in towers) Let F ⊂ K ⊂ M be field
extensions. Then we have
[M : F ] = [M : K][K : F ].
23
24 3. FIELD EXTENSIONS
Proof. Let {bi }i∈I be an F -basis for K and {aj }j∈J be a K-basis for M . We
claim that {ai bj }(i,j)∈I×J is an F -basis for M . This suffices, since then [K : F ] =
#I, [M : K] = #J, [M : F ] = #(I × J) = #I × #J.
Let c ∈ M .PThen there exist αj ∈ K, all but finitely many of which are zero,
such that c = j∈J αj aj . Similarly, for each j ∈ J, there exist βij ∈ F , all but
P
finitely many of which are zero, such that αj = i,j βij bj , and thus
X X
c= αj aj = βij ai bj ,
j∈J (i,j)∈I×J
so that {ai bj } spans K as an F -vector space. Now suppose the set {ai bj } were
linearly dependent. By definition, this means that there is some finite subset S ⊂
I × J such that {ai bj }(i,j)∈S is linearly dependent, and thus there exist βij ∈ F ,
not all zero, such that X
(βi jbj )ai = 0.
(i,j)∈S
Since
P the ai ’s are K-linearly independent elements of M , we have that for all i,
βij bj = 0, and then similarly, since the bj ’s are linearly independent elements of
K we have βij = 0 for all j.
Remark 3.1. In general the degree [L : K] of a field extension is a cardinal
number, and the statement of Theorem 3.1 is to be interpreted as an identity of
(possibly infinite) cardinals. On the other hand, when M/K and K/F are finite,
the argument shows that M/F is finite and the result reduces to the usual product
of positive integers. Moreover the finite case is the one that is most useful.
Let L/K be an extension of fields and α ∈ L. We say that α is algebraic over K
if there exists some polynomial P (t) = tn + an−1 tn + . . . + a1 t + a0 ∈ K[t] such that
P (α) = 0. If α is not algebraic over K it is said to be transcendental over K. A
complex number which is algebraic over Q is called an algebraic number.
Example 3.2. The element i is algebraic over R since it satisfies the equation
1
i2 +1 = 0. It is also algebraic over Q for the same reason. Indeed for any a ∈ Q, a n
1
is algebraic over Q. This is almost tautological, since by a n , one generally means
any complex number α such that αn = a, so α satisfies tn − a = 0.
The following exercise gives less trivial examples.
a
Exercise 3.1. Let b ∈ Q. Show cos( ab π) and sin( ab π) are algebraic.
Exercise 3.2. a) Show that the set of all algebraic numbers is countably infi-
nite.
b) More generally, let K be any infinite field and L/K be any field extension. Show
that the cardinality of the set of elements of L which are algebraic over K is equal
to the cardinality of K.
So “most” real or complex numbers are transcendental. This was observed by Can-
tor and stands as a famous early application of the dichotomy between countable
and uncountableP∞sets. Earlier Liouville had constructed particular transcendental
numbers, like n=1 10−n! : an application of the Mean Value Theorem shows that a
number which is “too well approximated” by rational numbers cannot be algebraic.
It is of course a different matter entirely to decide whether a particular, not obvi-
ously algebraic, number which is given to you is transcendental. Let us say only
1. INTRODUCTION 25
that both e and π were shown to be transcendental in the 19th century; that there
were some
√
interesting results in transcendence theory in the 20th century – e.g.
eπ and 2 2 are transcendental – and that to P this day the transcendence of many
∞
reasonable looking constants – e.g. π e , ζ(3) = n=1 n13 – is much beyond our reach.
Case 1: I = 0, i.e., Φ embeds K[t] into L. This means precisely that α satisfies
no polynomial relations with K-coefficients, so occurs iff α is transcendental over K.
Let us summarize:
Theorem 3.3. Let L/K be a field extension and α ∈ L.
a) The following are equivalent:
(i) The element α is algebraic of degree d over K.
(ii) The K-vector space K[α] is finite, of degree d.
(iii) The K-vector space K(α) is finite, of degree d.
b) If α is algebraic of degree d, then K[α] = K(α) ∼ = K[t]/(P (t)), where
P (t) ∈ K[t] is the unique monic polynomial of degree d such that P (α) = 0.
c) If α is transcendental over K, then K[t] ∼= K[α] ( K(α) ∼ = K(t).
P (π)
It follows that the set of all rational expressions Q(π) with P, Q ∈ Q[t] is isomorphic
to the rational function field Q(t)! In other words, there is no genuinely algebraic
distinction to be made between “fields of numbers” and “fields of functions.”
so
[L : K] = [L : K(α)][K(α) : K] ≥ ℵ0 .
The converse does not hold: many fields admit infinite algebraic extensions. A
detailed analysis of algebraic field extensions is still ahead of us, but it is easy to
S 1
see that the extension Q[ n≥2 2 n ] is an infinite algebraic extension, since it contains
subextensions of arbitrarily large finite degree.
Exercise 3.4. (Direct Limits) Let (I, ≤) be a directed set: recall that this
means that I is partially ordered under ≤ and for any i, j ∈ I there exists k ∈ I
with i ≤ k and j ≤ k. A directed system of sets is a family of sets {Xi }i∈I
together with maps ι(i, j) : Xi → Xj for all i ≤ j satisfying the natural compatibility
conditions: (i) ιi,i = 1Xi and (ii) for all i ≤ j ≤ k, ι(i, k) = ι(j, k) ◦` ι(i, j). By
definition, the direct limit limI X is the quotient of the disjoint union i∈I Xi by
the equivalence relation (x, Xi ) ∼ (ι(i, j)x, Xj ) for all i ≤ j.
a) Show: there are natural maps ιi : Xi → limI Xi . State and prove a
universal mapping property for the direct limit.
b) Suppose the maps ι(i, j) are all injective. Show that the maps ιi : Xi →
limI Xi are all injective. Explain why in this case limI Xi is often infor-
mally referred to as the “union” of the Xi ’s.
c) In any concrete category C – i.e., a category whose objects are sets, for
which the set of all morphisms from an object A to an object B is a subset
of the set of all functions from A to B, and for which composition and
identity of morphisms coincide with the usual notions of functions – one
has the notion of a directed system {Ai } of objects in C, i.e., we have
sets Ai indexed by the directed set (I, ≤) and for all i ≤ j, the function
ι(i, j) : Ai → Aj is a morphism in C. Give a definition of the direct limit
limI Ai in this more general context. Show that the direct limit exists
in the following categories: monoids, groups, commutative groups, rings,
commutative rings, fields.
d) Give an example of a concrete category in which directed limits do not
necessarily exist.2
e) Show that a field extension L/K is algebraic iff it is the direct limit of its
finite subextensions.
Thus any constructible number α lies in a field which is at the top of a tower
of quadratic field extensions, so [Q(α) : Q] is a power of 2. The impossibility of
three classically sought after constructions follows easily.
First we cannot double the cube: given a cube with sides of our unit length,
we cannot construct a cube whose √ volume√is twice that of the given cube, because
the length of a side would be 3 2, and [Q( 3 2) : Q] = 3. Similarly we can construct
angles that we cannot trisect; in particular, we can construct
√
an angle of 60 degrees
o 1 o 3
(i.e., we can construct cos 60 = 2 and sin 60 = 2 ), but we cannot construct
cos 20o since it satisfies an irreducible cubic polynomial over Q. Finally, we cannot
square the circle i.e., construct a square whose √ area is that of a unit circle, for
that would involve constructing a side length of π and π is not even algebraic!
Let us this result in a more general context, that of integral extensions of do-
mains. The generalized proof is not much harder and will be extremely useful for
any student of algebra. So: let R be a domain and S a domain which extends R,
i.e., there is an injective homomorphism R → S. We say that α ∈ S is integral
over R if α satisfies a monic polynomial with R-coefficients:
∃ an−1 , . . . , a0 ∈ R | αn + an−1 αn−1 + . . . + a1 α + a0 = 0.
We say that the extension S/R is integral if every element of S is integral over R.
28 3. FIELD EXTENSIONS
Let A be the n × n matrix αIn − (rij ); then recall from linear algebra that
AA∗ = det(A) · In ,
where A∗ is the “adjugate” matrix (of cofactors). If m = (m1 , . . . , mn ) (the row
vector), then the above equation implies 0 = mA = mAA∗ = m det(A) · In . The
latter matrix equation amounts to mi det(A) = 0 for all i. Thus • det(A) = •0 on
M , and by faithfulness this means det(A) = 0. Since so that α is a root of the
monic polynomial det(T · In − (aij )).
Lemma 3.7. Let R ⊂ S ⊂ T be rings. If α ∈ T is integral over R, then it is
also integral over S.
Proof. If α is integral over R, there exists a monic polynomial P ∈ R[t] such
that P (α) = 0. But P is also a monic polynomial in S[t] such that P (α) = 0, so α
is also integral over S.
Lemma 3.8. Let R ⊂ S ⊂ T be rings. If S is finitely generated as an R-
module and T is finitely generated as an S-module then T is finitely generated as
an R-module.
Proof. If α1 , . . . , αr generates S as an R-module and β1 , . . . , βs generates T
as an S-module, then {αi βj }{ 1 ≤ i ≤ r, 1 ≤ j ≤ s} generates T as an R-module:
for α ∈ T , we have X XX
α= bj βj = (aij αi )βj ,
j i j
4. DISTINGUISHED CLASSES 29
4. Distinguished Classes
Here is an organizing principle for classes of field extensions due to S. Lang.
(DC1) (Tower meta-property) For a tower M/K/F , then M/F ∈ C iff M/K ∈ C
and K/F ∈ C.
(DC2) (Base change meta-property) Let K/F be an element of C, let L/F be any
30 3. FIELD EXTENSIONS
extension such that K and L are contained in a common field. Then LK/L ∈ C.
Normal Extensions
Proof. By Proposition 4.1, if ClL (K) is not algebraically closed then there is
a monogenic finite degree extension K(α) ) K. Because α is algebraic over ClL (K)
and ClL (K) is algebraic over K, we have by Corollary 3.9 that α is algebraic over K.
Let f ∈ F [t] be the minimal polynomial of α. By Proposition 4.1, as a polynomial
over L[t] we have
f (t) = (t − α1 )(t − α2 ) · · · (t − αd )
for some α1 , . . . , αd ∈ L. Each αi is algebraic over K so lies in ClL (K). Moreover
the α1 , . . . , αd are the only roots of f in L, and thus for some i we have α = αi ∈
ClL (K), a contradiction.
Corollary 4.5. The field Q of all algebraic numbers is algebraically closed.
Proof. Since Q is the algebraic closure of Q in C, this follows from Theorem
4.2 and Proposition 4.4.
Let K be a field. An algebraic closure of K is a field extension K/K that is both
algebraic and algebraically closed. It follows from Proposition 4.1 an algebraic
closure of K is precisely a maximal algebraic extension of K, i..e., an algebraic
extension that is not properly contained in any other algebraic extension of K.
Exercise 4.2. Let K/F be an algebraic field extension. Let L/K be a field
extension. Show: L is an algebraic closure of K iff L is an algebraic closure of F .
enough to use that every proper ideal is contained in a prime ideal: this gives us a
domain, and we can take the fraction field. The assertion that every proper ideal
in a commutative ring is contained in a prime ideal is known to be equivalent to
the Ultrafilter Lemma (UL), which does not imply (AC).
It seems to be an open problem whether the existence of an algebraic closure of
every field implies (UL): cf. http://mathoverflow.net/questions/46566. How-
ever, it is known that (AC) is required for Theorem 4.7 to hold in the sense that
there is a model of Zermelo-Fraenkel set theory in which not every field admits an
algebraic closure [Je, Thm. 10.13].
The proof of Theorem 4.7 comes from E. Artin by way of Lang [LaFT]. It is
unnecessarily (though helpfully) slick in several respects. The use of polynomial
rings is a crutch to avoid some mostly set-theoretic unpleasantries: later we wil
see that an algebraic closure of F is essentially the direct limit of all finite degree
normal field extensions K/F : here the essentially means that we want each K/F
to appear exactly once up to F -isomorphism. It just happens that the easiest way
to do that is to realize each K inside a fixed algebraically closed field containing F !
But by the time the reader has made it to the end of this section, she may consider
trying to construct this direct limit directly.
Note the scare quotes around uniqueness. This is because we have shown that
the algebraic closure of F is unique up to F -algebra isomorphism, but given two
algebraic closures of F there is in general no canonical F -algebra isomorphism
between them. If ϕ, ψ : F1 ,→ F2 are two F -algebra isomomorphisms, then ψ −1 ◦ ϕ
is an F -algebra automorphism of F1 , and conversely: the ambiguity in the choice of
isomorphism is precisely measured by the group GF := Aut(F 1 /F ). This group is
called the absolute Galois group of F and is in general a very large, interesting
group. In fact, we should not speak of “the” absolute Galois group of F (though we
will: it is traditional to do so): it is well-defined up to isomorphism, but switching
from one isomorphism F1 → F2 to another gives rise to an inner automorphism
(i.e., a conjugation) of G. More on this later.
Remark 4.1. There are models of Zermelo-Fraenkel set theory – i.e., without
(AC) – in which a field F can admit non-F -isomorphic algebraic closures.
36 4. NORMAL EXTENSIONS
4. Conjugates
Let K/F be an algebraic field extension. We say that elements α, β ∈ K are con-
jugate over F if α and β have the same minimal polynomial over F . Thus if
[K(α) : K] = d, then the number of conjugates of α lying in K is at most d. We
will see later that whether α has d conjugates in K is is a crucial issue: it will be so
if and only if the extension K(α)/K is both normal (the property we are studying
in this chapter, with the definition to come soon) and separable (the property we
will study in the next chapter).
For the remainder of this section we fix an algebraic closure F of F and only
consider algebraic extensions K/F that are subextensions of F /F (again, every
algebraic extensions occurs this way up to F -algebra isomorphism). From this per-
spective, being conjugate over F is an equivalence relation on F . Moreover, if σ
is an F -algebra automorphism of F , then for all α ∈ F , we have that σ(α) is a
conjugate of α: indeed, for every polynomial f ∈ F [t], we have
f (α) = 0 ⇐⇒ f (σ(α)) = 0
and thus α and σ(α) have the same minimal polynomial. Conversely, if α, β ∈ F
are conjugate over F , then there is an F -algebra automorphism σ of F such that
σ(α) = β. Indeed, let f ∈ F [t] be the common minimal polynomial of α and β.
Then the field extensions F (α) and F (β) are both isomorphic to F [t]/(f (t)), so
there is an isomorphism
F (α) → F (β),
which by Corollary 5.2 extends to an automorphism of F . We sum up this discussion
as follows.
Lemma 4.13 (Extension Lemma). Let F be an algebraic closure of the field F .
For α, β ∈ F , the following are equivalent:
(i) The elements α and β are conjguate over F : that is, they have the same
minimal polynomial.
(ii) There is σ ∈ Aut(F /F ) such that σ(α) = β.
Remark 4.2. Recall that if a group G acts on a set X, we say that two elements
x, y ∈ X are conjugate if there is g ∈ X such that gx = y. As we just saw, the
terminology of conjugate elements of F is compatible with this: two elements of F
are conjugate iff they are conjugate under the action of Aut(F /F ).
Exercise 4.7. Let Q be an algebraic closure of Q. Show that Aut(Q/Q) is
infinite.
6. NORMAL EXTENSIONS 37
5. Splitting Fields
It follows from Proposition 4.8 that if K/F is an algebraic field extension such that
every nonconstant f ∈ F [t] splits in K, then K is an algebraic closure of F . This
view on algebraic closure opens the door to a natural and important generalization:
we go from “all polynomials” to “some polynomials.”
6. Normal Extensions
Lemma 4.15. Let F be an algebraic closure of F , let K be a subextension of
F /F , and let σ : K ,→ F be an F -algebra embedding. The following are equivalent:
(i) σ(K) ⊂ K.
(ii) σ(K) ⊃ K.
(iii) σ(K) = K.
Proof. Certainly (iii) implies both (i) and (ii). We will show (i) =⇒ (iii),
and it will be clear how to modify the argument so as to obtain (ii) =⇒ (iii).
(i) =⇒ (iii): Let α ∈ K, and let S be the set of F -conjugates of α that lie in
K. We observe that S is a finite set containing α. For β ∈ K, we have that β is a
conjugate of α iff σ(β) is a conjugate of α, so the set of F -conjugates of α that lie in
38 4. NORMAL EXTENSIONS
σ(K) is precisely σ(S). By hypothesis we have σ(S) ⊂ S; since both are finite sets
of the same cardinality we must have σ(S) = S and thus α ∈ σ(S) ⊂ σ(K).
Theorem 4.16. Let K/F be an algebraic field extension. Let F be an algebraic
closure of K (hence also of F ). The following are equivalent:
(i) For every F -algebra embedding σ : K ,→ F we have σ(K) = K.
(ii) K/F is the splitting field of a subset S ⊂ F [t].
(iii) Every irreducible polynomial f ∈ F [t] with a root in K splits in K.
(iv) For all α ∈ K, if β ∈ F is an F -conjugate of α, then β ∈ K.
An extension K/F satisfying these properties is called normal.1
Proof. (i) ⇐⇒ (iv): We saw above that for α ∈ K and β ∈ F , β is a
conjugate of α in F iff there is an F -algebra homomorphism σ : K ,→ F such
that σ(α) = β. It follows S that as we range over all F -algebra homomorphisms
σ : K ,→ F , we have that σ σ(K) is the set of all conjugates of all elements of K.
Condition
S (iv) holds iff the set of all conjugates of all elements of K is just K itself
iff σ σ(K) = K iff σ(K) ⊂ K for all σ iff σ(K) = K for all σ(K): condition (i).
(iii) ⇐⇒ (iv) is immediate.
(ii) ⇐⇒ (iv): Condition (ii) can be rephrased by saying that K is generated by
adjoining to F a subset S of F that is stable under conjugation. Thus if (iv) holds,
then (ii) holds with S = K. Conversely, suppose that K is obtained by adjoining to
F a set S that is stable under conjugation, and let x ∈ K. Then x = f (α1 , . . . , αn ) is
a rational function in elements α1 , . . . , αn ∈ S with F -coefficients. Every conjugate
of x in F is of the form σ(x) for some F -automorphism σ of F , and then
σ(x) = f (σ(α1 ), . . . , σ(αn )) ∈ K,
since S is closed under conjugation.
Exercise 4.10. Let F ⊂ K ⊂ L be field extensions. If L/F is normal, show
that L/K is normal.
Corollary 4.17. Let K/F be a normal field extension. Let L/K be any field
extension and let σ : L → L be any automorphism. Then σ(K) = K.
Proof. Let L be an algebraic closure of L. By Corollary 4.11, we may extend
σ to an automorphism σ : L → L. Let K = ClL (K), the unique algebraic closure
of K contained in L. Since K/F is algebraic, K is also an algebraic closure of F .
Since σ(K) is also an algebraic closure of K contained in L, by the aforementioned
uniqueness we have σ(K) = K. By Theorem 4.16 we have σ(K) = K.
√
Example 4.18. For each n ≥ 3, the extension K = Q[ n 2]/Q is a non-normal
extension of√ degree n. Indeed, let ζn = e2πi/n ; then the other roots of tn − 2 in
C are ζnk · n 2 with 0 ≤ k < n, which are not even real numbers unless k = 0 or
k = n2 . So tn − 2 does not split over K. In this case, any extension of K which
√
is normal over Q must contain all the roots of tn − 2, hence must contain n 2 and
ζn . Therefore
√ the smallest normal extension is the splitting field of tn − 2, which is
n
M = Q[ 2, ζn ].
Exercise 4.11. a) Suppose K/F is finite of degree at most 2. Show:
K/F is normal.
1Normal field extensions are by definition algebraic.
6. NORMAL EXTENSIONS 39
√ √
b) Use the example Q ⊂ Q( 2) ⊂ Q( 4 2) to show that if we have fields
F ⊂ K ⊂ L and K/F and L/K are both normal, then L/F need not be
normal. ((Thus normality does not satisfy the tower meta-property.)
Example 4.19. Suppose F has characteristic p > 0. Suppose a ∈ F is such
n
that f (t) = tp − a ∈ F [t] is irreducible. Let K = F [t]/(f (t)), and write α for the
n n
coset t+(f (t)): thus αp = a. Then as an element of K[t] we have f (t) = (t−α)p .
That is, despite the fact that f has degree pn , α is conjugate in F only to itself.
Thus K/F is a normal extension.
Exercise 4.12. Show: a direct limit of normal extensions is normal.
Exercise 4.13. a) Let L/F be an extension and K1 , K2 be subexten-
sions. Show: K1 ∩ K2 is again an extension field of F .
b) As above, but with any collection of intermediate field extensions {Ki }i∈I .
Proposition 4.20. Let L/F be an extension and {Ki }i∈I /F a collection of
algebraic
T subextensions. If each Ki /F is normal, then so is the intersection K =
i K i .
Proof. Without loss of generality we may replace L by the algebraic extension
ClL (F )/F . Let f ∈ F [t] be an irreducible polynomial with a root α in K. Then
for all i ∈ I, f has a root in Ki , thus each fi contains all the F -conjugates of α,
hence so does K, so f splits in K.
8. Isaacs’ Theorem
The goal of this section is to prove the following result of Isaacs.
Theorem 4.23. (Isaacs [Is80]) Let F be a field. For an algebraic extension
K/F , let P(K) be the set of polynomials f ∈ F [t] having a root in K. Then for
algebraic extensions K1 /F , K2 /F , the following are equivalent:
(i) The F -algebras K1 and K2 are isomorphic.
(ii) We have P(K1 ) = P(K2 ).
Exercise 4.15. a) Show: Isaacs’ Theorem implies that a field extension
L/F that contains a root of every nonconstant f ∈ F [t] contains an alge-
braic closure of F .
b) Deduce Gilmer’s Theorem (Theorem 4.9).
CHAPTER 5
Let K/F be an algebraic field extension. We have already explored one desirable
property for K/F to have: normality. Normality can be expressed in terms of
stability under F -homomorphisms into any extension field, and also in terms of
irreducible polynomials: every irreducible polynomial in F [t] with a root in K[t]
must split. There is another desirable property of an algebraic extension L/K called
separability. In some sense it is dual to normality, but this is hard to believe at
first because there is a large class of fields F for which all algebraic extensions K/F
are separable, including all fields of characteristic 0. (For that matter, there are
fields for which every algebraic extension is normal, like R and Fp .) Like normality,
separability can also be expressed in terms of polynomials and also in terms of
embedding conditions. We begin with a study of polynomials.
1. Separable Polynomials
A nonconstant polynomial P ∈ F [t] is separable if over an algebraic closure F ,
P (t) splits into distinct linear factors. Equivalently, if P has degree n, then there
are n distinct elements α1 , . . . , αn ∈ F such that P (αi ) = 0 for all i. Each condition
is easily seen to be independent of the chosen algebraic closure.
Exercise 5.1. Let F be a field and K/F be any extension. Show that a poly-
nomial P ∈ F [t] is separable as a polynomial over F iff it is separable when viewed
as a polynomial over K.
Lemma 5.1. Let F be a field of characteristic p > 0 and α ∈ F × \ F ×p . (That
n
is, α is not a pth power in F .) Then for all n ≥ 1, the polynomial tp − α is
irreducible.
Proof. We shall prove the contrapositive: suppose that for some n ∈ Z+ the
n
polynomial tp − α is reducible; we will show that α is a pth power in F . We may
n
write tp − α = f (t)g(t), where f (t) and g(t) are nonconstant monic polynomials.
n
Let K/F be an extension field containing a root β of tp − α, so that in K[t] we
have n n n n
tp − α = tp − β p = (t − β)p .
Since K[t] is a UFD and f (t) and g(t) are monic, we therefore have f (t) = (t − β)r
for some 0 < r < pn . Write r = pm s with gcd(p, s) = 1. Note that m < n. Then
m m
f (t) = (tp − β p )s ,
m m
so that the coefficient of tp (s−1) is −sβ p . This lies in F and – since s 6= 0 in F
m
– we conclude β p ∈ F . Thus
m n−m n−m
α = (β p )p ∈ Fp ∈ Fp
since m < n.
41
42 5. SEPARABLE ALGEBRAIC EXTENSIONS
Over any field F it is no trouble to come up with a polynomial that is not sepa-
rable: t2 . What is of more interest is whether there is an inseparable irreducible
polynomial in F [t]. Note that some authors define a polynomial to be separable if
all its irreducible factors are separable and others only discuss in/separability for
irreducible polynomials. Although these conventions certainly “work” as well, I find
the current definition to be more convenient and more thematic. First, Exercise 5.1
shows that with this definition, separability is faithfully preserved by base exten-
sion. Since the way one will check whether an irreducible polynomial is separable is
by considering it over the algebraic closure, where of course it is a product of sep-
arable (linear!) polynomials, our definition seems simpler. Moreover, in the theory
of algebras one does meet reducible polynomials: for any nonconstant P ∈ F [x],
we may consider the finite-dimensional F -algebra AP = F [x]/(P (x)). Then our
definition makes it true that P is separable iff AP is a separable algebra, i.e., an
algebra that is semisimple and remains semisimple after arbitrary base change.
In general it is far from obvious whether the field extension obtained by adjoin-
ing a root of an irreducible polynomial is normal. Fortunately, it is much easier to
determine whether a polynomial, especially an irreducible polynomial, is separable.
Exercise 5.2. Let k be a field.
a) Show: there is a unique k-linear endomorphism f 7→ f 0 of k[t] such that
for all n ∈ N we have (tn )0 = ntn−1 .
b) Show: for all f, g ∈ k[t] we have (f g)0 = f 0 g + f g 0 .
c) Show: for all f, g ∈ k[t] we have (f (g(t)))0 = f 0 (g(t))g 0 (t).
d) Suppose k has characteristic 0. Show: if deg(f ) = n ≥ 1, then deg(f 0 ) =
n − 1. Deduce that {f ∈ k[t] | f 0 = 0} = k.
e) Suppose k has characteristic p > 0. Show:
{f ∈ k[t] | f 0 = 0} = k[tp ].
Proposition 5.2 (Derivative Criterion). Let f ∈ F [t] be a nonconstant poly-
nomial. Then:
a) The polynomial f is separable iff gcd(f, f 0 ) = 1.
b) If f is irreducible, it is separable iff f 0 6= 0.
c) An irreducible polynomial is always separable in characteristic 0. In char-
acteristic p > 0, an irreducible polynomial is inseparable iff there exists
g ∈ F [t] such that f (t) = g(tp ).
Proof. a) Let d ∈ F [t] be a greatest common divisor of f and f 0 . This means
{αf + βf 0 mod α, β ∈ F [t]} = {γd | γ ∈ F [t]}.
If K/F is any field extension, it follows that
{αf + βf 0 mod α, β ∈ K[t]} = {γd | γ ∈ K[t]},
so d is again a greatest common divisor of f and f 0 in K[t]. Moreover, if F is an
algebraic closure of F and K is an algebraic closure of K, then the Magic Mapping
Theorem gives an F -algebra homomorphism F ,→ K, so f ∈ F [t] is separable iff
f ∈ K[t] is separable. Thus both of the conditions of part a) are stable under
replacing F by an extension field, so we may assume that F is algebraically closed
and thus f is split. If f is not separable, then for some α ∈ F we have
f = (t − α)2 g
1. SEPARABLE POLYNOMIALS 43
and thus
f 0 = (t − α)2 g 0 + 2(t − α)g = (t − α)h
so (t − α) | gcd(f, f 0 ). Conversely, if f is separable, then for every root α of f we
have
f = (t − α)g with g(α) 6= 0,
so
f 0 = (t − α)g 0 + g,
so f 0 (α) = g(α) 6= 0 and thus (t − α) - f 0 . Thus gcd(f, f 0 ) = 1.
b) If f is irreducible, then since gcd(f, f 0 ) | f , if gcd(f, f 0 ) 6= 1 then gcd(f, f 0 ) = f
so f | f 0 . Since deg(f 0 ) < deg(f ), this occurs if and only if f 0 = 0.
c) This follows from part b) and the previous exercise.
hence P (t) is not irreducible after all. Therefore F is perfect. Inversely, if the
Frobenius homomorphism is not surjective, then there exists some α ∈ F which is
not a pth power, and then by Lemma 9.20, the inseparable polynomial tp − α is
irreducible, so F is not perfect. This gives part a). As for part a), like any field
homomorphism, the Frobenius map is injective, and an injective map from a finite
set to itself is necessarily surjective. If F is algebraically closed, then for any α ∈ F
the polynomial tp − α has a root in F , i.e., α ∈ F p .
a
For any positive integer a, we may consider the pa , the map which takes α 7→ αp ,
which can also be described as the ath power of the Frobenius map. We write
a
F p = pa (F ). If F is not perfect then we get an infinite descending chain of proper
subfields
2
F ) Fp ) Fp ) ...
a−1 a−1 a
Indeed, if α ∈ F \ F p , then αp ∈ Fp \ F p . This gives another proof that an
imperfect field is infinite.
44 5. SEPARABLE ALGEBRAIC EXTENSIONS
For an extension K/F , the separable degree [K : F ]s is the cardinality of the set
of F -algebra embeddings σ : K → F .
Exercise 5.5. Show that the separable degree may be computed with respect to
embeddings into any algebraically closed field containing F .
Theorem 5.5. The separable degree is multiplicative in towers: if L/K/F is a
tower of finite degree field extensions, then [L : F ]s = [L : K]s [K : F ]s .
Proof. Let σ : F ,→ C be an embedding of F into an algebraically closed field.
Let {σi }i∈I be the family of extensions of σ to K, and for each i ∈ I let {τij }j∈Ji be
the family of extensions of σi to L. Each σi admits precisely [L : K]s extensions to
2. SEPARABLE ALGEBRAIC FIELD EXTENSIONS 45
Theorem 5.10.
Let L/F be an algebraic field extension. The following are equivalent:
(i) Every finite subextension of L/F is separable.
(ii) Every irreducible polynomial P ∈ F [t] which has a root in L is separable.
(iii) L is obtained by adjoining to F a set of roots of separable polynomials.
An extension satisfying these equivalent properties is called a separable algebraic
extension.
Exercise 5.7. Prove Theorem 5.10.
Corollary 5.11. Algebraic separable extensions are a distinguished class of
field extensions.
Exercise 5.8. Prove Corollary 5.11.
Corollary 5.12. For a family {Ki /F }i∈I of algebraic field extensions inside
a common algebraically closed field M , the following are equivalent:
(i) For all i ∈ I, Ki /F
Q is a separable algebraic field extension.
(ii) The compositum i Ki is a separable algebraic field extension.
Exercise 5.9. Prove Corollary 5.12.
Corollary 5.12 has the following important consequence: for any field extension
K/F , there exists a unique maximal separable algebraic subextension SepClK (F ),
the separable closure of F in K.
Proof. One may reduce to the case of a simple extension K = F [α], and then
a
α is purely inseparable over F so has minimal polynomial of the form tp − α for
some a ∈ Z+ .
Corollary 5.15. A purely inseparable extension is normal.
Proof. This follows immediately from condition (i) of Theorem 5.13.
The flavor of these results is that many formal properties are common to both
separable and purely inseparable extensions. The exceptions to this rule are the
following: first, purely inseparable extensions are always normal, whereas this is
most certainly not the case for separable extensions. A more subtle difference is
expressed in Theorem 5.13: if K/F is not purely inseparable, then it must have
a nontrivial separable subextension. However, if K/F is not separable, that does
not mean that it has a nontrivial purely inseparable subextension.
Example 5.16. [Mo96, p. 48]: Let k be a field of characteristic 2, F =
k(x, y) (rational function field), u a root in F of the separable irreducible quadratic
√
polynomial t2 +t+x, S = F (u) and K = S( uy). Clearly K/S is purely inseparable
and S/F is separable. But there is no nontrivial purely inseparable subextension of
K/F . Equivalently, we will show that if a ∈ K, a2 ∈ F , then already a ∈ F . An
√ √
F -basis for K is 1, u, uy, u uy. If a2 ∈ F , write
√ √
a = α + βu + γ uy + δu uy, α, β, γ, δ ∈ F.
Since a2 ∈ F , the coefficient of u = 0, i.e.,
β 2 + (γ + δ)2 y + δ 2 xy = 0.
If δ = 0 then β 2 + γ 2 y = 0, so γ = 0 since y is not a square in F . But then β = 0
and a ∈ F . If δ 6= 0, then
β 2 + (γ + δ)2 y γ β
x= 2
= ( + 1)2 + ( )2 y,
δ y δ δ
so that x ∈ F 2 (y), which is not the case. So δ = 0 and a ∈ F .
Theorem 5.19. For an algebraic extension K/F , let Fs and Fi be, respec-
tively, the separable and purely inseparable closures of F in K. The following are
equivalent:
(i) K = Fs Fi .
(ii) K is separable over Fi .
Proof. (i) =⇒ (ii): K is obtained by adjoining to Fi roots of separable
polynomials with coefficients in F , hence by polynomials with coefficients in Fs .
(ii) =⇒ (i): If K/Fi is separable, then K/Fi Fs is separable. Similarly, since K/Fs
is inseparable, K/Fi Fs is inseparable. By Proposition 5.17, K = Fi Fs .
Corollary 5.20. The equivalent conditions of Theorem 5.19 hold when K/F
∞
is normal. In particular they hold for F /F , giving F = F sep F 1/p .
Proof. Let α ∈ K \ Fi . The minimal polynomial P (t) ∈ Fi [t] of α has at least
one other distinct root, say β, in an algebraic closure F . Snce K/F is normal, also
K/Fi is normal, so we have β ∈ K, and the Extension Theorem (Theorem 4.22)
shows that there is an element s ∈ Aut(K/Fi ) such that s(α) = β. This shows that
the set of elements in F that are fixed by every element of Aut(K/Fi ) is Fi .
Later on we will see that we have proved that K/Fi is a Galois extension, which
implies that it is separable. For now we argue by hand: let α = α1 , . . . , αr be all
the Fi -conjugates of α in K, and consider the separable polynomial The
Yr
R(t) := (t − αi ) ∈ K[t].
i=1
The action of Aut(K/Fi ) on K extends to an action on K[t] by
Xn n
X
s( ai ti ) := s(ai )ti .
i=0 i=0
Since element s ∈ Aut(K/Fi ) permutes the roots αi of R, we have σ(R) = R, and
thus every coefficient of R is fixed by every element of Aut(K/Fi ), so R ∈ Fi [t]. It
follows that R = P , so P is separable, and thus K/Fi is (normal and) separable.
The second sentence of the Corollary follows immediately from the first.
Corollary 5.21. For a finite degree extension K/F , we have
[K : F ]s = [SepClK (F ) : F ] | [K : F ].
Proof. We have [K : F ]s = [K : SepClK (F ) : F ]s [SepClK (F ) : F ]s . But the
separable degree of a purely inseparable extension is 1, so the conclusion follows.
For a finite degree extension K/F one may therefore define the inseparable de-
gree [K : F ]i of a finite degree extension to be [K : F ]/[K : F ]s = [K : SepClK (F )].
Corollary 5.23. Let K/F be a normal algebraic extension. Then the sepa-
rable closure F s of F in K is also normal.
Proof. For any embedding σ of K into F , the image σ(F s ) lies in K (by
normality of K) and is evidently also a separable subextension of K/F . Therefore
we must have σ(F s ) = F s .
Corollary 5.24. A field F is perfect iff its separable closure is algebraically
closed.
Proof. If F is perfect then all algebraic extensions are separable, so the result
is clear. Inversely, suppose that F is not perfect, so there exists α ∈ F \ F p
and a corresponding purely inseparable field extension F [α1/p ]/F defined by the
irreducible inseparable polynomial P = tp − α. By Theorem 5.10, only a separable
irreducible polynomial can acquire a root in a separable field extension, so the
polynomial P remains irreducible over the separable closure of F .
Corollary 5.25. Let K/F be a finite degree field extension. Then F is perfect
iff K is perfect.
Proof. The result holds vacuously in characteristic 0, since all fields are per-
fect. So suppose that F (hence also K) has characteristic p > 0. If F is perfect,
then every algebraic extension of F is separable, hence every algebraic extension of
every algebraic extension of F is separable, and thus not only is every finite degree
extension of F perfect but every algebraic extension of F is perfect.
The new case is therefore when F is not perfect, so there is x ∈ F \ F p . We
−N −N
claim that there is a largest natural number N such that xp lies in K: if so, xp
lies in K \ K p , so K is not perfect. To establish the claim, we observe that since
x is not a pth power in F , Lemma 9.20 says that for all n ∈ Z+ the polynomials
n −n
tp − x ∈ F [t] are irreducible and thus we have [F (xp ) : F ] = pn . Thus when n
−n
is large enough so that [K : F ] < pn we cannot have xp ∈ K. This establishes
the claim and completes the proof.
CHAPTER 6
Choose a K-basis b1 , . . . , bn for L. With respect to such a basis, the linear trans-
formation x• is represented by an n × n matrix, say M (x).
Example 6.2. Take K = R, L = C, and the basis (1, i). Let x = a + bi. Then
x • 1 = a · 1 + b · i and x • i = −b · 1 + a · i. Therefore
a −b
M (x) = .
b a
51
52 6. NORMS, TRACES AND DISCRIMINANTS
Example 6.3. If x lies in K, then M (x) = mi,j is simply the scalar matrix
diag(x, . . . , x). Proposition 6.4 below gives a generalization.
We define the characteristic polynomial of x:
n
Y
Px (t) = det(tIn − M (x)) = (t − λi ).
i=1
Corollary 6.5. Let L/F be a finite degree field extension. Let α be an el-
ement of L, let f (t) be the minimal polynomial of α over F , and let g(t) be the
characteristic polynomial of α• ∈ EndF (L). Then g(t) = f (t)[L:F (α)] .
Proof. Put K = F (α). The minimal polynomial f of α over F is the char-
acteristic polynomial of x• ∈ EndF (K). So the result follows from Proposition
6.4.
Theorem 6.7. Let K/F be a field extension of degree n < ∞ and separable
degree ns . Put pe = nns = [K : F ]i . Let K be an algebraic closure of K. Let α ∈ K
and let f (t) be the characteristic polynomial of α• ∈ EndF (K). Let τ1 , . . . , τns be
the distinct F -algebra embeddings of K into K. Then
ns
Y e
f (t) = (t − τi (α))p .
i=1
It follows that
ns
Y e
(4) NK/F (α) = ( τi (α))p
i=1
and
m
X
(5) TrK/F (α) = pe τi (α).
i=1
be the minimal polynomial of α over F , and let g(t) be the characteristic polynomial
of α• on K, so by Corollary 6.5 we have
ds ds
!ni
ns
di n
Y Y
[K:L]
g(t) = f (t) = ( (t − σi (α)) d = ( (t − σi (α)) ds
i=1 i=1
ns
!pi
Y
= (t − τi (α)) .
i=1
Corollary 6.8. Let Fqd /Fq be an extension of finite fields. Then the norm
map N : F×
qd
→ F×
q is surjective.
symmetric in the sense that T (x, y) = T (y, x) for all x, y ∈ K. A natural question
is when the trace form is nondegenerate.
Theorem 6.10. Let K/F be a field extension of finite degree n. The following
are equivalent:
(i) The trace form T : K × K → F is nondegenerate.
(ii) There exists some x ∈ K such that Tr(x) 6= 0.
(iii) The trace function Tr : K → F is surjective.
(iv) The extension K/F is separable.
Proof. The implications (i) =⇒ (ii) =⇒ (iii) may safely be left to the
reader.
(iii) =⇒ (iv): we argue by contraposition. If K/F is not separable, then char(F ) =
p > 0, [K : F ]i = pe is divisible by p, and thus (5) shows that the trace function is
identically zero.
(iv) =⇒ (i): By the Primitive Element Corollary, we have K = F [α] for some
α ∈ K. Then (1, α, . . . , αn−1 ) is an F -basis of K. Let x ∈ K. By Proposition 6.9,
it is enough to show that the Gram matrix M (i, j) = Tr(αi−1 αj−1 ) = Tr(αi+j−2 ) is
nonsingular.
Pn To see this, let α1 , . . . , αn be the distinct F
P-conjugates of α in K. Then
n
Tr(α) = i=1 αi , so that for any N ∈ N, Tr(αN ) = i=1 αiN . Now we introduce
the Vandermonde matrix V = V (α1 , . . . , αn ): V (i, j) = αji−1 . Why? Well, we
Pn
compute that the (i, j) entry of V V T is k=1 αki−1 αkj−1 = M (i, j). Therefore
Y
det M = det V V T = (det V )2 = ( αi − αj )2 6= 0.
i>j
(ii) =⇒ (i): Suppose K = F [α], and let f (t) ∈ F [t] be the minimal polynomial
for α over F . For each subextension E of K/F , let gE (t) ∈ E[t] be the minimal
polynomial for α over E. Let E 0 be the subextension of K/F generated by the
coefficients of gE . So F ⊂ E 0 ⊂ E ⊂ K; since gE is irreducible over E, it is also
irreducible over E 0 , and thus [K : E 0 ] = [E 0 [α] : E 0 ] = [E[α] : E] = [K : E]. It
follows that E = E 0 . In other words, E can be recovered from gE and thus the
map E 7→ gE is bijective. However, we also have that gE divides f for all E, so gE
is a monic polynomial whose multiset of roots in any algebraic closure is a subset
of the multiset of roots of f . So there are only finitely many possibilities for E.
Since σi and σj are distinct F -algebra embeddings of K = F [α, β], we must have
either σi (α) 6= σj (α) or σi (β) 6= σj (β), so P is nonzero. By Alon-Tarsi (Lemma
7.1) there are s2 ∈ S such that P (s1 , s2 ) 6= 0, which means that for 1 ≤ i ≤ n the
elements σi (s1 α + s2 β) are all distinct, so we have
F ⊂ F [s1 α + s2 β] ⊂ F [α, β]
and
[F [s1 α + s2 β] : F ] ≥ n = F [α, β],
so
F [α, β] = F [s1 α + s2 β].
Step 2: If we run through the above argument with
Y
P (t) := ((σi (α) − σj (α)) + t(σi (β) − σj (β)) ∈ K[t]
i6=j
in place of P (s, t), we get the similar conclusion that there is s2 ∈ S such that
K[α, β] = K[α + s2 β]. Step 3: We now finish by an inductive argument: let
2 ≤ i ≤ n − 1 and suppose that we have found s2 , . . . , si ∈ S such that
F [x1 , . . . , xi ] = F [s1 x1 + s2 x2 + . . . + si xi ].
2. THE PRIMITIVE ELEMENT THEOREM AND ITS COROLLARY 59
Before we get there though, one remark: we have not yet seen a finite degree
field extension that is not monogenic, and Corollary 7.3 shows that such exam-
ples must be relatively exotic: in particular they must be inseparable and thus
can only occur in positive characteristic. Later we will see that in positive charac-
teristic such examples are relatively abundant: for instance, Exercise 13.10 shows
that if k is any field of characteristic p and k(x1 , . . . , xn ) is a rational function field
in n variables, then the field extension k(x1 , . . . , xn )/k(xp1 , . . . , xpn ) has degree pn
and requires n generators. Taking n = 2 for concreteness, we draw the following
remarkable conclusion.
Corollary 7.5. Let k be a field of characteristic p. The degree p2 field exten-
sion k(x, y)/k(xp , y p ) has infinitely many subextensions.
For me, the existence of a finite degree field extension with infinitely many subex-
tensions is the most startling result in these notes and, in fact, one of the most
counterintuitive mathematical results I have met in my adult mathematical life. (I
am fairly confident that I received my PhD in mathematics blissfully ignorant of
the fact that such field extensions existed. In fact I vaguely thought that standard
results in field theory rule them out...as is certainly the case in characteristic 0.)
Thus we find that the subfield lattice of a finite degree field extension, though both
60 7. THE PRIMITIVE ELEMENT THEOREM
Noetherian and Artinian – i.e., satisfying both the ascending and descending chain
conditions – may be infinite.
CHAPTER 8
Galois Extensions
1. Introduction
For any field extension K/F we define Aut(K/F ) to be the group of F -algebra
automorphisms of K, i.e., the set of all field isomorphisms σ : K → K such that
σ(x) = x for all x ∈ F . This is a group under composition.
K G = {x ∈ K | σ(x) = x ∀σ ∈ G}.
Note that the notation comes from representation theory: if R is a commutative
ring, M an R-module and G is a group, then one has the notion of an R-linear
representation of G on M , i.e., a homomorphism from G to the group of R-module
automorphisms of M . In such a situation one can “take invariants”, i.e., consider
the subset of M on which G acts trivially: this is denoted M G . The present defi-
nition is an instance of this with R = F , M = K.
so that the automorphism group of L/K has order 2, the nontrivial element being
the unique K-automorphism σ of L which sends α 7→ α. Since LAut(L/K ) is a
subextension of the degree 2 extension L/K, it could only be L or K, and since
σ(α) = α 6= α, we conclude that the fixed field is K and the extension is Galois. In
contrast the automorphism group of an inseparable quadratic extension is trivial, so
this extension is not Galois.
√
Example 8.3. Let K = Q[t]/(t3 −2) = Q[ 3 2]. Since K contains exactly one of
the three roots of t3 − 2 in Q, Aut(K/Q) is the trivial group and K/Q is not Galois.√
On the other hand, the automorphism group of the normal closure M = Q[ζ3 , 3 2]
of √
K/Q has order 6: since everything is separable, there are three embeddings of
Q[ 3 2] into M , and each of these extends in two ways to an automorphism of M .
Any automorphism s of M is determined by an i ∈ {0, 1, 2} and j ∈ {0, 1} such
that √ √ j
s : 2 7→ ζ3i 2, ζ3 7→ (ζ3 )(−1) .
3 3
Since there are six possibilities and six automorphisms, all of these maps must
indeed give √automorphisms.
√ In particular, there is an order 3 automorphism√σ
which takes 3 2 7→ ζ3 3 2 and fixes ζ3 and an order 2 automorphism τ which fixes 3 2
and maps ζ3 7→ ζ3−1 . One checks that τ στ = τ στ −1 = σ −1 , i.e., Aut(L/Q) ∼ = S3 ,
the symmetric group on three elements. Indeed, these three elements can be viewed
as the three roots of t3 − 2 √ in M . Finally, the subgroup fixed by {1, σ} is precisely
K, whereas the generator 3 2 of K/Q is not fixed by σ, so that we conclude that
M Aut(M/Q) = Q and M/Q is Galois.
These examples already suggest that a finite degree extension K/F is Galois iff it
is normal and separable, and in this case # Aut(K/F ) = [K : F ]. We will show in
the next section that these conditions are all equivalent.
Example 8.4. The extension Q/Q is Galois. We cannot show this by some sort
of direct computation of GQ := Aut(Q/Q): this group is uncountably infinite and
has a very complicated structure. Indeed, as an algebraic number theorist I am more
or less honorbound to inform you that the group GQ is the single most interesting
group in all of mathematics! We will see that the Galois theory of infinite algebraic
extensions cannot be developed in exactly the same way as in the finite case, but is,
in theory, easily understood by a reduction to the finite case.
Example 8.5. The extension C/Q is Galois, as is C/Q. In particular the
automorphism group of the complex field is (much) larger than just {1, c}. In fact we
will show that if F has characteristic zero and K is algebraically closed, then K/F
is Galois. These results are not part of “Galois theory” as it is usually understood,
but rather are facts about automorphism groups of transcendental extensions. These
results will be shown in §10.1.
Example 8.6. For any field F , Aut(F (t)/F ) is the group of linear fractional
a b
transformations: the group GL2 (F ) of 2×2 matrices with ad 6= bc acts by
c d
automorphisms on F (t), via t 7→ at+b
ct+d . Scalar matrices – those with b = d = 0, a =
c – act trivially, so the action factor through to the quotient P GL2 (F ) of GL2 (F )
by the subgroup F × of scalar matrices. It is a standard fact (more in the vein
of algebraic geometry than pure field theory) that this is the entire automorphism
group of F (t).
2. FINITE GALOIS EXTENSIONS 63
Remark Aside: I am not aware of a simple necessary and sufficient condition for
an extension K/F which is finitely generated, but of infinite degree, to be Galois.
When K/F is regular of transcendence degree 1 (two terms which we have not
yet defined), one can give such a criterion in terms of the Jacobian J(C) of the
corresponding algebraic curve C/F , namely K/F is Galois iff dim J(C) = 0 or
(dim J(C) = 1 and J(C)(F ) is infinite). In particular no such field of genus g ≥ 2
is Galois. One can give some examples of Galois extensions of higher transcendence
degree – e.g. the proof of Proposition XX easily adapts to show that F (t1 , . . . , tn )/F
is Galois if F is infinite – but the general problem seems to be a quite subtle one
in birational arithmetic geometry.
K-linearly independent. Therefore in the last equation we must have ci = 0 for all
i, a contradiction.
K/F , being a compositum of the separable extensions s(K) as s ranges over the
finite set of distinct F -algebra embeddings of K into F is separable and normal,
hence Galois. M/K is even the minimal extension of K which is normal over F , so
certainly it is the minimal such Galois extension.
In view of Corollary 8.11, it is reasonable to call the normal closure of a finite degree
separable field extension the Galois closure.
Theorem 8.12 (Natural Irrationalities). Let K/F be a finite degree Galois
extension, and let L/F be an arbitrary field extension. Then:
a) The field extension KL/L is Galois.
b) The restriction map r : Aut(KL/L) → Aut(K/K ∩ L) is an isomorphism.
c) We have [KL : L] = [K : K ∩ L].
Proof. a) This is the assertion that finite Galois extensions have the base
change meta-property. But all of the following properties have the base-change
meta property: being of finite degree, normality and separability. Alternately, since
K/F is finite Galois, it is the splitting field of the separable polynomial f ∈ F [x].
Then KL/L is the splitting field of the polynomial f ∈ L[x], which is still separable
because of the Derivative Criterion.
b) Let σ ∈ Aut(KL/L), and let r(σ) denote the restriction of σ to K. Since σ
fixes L pointwise and F ⊂ L, also σ fixes F pointwise. So for all x ∈ K, r(σ)(x)
is an F -conjugate of x; since K/F is normal, this implies r(σ)(x) ∈ K and thus
r(σ) ∈ Aut(K/F ). Indeed, because σ pointwise fixes L, r(σ) pointwise fixes K ∩ L
and r(σ) ∈ Aut(K/K ∩ L). This defines a map
r : Aut(KL/L) → Aut(K/K ∩ L).
That r is a group homomorphism is immediate. Moreover, the kernel of r consists
of the set of automorphisms α of KL that pointwise fix both K and L and thus
also pointwise fix KL: r is injective. Finally we must show that α is surjective.
Its image is a subgroup of Aut(K/K ∩ L), which by the Galois correspondence is
therefore of the form Aut(K/E) for some K ∩ L ⊂ E ⊂ K. Now observe that E
is pointwise fixed by every α ∈ Aut(KL/L), so hence E ⊂ (KL)Aut(KL/L) = L. It
follows that E ⊂ K ∩ L and thus E = K ∩ L and α is surjective.
c) By part b) we have
[KL : L] = #Aut(KL/L) = # Aut(K/K ∩ L) = [K : K ∩ L].
which is a subgroup of H. (We could define in the same way X S for any subset
S ⊂ G, but one checks immediately that if H is the subgroup generated by S,
X S = X H , so this extra generality leads nowhere.) To be very formal about it, we
have thus defined a map
Φ : Λ(X) → Λ(G), Y 7→ GY
and a map
Ψ : Λ(G) → Λ(X), H 7→ X H .
Let us explore what can be said about these two maps in this extreme level of gen-
erality. Statements that we do not prove are exercises in unwinding the definitions
and left to the reader. (We do recommend that the reader peform these exercises!)
First, both Φ and Ψ are anti-homomorphisms of the partially ordered sets, i.e.,
if Y1 ⊂ Y2 , then Φ(Y2 ) ⊂ Φ(Y1 ), and similarly if H1 ⊂ H2 then Ψ(H2 ) ⊂ H1 . This
implies that Ψ ◦ Φ : Λ(X) → Λ(X) and Φ ◦ Ψ : Λ(G) → Λ(G) are homomorphisms
of partially ordered sets:
Y1 ⊂ Y2 =⇒ X GY1 ⊂ X GY2 ,
H1 ⊂ H2 =⇒ GX H1 ⊂ GX H2 .
Moreover, for all Y ⊂ X and H ⊂ G we have
(GC) Y ⊂ X H ⇐⇒ H ⊂ GY .
Indeed, both containments assert precisely that every element of H acts trivially on
every element of Y . If H = GY we certainly have the second containment, therefore
by (GC) we have
(6) Y ⊂ X GY .
Dually with Y = X H we certainly have the first containment hence (GC) gives
(7) H ⊂ GX H .
Proposition 8.13. Let H be a subgroup of G, Y a subset of X and σ ∈ G.
We have:
a) σGY σ −1 = GσY .
−1
b) σX H = X σHσ .
Proof. We have g ∈ GσY ⇐⇒ ∀y ∈ Y, gσy = σy ⇐⇒ ∀y ∈ Y, σ −1 gσy =
y ⇐⇒ σ −1 gσ ∈ GY ⇐⇒ g ∈ σGY σ −1 . Similarly, y ∈ σX H ⇐⇒ σ −1 y ∈
X H ⇐⇒ ∀h ∈ h, hσ −1 y = σ −1 y ⇐⇒ ∀h ∈ H, (σhσ −1 )y = y ⇐⇒ y ∈
σHσ −1 .
Let us now introduce the following simplified (and symmetric) notation: for Y ⊂ X,
we write Y 0 for GY ; for H ⊂ G, we write H 0 for X H . Equations (6) and (7) now
read as Y ⊂ Y 00 and H ⊂ H 00 . Let us call a subset Y of X (resp. a subgroup H of
G) closed if Y 00 = Y ( resp. if H 00 = H).
Proposition 8.14. For any Y ∈ Λ(X) and H ∈ Λ(G), we have Y 0 = Y 000 and
H = H 000 . Hence Y 0 is a closed subgroup of G and X 0 is a closed subset of X.
0
3. AN ABSTRACT GALOIS CORRESPONDENCE 67
Exercise 8.2. Let Λ and Λ0 be two partially ordered sets. A Galois connec-
tion between Λ and Λ0 is a pair of order-reversing maps Φ : Λ → Λ0 , Ψ : Λ0 → Λ
satisfying the analogoue of identity (GC)2 above: for x ∈ Λ, y ∈ Λ0 , Φ(x) ≤ y ⇐⇒
x ≤ Ψ(y).
a) Check that the entire discussion (except for the bit about conjugation and nor-
mality) goes through in this level of generality: we get closure operators on Λ and
Λ0 such that Φ and Ψ give mutually inverse anti-automorphisms on the subsets of
∼ ∼
closed elements: Φ : Λc → Λ0c , Ψ : Λ0c :→ Λc .
b) Look for Galois connection. in your everyday (mathematical) life, paying spe-
cial attention to the closure process. For example, consider the polynomial ring
R = k[t1 , . . . , tn ] over an algebraically closed field k. Let Λ be the set of ideals I
of R, and let Λ0 be the set of algebraic
Tr subsets of affine n-space An over k: that
n −1
is, the subsets of k of the form i∈I Pi (0), where {Pi }i∈I is a set of elements
of R. Define Φ : Λ → Λ0 by I 7→ V (I), the set of points of (a1 , . . . , an ) such that
P (a1 , . . . , an ) for all P ∈ I. Define Ψ : Λ0 → Λ by S 7→ I(S), the ideal of all
elements of R which vanish at every (x1 , . . . , xn ) ∈ S. It is no problem to see that
this gives a Galois connection. What are the closed ideals? What are the closed
algebraic subsets?
Aut(K/L).
This theorem is probably the single most important result in field theory. It re-
duces the study of the lattice of subextensions of a finite Galois extension K/F
to the corresponding lattice of subgroups of the finite group Aut(K/F ), which is
much easier to study, e.g. is a priori finite. Indeed, if K/F is any finite separable
extension, then one may – and should! – apply the Galois correspondence to the
Galois closure M/F .
Exercise 8.3. Use the Galois Correspondence to give a more natural proof of
the Primitive Element Corollary (7.3).
When K/F is Galois, we write Gal(K/F ) for Aut(K/F ) and speak of Gal(K/F ) as
the Galois group of K/F . We note that some authors (e.g. Shifrin, Kaplansky)
use the notation Gal(K/F ) for the automorphism group of an arbitrary field exten-
sion, but from the perspective of infinite Galois theory (coming up!) and modern
number theory this seems dangerously misleading. Namely, it would then be tempt-
ing to call any automorphism group of a finite degree extension “a Galois group”
and this is most certainly at odds with contemporary terminology. Indeed, perhaps
the single outstanding problem in field theory is to decide whether, for any finite
group G, there is a Galois extension K/Q such that Gal(K/Q) ∼ = G. However, the
corresponding statement that any finite group is the automorphism group of some
finite extension K/Q – possibly with [K : Q] > #G – is a much weaker one, and
indeed this is a known theorem of E. Fried and J. Kollar [FK78]
This is in fact rather useful: let C be any class of finite groups which is closed under
formation of direct products and passage to subgroups, and suppose that Ki /F are
two C-Galois extensions, i.e., finite Galois extensions whose Galois groups lie in C.
Then the compositum K1 ∨ K2 is a C-Galois extension. E.g. we may profitably
take C to be the class of all finite abelian groups, or the class of all finite solvable
groups. When we turn to infinite Galois theory we will see that we are allowed to
take infinite composita as well, and this observation will show that any field admits
a maximal C-Galois extension.
Exercise 8.4. Let K1 , K2 /F be two finite Galois extensions, and K = K1 K2
their compositum. Let H be the image of the map ι : Gal(K/F ) → Gal(K1 /F ) ×
70 8. GALOIS EXTENSIONS
Every known proof of the existence of normal bases must negotiate a fundamental
dichotomy between finite fields and infinite fields. This dichotomy comes up several
times in field theory, algebra and algebraic geometry (another good example of a
theorem for which the finite field case must be taken separately is the Noether
Normalization Theorem), but often without much fanfare our explanation. To
our mind at least, the source of the trouble is the different behavior of the evalu-
ation map on polynomials over finite domains versus infinite integal domains. (A
geometer might point to the fact that for any n ∈ Z+ , a field K is infinite iff the
K-rational points of affine n-space over K are Zariski dense, but in fact this comes
down to the same algebraic observation.)
Lemma 8.21. Let R ⊂ S be an extension of domains and n ∈ Z+ . The following
are equivalent:
(i) For all f ∈ S[t1 , . . . , tn ], f (a1 , . . . , an ) = 0 for all (a1 , . . . , an ) ∈ Rn =⇒ f = 0.
(ii) R is infinite.
Proof. (i) =⇒ (ii): We prove the contrapositive. Note that any finite domain
is a field, so suppose R = Fq . Let f (t) = tq1 − t1 . Then for all a = (a1 , . . . , an ) ∈ Fnq ,
f (a) = aq1 − a1 = 0.
(ii) =⇒ (i): We go by induction on n.
Base Case (n = 1): suppose f ∈ S[t] is a polynomial which is not the zero polyno-
mial. Then it has degree d ≥ 0 and by the Root-Factor Theorem has at most d roots
in the fraction field of R, hence a fortori at most d roots in R. But #R ≥ ℵ0 > d,
so there exists a1 ∈ R with f (a1 ) = 0.
Induction Step: Suppose n > 1 and that every polynomial in n − 1 vari-
ables with S-coefficients which is not the zero polynomial has a R-rational root.
Let f (t1 , . . . , tn−1 , z) ∈ S[t1 , . . . , tn−1 , z]. Put S 0 = S[t1 , . . . , tn−1 ], so f may
be identified with a nonzero polynomial g(z) ∈ S 0 [z]. Applying the Base Case,
there exists A ∈ R0 such that 0 6= g(A) ∈ R0 . Now g(A) is a nonzero element
of S 0 = S[t1 , . . . , tn−1 ], so by induction there exist a1 , . . . , an−1 ∈ R such that
g(A(a1 , . . . , an−1 )) 6= 0. Putting an = A(a1 , . . . , an−1 ) we have
f (a1 , . . . , an−1 , an ) = g(A(a1 , . . . , an−1 )) 6= 0.
5. THE NORMAL BASIS THEOREM 71
Proposition 8.22. Any finite cyclic extension K/F admits a normal basis.
Proof. Let K/F be cyclic of degree n with Gal(K/F ) = hαi. We may en-
dow K with the structure of an F [t]-module extending its F -module structure by
putting t · x = σ(x) for all x ∈ K. Then tn − 1 annihilates K; moreover, by linear
independence of characters, no smaller degree polynomial does so. It follows that
as an F [t]-module, K is isomorphic to F [t]/(tn − 1). Thus there exists α ∈ K
such that ann(α) = (tn − 1) – take, e.g., the preimage of 1 (mod tn − 1) under an
isomorphism – so the elements α, σα, σ 2 α, . . . , σ n−1 α are F -linearly independent
and thus give a normal basis.
Lemma 8.23. Let K/F be a degree n Galois extension, and write Aut(K/F ) =
{σi }ni=1 . For α1 , . . . , αn ∈ K, the following are equivalent:
(i) α1 , . . . , αn is an F -basis of K.
(ii) The matrix A ∈ Mn (K) with Aij = σi αj is nonsingular.
Proof. (i) =⇒ (ii) follows almost immediately from the (K-)linear indepen-
dence of the characters σ1 , . . . , σn : details are left to the reader.
(ii) =⇒ (i): We argue by contraposition: suppose α1 , . . . , αn is not an F -basis for
K, so there exist a1 , . . . , an ∈ F , not all zero, with a1 α1 + . . . + an αn = 0. Then
for all i we have
X n Xn Xn
aj Aij = aj σi αj = σi ( ai αj ) = 0,
j=1 j=1 j=1
which shows that the columns of the matrix A are linearly dependent.
By linear independence of characters, for any field extension K/F , any finite set
of automorphisms σ1 , . . . , σn ∈ Aut(K/F ) is K-linearly independent. If K/F is a
Galois extension and F is infinite, we have the following significantly stronger
independence result.
Theorem 8.24. Let K/F be a finite degree Galois extension of infinite fields.
Then the elements σ1 , . . . , σn of Aut(K/F ) are algebraically independent – if
0 6= f (t1 , . . . , tn ) ∈ K[t1 , . . . , tn ], there exists α ∈ K such that f (σ1 (α), . . . , σn (α)) 6=
0.
Proof. As a matter of notation, for an n-tuple (x1 , . . . , xn ) ∈ K n , we will
denote by (x1 , . . . , xn )• the corresponding column vector, i.e., element of Mn,1 (K).
If it brings no confusion, we will suppress indices by writing x• for (x1 , . . . , xn )• .
Let α1 , . . . , αn be a basis for K/F . Define A ∈ Mn (K) by Aij = σi αj . By
Lemma 8.23, A is nonsingular. Now let c = (c1 , . . . , cn ) ∈ F n and put
n
X
α= cj α j .
j=1
6. Hilbert’s Theorem 90
Let G be a group, and let M be a G-module, i.e., commutative group on which
G acts Z-linearly: that is, we are given a homomorphism G → AutZ (M ). Let
Z 1 (G, M ) be the set of all maps f : G → M which satisfy the cocycle condition:
∀σ, τ ∈ G, f (στ ) = f (σ) + σ(f (τ )).
1
Let B (G, M ) be the set of maps f : G → M such that there is a ∈ M with
f (σ) = σ(a) − a for all σ ∈ G.
Exercise 8.8. a) Show that Z 1 (G, M ) and B 1 (G, M ) are commutative groups
under pointwise addition.
b) Show that B 1 (G, M ) ⊂ Z 1 (G, M ).
We may therefore define
H 1 (G, M ) = Z 1 (G, M )/B 1 (G, M ),
the first cohomology group of G with coefficients in M.
Exercise 8.9. Suppose that G acts trivially on M . Show that H 1 (G, M ) =
Hom(G, M ), the group of all homomorphisms from G to M .
Now observe that if K/F is a field extension and G = Aut(K/F ), then both K (as
an additive group) and K × (as a multiplicative group) are G-modules.
6. HILBERT’S THEOREM 90 73
Theorem 8.26. Let K/F be a finite Galois extension, with Galois group G =
Aut(K/F ).
a) H 1 (G, K) = 0.
b) H 1 (G, K × ) = 0.
Proof. a) Let f : G → K be a 1-cocycle. Since K/F is finite separable, by
Theorem X.X there is c ∈ K with TrK/F (c) = 1. Put
X
b= f (σ)σ(c),
σ∈G
so
X
τ (b) = τ (f (σ))(τ σ)(c)
σ∈G
X X X
= (f (τ σ) − f (τ )) (τ σ)(c) = f (τ σ)(τ σ)(c) − f (τ )(τ σ)(c)
σ∈G σ∈G σ∈G
!
X
= b − f (τ ) · τ σ(c) = b − f (τ ).
σ∈G
Proof. Step 1: Because Galois conjugate elements have the same norm and
trace, in both parts a) and b) the implications (ii) =⇒ (i) are immediate.
Step 2: Let c ∈ K be such that TrK/F (c) = 0. Since K/F is separable, by Theorem
6.10 there is b ∈ K with TrK/F (b) = 1.3
Put
a = cb + (c + σ(c))σ(b) + . . . + (c + σ(c) + . . . + σ n−2 (c))σ n−2 (b).
Then
σ(a) = σ(c)σ(b) + (σ(c) + σ 2 (c))σ 2 (b) + . . . + (σ(c) + . . . + σ n−1 (c))σ n−1 (b).
Since TrK/F (c) = c + σ(c) + . . . + σ n−1 (c) = 0, we have
a − σ(a) = cb + cσ(b) + . . . + cσ n−1 b = c TrK/F (b) = c.
Step 3: Let c ∈ K be such that NK/F (c) = 1. By Dedekind’s linear independence
of characters, there is b ∈ K with
a = b + cσ(b) + cσ(c)σ 2 (b) + . . . + cσ(c) · · · σ n−2 (c)σ n−1 (b) 6= 0.
Then
cσ(a) = cσ(b) + cσ(c)σ 2 (b) + . . . + cσ(c) · · · σ n−1 (c)b = a,
so
a
c= .
σ(a)
We will use Theorem 8.28 later on in our study of cyclic extensions.
Solvable Extensions
1. Cyclotomic Extensions
1.0.1. Basics. Let K be a field. An element x ∈ K × is a root of unity if there
is n ∈ Z+ such that xn = 1; equivalently, x lies in the torsion subgroup of K × . We
put
µn (K) = {x ∈ K | xn = 1},
S
the nth roots of unity in K. We put µ(K) = n≥1 µn (K). Thus µn (K) and µ(K)
are subgroups of K × and µ(K) = K × [tors].
Lemma 9.1. For any field K and n ∈ Z+ , we have #µn (K) ≤ n.
Proof. The elements of µn (K) are the roots of the polynomial tn − 1 over K,
and a nonzero polynomial over a field cannot have more roots than its degree.
Lemma 9.2. For any field K and n ∈ Z+ , the group µn (K) is finite cyclic.
Proof. By Lemma 9.2, µn (K) is finite. We use the Cyclicity Criterion
[CL-NT, Thm. B.9]: a finite group G is cyclic iff for all d ∈ Z+ there are at most
d element of order n in G. This holds in µn (K) since the polynomial td − 1 can
have no more than d roots.
2πki
Example 9.3. Fix n ∈ Z. For 0 ≤ k < n, the elements e n are distinct nth
roots of unity in C. So #µn (C) = n.
Exercise 9.1. Let K be an ordered field. Show that µ(K) = {±1}.
An element of K × of exact order n is called a primitive nth root of unity.
Proposition 9.4. Let K be an algebraically closed field. For n ∈ Z+ , the
following are equivlaent:
(i) char K - n.
(ii) #µn (K) = n.
(iii) K admits a primitive nth root of unity.
(iv) K admits precisely ϕ(n) primitive nth roots of unity.
Proof. (i) ⇐⇒ (ii): Let f (t) = tn − 1. Then f 0 (t) = ntn−1 . Thus char K -
n ⇐⇒ gcd(f, f 0 ) = 1 ⇐⇒ tn − 1 has n distinct roots ⇐⇒ #µn (K) = n.
(ii) ⇐⇒ (iii): By Lemma 9.2, µn (K) is a finite, cyclic n-torsion abelian group.
Thus it has order n iff it has an element of order n.
(ii) =⇒ (iv): (ii) holds ⇐⇒ µn (K) is cyclic of order n, in which case it has
precisely ϕ(n) generators.
(iv) =⇒ (iii): Since for all n ∈ Z+ , ϕ(n) ≥ 1, this is clear.
77
78 9. SOLVABLE EXTENSIONS
Exercise 9.3. Show that for any field K, µ(K sep ) = µ(K).
Henceforth we only consider µn for char K - n.
For a field K, we denote by K cyc the field obtained by adjoining to K all roots
of unity in a fixed algebraic closure K. Then K cyc is the splitting field of the set
{tn − 1}char K-n of separable polynomials, so is an algebraic Galois extension, the
maximal cyclotomic extension of K. For n ∈ Z+ with char K - n, let K(µn )
be the splitting field of the separable polynomial tn − 1, the nth cyclotomic ex-
tension. Thus K cyc = limK(µn ).
−→
Proof. a) Both sides of (9) are monic polynomials with C coefficients whose
roots are precisely the nth roots of unity in C. So they are equal.
b) By strong induction on n. The base case is clear: Φ1 (t) = t − 1. Now sup-
Q n > 1 and that Φd (t) ∈ Z[t] for all proper divisors d ofn n. Then Q(t) =
pose
d|n, d6=n Φd (t) ∈ Z[t] is a monic polynomial and Q(t)Φn (t) = t − 1. Now imagine
actually performing polynomial long division of tn − 1 by Q(t) to get Φn (t): since
tn − 1, Φn (t) ∈ Z[t] are monic, the quotient Φn (t) has Z-coefficients.
c) This follows from part a) by the Möbius Inversion Formula applied in the com-
mutative group Q(t)× .2
Theorem 9.7. Let n ∈ Z+ and let K be a field of characteristic p. Regard
Φn (t) ∈ Fp [t] ⊂ K[t]. Then Φn (t) is a separable polynomial whose roots in K[t] are
precisely the primitive nth roots of unity.
Proof. By the Derivative Criterion tn − 1 ∈ K[t] is separable; by (8) so is
Φn (t). It is clear that the ϕ(n) roots of Φn (t) in K are nth roots of unity; that they
are the ϕ(n) primitive nth roots of unity follows by an easy induction argument.
Exercise 9.6. Let p be a prime number and a ∈ Z+ .
a) Show that Φp (t) = 1 + t + . . . + tp−1 .
b) Show that Φ2p (t) = 1 − t + . . . + (−t)p−1 .
a−1
c) Show that Φpa (t) = Φp (tp ).
Exercise 9.7. For n ∈ Z+ , let r(n) = p|n p. Show:
Q
n
Φn (t) = Φr(n) (t r(n) ).
Exercise 9.8. Let n ∈ Z+ .
a) Show: for all n ≥ 2, the constant coefficient of Φn (t) is 1.
b) Show: for all n 6= 2, the product of the primitive nth roots of unity in C is 1.
1The use of C here is somewhere between tradition and psychology: any algebraically closed
field of characteristic zero – e.g. Q – would serve as well.
2In fact we don’t need this in what follows – it is just a pretty formula.
80 9. SOLVABLE EXTENSIONS
hσi = n = [L : K] = # Aut(L/K).
There is an important converse to Proposition 9.15. To prove it, we need first the
following result, which despite its innocuous appearance is actually quite famous.
Lemma 9.16. Let K be a field, ζn ∈ K a primitive nth root of unity. Let
L/K be a cyclic extension of degree n, with generator σ. There is α ∈ L such that
ζn = σ(α)
α .
2. CYCLIC EXTENSIONS I: KUMMER THEORY 83
so α M
∈ K and thus a M
= αM n ∈ K ×n , so m | M . It follows that
m = M = # Aut(L/K) = [L : K].
c) Indeed, we just saw that we may take b = αm .
84 9. SOLVABLE EXTENSIONS
3. The equation tn − a = 0
In this section we analyze the structure of the splitting field of a polynomial tn −a =
0 without assuming that the ground field contains a primitive nth root of unity.
We closely follow [LaFT, §VI.9].
Lemma 9.20. Let F be a field of characteristic p > 0 and a ∈ F × \ F ×p . Then
n
for all n ≥ 1, the polynomial tp − a is irreducible.
Proof. We shall prove the contrapositive: suppose that for some n ∈ Z+ the
n
polynomial tp − α is reducible; we will show that α is a pth power in F . We may
n
write tp − α = f (t)g(t), where f (t) and g(t) are nonconstant monic polynomials.
n
Let K/F be an extension field containing a root β of tp − α. Then in K[t] we have
n n n n
tp − α = tp − β p = (t − β)p .
Since K[t] is a UFD and f (t) and g(t) are monic, we therefore have f (t) = (t − β)r
for some 0 < r < pn . Write r = pm s with gcd(p, s) = 1. Note that m < n. Then
m m
f (t) = (tp − β p )s ,
m m
so that the coefficient of tp (s−1) is −sβ p . This lies in F and – since s 6= 0 in F
m
– we conclude β p ∈ F . Thus
m n−m n−m
α = (β p )p ∈ Fp ∈ Fp
3. THE EQUATION tn − a = 0 85
since m < n.
Theorem 9.21. Let n ≥ 2, let F be a field, and let a ∈ F × . We suppose:
• For all prime numbers p | n, we have a ∈ / F p , and
4
• If 4 | n, then a ∈
/ −4F .
Then f (t) := tn − a is irreducible in F [t].
Proof. We begin by establishing several special cases.
Step 1: Suppose n = pe is a prime power, a ∈ F \ F p and p is the characteristic of
F . This case is covered by Lemma 9.20.
Step 2: Suppose n = pe is a prime power, a ∈ F \ F p and p is not the characteristic
of F . First we claim that tp − a is irreducible. Otherwise, there is some root α ∈ F
of tp − a such that [F (α) : F ] = d < p. Let N denote the norm map from F (α) to
F : since αp = a, we have
N (α)p = N (a) = ad .
Since gcd(d, p) = 1, there are x, y ∈ Z such that xd + yp = 1, and thus
a = axd aup = (N (α)x au )p ∈ F,
contradiction. Now write
p
Y
tp − a = (t − αi ),
i=1
with α1 , . . . , αp ∈ F and α1 = α. We may thus also write
p
e Y e−1
tp − a = (tp − αi ).
i=1
e−1
Suppose first that α ∈/ F (α)p . Let A be root of tp − α. If p is odd, then by
induction A has degree pe−1 over F (α) and thus degree pe over F and it follows
e
that tp − a is irreducible. If p = 2, suppose α = −4β 4 for some β ∈ F (α). Again
let N be the norm from F (α) to F . Then
−a = N (α) = 16N (β)4 ,
√ √
so −a ∈ F 2 . Since p = 2 it follows that −1 ∈ F (α) but α = ( −12β 2 )2 , a
contradiction. By induction, A has degree pe over F . So we may assume that there
is β ∈ F (α) such that β p = α. . . .
The following is an immediate consequence.
Corollary 9.22. Let p be a prime number, F a field, and a ∈ F \ F p . If p is
either odd or equal to the characteristic of F , then for all n ∈ Z+ the polynomial
n
tp − a is irreducible in F [t].
Let F be a field. Let n be a positive integer that is not divisible by the characteristic
of F , let a ∈ F × , and let K be the splitting field of the separable polynomial p(t) =
tn −a. We address the following question: what is the Galois group G := Aut(K/F )?
Let α be a root of p(t) in K, so K = (α, ζn ). Then an element σ ∈ G is determined
by its action on α and ζn , and we have
σ(α) = ζ b(σ) α, b(σ) ∈ Z/nZ,
σ(ζn ) = ζnd(σ) , d(σ) ∈ (Z/nZ)× .
86 9. SOLVABLE EXTENSIONS
Exercise 9.13.
√ a) Let f (t) = x8 − 2 ∈ Q[t]. Show: the splitting field of
8
f is Q( 2, ζ4 ).
b) Observe that f satisfies all of the hypotheses of Theorem 9.24 except that
8 is not odd, and that the conclusion does not hold: [K : F ] = nϕ(n)
2 , not
nϕ(n).
We remark that the essential content content of Theorem 9.24 lies in the assertion
that (under the hypotheses and notation used therein) F (ζn ) ∩ F (α) = F , since
by Natural Irrationalities this implies that [F (ζn , α) : F (ζ)] = n. It is also natural
to think in terms of linear disjointness (cf. §12): because F (ζn )/F is Galois, the
identity F (ζn ) ∩ F (α) = F holds iff F (ζn ) and F (α) are linearly disjoint over F .
Since this holds iff
F (ζn ) ⊗F F (α) = F (ζn )[t]/(tn − a)
is a field, another equivalent condition is that the polynomial tn − a remains irre-
ducible over F (ζn ). In the situation of the above exercise we have
√
8
√
Q(ζ8 ) ∩ Q( 2) = Q( 2)
88 9. SOLVABLE EXTENSIONS
and thus the polynomial t8 − 2, which is irreducible over Q, becomes reducible over
Q(ζ8 ): indeed we have √ √
t8 − 2 = (t4 − 2)(t4 + 2).
91
CHAPTER 11
Many find this terminology confusing at first, because it is so close to the ter-
minology “algebraically closed” and yet means something quite different. There is
however some connection:
Exercise 11.1. For a field K, show that the following are equivalent:
(i) For every field extension L/K we have that K is algebraically closed in L.
93
94 11. STRUCTURE OF TRANSCENDENTAL EXTENSIONS
Two matrices determine the same automorphism iff they are scalar multiples of
each other, so if we define PGL2 (K) as the quotient
of GL2(K) modulo the (central,
a 0
hence normal) subgroup K × of scalar matrices { | a ∈ K × } then we get
0 a
a canonical bijection from Aut(K(t)/K) to PGL2 (K). In the following theorem we
summarize the above discussion and add one more piece.
Theorem 11.3. The map Aut(K(t)/K) → PGL2 (K) given by
at + b a b
s ∈ Aut(K(t)/K) 7→ s(t) = 7→ (mod K × )
ct + d c d
is an isomorphism of groups.
Exercise 11.2. Complete the proof of Theorem 11.3 by showing that the above
bijection from Aut(K(t)/K) to PGL2 (K) is an isomorphism of groups.
Exercise 11.3. Let K be a field.
a) Suppose that K ∼
= Fq . Show: # Aut(K(t)/K) = (q + 1)(q 2 − q).
b) Suppose that K is infinite. Show: # Aut(K(t)/K) = #K.
1.3. Lüroth’s Theorem.
Theorem 11.4 (Lüroth’s Theorem). Let K be a field, and let L be a field such
that K ( L ⊂ K(t). Then there is f ∈ K(t) such that L = K(f ).
Proof. Let v ∈ L \ K. By Theorem 11.1 we know that K(t)/K(v) has finite
degree, so [K(t) : L] is finite. Let
f (x) := xn + `n−1 xn−1 + . . . + `1 x + `0 ∈ L[x]
be the minimal polynomial of t over L. Since t is transcendental over K, there is
0 ≤ j ≤ n − 1 such that lj lies in L \ K: we fix one such index and call it J. Put
u := `J ;
we will in fact show that L = K(u). Put
m := [K(t) : K(u)].
Since K(u) ⊂ L we have m ≥ n, so to show L = K(u) it suffices to show n ≥ m.
We may scale the coefficients of f so as to get a primitive polynomial in K[t][x]:
choose c0 (t), . . . , cn−1 (t), d(t) ∈ K[t] such that gcd(c0 , . . . , cn−1 , d) = 1 and for all
c
0 ≤ j ≤ n − 1 we have `j = dj . In particular we have
cJ
u = `J = .
d
Put
M := max(deg(cJ ), deg(d)).
By Theorem 11.1 we have
m = [K(t) : K(u)] ≤ M.
The inequality is because of the possibility that cJ and d are not coprime. Put
F (x, t) := d(t)f (x) = d(t)xn + cn−1 (T )xn−1 + . . . + c0 (T ) ∈ K[x, t].
96 11. STRUCTURE OF TRANSCENDENTAL EXTENSIONS
These higher dimensional counterexamples to the Lüroth Problem are far beyond
our means in these notes. Indeed the methods of pure field theory are not even
1When n = 1 it follows from Theorem 11.1 that for any L with K ( L ⊂ K(t), then K(t)/L
must have finite degree. This is clearly false when n ≥ 2, hence we impose that hypothesis
explicitly.
2. TRANSCENDENCE BASES AND TRANSCENDENCE DEGREE 97
well equipped to give an example of a finite degree field extension K/C(t) that is
not isomorphic to C(t), although such examples are given by meromorphic function
fields of any compact Riemann surface of genus g ≥ 1. The situation is highly
analogous (and actually more than analogous!) to what one encounters in general
topology, where the methods developed are “too general” to provably exhibit a
compact, connected topological surface that it not homeomorphic to the 2-sphere–
even though the torus presents itself as a very plausible candidate, one needs the
methods of a different subject, algebraic topology, to confirm this.
Theorems 11.1 and 11.4 give a good picture of the subfield lattice of K(t)/K for any
field K. Namely, this lattice is Noetherian – for any subextension K ( L ⊂ K(t)
we have [K(t) : L] is finite, so there are no infinite ascending chains. On the other
hand, the lattice is not Artinian, as there are infinite descending chains, e.g.
n
K(t) ) K(t2 ) ) K(t4 ) ) . . . ) K(t2 ) ) . . . .
This observation can be generalized.
Exercise 11.4. Let L/K be a transcendental field extension. Show: the subfield
lattice of L/K is not Artinian.
It turns out that the subfield lattice of a field extension L/K is Noetherian iff L/K
is finitely generated: this is Exercise 11.11, and at the point at which it is assigned
it will be more clear how to show it.
Exercise 11.5. Show that there is an algebraic extension L/K that is not
finitely generated (equivalently, is of infinite degree) such that the subfield lattice is
Artinian.
Exercise 11.6. Combining Exercise 11.3 with Lüroth’s Theorem (Theorem
11.4) we deduce the following intriguing consequence: for a finite field Fq , there is
a function f ∈ Fq (t) such that
Fq (t)Aut(Fq (t)/Fq ) = Fq (f )
and Fq (t)/Fq (f ) is finite degree Galois with automorphism group PGL2 (Fq ). Can
you explicitly write down this rational function f ?2
Step 1: We must check the result for Fp and Q. In the former case we have identified
the automorphism group as Ẑ, which indeed has cardinality c = 2ℵ0 = 2#Fp . In the
latter case we can by no means “identify” Aut(Q), but to see that it has continuum
cardinality it suffices, by the automorphism extension theorem, to exhibit a simpler
Galois extension K/Q which has continuum cardinality. Indeed one can take K to
be quadratic closure of Q, i.e., the compositum of all quadratic field extensions of
100 11. STRUCTURE OF TRANSCENDENTAL EXTENSIONS
analogy in the back of our mind in case we need it later? Depending on taste, this
is a reasonable approach to take, perhaps more reasonable for the mind which is
able to quickly remember what it once knew. As for myself, I would at the same
time worry that it would take me some time and energy to recreate the analogy if
I hadn’t written it down, and I would also be curious whether A and B might be
common instances of a more general construction that it might be interesting or
useful to know explicitly. So we shall follow the second course here, with apologies
to those with different tastes.
(A) Maximal independent sets exist, and every independent set is contained in
some maximal independent set.
Could it be that (LI1) through (LI3) imply the following desirable property?
Unfortunately this is not the case. Suppose we have a set X which is partitioned
into disjoint subsets:
a
X= X.
i
A set X equipped with a family of subsets {Si } satisfying axioms (LI1) through
(LI4) is called an independence space.
102 11. STRUCTURE OF TRANSCENDENTAL EXTENSIONS
In an independence space, if S1 and S2 are independent sets with #S1 < #S2 ,
then S1 is non-maximal. Therefore a maximal independent set has cardinality at
least as large as any other independent set, so by symmetry all maximal indepen-
dent sets have the same cardinality: independence spaces satisfy (B). Conversely,
(LI1) through (LI3) and (B) clearly imply (LI4).
(SO1) = (CL1) S ⊂ S
(SO2) = (CL2) S ⊂ S 0 =⇒ S ⊂ S 0
(SO3) = (CL3) S = S.
But the linear span satisfies two other properties, the first of which is not sur-
prising in view of what has come before:
Note that it is immediate to show that the independent sets for a spanning op-
erator satisfy (LI1) through (LI3). In particular, we have (A), that bases exist and
any independent set is contained in a basis. Again it is not obvious that (LI4) is
satisfied. Rather we will show (B) directly – which is what we really want anyway
– and by the above remarks that implies (LI4).
Case 2: We may now suppose that B and B 0 are both infinite. For every x ∈ X,
we claim the existence of a subset Ex with the property that x ∈ Ex and for any
subset E of B such that x ∈ E, Ex ⊂ E. Assuming S the claim for the moment,
we complete the proof. Consider the subset S = x∈B 0 Ex of B. Since each Ex
is finite, #S ≤ #B 0 . On the other hand, for all x ∈ B 0 , x ∈ Ex ⊂ S, so B 0 ⊂ S
and therefore S ⊃ B 0 = X. Therefore S is a spanning subset of the basis B, so
S = B and thus #B ≤ #B 0 . By reversing the roles of B and B 0 in the argument
we conclude #B = #B 0 .
It remains to prove the claim on the existence of Ex . In turn we claim that if
E 0 and E 00 are two subsets of B such that x ∈ E 0 ∩ E 00 and x is not in the span
of any proper subset of E 0 , then E 0 ⊂ E 00 ; this certainly suffices. Assuming to the
contrary that there exists y ∈ E 0 \ E 00 . Then x is not in the span of E 0 \ y and is in
the span of (E 0 \ y) ∪ y, so by (SO5) y is in the span of (E 0 \ y) ∪ x. Since x is in
the span of E 00 , we get that y is in the span of (E 0 \ y) ∪ E 00 . But this contradicts
the fact that the (E 0 \ y)cupE 00 ∪ {y}, being a subset of B, is independent.
For future reference, for a field extension L/K, we will refer to the matroid with
sets the subsets of L, spanning operator S 7→ S the algebraic closure of K(S) in
L and (it follows) with independent sets the algebraically independent subsets the
transcendence matroid of L/K.
For finite matroids, combinatorialists know at least half a dozen other equivalent
axiomatic systems: e.g. in terms of graphs, circuits, “flat” subspaces and pro-
jective geometry. As above, demonstrating the equivalence of any two of these
systems is not as easy as one might expect. This phenomenon of multiple nonob-
viously equivalent axiomatizations has been referred to, especially by G. Rota, as
5. MORE ON TRANSCENDENCE DEGREES 105
6. Digging Holes
Lemma 11.24. Let K/k be an extension of commutative rings, and let S be a
subset of K that is disjoint from k. Then there is a ring l such that k ⊂ l ⊂ K, l
is disjoint from S and l is maximal with respect to these properties.
Exercise 11.12. Prove Lemma 11.24. (Use Zorn’s Lemma!)
In this section we are interested in the case of Lemma 11.24 in which K/k is a a field
extension, and K is algebraically closed, so henceforth we impose these hypotheses.
Proposition 11.25 (Quigley-McCarthy [Qu62] [Mc67]). Let K/k be an ex-
tension of fields with K algebraically closed, let S ⊂ K be a finite subset that is
disjoint from S, and let l be a subextension of K/k that is maximal with resepect
to the exclusion of S. Then K/l is an algebraic field extension.
Proof. Seeking a contradiction, let t ∈ K be transcendental over l. By our
hypothesis on l, there is an element a1 ∈ l(t) ∩ S. The element a1 is transcendental
over l – otherwise the rational function field l(t) contains a nontrivial finite degree
field extension l(a1 )/l, and tensoring the inclusion l(a1 ) ,→ l(t) with l(a1 ) gives
a contradiction. It follows that l ( l(a21 ) and a2 ∈ / l(a21 ), so l(a21 ) contains some
6. DIGGING HOLES 107
Linear Disjointness
By Exercise 12.3 this forces αij = 0 for all i, j and thus βj = 0 for all k, so the ki ’s
are L-linearly independent.
(i) =⇒ (ii0 ): The above proof works with the roles of K and L reversed.
(ii) =⇒ (i): By Exercise 12.3, it is enough to fix m, n ∈ Z+ let k1 , . . . , km
be F -linearly independent elements of K and l1 , . . . , ln be F -linearly independent
elements of L and show that {ki lj P } are F -linearly independent elements of E.
Suppose that αij ∈ F are such that i,j αij ki lj = 0. But we may rewrite this as
(α11 l1 + . . . + α1n ln )k1 + . . . + (αm1 l1 + . . . + αmn ln )km = 0.
By hypothesis the ki ’s are L-linearly independent, so this forces all the coefficients
of the above equation to be equal to zero, which in turn, since the lj ’s are F -linearly
1. DEFINITION AND FIRST PROPERTIES 111
K[L] are linearly disjoint over K, which by Lemma 12.3 implies that M and KL
are linearly disjoint over F .
(ii) =⇒ (i): If K and L are linearly disjoint over F , then ϕ1 is injective. If M and
KL are linearly disjoint over K then by Exercise 12.5 we have that M and K[L]
are linearly disjoint over K, so ϕ2 is injective. Therefore ϕ is injective, so M and
L are linearly disjoint over F .
Proof. (i) =⇒ (ii): In order to show that the (evidently nonzero, since it
contains F ) ring R = K ⊗F L is a field, it suffices to show that the only maximal
ideal is (0). So let m be a maximal ideal of R. Then E = R/m is a field extension
of K, L and the induced map K ⊗F L → E is precisely the quotient map R → R/m.
Since this map is injective, m = (0).
(ii) =⇒ (i): If R = K ⊗F L is a field, then every homomorphism into a nonzero
ring – and in particular, any F -algebra homomorphism – is injective.
Theorem 12.8. Let K, L be field extensions of F .
a) Suppose that K, L are everywhere F -linearly disjoint. Then at least one
of K, L is algebraic over F .
b) Conversely, suppose that at least one of K, L is algebraic over F . Then
K, L are somewhere F -linearly disjoint iff they are everywhere F -linearly
disjoint.
Proof. a) If K and L are transcendental over F , then they admit subexten-
sions K 0 = F (a), L0 = F (b). By Exercise 12.5, it suffices to show that F (a) and
F (b) are not everywhere F -linearly disjoint over F . To see this take E = F (t) and
map K 0 → E by a 7→ t and L0 → E by b 7→ t and apply Lemma 12.2.
b) Because every algebraic extension is a direct limit of finite degree extensions, by
Exercise 12.4 it is no loss of generality to assume that K/F is finite, and in light
of Propositions 12.6 and 12.7, we must show that if K ⊗F L is a domain then it is
a field. But if {k1 , . . . , kn } is a basis for K/F , then k1 ⊗ 1, . . . , kn ⊗ 1 is a basis for
K ⊗F L over L, so K ⊗F L is a domain and a finitely generated L-module. Therefore
it is a field, by an elementary argument which we have seen before (and which is a
special case of the preservation of Krull dimension in an integral extension).
In conclusion: the notion of F -linear disjointness of two field extensions K, L is in-
trinsic – independent of the embeddings into E – iff at least one of K, L is algebraic
over F . In most of our applications of linear disjointness this hypothesis will be
satisfied, and when it is we may safely omit mention of the ambient field E.
Theorem 12.12. Let E/F be a field extension, and let K, L be two algebraic
subextensions such that K/F is Galois. Then K, L are F -linearly disjoint (in E,
but by Theorem 12.8 this does not matter) iff K ∩ L = F .
Proof. By a now familiar argument involving Exercise 12.5 and Proposition
12.7, we reduce to the case in which K/F is finite Galois. Now by the theorem of
Natural Irrationalities, we have
[KL : F ] = [KL : L][L : F ] = [K : K ∩ L][L : F ],
so
[KL : F ] = [K : F ][L : F ] ⇐⇒ K ∩ L = F.
It is clear that separating transcendence bases need not exist, e.g. an insepara-
ble algebraic extension will not admit a separating transcendence basis. On the
other hand, it is clear that separable algebraic extensions and purely transcenden-
tal extensions both admit separating transcendence bases: as with being linearly
disjoint from the perfect closure, this is something that these apparently very dif-
ferent classes of extensions have in common.
116 12. LINEAR DISJOINTNESS
b) Conclude that separably generated extensions – and thus also separable extensions
– do not form a distinguished class in the sense of Lang.
c) Prove or disprove: the compositum of two separably generated extensions is sep-
arably generated.
d) Prove or disprove: the compositum of two separable extensions is separable.
5. Interlude
For later use we record the following result, whose proof will use methods of
commutative algebra rather than field theory.1
Theorem 12.19 (Integrality of Products). Let k be an algebraically closed field,
and for i = 1, 2, let Ri be a domain that is finitely generated as a k-algebra. Then
R1 ⊗k R2 is a domain.
Proof. The statement holds trivially if either R1 = k or R2 = k, so assume
that each of R1 and R2 have positive transcendence degree over k. Seeking a
contradiction, suppose that we have nonzero elements
m
X n
X
x= ai ⊗ bi , y = cj ⊗ dj ∈ R1 ⊗k R2
i=1 j=1
which exhibits a product of two nonzero elements in the domain R2 being zero, a
contradiction.
Although our motivation for proving Theorem 12.19 is field theoretic – it will be
used in the proof of Theorem 12.22 giving an important equivalent condition on a
class of field extensions – for those with some familiarity with algebraic geometry
it is hard not to notice that it is saying precisely that over an algebraically closed
field, the product of two integral affine varieties remains integral.
Exercise 12.14. Show that if the conclusion of Theorem 12.19 holds for a field
k, then k is algebraically closed.
6. Regular Extensions
A field extension K/F is regular if for all field extensions L/F , the F -algebra
K ⊗F L is a domain.
Lemma 12.20. If K/F is regular, then for all algebraic extensions L/F , we
have that K ⊗F L is a field.
Proof. We have by definition that K ⊗F L is a domain, so by Proposition
12.6 we have that K and L are somewhere linearly disjoint over F . Since L/F is
algebraic, Theorem 12.8 gives that K and L are everwhere linearly disjoint over F ,
and then Proposition 12.7 gives that K ⊗F L is a field.
Theorem 12.22. For a field extension K/F , the following are equivalent:
(i) We have that K/F is regular.
(ii) For every algebraic extension L/F , we have that K ⊗F L is a field.
(iii) For any algebraic closure F of F , we have that K ⊗F F is a field.
(iv) We have that F is algebraically closed in K and K/F is separable.
Proof. (i) =⇒ (ii): This follows from Lemma 12.20.
(ii) =⇒ (iii) is immediate.
(iii) =⇒ (ii): Let L/F be an algebraic field extension. We have by assumption
that K ⊗F F = (K ⊗F L)⊗L F is a field, so by Lemma 12.21 we deduce that K ⊗F L
is a field.
(ii) =⇒ (iv): Let L be the algebraic closure of F in K, so by assumption we have
that K ⊗F L is a field. Thus the F -subalgebra L ⊗F L is a domain, and since L/F
is algebraic this implies L = F . The separability is automatic in characteristic 0
−1
and in characteristic p we have that K ⊗F F p is a field, so K/F is separable by
Theorem 12.15.
(iv) =⇒ (iii): Let F sep be the maximal separable subextension of F /F . Then
F sep /F is Galois and F sep ∩K = F , so K ⊗F F sep is a field. Since K/F is separable,
K ⊗F F sep /F sep is separable (indeed one can choose the same separating transcen-
dence basis), hence K is F -linearly disjoint from the maximal purely inseparable
extension of F sep , which is F .
(iii) =⇒ (i): We must show that K ⊗F L is a domain for every field extension
L/F . For this it is harmless to replace L with a larger field extension, so suppose
that L is algebraically closed, and let F be the algebraic closure of F in L (which is
indeed algebraically closed!). By assumption, we have that K ⊗F F is a field, and
since
K ⊗F L = f (K ⊗F F ) ⊗F L,
120 12. LINEAR DISJOINTNESS
1. Derivations
1.1. Definitions and First Results. Let R be a commutative ring, and let
M be an R-module. A derivation of R into M is a map D : R → M satisfying
both of the following:
Proposition 13.3. Let L/K be a field extension, and let D ∈ DerK (L). Let
f ∈ K[t1 , . . . , tn ] and a = (a1 , . . . , an ) ∈ Ln . Then
n
X
D(f (a)) = ∂i f (a1 , . . . , an )D(ai ).
i=1
Proposition 13.4. Let L/K be a field extension, and let S ⊂ L be such that
L = K(S).
a) We have
dimL DerSK (L) ≤ #S.
b) In particular if L can be generated as a field extension by n < ∞ elements, then
dimL DerK (L) ≤ n.
Proof. Let L(S) be the set of all finitely nonzero functions from S to L. This
is an L-vector space with basis canonically in bijection with S: indeed, for s ∈ S,
let δs be the function which takes the value 1 at s and zero elsewhere. Then {δs }s∈S
is an L-basis for L(S) .
The natural restriction map DerSK (L) → L(S) is L-linear and injective. The L-
linearity is a triviality: the injectivity follows from the fact that every element
of L is a rational function in the elements of S with coefficients in K. Since
dim L(S) = #S, part a) follows immediately! Part b) is also immediate from the
observation that S-finiteness is a vacuous condition when S itself is a finite set.
124 13. DERIVATIONS AND DIFFERENTIALS
n
! n
!
X X
fjD1 (x1 , . . . , xn ) + D
(∂i fj1 (x1 , . . . , xn )ui + fj2 (x1 , . . . , xn ) + (∂i fj2 (x1 , . . . , xn )ui
i=1 i=1
= 0 + 0 = 0.
For j ∈ J and g ∈ K[t1 , . . . , tn ], we have
n
X
D
(gfj ) (x1 , . . . , xn ) + (∂i (gfj ))(x1 , . . . , xn )ui
i=1
n
!
X
= g(x1 , . . . , xn ) fjD (x1 , . . . , xn ) + (∂i fj )(x1 , . . . , xn )ui
i=1
n
!
X
D
+fj (x1 , . . . , xn ) g (x1 , . . . , xn ) + (∂i g)(x1 , . . . , xn )ui
i=1
n
!
X
D
= g(x1 , . . . , xn ) · 0 + 0 · g (x1 , . . . , xn ) + (∂i g)(x1 , . . . , xn )ui = 0.
i=1
This shows that (16) holds for all f ∈ I(x1 , . . . , xn ). Now we define D on K[x1 , . . . , xn ]
by putting, for g ∈ K[t1 , . . . , tn ],
n
X
D(g(x1 , . . . , xn )) := g D (x1 , . . . , xn ) + (∂i g)(x1 , . . . , xn )ui .
i=1
The point here is that the same element of K[x1 , . . . , xn ] can in general by expressed
as g(x1 , . . . , xn ) for several polynomials g ∈ K[t1 , . . . , tn ], but any two such polyno-
mials differ by an element of I(x1 , . . . , xn ), so our calculation just above has checked
that if g1 (x1 , . . . , xn ) = g2 (x1 , . . . , xn ) then D(g1 (x1 , . . . , xn )) = D(g2 (x1 , . . . , xn )).
The fact that D is a derivation now follows from the fact that g 7→ g D and g 7→ ∂i g
are derivations. Thus we have succeeded in extending D to a derivation D of
K[x1 , . . . , xn ]. By Theorem 13.2a), there is a unique extension of D to a derivation
on K(x1 , . . . , xn ). Finally, for 1 ≤ i ≤ n, taking g = ti gives
n
X
D(xi ) = D(g(x1 , . . . , xn )) = g D (x1 , . . . , xn ) + (∂i g)(x1 , . . . , xn )ui
i=1
= 0 + 0 · u1 + . . . + 0 · ui−1 + 1 · ui + 0 · ui+1 + . . . + 0 · un = ui .
1.3. The structure of DerK (L) for a finitely generated extension L/K.
Let L/K be any finitely generated field extension. In this section we will apply
Theorem 13.5 to give a good description of the L-vector space DerK (L). In par-
ticular its dimension is an interesting invariant of the extension L/K and will be
computed. What we know so far is that, by Proposition 13.4, dimL DerK (L) is
bounded above by the minimal number of generators of L/K, and (as we are about
to nail down explicitly) it follows easily from Exercise 13.5 that equality holds if
L/K is purely transcendental. It turns ou that in characteristic 0 the dimension
is always equal to trdeg(L/K) whereas in positive characteristic separability issues
intervene to make things more interesting.
K(t1 , . . . , tn ) by freely choosing the values of D(t1 ), . . . , D(tn ). We have met spe-
cial cases of this result before: Exercise 13.7 treats the case D(ti ) = 0 for all i and
Exercise 13.5 fixes 1 ≤ j ≤ n and treats the case D(ti ) = δ(i, j) (Kronecker δ).
This gives part a) of the following result, and part b) will follow easily.
Corollary 13.6. Let L = K(t1 , . . . , tn ) be a purely transcendental extension.
a) Let D ∈ Der K. For each 1 ≤ i ≤ n, there is a unique Di ∈ Der L that
extends D and such that for all 1 ≤ i ≤ n we have
(
1 i=j
Di (tj ) = .
0 otherwise
Example 13.8. Let K have characteristic p > 0, and let L/K be a monogenic
purely inseparable extension, so there is a ∈ K, x ∈ L and n ∈ Z+ such that L =
n n
K(x) and xp = a. Let D ∈ Der(K). The ideal I(x) is generated by f (t) = tp − a.
n
We have f D (x) = D(1)tp − D(a) = −D(a), so by Theorem 13.5 there is a unique
extension of D to a derivation on L with D(x) = u for each u ∈ L such that
0 = f D (x) + f 0 (x)u = −D(a) + 0 · u = −D(a).
Thus: if D(a) 6= 0, there is no extension of D to L. If D(a) = 0, then such
extensions correspond to elements of L. In particular, we have dimL DerK L = 1:
a basis is given by {Dx }, where Dx is the unique K-derivation with Dx (x) = 1.
Exercise 13.9. Let L = K(x) be a monogenic inseparable algebraic extension.
Show that once again dimL DerK L = 1.
Exercise 13.10. Let k be a field of characteristic p > 0, let x1 , . . . , xn be inde-
pendent indeterminates over k, let L := k(x1 , . . . , xn ) and let K := k(xp1 , . . . , xpn ).
a) Show: [L : K] = pn .
(Suggestion: use the tower k(xp1 , . . . , xpn ) ( k(x1 , xp2 , . . . , xpn ) ( . . . (
k(x1 , . . . , xn ).)
b) Let I = I(x1 , . . . , xn ) be the ideal of all f ∈ K[t1 , . . . , tn ] such that
f (x1 , . . . , xn ) = 0. Show that K[t1 , . . . , tn ]/I is isomorphic to L and thus
#K[t1 , . . . , tn ]/I = pn .
c) Let J be the ideal of K[t1 , . . . , tn ] generated by tp1 − xp1 , . . . , tpn − Lxpn . Show
that # dim K[t1 , . . . , tn ]/J ≤ pn . Deduce that I = J.
d) Let S = {x1 , . . . , xn }. Use Theorem 13.5 to show that the restriction-to-S
map r : DerK (L) → LS is an L-vector space isomorphism.
e) Deduce that the minimal number of generators for the degree pn field ex-
tension L/K is n.
f) Let s1 , . . . , sm be independent indeterminates over L, and put
M := L(s1 , . . . , sm ) = K(x1 , . . . , xn , s1 , . . . , sm ).
Let T := {x1 , . . . , xn , s1 , . . . , sm }. Use Theorem 13.5 to show that the
restriction-to-T map r : DerK (M ) → T M is an M -vector space isomor-
phism. Deduce the that the minimal number of generators of the field
extension M/K is m + n.
Taking m = 1 in Exercise 13.10 answers a question I raised in my Fall 2020 course
on algebraic function fields: whereas any separable finitely generated field extension
of transcendence degree 1 can be generated by 2 elements, for all n ∈ Z+ there is
a finitely generated field extension M/K of transcendence degree 1 in which the
minimal number of generators is 1. As of now I still wonder whether such examples
exist with K algebraically closed in M .
Lemma 13.9. Let K be a field of characteristic p > 0, and let L be a field with
K ⊂ L ⊂ K 1/p . If S is a set of generators for L/K and D ∈ Der(K) is such that
D(xp ) = 0 for all x ∈ S, then D can be extended to a derivation on L.
Proof. Consider the set S of pairs (M, DM ) such that M is a field with
K ⊂ M ⊂ L and DM ∈ Der(M ) is such that DM |K = D. On S we define the
relation (M, DM ) ≺ (M 0 , DM 0 ) if M ⊂ M 0 and DM 0 |M = DM . This makes S into
a partially ordered set in which the union over any chain is an upper bound, so by
128 13. DERIVATIONS AND DIFFERENTIALS
Zorn’s Lemma the set S has a maximal element (M, DM ). If M 6= L, then there
is x ∈ S \ M . Then xp ∈ K and DM (xp ) = D(xp ) = 0, so by Example 13.8 we
may extend DM to a derivation DM (x) on M (x), contradicting the maximality of
(M, DM ). Thus M = L and DM is the desired extension of D to L.
Corollary 13.10. Let K be a field of characteristic p > 0, and let L/K be a
finite degree extension such that K ⊂ L ⊂ K 1/p . Put [L : K] = ps . Then:
a) We have dimL DerK (L) = s.
b) There are elements x1 , . . . , xs ∈ L and D1 , . . . , Ds ∈ DerK (L) such that
L = K(x1 , . . . , xs ), for all 1 ≤ i ≤ n we have xi ∈ / K(x1 , . . . , xi−1 ) and
(
1 i=j
∀1 ≤ i, j ≤ s, Di (xj ) = .
0 i 6= j
Proof. Of course the result holds trivially if L = K, so we may assume
that K ( L and choose x1 ∈ L \ K. Then xp1 ∈ K, so [K(x1 ) : K] = p.
If K(x1 ) ( L then we may choose x2 ∈ L \ K(x1 ) and once again we have
xp2 ∈ K ⊂ K(x1 ), so [K(x1 , x2 ) : K(x1 )] = p and thus [K(x1 , x2 ) : K] = p2 .
Since [L : K] = ps , this process terminates precisely after we have chosen x1 , . . . , xs
such that [K(x1 , . . . , xi ) : K(x1 , . . . , xi−1 )] = p for all 1 ≤ i ≤ s.
It follows from Example 13.12 that for any K ⊂ M ⊂ L, if D ∈ DerK (M )
and xi ∈ / M , then for all α ∈ M (xi ) there is a unique extension of D to a
derivation on M (xi ) such that D(xi ) = α. Fix 1 ≤ j ≤ s. Applying this
observation with M = K(x1 , . . . , xj−1 ) and D = 0, we get a unique derivation
Dj0 ∈ DerK(x1 ,...,xj−1 ) (K(x1 , . . . , xj )) with Dj0 (xj ) = 1 and then applying the ob-
servation again several more times we get a unique extension of Dj0 to Dj ∈ Der(L)
such that Dj (xj+1 ) = . . . = Dj (xs ) = 0. These elements D1 , . . . , Ds ∈ DerK (L)
indeed have the desired property that Di (xj ) = δ(i, j) (Kronecker delta) for all
1 ≤ i, j ≤ s, which establishes part b).
Upon restriction to S = {x1 , . . . , xj }, the derivations D1 , . . . , Ds yield the stan-
dard basis for the L-vector space LS of all functions from S to L, so they are cer-
tainly linearly dependent as functions from L to L, which shows that dimL DerK (L) ≥
s. Because L = K(x1 , . . . , xs ) we have dimL DerK (L) ≤ s, and thus we have
dimL DerK (L) = s, establishing part a).
Proposition 13.11. For a finitely generated field extension L/K, the following
are equivalent:
(i) The extension L/K is separable algebraic.
(ii) We have DerK (L) = 0.
Proof. (i) =⇒ (ii): This is Corollary 13.7.
(ii) =⇒ (i): If L/K is not separably generated (equivalently, is not separable),
then there is a subextension F1 of L/K such that L/F1 is finite degree inseparable.
By decomposing L/F1 is purely inseparable over separable, one sees that there is
a subextension M of L/F1 such that L/M is purely inseparable monogenic. By
Example 13.12 we have
(0) ( DerM (L) ⊂ DerK (L).
Otherwise L/K is separably generated and transcendental; let x1 , . . . , xd be a sep-
arable transcendence basis, so L/K(x1 , . . . , xd ) is separable algebraic (in fact of
1. DERIVATIONS 129
finite degree, though this is not needed). Then δ1 ∈ DerK (L(x1 , . . . , xn )) \ {0}, and
by Corollary 13.7 δ1 extends (uniquely) to a nonzero element of DerK (L).
The following result is an interesting instance of derivations working for us rather
than we for them.
Corollary 13.12. Let L = K(x1 , . . . , xn ) be a finitely generated field etension.
Suppose there are polynomials f1 , . . . , fn ∈ I(x1 , . . . , xn ) ⊂ K[x1 , . . . , xn ] such that
the matrix (∂i fj )(x1 , . . . , xn ) ∈ Mn (L) is nonsingular. Then L/K is separable
algebraic.
Proof. Let D ∈ DerK (L). Since fj (x) = 0, we have
n
X n
X
0 = D(fj (x)) = f D (x) + (∂i fj )(x)D(xi ) = (∂i fj )(x)D(xi ).
i=1 i=1
Finally we come to the general computation of dimL DerK (L) that was promised.
pd = [L : KLp ].
1As should be clear, this argument is valid whenever K/F is separably generated.
132 13. DERIVATIONS AND DIFFERENTIALS
p
P D ∈ Der(F ). Every element x ∈ F [K ] has a unique expression as a finite
Let
sum i∈I xi ui with xi ∈ F . We put
X
D(x) := D(xi )ui .
i∈I
p
P
The map D : F [K ] → K is clearly additive. Moreover, for x = i∈I xi ui , y =
p p
P
i∈I yi ui ∈ F [K ], then since D(γi,j,k ) = 0 for all γi,j,k ∈ F we have
X
D(xy) = D( xi yj γi,j,k uk )
i,j,k
X
= (D(xi )yj + si D(yj ))γi,jk )uk = xD(y) + D(x)y.
i,j,k
2. Kähler Differentials
Let B ⊂ A be an extension of commutative rings. We define the module of
Kähler differentials ΩA/B to be the quotient of the free A-module à on the set
{da | a ∈ A} (here we understand that for each a ∈ A we have a formal symbol da,
such that the map a 7→ da is injective) via the following submodule R of relations:
∀a, b ∈ A, d(a + b) − da − db,
∀a, b ∈ A, d(ab) − (adb + bda),
∀α ∈ B, dα.
The effect of this is that we get a map
d : A → ΩA/B , a 7→ da.
It is easy to check that d is a B-derivation. Moreover it is universal among all
B-derivations into an A-moduule M in the following sense.
Proposition 13.19. If M is an A-module and D : A → M is a B-derivation,
then there is a unique A-module homomorphism f : ΩA/B → M such that D = f ◦d.
Proof. There is a unique A-module homomorphism f˜ : Ã → M such that for
all a ∈ A, f˜(da) = D(a). For all a, b ∈ A, we have
f˜(d(a+b)−d(a)−d(b)) = f˜(d(a+b))−f˜(d(a))−f˜(d(b)) = D(a+b)−D(a)−D(b) = 0,
f˜(d(ab) − adb − bda) = f˜(d(ab)) − af˜(db) − bf˜(da) = D(ab) − aD(b) − bD(a) = 0,
and for all α ∈ B we have
f˜(dα) = D(α) = 0.
Thus f˜ factors through f : ΩA/B → M and has the property that for all a ∈ A,
f (da) = D(a), so D = f ◦ d. Conversely, any such f satisfies f (da) = D(a) for all
a ∈ A, and since {da}a∈A generate ΩA/B as an A-module, the map is unique.
Exercise 13.15. Let B ⊂ A be a ring extension.
(i) If S is a set of generators for B as an A-algebra (that is, every element
of B is an A-linear combination of products of elements of S), then {ds |
s ∈ S} is a set of generators for ΩA/B as an A-module.
(ii) Suppose that A and B are fields and A = B(S). Show that the elements
{ds | s ∈ S} span ΩA/B as an A-vector sapce.
We can restate Proposition 13.19 as giving a natural A-module isomorphism
DerB (A, M ) = HomA (ΩA/B , M ).
In particular, taking M = A, we get
(19) DerB (A) = HomA (ΩA/B , A) = Ω∨
A/B .
That is, the B-derivations of A are the A-linear functionals on ΩA/B . In particular:
Corollary 13.20. a) Suppose ΩA/B is finitely generated and free as an
A-module. Then DerB (A) is finitely generated and free as an A-module
and rank DerB (A) = rank ΩA/B .
134 13. DERIVATIONS AND DIFFERENTIALS
Theorem 13.21. Let L/K be a finitely generated field extension, and let x1 , . . . , xn ∈
L.
a) If {x1 , . . . , xn } is a separating transcendence basis for L/K, then {dx1 , . . . , dxn }
is an L-basis for ΩL/K .
b) Suppose that L/K is separably generated. If {dx1 , . . . , dxn } is an L-basis
for ΩL/K , then {x1 , . . . , xn } is a separating transcendence basis for L/K.
Proof. a) If x1 , . . . , xn is a separating transcendence basis for L/K, then for
all 1 ≤ i ≤ n there is a unique Di ∈ DerK (L) such that Di (xj ) = δ(i, j) and
D1 , . . . , Dn is an L-basis of DerK (L). Since DerK (L) = Ω∨ L/K , this basis is the dual
basis of a unique L-basis b1 , . . . , bn of ΩL/K : that is, for all 1 ≤ i, j ≤ n, we have
Di (bj ) = δ(i, j). We have identified Di with an L-linear functional `i on ΩL/K :
this functional is the unique one such that for all x ∈ L we have λi (dx) = Di (x).
Thus for all 1 ≤ i, j, ≤ n we have λi (dxj ) = Di (xj ) = δ(i, j), and it follows that we
have b1 = dx1 , . . . , bn = dxn .
b) Suppose dx1 , . . . , dxn is an L-basis for ΩL/K = DerK (L)∨ . First of all this
gives that dim DerK (L) = n, and since L/K is finitely generated and separable, by
Theorem 13.13 we get that trdeg(L/K) = n.
Now, let D ∈ DerK (L) be such that for all 1 ≤ i ≤ n we have 0 = dxi (D) =
D(xi ). Since the dxi ’s span DerK (L)∨ this implies D = 0, which shows that
DerK (L) = DerK(x1 ,...,xn ) (L). Using Proposition 13.11 we get that L/K(x1 , . . . , xn )
is separable algebraic, so x1 , . . . , xn is a separating transcendence basis for L/K.
Exercise 13.16. Show that Theorem 13.21 holds without the hypothesis that
L/K is finitely generated.
3. APPLICATIONS TO ONE VARIABLE FUNCTION FIELDS 135
0 = dy(δx ) = δx (y).
Exercise 13.17. Let K be a perfect field of chracteristic p > 0, and let L/K
be a finitely generated field extension.
a) Show: [L1/p : L] is finite.
b) Suppose that L/K has transcendence degree 1. Show: [L1/p : L] = p.
136 13. DERIVATIONS AND DIFFERENTIALS
4. p-Bases
Throughout this section we work with fields of a fixed characteristic p > 0.
Lemma 13.25. Let F be a field of characteristic p, and let S ⊂ F . Let S be the
set of all F p -linear combinations of monomials xi11 · · · xinn with x1 , . . . , xn ∈ S and
0 ≤ i1 , . . . , in < p. Then S = F p (S).
Proof. We have F p (S) = lim F p (T ) as T ranges over finite subsets of S
−→
and S = lim T as T ranges over all finite subets of S. So it suffices to show
−→
that S = F p (S) for all finite subsets S ⊂ F . The inclusion S ⊂ F p (S) is clear.
Conversely, because S is finite and consists of elements algebraic over F p , we have
F p (S) = F p [S]: that is, each element of F p (S) is an F p -linear combination of
monomials xa1 1 · · · xann with x1 , . . . , xn ∈ S and a1 , . . . , an ∈ N. For 1 ≤ i ≤ n, write
ai = pbi + ri with 0 ≤ ri < p. Then
xa1 1 · · · xann = (xb11 · · · xbnn )p xr11 · · · xrnn = αxr11 · · · xrnn
with α ∈ F p . This completes the proof.
A p-spanning subset is a subset S ⊂ F such that S = F . On the other hand:
Lemma 13.26. For a subset S ⊂ F , the following are equivalent:
(i) For every finite subset {s1 , . . . , sn } ⊂ S, the set of monomials si11 · · · sinn
with 0 ≤ i1 , . . . , in < p is F p -linearly independent.
(ii) For all s ∈ S, we have s ∈ / S \ {s}.
4. p-BASES 137
Probably the reader suspects what is coming next: we claim that for subsets S of
F , the mapping S 7→ S = F p (S) ⊂ F is a spanning operator in the sense of §11.4.
The properties (SO1) through (SO4) hold immediately: in fact, for any subfield A
of a field F , the operator S ⊂ F 7→ A(S) satisfiies these properties. We now check
(SO5) (“Exchange Lemma”): for x, y ∈ F and S ⊂ F , if y ∈ S ∪ {x} \ S, we must
show that x ∈ S ∪ {y}.
Because y ∈ / S, there is at least one I with in+1 6= 0 such that αI 6= 0 and this
shows that x satisfies a polynomial relation of degree less than p with coefficients
in F p (S, y). In other words, x is separable algebraic over F p (S, y), but it is also
purely inseparable over F p (S, y), so x ∈ F p (S, y) = S ∪ {y}.
Proof. (i) =⇒ (ii): Suppose that S is p-independent, and let S be the set
of subsets T ⊂ S such that for all t ∈ T there is DT,t ∈ DerF p (F p (T )) such that
DT,t (t) = 1 and DT,t (t0 ) = 0 for all t0 ∈ST \ {t}, partially ordered under inclusion.
For any chain {Ti }i∈I in S, the union i∈I Ti is an upper bound: the derivation
DT,t is uniquely determined by its properties, S so if t ∈ T1 ⊂ T2 then the restriction
of DT2 ,t to F p (T1 ) is DT1 ,t . For any t ∈ i∈I Ti , the existence and uniquenessf of
the derivation DSi∈I Ti ,t follows easily from this. By Zorn’s Lemma the set S has
a maximal element T . If T ( S, choose s ∈ S \ T ; by the p-independence of S we
have s ∈ / F p (T ) and sp ∈ F p ⊂ F p (T ), so by Example 13.8, for all α ∈ F P (T ∪{s}),
each derivation D ∈ DerF p (F p (T )) admits a unique extension to F p (T ∪ {s}) such
that D(s) = α. For all t ∈ T , extending DT,t by D(s) = 0 yields the derivation
DS,t , while extending the 0 derivation on F p (T ) to F p (T ∪ {s}) by D(s) = 1 yields
the derivation DS,s . This shows that T ∪ {s} ∈ S, contradicting the maximality of
T . It follows that T = S.
¬ (i) =⇒ ¬ (ii): If the subset S is not p-independent, then there is s ∈ S such
that s ∈ F p (S \ {s}). It follows that every D ∈ DerF p (F ) such that D(t) = 0 for
all t 6= s also satisfies D(s) = 0.
Corollary 13.29. Let F be a field of characteristic p > 0. For S ⊂ F , the
following are equivalent:
(i) The subset S is a p-basis for F .
(ii) The map s 7→ ds is injective and {ds | s ∈ S} is a basis for ΩF/F p .
Proof. (i) =⇒ (ii): Let S be a p-basis for F .
Since S is a p-spanning subset of F we have F = F p (S), so by Exercise 13.15
the set {ds | s ∈ S} spans ΩF/F p .
Since S is a p-independent subset, applying Theorem 13.28 gives a family of
derivations {Dx | x ∈ S} such that for all y ∈ S, we have
(
1 y=x
Dx (y) = .
0 y 6= x
Regarding DerF p (F ) as the F -vector space dual of ΩF/F p , we have
(
1 y=x
∀x, y ∈ S, Dx (dy) = Dx (y) = .
0 y 6= x
Thus if dx were an F -linear combination of elements dy with y ∈ S \{x}, evaluating
at Dx gives 1 = 0, a contradiction. So the dx’s are distinct elements and form an
F -linearly independent subset of ΩF/F p .
(ii) =⇒ (i): Suppose that x ∈ S 7→ dx is an injection and that {dx | x ∈ S}
is an F -basis for ΩF/F p . Because DerF/F p = Ω∨F/F p , there is a unique family of
derivations {Dx | x ∈ S} such that Dx ∈ DerF p (F ) and for all x, y ∈ S we have
4. p-BASES 139
(
1 y=x
Dx (y) = . Theorem 13.28 implies that S is a p-independent subset. If it
0 y 6= x
were not a p-basis, there would be a strictly larger p-independent subset S̃ = S∪{s},
and Theorem 13.28 implies that there is Ds ∈ DerF p (F ) such that Ds (s) = 1 and
Ds (x) = 0 for all x ∈ S. But then on the one hand we would have Ds (ds) =
Ds (s) = 1, and on the other hand, since ds is an F -linear combination of dx for
x ∈ S, we would have Ds (ds) = 0, a contradiction.
Corollary 13.29 is an example of how Kähler differentials behave better than deriva-
tions in non-omega finite extensions. Indeed, when ΩK/F is infinite-dimensional as
a K-vector space, its dual vector space DerF (K) is an infinite vector space of larger
dimension. If we choose a differential basis for K/F – i.e., a subset S of K such
that s 7→ ds ∈ ΩK/F is injective and {ds | s ∈ S} is a K-basis for ΩK/F – then there
is a “dual basis” {D
( s | s ∈ S} of F -derivations of K characterized by the usual
1 x=y
relation Dx (y) = for all x, y ∈ S. However this dual basis is a basis
0 x 6= y
not for DerF (K) but for the subspace DerSF (K) of S-finite derivations, a subspace
which depends strongly on the choice of S.
Exercise 13.20. An extension of characteristic p fields F ⊂ K is p-radical if
K ⊂ F 1/p : in other words, K is obtained by adjoining pth roots of elements of F .
For any p-adical extension K/F and any subset S ⊂ K, define S :== F (S).
a) Show that S is the F -span of monomials si11 · · · sinn with s1 , . . . , sn ∈ S and
0 ≤ i1 , . . . , in < p.
b) Extend the theory of p-spanning subsets, p-independent subsets and p-bases
to p-radical extensions K/F . Show in particular:
(i) The extension K/F has a p-basis – a subset S such that K = F (S)
and for all s ∈ S, s ∈ / F (S \ {s}) – and any two p-bases have the
same cardinality.
(ii) If S is a finite p-basis for K/F , then [K : F ] = p#S , while if S is an
infinite p-basis for K/F , then [K : F ] = #S.
(iii) A subset S ⊂ F is a p-basis iff s 7→ ds ∈ ΩK/F is injective and
{ds | s ∈ S} is an F -basis for ΩK/F .
(iv) For x ∈ F , we have dx = 0 ∈ ΩK/F iff x ∈ F .
CHAPTER 14
Ordered Fields
A homomorphism of ordered abelian groups f : (G, <) → (H, <) is a group homo-
morphism which is isotone: for all x1 ≤ x2 , f (x1 ) ≤ f (x2 ).
Lemma 14.1. For x, y, z in an ordered abelian group G, if x < y then x + z <
y + z.
Proof. Since x < y, certainly x ≤ y, so by (OAG) we have x + z ≤ y + z. If
x + z = y + z then adding −z to both sides gives x = y, a contradiction.
Lemma 14.2. Let x1 , x2 , y1 , y2 be elements of an ordered abelian group G with
x1 ≤ x2 and y1 ≤ y2 . Then x1 + y1 ≤ x2 + y2 .
Proof. Applying (OAG) with x1 , x2 , y1 gives x1 + y1 ≤ x2 + y1 . Applying
(OAG) with y1 , y2 , x2 gives x2 + y1 = y1 + x2 ≤ y2 + x2 . By transitivity we have
x1 + y1 ≤ x2 + y2 .
To an ordering on a commutative group we associate its positive cone:
G+ = {x ∈ G | x > 0}.
Elements of G+ are called positive. We also define
G− = {x ∈ G | x < 0}.
Elements of G− are called negative.
Lemma 14.3. Let x be a nonzero element of`the ordered abelian group G. Then
exactly one of x, −x is positive. Thus G = {0} G+ G− .
`
b) If x1 < 0 and x2 < 0, then by Lemma 14.3 we have −x1 , −x2 > 0. Now part a)
gives −x1 − x2 = −(x1 + x2 ) > 0, so by Lemma 14.1 again we have x1 + x2 < 0.
In an ordered abelian group we define |x| to be x if x ≥ 0 and −x otherwise.
Exercise 14.1. Let x, y be elements of an ordered abelian group G.
a) Suppose x ≤ y and n ∈ N. Show that nx ≤ ny.
b) Suppose x ≤ y and n is a negative integer. Show that nx ≥ ny.
Example 14.5. Let (G, <) be an ordered abelian group and H a subgroup of
G. Restricting < to H endows H with the structure of an ordered abelian group.
Example 14.6 (Lexicographic Ordering). Let {Gi }i∈I be a nonempty indexed
family of ordered abelian groups. Suppose that we are given Q a well-ordering on the
index set I. We may then endow the direct product G = i∈I Gi with the structure
of an ordered abelian group, as follows: for (gi ), (hi ) ∈ G, we decree (gi ) < (hi ) if
for the least index i such that gi 6= hi , gi < hi .
Theorem 14.7. (Levi [Lev43])
For an abelian group G, the following are equivalent:
(i) G admits at least one ordering.
(ii) G is torsionfree.
Proof. (i) =⇒ (ii) Let < be an ordering on G, and let x ∈ G• . By Lemma
14.4 we have nx 6= 0 for all n ∈ Z+ .
(ii) =⇒ (i): Let G be a torsionfree abelian group. Then G is a flat Z-module.
Tensoring the injection Z ,→ Q gives us an injection G ,→ L G ⊗ Q. Since Q is a
field, the Q-module G ⊗ Q is free, i.e., it is isomorphic to i∈I Q. Choose a total
ordering on I. Give each copyL of Q its standard ordering as a subfield of R and put
the lexicographic ordering on i∈Q Q ∼ = G ⊗ Q. Via the injection G ,→ G ⊗ Q this
induces an ordering on G.
An anti-isomorphism of abelian groups is an order-reversing group isomophism.
For every ordered abelian group (G, <), the inversion map x ∈ G 7→ −x is an
anti-isomorphism of G.
Exercise 14.2. a) Show that the abelian group Z admits exactly two orderings
∼
<1 and <2 , such that inversion gives an isomorphism (Z, <1 ) → (Z, <2 ).
b) Give an example of an abelian group G admitting orderings <1 and <2 such that
(G, <1 ) is not isomorphic or anti-isomorphic to (G, <2 ).
The comparability quasi-ordering: For x, y ∈ G, we write x ≺ y if there exists
n ∈ Z+ such that |x| ≤ n|y|. We claim that ≺ is a quasi-ordering on G, i.e., a
reflexive, transitive but not necessarily anti-symmetric binary relation. Indeed the
reflexivity is immediate; if x ≺ y and y ≺ z then there exist n1 , n2 ∈ Z+ such that
|x| ≤ n1 |y| and |y| ≤ n2 |z|, and thus |x| ≤ n1 n2 |z|.
In any ordered abelian group G, {0} is its own ≈-equivalence class, hence any non-
trivial ordered abelian group has at least two ≈-equivalence classes. We refer to
nonzero ≈-equivalence classes as Archimedean equivalence classes and denote
the set of all such equivalence classes as Ω(G).
An ordered abelian group with #Ω(G) ≤ 1 is called Archimedean. Equiva-
lently, for all x, y ∈ G• , there are n1 , n2 ∈ Z+ such that |x| ≤ n1 |y| and |y| ≤ |x|.
Example 14.8. The group (R, +) is Archimedean. That is, for any x ∈ R>0
there are positive integers n1 and n2 such that n11 ≤ x ≤ n2 . Indeed the second
inequality follows from the least upper bound axiom: if this were not the case then
the set Z+ of positive integers would be bounded above in R, and this set cannot
have a least upper bound. The first inequality follows from the second upon taking
reciprocals.
Example 14.9. A subgroup of an Archimedean ordered abelian group is Archimedean.
In particular, any subgroup of (R, +) is Archimedean in the induced ordering.
Rather remarkably, the converse is also true.
Theorem 14.10. (Hölder [Hö01]) Let (G, +) be an ordered abelian group. If
G is Archimedean, there exists an embedding (G, +) ,→ (R, +).
Proof. We may assume G is nontrivial. Fix any positive element x of G. We
will construct an order embedding of G into R mapping x to 1.
Namely, let y ∈ G. Then the set of integers n such that nx ≤ y has a maximal
element n0 . Put y1 = y − n0 x. Now let n1 be the largest integer n such that
nx ≤ 10y1 : observe that 0 ≤ n1 < 10. Continuing in this way we get P a set of integers
∞ nk
n1 , n2 , . . . ∈ {0, . . . , 9}. We define ϕ(y) to be the real number n0 + k=1 10 k . It is
not hard to show that ϕ is isotone – y ≤ y 0 =⇒ ϕ(y) ≤ ϕ(y 0 ) – and also that ϕ is
injective: we leave these tasks to the reader.
But let us check that ϕ is a homomorphism of groups. For y ∈ G, and r ∈ Z+ ,
let 10nr be the rational number obtained by truncating ϕ(y) at r decimal places.
The numerator n is characterized by nx ≤ 10r y < (n + 1)x. For y 0 ∈ G, if
n0 x ≤ 10r y 0 ≤ (n0 + 1)x, then
(n + n0 )x ≤ 10r (y + y 0 ) < (n + n0 + 2)x,
so
2
ϕ(y + y 0 ) − (n + n0 )10−r <
10r
and thus
4
|ϕ(y + y 0 ) − ϕ(y) − ϕ(y 0 )| < .
10r
Since r is arbitrary, we conclude ϕ(y + y 0 ) = ϕ(y) + ϕ(y 0 ).
Proposition 14.11. Let G be a nontrivial Archimidean ordered abelian group.
Then exactly one of the following holds:
(ii) G is order-isomorphic to Z.
(ii) The ordering on G is dense.
Proof. Step 1: Suppose G+ has a least element x. Let y ∈ G+ . Since
the ordering is Archimedean there is a largest n ∈ Z+ such that nx ≤ y. Then
y − nx ≥ 0; if y > 0 then y − nx ≥ x so y ≥ (n + 1)x, contradicting the maximality
of n. Thus y = nx, i.e., every positive element of G+ is a multiple of x. It follows
144 14. ORDERED FIELDS
(OR) ∀x, y ≥ 0, xy ≥ 0.
In these notes the ordered rings we will study are ordered fields.
Example 14.14. The real numbers R with the standard < form an ordered
field.
√
Example 14.15. Let F = Q( 2). There are two embeddings F ,→ R √ which
differ
√ from each other by the nontrivial
√ automorphism of F , which carries 2 7→
− 2. In one of these embeddings, 2 goes to the positive real number whose square
is 2, and in the other one it goes to the negative real number whose square is 2.
Thus the two embeddings give different orderings, and it is easy to check that these
are the only two orderings of F .
A homomorphism f : (F, <) → (F 0 , <0 ) is a monotone field homomorphism: i.e., a
field homomorphism such that x < y =⇒ f (x) < f (y).
Exercise 14.4. Let (K, <) be an ordered field and let F be a subfield of K.
Denote by <F the restriction to F of <. Show that (F, <F ) is an ordered field and
the inclusion of F into K is an homomorphism of ordered fields.
We denote by X(K) the set of all field orderings on K.
Exercise 14.5. Show that there is a natural action of Aut(K) on X(K). Give
an example where the orbit space Aut(K)\X(K) consists of more than one element.
2. INTRODUCING ORDERED FIELDS 145
(PO1) P + P ⊂ P , and P P ⊂ P .
(PO2) Σ (K) = {x21 + . . . + x2n | xi ∈ K} ⊂ P .
(PO3) −1 ∈ / P.
(PO30 ) P ∩ (−P ) = {0}.
(PO300 ) P • + P • ⊂ P • .
(PO3000 ) P 6= K.
(PO4) P ∪ (−P ) = K.
Exercise 14.6. Let P ⊂ K satisfy (PO1) and (PO2).
a) Show: (PO3), (PO30 ), and (PO300 ) are equivalent conditions on P .
b) Suppose char K 6= 2. Show: (PO3000 ) and (PO3) are equivalent conditions on P .
2 2
(Hint: x = x+1
2 − x−12 .)
c) Suppose char K = 2. Show: P satisfies (PO1) and (PO2) iff P is a subfield of
K containing K 2 .
Exercise 14.7. Let P ⊂ K satisfy (PO1) and (PO4). Show: P satisfies
(PO2).
Lemma 14.17. Let K be a field.
a) If ≤ is a field ordering on K, put P = K ≥0 . Then K satisfies (PO1) through
(PO4) above, and also 1 ∈ P .
b) Let P ⊂ K satisfy (PO1) through (PO4). Define a relation ≤ on K by x ≤
y ⇐⇒ y − x ∈ P . Then ≤ is a field ordering on K.
Proof. a) By Proposition X.X, K has characteristic 0. Lemma 146 implies P
satisfies (PO1), and Lemma 147 implies P satisfies (PO30 ) and (PO4). By Exercise
X.X, P satisfies (PO2). Finally, by (PO4), exactly one of 1, −1 lies in P . But if
−1 ∈ P , then (−1)2 = 1 ∈ P , so 1 ∈ P .
b) By (PO3) and (PO4), ≤ is a total ordering on K. Given x, y, z ∈ K with x ≤ y,
then (y + z) − (x + z) = y − x ∈ P , so x + z ≤ y + z: (K, ≤) is an ordered abelian
group. Finally, if x, y ≥ 0 then x, y ∈ P , so by (PO1) xy ∈ P , i.e., xy ≥ 0.
In view of this result, we refer to a subset P ⊂ K satisying (PO1), (PO2) and
(PO3) as being an ordering on K, and we often refer to the ordered field (K, P ).
Exercise 14.8. Let P1 , P2 be two orderings on a field K. Show: P1 ⊂ P2 =⇒
P1 = P2 .
The alert reader may now be wondering why we have introduced (PO2) at all since
it is implied by the other axioms for an ordering.1 The reason is that it is a key
idea to entertain a more general structure.
of K. (Note that our choice of (PO3) instead of (PO3000 ) shows that a field of
characteristic 2 admits no preorderings.)
Exercise 14.9. Let T be a preordering on F and x, y ∈ T .
Show that x, y ∈ T, x + y = 0 =⇒ x = y = 0.
A field K is formally real if −1 ∈
/ Σ (K).
Lemma 14.18. Suppose char K 6= 2. Then the following are equivalent:
(i) K is formally real.
(ii) Σ (K) ( K. (iii) Σ (K) is a preordering on K.
(iv) For all n ≥ 1 and all x1 , . . . , xn ∈ K, x21 +. . .+x2n = 0 =⇒ x1 = . . . = xn = 0.
Exercise 14.10. Prove it.
Remark: Condition (iv) above makes a connection with quadratic form theory. A
quadratic form q(x) = a1 x21 + . . . + an x2n over a field K is isotropic if there exists
some nonzero x ∈ K n with q(x) = 0 and otherwise anisotropic. Thus by Lemma
X.X, a field is formally real iff for each n the sum of n squares form is anisotropic.
Lemma 14.19. Let F be a field such that Σ (F ) ∩ (−Σ (F )) = ∅ and Σ (F ) ∪
(−Σ (F )) = F × . Then P = Σ (F ) is the unique ordering on F .
Exercise 14.11. Prove Lemma 14.19.
Exercise 14.12. Use Lemma 14.19 to show that each of the following fields
admits a unique ordering: R, Q, the field of constructible numbers.
Proposition 14.20. If (F, P ) is an ordered field, F is formally real.
Proof. The contrapositive is clear: if F is not formally real, then −1 is a sum
of squares, so it would be – along with 1 – in the positive cone of any ordering.
It follows that any ordered field has characteristic 0.
Much more interestingly, the converse of Proposition 14.20 is also true. In order to
prove this celebrated result we need the following innocuous one.
Lemma 14.21. Let F be a field, T ⊆ F a preordering on F , and a ∈ F × . The
following are equivalent:
(i) The set T [a] := {x + ya | x, y ∈ T } is a preordering.
(ii) a is not an element of −T .
Proof. Since by (PC4) no preordering can contain both a and −a, (i) =⇒
(ii) is clear. Conversely, assume (ii). It is immediate to verify that T [a] satisfies
(PC1), (PC2) and (PC5), so it suffices to show that there is no x ∈ F × such that
x and −x both lie in T [a]. If so, we deduce
1
−1 = −x · x · ( )2 ∈ T [a].
x
But now suppose −1 = x + ya for x, y ∈ T . Then −ya = 1 + x is a nonzero element
of T , so a = (−y)−2 (−y)(1 + x) ∈ −T , a contradiction.
Theorem 14.22. Let t be a preordering on a field K. Then:
a) The set t is the intersection of all orderings P ⊃ t.
b) (Artin-Schreier) If K is formally real, then it admits an ordering.
3. EXTENSIONS OF FORMALLY REAL FIELDS 147
Remark: Corollary 14.23 does not extend to all fields of characteristic 2. Indeed,
for a field F of characteristic 2, we simply have Σ (F ) = F 2 , so every element of
F is a sum of squares iff F is perfect. (In no case are there any orderings on F .)
In general the extension problem for orderings is a rich one with a large liter-
ature. But we will give one fundamental and useful result, an extension of the
Artin-Schreier Theorem. First:
Lemma 14.24. For an ordered field (K, p), an extension L/K, and c ∈ L, the
following are equivalent:
(i) There are a1 , . . . , an ∈ p• and x1 , . . . , xn ∈ L such that
c = a1 x21 + . . . + an x2n .
2particular, a subfield of a formally real field is formally real. But that was clear anyway
from the definition.
148 14. ORDERED FIELDS
T
(ii) c ∈ P ⊃p P, the intersection being over all orderings of L extending p.
Proof. Let
t = {a1 x21 + . . . + an x2n | ai ∈ p, xi ∈ L},
T
and note that the desired equivalence can be rephrased as t = P ⊃p P . Moreover
t satisfies (PO1) and (PO2), and an ordering P of L contains t iff it contains
T p.
Case 1: Suppose −1 ∈ / t. Then t is a preordering, and by Theorem X.X, t = P ⊃p P .
Case 2: If −1 ∈ t, there is no ordering on L extending p. Then – since KThas ordered
and thus not of characteristic 2! – by Exercise X.X, we have t = K = P ⊃p P .
We are now ready for one of our main results.
Theorem 14.25. For an ordered field (K, p) and an extension field L/K, the
following are equivalent:
(i) There is an ordering on L extending p.
(ii) For all a = (a1 , . . . , an ) ∈ pn , the quadratic form
qa (x) = a1 x21 + . . . + an x2n
is anisotropic over L: if x = (x1 , . . . , xn ) ∈ Ln is such that q(x) = 0, then x = 0.
Proof. (i) =⇒ (ii) is immediate.
(ii) =⇒ (i): If for any a ∈ pn the quadratic form qa (x) represents −1, then
the form qa,1 (x) = a1 x21 + . . . + an x2n + x2n+1 would be isotropic, contrary to our
hypothesis. It follows that
/ t = {a1 x21 + . . . + an x2n | ai ∈ p, xi ∈ L},
−1 ∈
so – as in the proof of the previous result – t is a preordering of L containing p. By
Theorem X.X, t must extend to at least one ordering of L.
Exercise 14.13. Deduce from Theorem 14.25 that every formally real field L
admits an ordering. (Hint: we wrote L, not K!)
We will now deduce several sufficient conditions for extending orderings.
√
Theorem 14.26. Let (K, p) be an ordered field, and let L = K({ x}x∈p ) be
the extension obtained by adjoining all square roots of positive elements. Then the
ordering p extends to L.
Proof. By Theorem 14.25, it suffices√ to show √ that for any n, r ∈ Z+ and any
b1 , . . . , br , c1 , . . . , cn ∈ p, if x1 , . . . , xn ∈ F ( b1 , . . . , br ) are such that
(20) c1 x21 + . . . + cn x2n = 0,
then x1 = . . . = xn = 0. For any fixed n, we prove this by induction on r. Suppose
by induction that the equation c1 x21 + . . . + cn x2n = 0 has no nontrivial solutions
√
over Kr−1 , and let (z1 , . . . , zn ) ∈ Krn be a solution to (20). Write zi = xi + br yi ,
with xi , yi ∈ Kr−1 . Then equating “rational parts” in the equation
X X X X p
0= ci zi2 = ci x2i + br ci yi2 + 2 ci xi yi br
2n
shows that (x1 , . . . , xn , y1 , . . . , yn ) ∈ Kr−1 is a solution of
c1 t21 + . . . + cn t2n + br c1 t2n+1 + . . . + br cn t22n = 0.
By induction, x1 = . . . = xn = y1 = . . . = yn = 0, i.e., z1 = . . . = zn = 0.
3. EXTENSIONS OF FORMALLY REAL FIELDS 149
To obtain further results we take a perspective arising from quadratic form theory.
Let us say a field extension L/K is anistropic if every anisotropic quadratic form
q(x1 , . . . , xn ) ∈ K[x1 , . . . , xn ] remains aniostropic when extended to L. (In the
algebraic theory of quadratic forms one studies the Witt kernel of a field extension:
the kernel of the natural ring homomorphism W (K) → W (L). An anisotropic
extension is precisely one in which the Witt kernel is trivial.) From Theorem 14.25
we immediately deduce the following result.
Corollary 14.27. If (K, p) is an ordered field and L/K is an anisotropic
extension, then the ordering p extends to L.
Exercise 14.14. a) Let K be a field and let {Li }i∈I be a directed system of
anisotropic extensions of K. Show that lim Li /K is an anisotropic extension.
−→
b) Let (K, p) be an ordered field and L/K a field extension. Suppose that p extends
to an ordering on any finitely generated subextension of L/K. Show that p extends
to an ordering on L.
The next results give the two basic examples of anisotropic extensions.
Theorem 14.28. A purely transcendental extension L/K is anisotropic.
Proof. Step 0: It suffices to prove that K(t)/K is anisotropic. Indeed, if
so then an immediate induction gives that K(t1 , . . . , tn )/K is anisotropic, and we
finish by applying Exercise X.X.
Step 1: Let K be any field, and let (f1 , . . . , fn ) ∈ K(t)n be an n-tuple of ra-
tional functions, not all zero. Then there exists a nonzero rational function f
such that (f f1 , . . . , f fn ) is a primitive vector in K[t], i.e., each f fi ∈ K[t] and
gcd(f f1 , . . . , f fn ) = 1. Indeed this holds with K[t] and K(t) replaced by any UFD
and its fraction field.
Remark: The proof of Proposition X.X used only that q was a form – i.e., a homo-
geneous polynomial – not that it was a quadratic form. Indeed any system of ho-
mogeneous polynomials would work as well, so the argument really shows: if V/K is
a projective variety which has a K(t)-rational point, then it has a K-rational point.3
3The same conclusion holds for arbitrary varieties over any infinite field, or for complete
varieties over a finite field. But taking the projective line over Fq and removing its Fq -rational
points shows that some hypothesis is necessary!.
150 14. ORDERED FIELDS
The following was conjectured by Witt in 1937 and proven by Springer in 1952.4
Theorem 14.29. (Springer [Sp52]) Let L/K be a field extension of finite odd
degree d. Then L/K is anisotropic.
Proof. We go by induction on the degree, the case d = 1 being trivial. Sup-
pose the result holds for all field extensions of odd degree less than d, and L/K
be an extension of odd degree d. If L/K had any proper subextension, then we
would be done by a dévissage argument. So we may assume in particular that L is
monogenic over K: L = K[x]. Let p(t) ∈ K[t] be the minimal polynomial of x. Let
q be an anisotropic quadratic form over K which becomes isotropic over L: i.e.,
there exists an equation
(21) q(g1 (t), . . . , gn (t)) = h(t)p(t)
with polynomials gi , h ∈ K[t], not all gi = 0, and M := max deg gi ≤ d − 1.
As in the proof of Proposition 14.28, we may also assume that (g1 , . . . , gn ) is a
primitive vector in K[t]. Since q is anisotropic, the left hand side of (21) has degree
2M ≤ 2d − 2, so deg h is odd and at most d − 2. In particular, h has an irreducible
factor h̃ of odd degree at most d − 2; let y be a root of h̃ in K. Taking t = y in (21),
we see that q(g1 (y), . . . , gn (y)) = 0. Note that since K[t] is a PID, the condition
gcd(g1 , . . . , gn ) = 1 is equivalent to the fact that 1 ∈ hg1 , . . . , gn i, which implies
that the polynomials g1 , . . . , gn remain setwise coprime as elements of K[y][t]. In
particular, not all gi (y) are equal to 0, so that qK[y] is isotropic. By induction, this
implies that q was isotropic, contradiction!
Exercise 14.15. Let (K, p) be an ordered field. Show that the formal power
series field K((t)) admits a unique ordering extending p in which 0 < t < x for all
x ∈ K.
Exercise 14.16. In the algebraic theory of quadratic forms it is shown that the
√
Witt kernel of a quadratic extension L = K( p)/K is the principal ideal generated
by α = h1, −pi: [Cl-QF, Thm. II.20]. In other words, it consists of quadratic
forms a1 x21 + . . . + an x2n − pa1 x21 − . . . − pan x2n for a1 , . . . , an ∈ K × . Use this (and
induction) to give another proof of Theorem 14.26.
Step 5: Now suppose that #G = 4. Then there exists at least one subexten-
sion K of F /F with [F : K]. Then the above reasoning shows that i 6∈ K, hence
not in F , but then F (i) is a subfield of F with [F : F (i)] = 2 and containing a 4th
root of unity, contradicting the above analysis.
In summary, we have shown so far that if 1 < [F : F ] < ∞, then F does not
have characteristic 2 and F = F (i). It remains to be shown that F is formally real,
and this is handled by the following result.
Lemma 14.32.
√ which −1 is not a square and such that every
Let F be a field in √
element of F ( −1) is a square in F ( −1). Then:
a) Σ (F ) = F 2 ,
b) char(F ) = 0, and
c) F is formally real.
√
Proof. Put i = −1. To show part a), it is enough to see that the sum of
two squares in F (i) is again a square in F (i). Let a, b ∈ F . By hypothesis, there
are c, d ∈ F such that (a + bi) = (c + di)2 , so a = c2 − d2 and b = 2cd and thus
a2 + b2 = (c2 + d2 )2 .
b) If F had positive characteristic p, then −1 is a sum of p − 1 squares but not
itself a square, contradicting part a).
c) Since −1 is not a square, F does not have characteristic 2, and thus by part a)
−1 is not a sum of squares and F is formally real.
Exercise 14.17. (E. Fried): Let F be a field. Suppose that there exists a
positive integer d such that for every irreducible polynomial P ∈ K[t], deg(P ) ≤ d.
Show that F is real-closed or algebraically closed.
Exercise 14.19. (K. Conrad): A field K is real-closed iff 1 < [K sep : K] < ∞.
6. Real Closures
Proposition 14.37. For every formally real field K, there exists an algebraic
extension K rc which is real-closed.
Proof. Let K be an algebraic closure of K, and consider the partially ordered
set of formally real subextensions of K/K. Since the union of a chain of formally
real fields is formally real, Zorn’s Lemma applies to give a maximal formally real
subextension, which is by definition real-closed.
Definition: A real closure of a formally real field K is a real-closed algebraic
extension of K.
Lemma 14.38. Let K be a field, let R/K be a real-closed extension field of K,
and let R0 be the algebraic closure of K in R. Then R0 is a real closure of K.
Exercise 14.23. Prove Lemma 14.38.
Thus we have shown the existence of real closures for formally real fields. What
about uniqueness? By comparison with the case of algebraically closed fields, one
might guess that any two real closures of a given formally real field K are isomorphic
as K-algebras. However, this is in general very far from being the case!
Example 14.39. Let K = Q(t). There is a unique embedding ι : K → R
in which t gets sent to π. Let K1 be the algebraic closure of ι(K) in R. On the
S 1
other hand, let ι2 : K → n R(t n ) be the natural embedding of K into the Puiseux
S 1
series field, and let K2 be the algebraic closure of ι2 (K) in n R(t n . By Corollary
14.33, K1 and K2 are both real-closed fields. In particular, they each admit a unique
ordering, in which the positive elements are precisely the nonzero squares. However,
the ordering on K1 is Archimedean and the ordering on K2 is not, since t is an
infinitesimal element. Therefore K1 and K2 are not isomorphic as fields, let alone
as K-algebras.
Theorem 14.40. Let (F, p) be an ordered field. Then there is an algebraic
extension R/F which is real-closed and such that the unique ordering on R extends
p.
√
Proof. Let K = F ({ x}x∈p ). By Theorem 14.26, K is formally real, and
now by Proposition 14.37, there exists a real-closed algebraic extension R of K.
Let P = {x2 | x ∈ R× } be the unique ordering on R. Every x ∈ p is a square
in K and hence also in R: that is, p ⊂ P ∩ F . Conversely, if x ∈ F × \ p, then
−x ∈ p ⊂ P ∩ F ⊂ P , so that x 6∈ P , hence x 6∈ P ∩ F . Thus P ∩ F = p.
In the above situation, we say R is a real closure of the ordered field (F, p).
Exercise 14.24. Use Theorem 14.40 to give a third proof of X.X and X.X.
Theorem 14.41. (Sylvester) Let (K, p) be an ordered field, and let (R, P ) be a
real-closed extension. Let f ∈ K[t] be a nonzero monic separable polynomial, and
put A = K[t]/(f ). Let Bf be the trace form on the K-algebra A, i.e., the bilinear
156 14. ORDERED FIELDS
√
form hx, yi = TrA/K (x, y). Let C = R( −1). Then:
a) The number of roots of f in R is equal to the signature of Bf .
b) Half the number of roots of f in C \ R is equal to the number of hyperbolic planes
appearing in the Witt decomposition of Bf .
Proof. Let f (t) = f1 (t) · · · fr (t) be the factorization
√ of f over R[t]. Since f is
separable, the polynomials fi are distinct, and since R( −1) is algebraically closed,
each fi has degree 1 or 2. Since A ⊗K R ∼ = R[t]/(f ), the trace form of A ⊗K R
is simply the scalar extension to R of the trace form Bf . Further, by the Chinese
Remainder Theorem
r
R[t]/(f ) ∼
Y
= R[t]/(fi ),
i=1
so
r
(Bf )/R ∼
M
= Bfi .
i=1
It is easy to see that if deg fi = 1 then the trace form is just h1i, whereas the
computation at the end of Section 7 shows that when deg fi = 2 – so that R[t]/(fi ) ∼
=
C – the trace form is congruent to h2, −2i ∼ = h1, −1i = H, the hyperbolic plane.
Both parts of the theorem follow immediately.
Sylvester’s Theorem may look rather specific and technical at first glance. Let
us explicitly extract from it the following key consequence: let f ∈ K[t] by a
polynomial defined over an ordered field (K, P ). Then if f has a root in one
real-closed field extending (K, P ), it has a root in every real-closed field extending
(K, P ). This is a very special case of Tarski’s transfer principle, which a logician
would express in the form “The theory of real-closed fields is model complete.”
Although it is a very special case, it has enough teeth to be the driving force
behind the powerful theorems we will now establish.
Theorem 14.42. Let (E, P )/(K, p) be an algebraic extension of ordered fields.
Let R be a real-closed field, and let σ : K → R be an ordered field embedding. Then
there is a unique order embedding ρ : E ,→ R extending σ.
Corollary 14.43. Let (K, P ) be an ordered field, and for i = 1, 2 let σ: (K, P ) →
Ri be real closures. There is a unique K-algebra isomorphism ρ : R1 → R2 .
Proof. Applying Theorem 14.42 with R1 = E and R2 = R, σ2 = σ, there is a
unique order embedding ρ : R1 → R2 extending σ2 . Since R2 /ρ(R1 ) is an algebraic
extension of real-closed fields we must have ρ(R1 ) = R2 . Finally, if τ : R1 → R2 is
any K-algebra homomorphism, then for all α > 0 in R1 , we have
√
τ (α) = τ ( α)2 > 0.
Thus τ is order-preserving, so τ = ρ.
be the minimal polynomial for y over K(x). Since y is integral over K[x], we have
ci (x) ∈ K[x] for all i. For a ∈ K, put fa (Y ) = f (a, Y ) ∈ K[Y ]. We look for roots of
fa in K. For if b ∈ K is such that fa (b) = f (a, b) = 0, there is a unique K-algebra
map ϕ : K[x, y] → K with ϕ(x) = a, ϕ(y) = b. So it is enough to show: there are
infinitely many a ∈ K such that there is b ∈ K with fa (b) = 0.
Step 4: Finally we use that E is formally real! Let P be an ordering on E and
let R be a real-closure of (E, P ). Then f (x, Y ) ∈ K[x][Y ] has a root in R, namely
y ∈ E ⊂ R. By Sylvester’s Theorem, sgn(Bf )/K(x) > 0. If we can show that
there are infinitely many a ∈ K such that sgn((Bf )a )/K) > 0, then applying
Sylvester’s Theorem again we will get infinitely many a such that fa (Y ) has a root
in K and be done. We may diagonalize the quadratic form corresponding to Bf
as hh1 (x), . . . , hn (x)i, say. Staying away from the finitely many a such that hi (a)
is zero or undefined for some i, we have that Bfa ∼ = hh1 (a), . . . , hn (a)i. By Lemma
14.44 there are infinitely many a such that sgn Bfa = sgn Bf > 0, and we’re done.
Actually Lang proved a stronger result, giving in particular a necessary and suffi-
cient condition for E to be formally real. His result uses the language of arithmetic
geometry, so unfortunately will probably not be accessible to all readers of these
notes, but here it is anyway.
Theorem 14.46 (Lang [La53]). Let V/R be a geometrically integral algebraic
variety over a real-closed field R, with function field E = R(V ). Then E is formally
real iff V has a nonsingular R-point.
The Artin-Lang homomorphism theorem is powerful enough to yield a quick proof
of the following result, which when one takes K = R = R, was the 17th of Hilbert’s
Problems proposed to the worldwide mathematical community in 1900.
Theorem 14.47 (Artin). Let K be a formally real field admitting a unique
ordering, and let R be a real closure of K. If f ∈ K[t1 , . . . , tm ] is such that
f (a1 , . . . , an ) ≥ 0 ∀(a1 , . . . , an ) ∈ Rn ,
then f is a sum of squares in K(t1 , . . . , tm ).
Proof. We argue by contraposition: suppose f ∈ K[t1 , . . . , tm ] is not a sum
of squares in K(t1 , . . . , tm ). By Corollary 14.23, there is an ordering P on E =
K(t1 , . . . , tm ) such that f <P 0. Let R be a real closure of (E, P ). Then f < 0
in R, so there is w ∈ R with w2 = −f . By Lemma 14.38, the algebraic closure
R0 of K in R is real-closed, hence is a real-closure of the ordered field K since K
admits exactly one ordering. By uniqueness of real closures R0 = R. The field
R(t1 , . . . , tm , w) is a subfield of the real-closed field R, hence by Artin-Lang there
is an R-algebra map
1
ϕ : R[t1 , . . . , tn , w, ] → R.
w
Note that the effect of including w1 is that ϕ(w)ϕ( w1 ) = 1, hence ϕ(w) 6= 0. For
1 ≤ i ≤ n, put ai = ϕ(ti ); then (a1 , . . . , an ) ∈ Rn and
f (a1 , . . . , an ) = ϕ(f ) = −ϕ(w)2 < 0.
8. ARCHIMEDEAN AND COMPLETE FIELDS 159
Famously, R satisfies the least upper bound axiom, i.e., its ordering is Dedekind
complete. So by Proposition 14.49 the ordering on R is Archimedean. (Probably
the reader was not in doubt of this, but this is an especially clean approach.) Thus
every subfield of R is Archimedean.
The order topology: let (S, ≤) be any linearly ordered space. Recall that we can
use the ordering to endow S with a topology, the order topology, in which a base
of open sets consists of all open intervals.6 Order topologies have several pleasant
properties: for instance, any order topology is a hereditarily normal space (i.e.,
every subspace is normal: for us, this includes Hausdorff).
Proposition 14.50. Let K be an ordered field. Then the order topology endows
K with the structure of a topological field. That is, the addition and multiplication
operations are continuous as functions from K × K to K.
Exercise 14.31. Prove Proposition 14.50. (Suggestion: use the characteriza-
tion of continuous functions as those which preserve limits of nets.)
Proposition 14.51. For any Archimedean ordered field F , Q is dense in the
order topology on F .
Proof. It is sufficient to show that for a, b ∈ F with 0 < a < b, there exists
x ∈ Q with a < x < b. Because of the nonexistence of infinitesimals, there exist
x1 , x2 ∈ Q with 0 < x1 < a and 0 < x2 < b−a. Thus 0 < x1 +x2 < b. Therefore the
set S = {n ∈ Z+ | x1 + nx2 < b} is nonempty. By the Archimedean property S is
finite, so let N be the largest element of S. Thus x1 + N x2 < b. Moreover we must
have a < x1 + N x2 , for if x1 + N x2 ≤ a, then x1 + (N + 1)x2 = (x1 + N x2 ) + x2 <
a + (b − a) = b, contradicting the definition of N .
Exercise 14.32. Deduce from Proposition 14.51 that the order topology on any
Archimedean ordered field is second countable. (Hint: show in particular that open
intervals with rational endpoints form a base for the topology.) From the normality
of all order topologies cited above and Urysohn’s Metrization Theorem, it follows
that the order topology on an Archimedean ordered field is metrizable.7
The order topology on K endows (K, +) with the structure of a commutative
topological group. In such a situation we can define Cauchy nets, as follows: a
net x• : I → G in a commutative topological group G is Cauchy if for each
neighborhood U of the identity 0 ∈ G there exists i ∈ I such that for all j, k ≥ i,
xj − xk ∈ U . A topological group is complete if every Cauchy net converges.
Let F be an ordered field. We define the absolute value function from F to
F ≥0 , of course taking |x| to be x if x ≥ 0 and −x otherwise.
Exercise 14.33. Let F be an ordered field. Show that the triangle inequality
holds: for all x, y ∈ F , |x + y| ≤ |x| + |y|.
6If there is a bottom element b of S, then the intervals [b, b) are deemed open. If there is a
top element t of S, then the intervals (a, t] are deemed open. Of course, no ordering on a field has
either top or bottom elements, so this is not a relevant concern at present.
7However, we are not going to use this fact in our discussion. Rather, as will become clear,
an ordered field K comes with a canonical “K-valued metric”, which will be just as useful to us
as an “R-valued metric” – a special case!
8. ARCHIMEDEAN AND COMPLETE FIELDS 161
Thus for any ordered field F , one can define the function ρ : F × F → F ≥0 by
ρ(x, y) = |x − y| and this has all the formal properties of a metric except that it
is F -valued. In particular, for any net x• in F we have x• → x iff |x• − x| → 0.
In general it can be of some use to consider “F -valued metrics” where F is a
non-Archimedean ordered field. But here is the key point: if the ordering on F
is Archimedean, then the convergence can be expressed by inequalities involving
rational numbers (rather than the infinitesimal elements that would be required
in the non-Archimedean case): namely, for an Archimedean ordered field F , a net
x• : I → F converges to x ∈ F iff for all n ∈ Z+ , there exists in ∈ I such that
j ≥ i =⇒ |xj − x| < n1 . Topologically speaking, we are exploiting the fact that
the topology of an Archimedean ordered field has a countable neighborhood base
at each point. Thus it is sufficient to replace nets by sequences. In particular we
have the following simple but important result.
Lemma 14.52. Let K be an Archimedean ordered field. Then the following are
equivalent:
(i) Every Cauchy net in K is convergent.
(ii) Every Cauchy sequence in K is convegent.
Proof. Of course (i) =⇒ (ii). Now suppose that every Cauchy sequence in
K converges, and let x• : I → K be a Cauchy net. We may assume that I has no
maximal element, for otherwise the net is certainly convergent. Choose i1 ∈ I such
that j, k ≥ i1 implies |xj − xk | < 1. Now pick i2 ∈ I such that i2 > I1 and j, k ≥ i2
implies |xj − xk | < 12 . Continuing in this manner we get an increasing sequence
{in } in I such that for all n, if j, j ≥ in , |xj − xk | < n1 . Thus from the net we have
extracted a Cauchy subsequence, which by hypothesis converges, say to x. From
this it follows immediately that the net x• converges to x.
Remark: The proof here is based on [Wi, Thm. 39.4], which asserts that the
uniform structure associated to a complete metric is a complete uniform structure
iff the metric is a complete metric.
Theorem 14.53. For an Archimedean ordered field K, the following are equiv-
alent:
(i) The ordering on K is Dedekind complete: every nonempty subset which
is bounded below has a greatest lower bound.
(ii) (K, +) is a Cauchy-complete topological group: every Cauchy net con-
verges.
Proof. (i) =⇒ (ii): Dedekind complete implies Archimedean implies second
countable implies first countable implies it is enough to look at Cauchy sequences.
The argument is then the usual one from elementary real analysis: suppose K is
Dedekind complete, and let xn be a Cauchy sequence in K. Then the sequence is
bounded, so there exists a least upper bound x. We can construct a subsequence
converging to x in the usual way: for all k ∈ Z+ , let xnk be such that |xnk − x| < n1 .
(That this implies that the subseuence converges is using the the Archimedean
property that for all x > 0, there exists n ∈ Z+ with n1 < x.) Then, as usual, a
Cauchy sequence with a convergent subsequence must itself be convergent.
(ii) =⇒ (i): let S ⊂ K be nonempty and bounded below. Let B be the set of all
lower bounds of S, with the ordering induced from K. What we want to show is
that B has a greatest element: we will prove this by Zorn’s Lemma. Let C be a
162 14. ORDERED FIELDS
the same argument applies to show that any two isomorphisms ϕ1 , ϕ2 from R1 to
R2 are inverses of the isomorphism $, so ϕ1 = ϕ2 : there is only one isomorphism
from R1 to R2 .
We have already identified the real numbers R as a complete Archimedean field, so
we know that the final object referred to in Theorem 14.56 indeed exists. Let us
restate things in a more concrete fashion using R.
Corollary 14.57. For any Archimedean ordered field K, there is a unique
embedding of ordered fields K ,→ R. Thus we may identify the Archimedean ordered
fields – up to unique isomorphism – as subfields of R with the inherited ordering.
It may be worth asking at this point: exactly how do we know that this field “of
real numbers” we’ve heard so much about actually exists? We’ve proven some fairly
remarkable facts about it: maybe rumors of its existence are greatly exaggerated!
and (and hence also all the basis elements) are closed as well open: this implies
that X(F ) is totally disconnected and Hausdorff. It is also compact. To see this,
×
note that an ordering P of F gives rise to an element of Y = {±1}F , namely for
each nonzero element a, we assign +1 if a ∈ P and −1 if −a ∈ P . Giving {±1}
the discrete topology and Y the product topology, it is a compact Hausdorff totally
disconnected space by Tychonoff’s theorem. It remains to be shown first that the
topology on X(F ) defined above is the same as the topology it gets as a subspace
of Y 8, and second that X(F ) is closed as a subspace of Y . Neither of these is very
difficult and we leave them to the reader.
Exercise 14.36. Let F/Q be a (possibly infinite) formally real Galois exten-
sion. Show that Aut(F ) = Gal(F/Q) acts continuously and simply transitively on
X(F ), and conclude that in this case X(F ) is homeomorphic to the underlying
topological space of a profinite group. In particular, if F/Q is infinite, X(F ) is
an infinite profinite space without isolated points and with a countable basis, so is
√
homeomorphic to the Cantor set. (A good example is F = Q({ p}) as p ranges
over all the prime numbers: here Aut(F ) ∼ = (Z/2Z)ℵ0 really looks like the Cantor
set.)
Exercise 14.37. Use weak approximation of valuations to show that any in-
verse system
. . . → Sn+1 → Sn → . . . → S1
of finite sets can be realized as the system of X(Fn )’s where
F1 . . . ,→ Fn ,→ Fn+1 ,→ . . .
is a tower of number fields. Conclude that any profinite space with a countable basis
arises as the space of orderings of an algebraic field extension of Q.
8One might wonder why we didn’t save ourselves the trouble and define the topology on X(F )
in this latter way. It turns out that the sets H(a), called the Harrison subbasis, are important in
their own right.
Bibliography
[Ar44] E. Artin, Galois Theory. Second edition. Notre Dame Mathematical Lectures, no. 2.
University of Notre Dame, Notre Dame, Ind., 1944.
[BAI] N. Jacobson, Basic algebra. I. Second edition. W. H. Freeman and Company, New York,
1985.
[BAII] N. Jacobson, Basic algebra. II. Second edition. W. H. Freeman and Company, New
York, 1989.
[Bar51] D. Barbilian, Solution exhaustive du problème de Steinitz. Acad. Repub. Pop. Romn̂e.
Stud. Cerc. Mat. 2, (1951). 195–259 (misprinted 189–253).
[BJ01] F. Borceux and G. Janelidze, Galois theories. Cambridge Studies in Advanced Mathe-
matics, 72. Cambridge University Press, Cambridge, 2001.
[CG72] C.H. Clemens and P.A. Griffiths, The intermediate Jacobian of the cubic threefold.
Ann. of Math. (2) 95 (1972), 281–356.
[Ch] P.M. Cohn, Basic algebra. Groups, rings and fields. Springer-Verlag London, Ltd.,
London, 2003.
[Ch70] A. Charnow, The automorphisms of an algebraically closed field. Canad. Math. Bull.
13 (1970), 95–97.
[Cl-CA] P.L. Clark, Commutative Algebra. http://math.uga.edu/~pete/integral2015.pdf
[CL-NT] P.L. Clark, Number Theory: A Contemporary Introduction. http://math.uga.edu/
~pete/4400FULL.pdf
[Cl-NTII] P.L. Clark, Algebraic Number Theory II: Valuations, Local Fields and Adeles. http:
//math.uga.edu/~pete/8410FULL.pdf
[Cl-QF] P.L. Clark, Lecture notes on quadratic forms. http://alpha.math.uga.edu/~pete/
quadraticforms0.pdf
[Cr75] T.C. Craven, The Boolean space of orderings of a field. Trans. Amer. Math. Soc. 209
(1975), 225–235.
[CW50] J.W.S. Cassels and G.E. Wall, The normal basis theorem. J. London Math. Soc. 25
(1950), 259–264.
[DG94] M. Dugas and R. Göbel, Automorphism groups of fields. Manuscripta Math. 85 (1994),
no. 3-4, 227–242.
[DG97] M. Dugas and R. Göbel, Automorphism groups of fields. II. Comm. Algebra 25 (1997),
3777–3785.
[Di74] J. Dieudonné, Sur les automorphismes des corps algébriquement clos. Bol. Soc. Brasil.
Mat. 5 (1974), 123–126.
[FK78] E. Fried and J. Kollar, Automorphism groups of algebraic number fields. Math. Z. 163
(1978), 121–123.
[Gi68] R. Gilmer, Classroom Notes: A Note on the Algebraic Closure of a Field. Amer. Math.
Monthly 75 (1968), 1101—1102.
[Hö01] O. Hölder, Die Axiome der Quantitat und die Lehre vom Mass. Ber. Verh. Sachs. Ges.
Wiss. Leipzig, Math.-Phys. Cl. 53 (1901), 1–64.
[Ho20] X.-D. Hou, PGL2 (Fq ) acting on Fq (x). Comm. Algebra 48 (2020), 1640–1649.
[Is80] I.M. Isaacs, Roots of polynomials in algebraic extensions of fields. Amer. Math. Monthly
87 (1980), 543–544.
[Ja1] N. Jacobson, Basic algebra. I. Second edition. W. H. Freeman and Company, New York,
1985.
[Ja2] N. Jacobson, Basic algebra. II. Second edition. W. H. Freeman and Company, New
York, 1989.
165
166 BIBLIOGRAPHY
[Je] T.J. Jech, The axiom of choice. Studies in Logic and the Foundations of Mathematics,
Vol. 75. North-Holland Publishing Co., New York, 1973.
[Kap95] I. Kaplansky, Fields and rings. Reprint of the second (1972) edition. Chicago Lectures
in Mathematics. University of Chicago Press, Chicago, IL, 1995.
[Ka89] G. Karpilovsky, Topics in field theory. North-Holland Mathematics Studies, 155. Notas
de Matematica [Mathematical Notes], 124. North-Holland Publishing Co., Amsterdam,
1989.
[Kr53] W. Krull, Über eine Verallgemeinerung des Normalkörperbegriffs. J. Reine Angew.
Math. 191 (1953), 54–63.
[Ku65] W. Kuyk, The construction of fields with infinite cyclic automorphism group. Canad.
J. Math. 17 (1965), 665–668.
[La53] S. Lang, The theory of real places. Ann. of Math. (2) 57 (1953), 378–391.
[LaFT] S. Lang, Algebra. Revised third edition. Graduate Texts in Mathematics, 211. Springer-
Verlag, New York, 2002.
[Las97] D. Lascar, The group of automorphisms of the field of complex numbers leaving fixed
the algebraic numbers is simple. Model theory of groups and automorphism groups
(Balubeuren, 1995), 110-114, London Math. Soc. Lecture Note Ser., 244, Cambridge
Univ. Press, Cambridge, 1997.
[Le55] H. Leptin, Ein Darstellungssatz für kompakte, total unzusammenhngende Gruppen.
Arch. Math. (Basel) 6 (1955), 371–373.
[Lev43] F.W. Levi, Contributions to the theory of ordered groups. Proc. Indian Acad. Sci., Sect.
A. 17 (1943), 199–201.
[LoI] F. Lorenz, Algebra. Vol. I. Fields and Galois theory. Translated from the 1987 German
edition by Silvio Levy. With the collaboration of Levy. Universitext. Springer, New
York, 2006.
[LoII] F. Lorenz, Algebra. Vol. II. Fields with structure, algebras and advanced topics. Trans-
lated from the German by Silvio Levy. With the collaboration of Levy. Universitext.
Springer, New York, 2008.
[Mac39] S. Mac Lane, Steinitz field towers for modular fields. Trans. Amer. Math. Soc. 46,
(1939). 23–45.
[Mc67] P.J. McCarthy, Maximal fields disjoint from certain sets. Proc. Amer. Math. Soc. 18
(1967), 347–351.
[Mo96] P. Morandi, Field and Galois theory. Graduate Texts in Mathematics, 167. Springer-
Verlag, New York, 1996.
[Mu61] D. Mumford, Pathologies of modular algebraic surfaces. Amer. J. Math. 83 (1961),
339–342.
[Mu62] D. Mumford, Further pathologies in algebraic geometry. Amer. J. Math. 84 (1962),
642–648.
[Mu67] D. Mumford, Pathologies. III. Amer. J. Math. 89 (1967), 94–104.
[Mu75] D. Mumford, Pathologies IV. Amer. J. Math. 97 (1975), 847–849.
[Pa74] T. Parker, Some applications of Galois theory to normal polynomials. Amer. Math.
Monthly 81 (1974), 1009–1011.
[Pi09] D. Pierce, https://dialinf.wordpress.com/2009/08/28/
completions-and-the-archimedean-property/
[Qu62] F. Quigley, Maximal subfields of an algebraically closed field not containing a given
element. Proc. Amer. Math. Soc. 13 (1962), 562–566.
[Rom06] S. Roman, Field theory. Second edition. Graduate Texts in Mathematics, 158. Springer,
New York, 2006.
[Rot98] J. Rotman, Galois theory. Second edition. Universitext. Springer-Verlag, New York,
1998.
[Sc92] B. Schnor, Involutions in the group of automorphisms of an algebraically closed field.
J. Algebra 152 (1992), 520–524.
[Si] J.H. Silverman, The arithmetic of elliptic curves. Second edition. Graduate Texts in
Mathematics, 106. Springer, Dordrecht, 2009.
[Sp52] T.A. Springer, Sur les formes quadratiques d’indice zéro. C. R. Acad. Sci. Paris 234
(1952), 1517–1519.
[St10] E. Steinitz, Algebrasiche Theorie der Körper. Journal für die reine und angewandte
Mathematik, 1910.
BIBLIOGRAPHY 167
[Su99] B. Sury, On an example of Jacobson. Amer. Math. Monthly 106 (1999), 675-–676.
[Wa74] W.C. Waterhouse, Profinite groups are Galois groups. Proc. Amer. Math. Soc. 42
(1974), 639–640.
[Wa94] W.C. Waterhouse, The normal closures of certain Kummer extensions. Canad. Math.
Bull. 37 (1994), 133–139.
[Wa04] W.C. Waterhouse, Intersections of two cofinite subfields. Arch. Math. (Basel) 82 (2004),
298–300.
[We09] S.H. Weintraub, Galois theory. Second edition. Universitext. Springer, New York, 2009.
[Wi] S. Willard, General topology. Reprint of the 1970 original. Dover Publications, Inc.,
Mineola, NY, 2004.
[Ya66] P.B. Yale, Automorphisms of the Complex Numbers. Math. Magazine 39 (1966), 135–
141.
[Za58] O. Zariski, On Castelnuovo’s criterion of rationality pa = P2 = 0 of an algebraic
surface. Illinois J. Math. 2 (1958), 303–315.
[ZSI] O. Zariski and P. Samuel, Commutative algebra. Vol. 1. With the cooperation of I. S.
Cohen. Corrected reprinting of the 1958 edition. Graduate Texts in Mathematics, No.
28. Springer-Verlag, New York-Heidelberg-Berlin, 1975.