Knebusch M Scheiderer C Real Algebra A First Course
Knebusch M Scheiderer C Real Algebra A First Course
Knebusch M Scheiderer C Real Algebra A First Course
Manfred Knebusch
Claus Scheiderer
Real Algebra
A First Course
Translated and with Contributions by
Thomas Unger
Universitext
Series Editors
Carles Casacuberta, Universitat de Barcelona, Barcelona, Spain
John Greenlees, University of Warwick, Coventry, UK
Angus MacIntyre, Queen Mary University of London, London, UK
Claude Sabbah, École Polytechnique, CNRS, Université Paris-Saclay, Palaiseau,
France
Endre Süli, University of Oxford, Oxford, UK
Universitext is a series of textbooks that presents material from a wide variety of
mathematical disciplines at master’s level and beyond. The books, often well class-
tested by their author, may have an informal, personal even experimental approach
to their subject matter. Some of the most successful and established books in the
series have evolved through several editions, always following the evolution of
teaching curricula, into very polished texts.
Thus as research topics trickle down into graduate-level teaching, first textbooks
written for new, cutting-edge courses may make their way into Universitext.
Manfred Knebusch • Claus Scheiderer
Real Algebra
A First Course
Translated by
Thomas Unger
School of Mathematics and Statistics
University College Dublin
Dublin, Ireland
Translation from the German language edition: “Einführung in die reelle Algebra” by Manfred Knebusch
et al., © Friedr. Vieweg & Sohn Verlagsgesellschaft mbH 1989. Published by Friedr. Vieweg & Sohn,
Braunschweig/Wiesbaden. All Rights Reserved.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
More than 30 years after its publication, we are pleased and surprised with the
ongoing interest in Einführung in die reelle Algebra. Given the vibrant development
of real algebra and geometry in the previous decades, this seems by no means self-
evident.
Real algebra has grown in many directions, partially from within itself, and
exceedingly also through nudges and stimuli from without. While some of these
developments were already discernible at the end of the 1980s, many modern ones
that currently belong to the core of the area—such as the connections with tropical
geometry, or with semidefinite optimization—would have been difficult to predict.
Hence, it is not obvious that a text that was timely and modern in 1989 still serves
current needs in a reasonable manner.
Nevertheless, we are convinced that also from today’s perspective the choice
and exposition of the material in Einführung in die reelle Algebra still constitute a
solid first course on the basic principles and important techniques of real algebra.
For this reason, we decided to keep the text essentially unchanged in the English
translation—even though it was tempting to change the wording here and there, and
to drop some sections in favour of new ones.
A small number of typos, omissions and errors that were present in the German
original have been corrected, and some notation and terminology have been brought
up to date. Several new examples (some in the form of exercises) have been added.
All definitions and propositions have been numbered in a consistent manner. The
original bibliography has been augmented with a number of newer references.
A feature of this translation is the addition of a short fourth chapter that provides
a succinct overview of the most important developments and advances in those parts
of real algebra that are directly related to material covered in Einführung in die reelle
Algebra.
We are grateful to Springer Nature, and Rémi Lodh in particular, for their
endorsement and friendly and professional assistance.
We received support and encouragement from many quarters, and are pleased to
extend special thanks to Oliver Hien and Marcus Tressl, who made valuable remarks
on the German original.
v
vi Preface
Finally, our very special and heartfelt thanks go to Thomas Unger, who translated
the original text in a conscientious and competent manner, and co-authored the final
chapter. It was always a great pleasure to work with him.
Algebra textbooks that are currently in common use present real algebra only in
later chapters and then usually rather tersely. This is so for the influential works
of van der Waerden [113], Jacobson [55] and Lang [73], although Jacobson covers
a bit more than the others. Bourbaki’s substantial multivolume text Eléments de
Mathématique shows a similar picture: the volume Algèbre contains just a short
chapter (Chapter 6, Groupes et corps ordonnés) on real algebra. In contrast, a
complete volume (currently counting nine chapters) is dedicated to commutative
algebra, even though on the whole only the elementary part is covered, and not
more than what is absolutely essential for the foundation of the theory, both by
current standards as well as Bourbaki’s own standards.
That being the case, not too many algebraists today seem to perceive real algebra
as a proper branch of algebra at all. This has not always been so. In the nineteenth
century real algebra flourished. The study of real zeroes of a real polynomial in one
variable was at the centre of algebraic interest during the whole century, and was an
indispensable part of any higher mathematical education.
Heinrich Weber’s large three-volume textbook on algebra [114] provides evi-
dence for this fact. As Weber’s research interests were primarily in number theory,
especially in complex multiplication and class field theory, he focussed his textbook
mostly on these topics. Nevertheless he devoted well over a hundred pages in the
first volume to real zeroes of real polynomials.
The twentieth century witnessed a dramatic decline in interest in real algebra.
This trend only seems to have changed since the late 1970s, which is all the more
astonishing since the essential seeds of a modern real algebra, as we understand it
today, are already present in two works by Artin and Schreier from the 1920s [3, 4].
So, what is real algebra? Instead of giving a formal definition—which would
be difficult—we rather answer the question with an analogy that puts real algebra
in parallel with commutative algebra. Commutative algebra can be seen as that
part of algebra that contains the algebraic foundations that are typically important
for algebraic geometry (in particular in its modern, abstract form); and algebraic
geometry is ultimately the study of solution sets of systems of polynomial equations
F (x1 , . . . , xn ) = 0 and non-equations F (x1 , . . . , xn ) = 0. Correspondingly,
vii
viii Preface to Einführung in die reelle Algebra
real algebra provides algebraic methods that typically function as tools for real
algebraic geometry, and in particular semialgebraic geometry, which is the study
of the solution sets of systems of polynomial inequalities F (x1 , . . . , xn ) > 0 or
F (x1 , . . . , xn ) ≥ 0, where the coefficients classically come from the field of real
numbers and more generally from any ordered field.
This analogy provides an argument for why the interest has been so much
stronger in commutative than in real algebra during our century thus far. Indeed,
algebraic geometry experienced a continuous and ultimately triumphant upswing in
the twentieth century, while real algebraic geometry was only practiced in isolation
and then mostly with transcendental methods. Thus, from the geometric point of
view, there was for many decades no engine available that could have pushed real
algebra forward.
It was not until 1987 that the first textbook [11] on real algebraic geometry
was published. Its authors’ report (in the last section of the introduction to this
commendable work [11, p. 4 ff.]) on the strange sleeping beauty slumber of real
algebraic geometry (insofar as it was practiced with algebraic methods) is worth
considering.
In the meantime this slumber has given way to a lively development. We
consider the introduction—or better: discovery—of the real spectrum Sper A of a
commutative ring A by Michel Coste and Marie-Françoise Roy around the year
1979 as the most important trigger.
One can view Sper A as an analogue of the Zariski spectrum Spec A, intro-
duced by Grothendieck. It is well-known that the Zariski spectrum is the key
to Grothendieck’s abstract algebraic geometry. Likewise, the real spectrum is the
key to an abstract semialgebraic geometry. (There is one difference: in contrast to
Spec A, the real spectrum Sper A seems to carry (at least) two important structure
sheafs, the sheaf of abstract Nash functions, introduced by Coste and Roy [29], as
well as the sheaf of abstract semialgebraic functions, introduced by G. Brumfiel and
N. Schwartz [19, 107, 108].)
Real algebra is necessary to understand the latest developments in real algebraic
geometry and to meet the future intellectual challenges in this area. This brings us
to the objective of this book.
Our book is based on two insights: real algebra is a branch of algebra that is
largely autonomous in its foundations, with methods specific to this branch. These
foundations can be successfully taught with little more preparation than the standard
background from a typical one-semester algebra course on linear algebra, group
theory, field theory and ring theory.
We differentiate more precisely between an elementary and a higher real algebra.
The former can be developed without any special prior knowledge and used with
benefit in real algebraic geometry. The latter makes serious use of resources from
other branches of mathematics, especially real algebraic geometry, commutative
algebra, algebraic geometry, model theory and the theory of quadratic forms, but
occasionally also algebraic topology, real analysis and complex analysis. (This list
can certainly be extended.) An analogous distinction can be made in commutative
algebra. A demarcation between “elementary” and “higher” in either area is not
Preface to Einführung in die reelle Algebra ix
entirely objectively possible, but rather is subject to personal points of view and
taste. Furthermore, the higher one goes, the more fluid and arbitrary the boundary
between the two kinds of algebra and their corresponding geometries becomes.
Our book is dedicated to elementary real algebra in the above sense. Another
book on higher real algebra is planned.1 In the current book we do get by with
previous knowledge of the extent sketched above. By adding another twenty to
thirty pages, we could have reduced the requirements further and, for example,
developed everything that is needed from commutative algebra and the theory of
quadratic forms. We have not done so however, since students who are interested in
real algebra more than likely already mastered almost all of the prerequisites for the
current book, and would easily be able to find the few things they may be missing
elsewhere.
One specific feature of real algebra should be pointed out in particular: the major
role that general (Krull) valuation rings play. In most parts of commutative algebra,
valuation rings that are not discrete are only viewed as an aid that can often be
done without. In real algebra on the other hand, general valuation rings are a natural
and even central concept throughout. The reason is that every convex subring of
an ordered field is a valuation ring, but only in rare cases a discrete valuation
ring. This fact was already observed by Krull in the introduction to his pioneering
work Allgemeine Bewertungstheorie [67]. (It can already be found in embryonic
form, without the concept of valuation ring, in the work of Artin and Schreier,
cf. [4, p. 95].) Krull also identified the theory of ordered fields as an important
application area of general valuation theory, but then devoted only one section—
albeit substantial—of his great work to ordered fields [67, §12].
Our book is divided into three chapters. In addition to the Artin-Schreier theory
of ordered fields and the elementary relationships between orderings and quadratic
forms, the first chapter covers some aspects of real algebra from the nineteenth
century. Various methods for determining the number of real zeroes of a real
polynomial are treated (Sturm’s algorithm, Hermite’s method using quadratic forms,
Hurwitz’ Theorem). Another section is devoted to the relationship between the
Cauchy index and the Hankel form, as well as the Bézoutian of a rational function.
The second chapter deals with real valuation theory. Everything we need from
general valuation theory is developed from scratch. The chapter culminates in a
presentation of Artin’s solution of Hilbert’s 17th Problem.
Finally, the third chapter is devoted to the real spectrum. After a short crash
course on the Zariski spectrum, the real spectrum is examined in detail, but only
as a topological space (i.e., without the introduction of a structure sheaf). In the
geometric setting (affine algebras over real closed fields), the points as well as
certain subsets of the real spectrum are described in terms of filter sets. In the last
five sections of the chapter we come to some parts of real algebra in which the real
spectrum proves to be clarifying and helpful, such as the reduced Witt ring of a field,
Regensburg
January 1989
Contents
xi
xii Contents
References .. .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 195
Symbol Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 201
Index . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . 203
Chapter 1
Ordered Fields and Their Real Closures
Our starting point is the basic notion for the entire book, the general concept of
orderings of arbitrary fields. Conceived by Artin and Schreier in their foundational
1927 paper [4], it was successfully used by Artin in his solution of Hilbert’s 17th
Problem in the same year [3]. We introduce real closed fields and show that they
have the same algebraic properties as the field of real numbers. Moreover we prove
the existence and uniqueness of a real closure for any ordered field.
The second main topic in this chapter concerns methods for counting real roots
of real polynomials. Instead of trying to calculate or estimate the roots numerically,
we discuss purely algebraic methods that extend from the real numbers to any real
closed base field. These results were mostly found in the nineteenth century or even
earlier, and are connected to famous names such as Descartes, Sturm, Sylvester and
Hermite.
Parallel to these two subjects we give an introduction to the basic notions of
quadratic forms over fields and their algebraic theory. They will find applications in
Chaps. 2 and 3.
Let K be a field.
Definition 1.1.1 An ordering of K is a subset P of K that satisfies the properties
(O1) P + P ⊆ P , P P ⊆ P ,
(O2) P ∩ (−P ) = {0},
(O3) P ∪ (−P ) = K,
where P + P := {a + b : a, b ∈ P } and P P := {ab : a, b ∈ P }. The pair (K, P )
is called an ordered field.
Remark 1.1.2 If we assume (O1) and (O3), then (O2) is equivalent with
(O2’) −1 ∈ P .
a ≤P b :⇔ b − a ∈ P (a, b ∈ K)
defines a total order relation ≤P on the set K which satisfies the properties
(i) a ≤P b ⇒ a + c ≤P b + c,
(ii) a ≤P b, c ≥ 0 ⇒ ac ≤P bc
for all a, b, c ∈ K. Conversely, let ≤ be a total order relation on K which satisfies
(i) and (ii). Show that P := {a ∈ K : a ≥ 0} is an ordering of K.
Thus we clearly have a one-one correspondence between orderings of K and total
order relations on K that satisfy (i) and (ii). Usually the latter are called orderings
of K as well. When P is clear we simply write a ≤ b instead of a ≤P b.
Definition 1.1.4 A preordering of K is a subset T of K that satisfies the properties
(P1) T + T ⊆ T , T T ⊆ T ,
(P2) T ∩ (−T ) = {0},
(P3) a 2 ∈ T for all a ∈ K.
a ≤T b :⇔ b − a ∈ T (a, b ∈ K)
for all a ∈ K.
4 1 Ordered Fields and Their Real Closures
P = {f ∈ Q(t) : f (ϑ) ≥ 0}
and
Pa,− := {0} ∪ (a − t)r f (t) : r ∈ Z, f ∈ F (t) with f (a) = ∞ and f (a) > 0
are orderings of F (t) for every a ∈ F . Both orderings extend the ordering of F .
The reason for this notation is that a < t < b with respect to Pa,+ holds for all
b ∈ F with b > a. Thus F (t) is ordered in such a way that the transcendental t
is located “immediately to the right of a” on the “line” F . Similarly, t is located
“immediately to the left of a” with respect to Pa,− .
Notation 1.1.14 If (K, P ) is an ordered field, we denote the sign function with
respect to P by signP : K → {−1, 0, 1}. Thus, signP (0) = 0, and for a ∈ K ∗ we
have signP (a) = 1 if a ∈ P and signP (a) = −1 if a ∈ P .
As usual, | · |P : K → P , |a|P := a · signP (a) denotes the absolute value. If P
is clear, the index P may be omitted.
The meaning of the generalized intervals [a, b]P , [a, b[P , ]a, b]P , ]a, b[P for
a, b ∈ K ∪ {−∞, +∞} is also clear, namely [a, b[P = {x ∈ K : a ≤P x <P b},
]a, ∞[P = {x ∈ K : x >P a}, etc. Here too we will usually drop the index P , and
instead often write [a, b]K , etc. if several fields are considered.
1.2 Quadratic Forms, Witt Rings, Signatures 5
n
qB (x) = bij xi xj x = (x1 , . . . , xn ) ∈ K n .
i,j =1
a1 , . . . , am ⊥ b1 , . . . , bn ∼
= a1 , . . . , am , b1 , . . . , bn
6 1 Ordered Fields and Their Real Closures
and
a1 , . . . , am ⊗ b1 , . . . , bn ∼
= ⊥a b .
i,j
i j
spaces that are Witt equivalent to ϕ by [ϕ] (the Witt class of ϕ), and the set of all
Witt classes of (nondegenerate) quadratic spaces over K by W (K).
Theorem 1.2.4 The set W (K) is made into a commutative unitary ring with unit
[1] by the (well-defined) pairings
Example 1.2.7 (Milnor’s Exact Sequence) For every prime number p, we denote
the finite field with p elements by Fp . Then W (Q) can be computed explicitly as an
additive group by the split exact sequence
where i : Z → W (Q) is the unique homomorphism that sends 1 to 1, and ∂2 and
the ∂p for odd primes p are certain group homomorphisms, cf. [72, VI, §4] for the
details. Witt groups of number fields are considerably harder to understand, cf. [72,
VI, §3]. There is a similar split exact sequence for Witt groups of rational function
fields in one variable, cf. [72, IX, §3].
Next we study the relations between the Witt ring of K and orderings of K.
Definition 1.2.8 Let (K, P ) be an ordered field and let ϕ = (V , q) be a quadratic
space over K. Then ϕ (or q) is called positive definite with respect to P if q(v) >P 0
for all 0 = v ∈ V . If −q is positive definite, then q is called negative definite.
The following theorem is well-known:
Theorem 1.2.9 (Sylvester’s Inertia Theorem) Let ϕ be a possibly degenerate
quadratic space over K and P an ordering of K. Then there is an orthogonal
decomposition
ϕ∼
= ϕ+ ⊥ ϕ− ⊥ Rad(ϕ),
Lemma 1.2.11 Let (K, P ) be an ordered field. Then [ϕ] → signP ϕ defines a ring
homomorphism signP from W (K) to Z.
Proof The following equalities are easily verified:
a, b ∼
= a + b, (a + b)ab .
Proof Since a + b is represented by a, b, there exists c ∈ K ∗ such that a, b ∼
=
a + b, c. Comparing determinants gives abK ∗2 = (a + b)cK ∗2, i.e., cK ∗2 =
(a + b)abK ∗2. The statement follows.
Theorem 1.2.12 justifies:
Definition 1.2.14 A signature of a field K is a ring homomorphism W (K) → Z.
10 1 Ordered Fields and Their Real Closures
Remark 1.2.15 We see that a field admits a signature if and only if it is real.
Note that for every field K (with char K = 2) there exists a ring homomorphism
e : W (K) → Z/2Z, defined by e[ϕ] := dim(ϕ) + 2Z. Every signature σ yields a
commutative diagram
Definition 1.2.16 The map e : W (K) → Z/2Z is called the dimension index. Its
kernel is denoted I (K) and is called the fundamental ideal of W (K). (Thus I (K)
consists of the Witt classes of even dimensional quadratic spaces.)
In the following results we consider field extensions for which every ordering
extends:
Proposition 1.3.5 Let L/K be a finite field extension of odd degree. Then every
ordering of K extends to L.
By Proposition 1.3.2, this proposition follows from
Theorem 1.3.6 (T.A. Springer) If L/K is a finite field extension of odd degree and
q is an anisotropic quadratic form over K, then qL is anisotropic over L.
Proof We may assume without loss of generality that L = K(α) is a simple field
extension. Let f ∈ K[t] be the minimal polynomial of α over K. We proceed
by induction on n = [L : K] = deg f . Assume that n > 1 is odd and that the
statement is true for all smaller degrees. Let q = a1 , . . . , am be an anisotropic
form over K. Assume for the sake of contradiction that qL is isotropic. Then there
exist g1 , . . . , gm , h ∈ K[t] with deg gi < n (i = 1, . . . , m) and not all gi = 0, such
that
Remark 1.3.9 From Theorem 1.3.6 and Proposition 1.3.8 it follows in particular
that the homomorphism iL/K : W (K) → W (L) is injective in case L/K is finite of
odd degree or purely transcendental.
1.4 The Prime Ideals of the Witt Ring 13
σ
that p = Ker W (K) − → Z → Z/pZ , i.e., that p = pσ + p · W (K). It is clear that
σ and p are uniquely determined by p. This establishes the proof.
Since in every (commutative) ring the nilradical (i.e., the set of all nilpotent
elements) is equal to the intersection of all prime ideals ([55, 68, 73], see also
Proposition 3.1.11), we obtain immediately
14 1 Ordered Fields and Their Real Closures
Having determined the prime ideals of W (K), we can state a further criterion for
when orderings extend:
Proposition 1.4.4 Let L/K be a field extension. A signature of K extends to L if
and only if it vanishes on the kernel of iL/K : W (K) → W (L).
For the proof we require a simple fact from commutative algebra:
Lemma 1.4.5 If φ : A → B is an injective ring homomorphism and p a minimal
prime ideal of A, then there exists a prime ideal q of B such that p = φ −1 (q).
Proof Since S := φ(A\p) does not contain zero, we have that S −1 B = 0 and every
prime ideal q of S −1 B satisfies p = φ −1 j −1 (q ) , where j : B → S −1 B denotes
the canonical homomorphism.
Proof of Proposition 1.4.4 Let I := Ker iL/K and let σ be a signature of K. If σ
has an extension τ , then σ = τ ◦ iL/K , i.e., σ (I ) = 0.
Conversely, assume that σ (I ) = 0, i.e., I ⊆ pσ := Ker σ . Then pσ /I is a
minimal prime ideal of W (K)/I by Theorem 1.4.1. By Lemma 1.4.5 (applied to
−1
W (K)/I → W (L)) there exists a prime ideal q of W (L) such that pσ = iL/K (q).
Since Z ∼= W (K)/pσ → W (L)/q and again by Theorem 1.4.1, it follows that
W (K)/pσ → W (L)/q is an isomorphism, i.e., q determines an extension τ of σ .
In Sect. 1.3 we proved the injectivity of the map iL/K for extensions L/K that are
of odd degree or that are purely transcendental. Next, we will determine the kernel
of iL/K for quadratic extensions (and so with Proposition 1.4.4 give a new proof of
Proposition 1.3.3):
√
Proposition 1.4.6 Let a ∈ K ∗ \ K ∗2 and L = K( a). Then the kernel of iL/K is
the ideal of W (K) generated by 1, −a.
A more precise statement is:
Proposition 1.4.7 Let q be an anisotropic quadratic form over K. There exist
quadratic forms q , q over K with q ∼ = q ⊥ 1, −a ⊗ q such that qL is
anisotropic.
1.5 Real Closed Fields and Their Field Theoretic Characterization 15
Proof By induction on dim q. The case dim q = 1 is clear. Assume thus that
dim q ≥ 2 and without loss of generality that qL is isotropic. If (V , b) denotes
the bilinear space associated to q, then there exist v, w ∈ V , not both zero, such that
√ √
0 = qL (v + a w) = q(v) + a · q(w) + a · b(v, w).
It follows that q(v)+a q(w) = 0 = b(v, w). Since q is anisotropic, q(v), q(w) = 0.
Let W := Kv + Kw ⊆ V . Then dim W = 2 (if v, w were linearly dependent, then
we would have b(v, w) = 0), and q|W ∼= −ac, c = 1, −a⊗c for c := q(w). In
particular, q|W is nondegenerate. Since V = W ⊥ W ⊥ , we can apply the induction
hypothesis to q|W ⊥ and so conclude the proof.
After our excursion into the theory of quadratic forms we return to real algebra in
the narrower sense. This section is fundamental for everything that follows.
Definition 1.5.1 A field is called real closed if it is real and has no proper algebraic
extension that is real.
The field R of real numbers is a well-known example of a real closed field.
Proposition 1.5.2 Let K be a field. The following statements are equivalent:
(i) K is real closed;
(ii) there exists an ordering P of K that cannot be extended to any proper algebraic
extension of K;
(iii) K ∗2 ∪ {0} = {a 2 : a ∈ K} is an ordering of K, and every polynomial (in one
variable) of odd degree over K has a zero in K.
Moreover, if these conditions are satisfied, then K ∗2 ∪ {0} is the only ordering of K.
Proof The final comment is clear by (iii) (every ordering P satisfies K ∗2 ∪{0} ⊆ P ).
(i) ⇒ (ii) is clear since a field has an ordering if (and only if) it is real.
(ii)√⇒ (iii): If it were true that K ∗2 ∪ √{0} = P , then there would exist d ∈ P
with d ∈ K, and P would extend to K( d) (Proposition 1.3.3), a contradiction.
Furthermore, K has no proper extensions of odd degree by Proposition 1.3.5.
(iii) ⇒ (i): Let L ⊃ K be a finite proper field extension of K. We will show that
L is not real. By (iii), [L : K] is a power of 2. Since K has an ordering by (iii), it
follows that char K = 0 and in particular that L/K is separable. Let L1 be the Galois
L over K, G = Gal(L1 /K) and H = Gal(L1 /L). Then H is a subgroup
closure of
of G, and g∈G H g = {1} by Galois theory, −1
where H g := g H g. By elementary
g
group theory it follows that |G| divides g∈G [G : H ] . By assumption (iii) this
number is a power of 2, and so G is a 2-group. From Sylow’s Theorems (elementary
16 1 Ordered Fields and Their Real Closures
group theory again) it follows that there exists a subgroup H of G of index 2 with
H ⊂ H . Let F be the fixed field of H . Then K ⊂ F ⊂ L and [F : K] = 2.
√
There exists a ∈ K with F = K( a). Since K ∗2 ∪ {0} is an √ ordering of K and
a∈ / K ∗2 ∪{0}, there exists b ∈ K ∗ with a = −b2 . Hence −1 = ( a/b)2 is a square
in F , and so F is not real. We conclude that L is not real either.
Note that ψ is additive, i.e., ψ(b + b ) = ψ(b) + ψ(b ). Consider the trace
tr = trK/L : K → L. Then tr ◦ψ = (ψ|L )◦tr. Indeed, if b ∈ K and if b1 , . . . , bp are
the L-conjugates of b in K, then ψ(b1 ), . . . , ψ(bp ) are the L-conjugates of ψ(b),
and it follows that tr (ψ(b)) = ψ(bi ) = ψ( bi ) = ψ(tr b). Hence the diagram
i i
commutes. Since ψ and tr are surjective (the second map since K/L is separable),
the map ψ|L : L → L is also surjective, a contradiction.
(3) q = 2 (and char K = 2).
L contains the q-th roots of unity since [K : L] = q and the q-th cyclotomic
polynomial is of degree q − 1. Hence, since char K = q, there exists a ∈ L with
2
K = L(a 1/q ) (see for instance [55, Vol. I, §4.7] or [73, VI, §6]). For β := a 1/q we
obtain as in (1) that (NK/L (β))q = (−1)q−1 a. Since a is not a q-th power in L, q
must be even, thus q = 2.
√
(4) K = L( −1).
√
Let i = −1 ∈ K, and let a ∈ L and β ∈ K as in (3), thus with √ β = a
4
and K = L(β ). Since NK/L (β) = −β , we have β /NK/L (β) = −1, and so
2 2 4 2
√
K = L( −1).
(5) L = K, thus K = K(i).
If K = L, then also K(i) = K. Carrying out steps (1) to (4) for K(i) instead of
K leads to a contradiction by (4).
(6) K is real (and thus real closed).
Since i ∈ K it suffices to show that the sum of two squares in K is a square
(Proposition 1.5.2). Thus, let a, b ∈ K ∗ . Since K is algebraically closed, there are
c, d ∈ K with a+bi = (c+di)2. Hence, a−bi = (c−di)2 and a 2 +b 2 = (c2 +d 2 )2 .
This proves the theorem.
Let Ks denote the separable closure of K.
Corollary 1.6.2 If char K = 0 and [Ks : K] < ∞, then K = Ks .
Proof Steps (2)–(6) of the proof can be carried out for Ks just as for K.
We finish this section with a consequence of the theorem of Artin and Schreier
for absolute Galois groups of fields. These are profinite groups (i.e., projective
limits of finite groups). For the basic aspects of infinite Galois theory, [54, Vol. III,
§IV.2], [55, Vol. II, §8.6] or [12, Ch. V, §10], for example, can be consulted, but for
understanding what follows this is not required.
1.7 Counting Real Zeroes of Polynomials (without Multiplicities) 19
Corollary 1.6.3 Let K be a field with separable algebraic closure Ks and absolute
Galois group = Gal (Ks /K). All elements of finite order > 1 in are involutions
(i.e., have order 2), and the map τ → Fix (τ ) (the fixed field of τ in Ks = K) is a
bijection from the set of involutions in to the set of real closed overfields of K in
K. In particular, contains elements of finite order > 1 if and only if K is real.
One of the oldest problems in real analysis consists of determining the number
and location of the real zeroes of a polynomial f (t) with real coefficients. The
first general answer was formulated by J. C. F. Sturm (1803–1855) in the form of
an algorithm, the so-called Sturm sequence, that determines the number of zeroes
of f in any given interval, counted without multiplicities. Another solution of
the problem, based on the examination of certain quadratic forms, was found by
Ch. Hermite (1822–1901). Both methods will be presented in this section.
It should be remarked that these results used to be part of the mathematics
curriculum at the turn of the nineteenth century (see for example H. Weber’s
Lehrbuch der Algebra [114]).
Sturm’s solution, found in 1829, initiated an historically very interesting and to
some extent stormy period in real algebra. (Sturm published the full proof only
in 1835.) Hermite’s theorem (Theorem 1.7.17) was anticipated by C.G.J. Jacobi
(1804–1851) and his student and friend C.W. Borchardt (1817–1880) (Theo-
rem 1.7.15), and was proved in 1853 by Hermite and—essentially independently—
also by J.J. Sylvester (1814–1897). In addition, Sylvester, A. Cayley (1821–1895)
and Sturm established interesting links between the theorems of Sturm and Hermite.
In addition to the original literature, the reader can find full particulars of this history
in the substantial text [65] of Krein and Naimark, as well as in [39].
We start with a number of classical results, well-known from real analysis, that
stay valid for rational functions over arbitrary real closed fields.
In the remainder of this section R denotes a real closed field.
Proposition 1.7.1 Let f (t) = t n +a1 t n−1 +· · ·+an ∈ R[t] be a monic polynomial.
Then all real zeroes of f in R) are in the interval [−M, M],
of f (i.e., all zeroes
where M := max 1, |a1 | + · · · + |an | .
Proof If 0 = a ∈ R is such that f (a) = 0, then a = −(a1 + a2a −1 + · · ·+ an a 1−n ),
and so |a| ≤ 1 or |a| ≤ |a1 | + · · · + |an | by the triangle inequality.
20 1 Ordered Fields and Their Real Closures
f (a)
r
a − aj
−1 = sign = sign .
f (b) b − aj
j =1
Proposition 1.7.4 Let 0 = f ∈ R(t) and a, b ∈ R with a < b and f (a) = f (b) =
0. If f has neither poles nor zeroes in (a, b), then the number of zeroes of f in
]a, b[, counted with multiplicities, is odd.
f (a + ε) f (b − ε)
· < 0,
f (a + ε) f (b − ε)
f0 = q1 f1 − f2 ,
f1 = q2 f2 − f3 ,
..
.
fr−2 = qr−1 fr−1 − fr ,
fr−1 = qr fr ,
For x ∈ R let W (x) := Var (g0 (x), . . . , gr (x)). We claim that g0 has exactly W (a)−
W (b) distinct zeroes in [a, b].
To prove the claim, we first note that for every c ∈ R with g0 (c) · · · gr (c) = 0
the function W (x) is constant on a neighbourhood of x = c. Let g := g0 · · · gr , and
let c ∈ [a, b] be a zero of g. We distinguish three cases:
If g0 (c) = 0, then a < c < b, and the function x → Var (g0 (x), g1 (x)) has value
1 on ]c − ε, c[ and value 0 on [c, c + ε[, for some ε > 0. This follows from (c).
If gi (c) = 0, 0 < i < r, then by (d) there exists ε > 0, such that gi−1 (x)gi+1 (x)
is negative on ]c − ε, c + ε[. Hence, x → Var (gi−1 (x), gi (x), gi+1 (x)) is constant
equal to 1 on ]c − ε, c + ε[ (note that for every α ∈ R, Var(−1, α, 1) =
Var(1, α, −1) = 1 !).
If gr (c) = 0, and g0 (c) = 0 or r > 1, then gr−1 (c) = 0 by (d), and x →
Var (gr−1 (x), gr (x)) is constant on {x : 0 < |x − c| < ε} for some ε > 0, since gr
is semidefinite on [a, b].
From these three cases we obtain for all c ∈ [a, b]: If g0 (c) gr (c) = 0, then W (x)
is constant on a neighbourhood of x = c; if g0 (c) = 0, then W (x) is constant on
a punctured neighbourhood of x = c; if g0 (c) = 0, there exists ε > 0 and N ∈ Z
with W (x) = N for c − ε < x < c and W (x) = N − 1 for c < x < c + ε.
The claim then follows since g0 and gr do not vanish at the boundary points a, b.
(2) Now let f be as in the statement of the theorem and let (f0 , . . . , fr ) be the
Sturm sequence of f . We let gi := fi /fr ∈ R[t] (i = 0, . . . , r) and W (x) :=
Var (g0 (x), . . . , gr (x)) (x ∈ R). Then the sequence (g0 , . . . , gr ) satisfies (a)–(d)
from (1). Indeed, since gi−1 = hi gi − gi+1 (i = 1, . . . , r − 1) and gr−1 = hr gr for
suitable hi ∈ R[t], and since gr = 1, it follows that gi−1 and gi have no zeroes in
common (i = 1, . . . , r), and so (d) follows recursively. (a) and (b) are clear, and (c)
follows from Lemma 1.7.3 since g0 g1 = ff /fr2 .
Furthermore, g0 and f0 = f have the same zeroes (if fr (c) = 0, then also
f (c) = 0, and the zero c occurs with larger multiplicity in f than in fr !). It thus
follows from (1) that f has precisely W (a) − W (b) distinct zeroes in [a, b]. The
proof then follows from the observation that V (a) = W (a) and V (b) = W (b) since
fr (a) fr (b) = 0.
We state a more general version of Sturm’s Theorem, also due to Sturm. We make
use of
Definition 1.7.9 Let f ∈ R[t] and a, b ∈ R with a < b and f (a)f (b) = 0. A
generalized Sturm sequence of f on [a, b] is a sequence (f0 , . . . , fr ) in R[t] \ {0}
with r ≥ 1 that satisfies the properties
(1) f0 = f ;
(2) fr (a)fr (b) = 0, and fr is semidefinite on [a, b];
(3) for 0 < i < r and c ∈ [a, b] with fi (c) = 0, we have fi−1 (c)fi+1 (c) < 0.
(Warning: The Sturm sequence of f is usually not a generalized Sturm sequence of
f in the sense of Definition 1.7.9!)
1.7 Counting Real Zeroes of Polynomials (without Multiplicities) 23
The statement is clear by the proof of Theorem 1.7.8! (We only have to count the
different types of sign changes of ff1 .)
Remark 1.7.11 From the proof we see that simplifications of the original Sturm
algorithm are possible. For example:
(a) The original Sturm sequence f0 = f, f1 = f , . . . may be terminated at fs
provided that fs (a) fs (b) = 0 and fs is semidefinite on [a, b].
(b) For a remainder fi+1 in the modified Euclidean algorithm (Definition 1.7.6) it
is allowed to leave out those factors that are semidefinite on [a, b] and that do
not vanish at a, b.
n
n
f = tn + ai t n−i = (t − ξj )
i=1 j =1
be a monic polynomial in K[t] that splits over K (with ai , ξj ∈ K). Prove that
ξ1 , . . . , ξn ≥ 0 ⇔ (−1)i ai ≥ 0 for i = 1, . . . , n,
sj +k−2 (f ) 1≤j,k≤n
.
n
In other words, S(f )(x1 , . . . , xn ) = sj +k−2 (f )xj xk .
j,k=1
Theorem 1.7.15 (Jacobi, Borchardt, Hermite) Let R be a real closed field and
f ∈ R[t] a monic non-constant polynomial. Then the rank of √ S(f ) equals the
number of distinct roots of f in the algebraic closure C = R( −1) of R, and
the signature of S(f ) equals the number of distinct real roots of f .
Note that therefore S(f ) always has nonnegative signature!
Proof Let β1 , . . . , βr be the pairwise distinct real roots of f , and γ1 , γ1 , . . . , γs , γs
the pairwise distinct non-real roots of f , where r, s ≥ 0 and α → α denotes
the nontrivial R-automorphism of√C. Denote the multiplicity of βj by mj and the
multiplicity of γj by nj . Let i = −1 ∈ C, and let Re α, Im α for α ∈ C have their
usual meaning. Then
S(f ) (x1 , . . . , xn ) = αjk+l−2 xk xl
1≤j,k,l≤n
n
n 2
= αjk−1 xk
j =1 k=1
r 2
s 2 2
= mj βjk−1 xk + nj γjk−1 xk + γj k−1 xk
j =1 k j =1 k k
r 2
= mj βjk−1 xk
j =1 k
s 2 2
+2 nj Re(γjk−1 )xk − Im(γjk−1 )xk ,
j =1 k k
we have
r
s
S(f ) = mj u2j + 2 nj (vj2 − wj2 ),
j =1 j =1
S(f ) ∼
= r × 1 ⊥ s × H ⊥ (n − r − 2s) × 0,
from which we immediately see that rank S(f ) = r + 2s, sign S(f ) = r.
The following modification makes it possible to also count real zeroes in
intervals:
Definition 1.7.16 Let K be a field, f ∈ K[t] a monic polynomial of degree n ≥ 1,
and λ ∈ K. The Sylvester form of f with parameter λ is the quadratic form
(x1 , . . . , xn ) → λ sj +k−2 (f ) − sj +k−1 (f ) xj xk
1≤j,k≤n
over K, denoted by Sλ (f ).
Theorem 1.7.17 (Hermite, Sylvester, 1853) Let R be real closed and f ∈ R[t] a
monic polynomial of degree n ≥ 1. For λ ∈ R we have
√
rank Sλ (f ) = # a ∈ R( −1) : f (a) = 0 and a = λ ,
sign Sλ (f ) = # a ∈ R : f (a) = 0, a < λ − # a ∈ R : f (a) = 0, a > λ .
n 2
= (λ − αj ) αjk−1 xk
j =1 k
r
= mj (λ − βj )u2j
j =1
s
+ nj (λ − γj )(vj + iwj )2 + (λ − γj )(vj − iwj )2 .
j =1
26 1 Ordered Fields and Their Real Closures
We rewrite [· · · ] as follows:
Hence,
r
s
Sλ (f ) = mj (λ − βj )u2j + 2nj ϕj (vj , wj ).
j =1 j =1
Sλ (f ) ∼
= λ − β1 , . . . , λ − βr ⊥ s × H ⊥ (n − r − 2s) × 0.
r
trA/K (a1 , . . . , ar ) = trAi /K (ai ) for all (a1 , . . . , ar ) ∈ A.
i=1
Notation 1.8.3
(a) Let A be a K-algebra and a ∈ A. Then [A, a]K denotes the quadratic space
over K defined by the quadratic form x → trA/K (ax 2 ) on A.
(b) For polynomials f, g ∈ K[t], f = 0, let
and
[A, a] ∼
= A , ϕ(a) for all a ∈ A.
28 1 Ordered Fields and Their Real Closures
[A, a]K ⊗K K ∼
= [A ⊗K K , a ⊗ 1]K (over K ).
(c) A1 × · · · × Ar , (a1 , . . . , ar ) ∼
= [A1 , a1 ] ⊥ · · · ⊥ [Ar , ar ] (ai ∈ Ai ).
(d) Nil A ⊆ Rad[A, a] for all a ∈ A.
SylK (f ; g) ⊗K K ∼
= SylK (f ; g) (over K ).
Syl(f1 · · · fr ; g) ∼
= Syl(f1 ; g) ⊥ · · · ⊥ Syl(fr ; g).
(c) Syl(f e ; g) ∼
= e ⊗ Syl(f ; g) ⊥ 0, . . . , 0 (e ∈ N).
(d) Syl(af ; g) = Syl(f ; g) and Syl(f ; ag) ∼ = a ⊗ Syl(f ; g) for any a ∈ K ∗ .
(e) Syl (t − a; g(t)) = g(a) for any a ∈ K.
Proof (a) and (b) follow from Lemma 1.8.4, and (d) and (e) are clear. We show (c):
let A = K[t]/(f e ), A = K[t]/(f ), and π : A → A the residue class map. By
Proposition 1.8.2 we have trA/K (a) = e · trA /K (πa) for all a ∈ A, from which the
proof follows.
r
∼
= ⊥
j =1
ej ⊗ Syl t − αj ; g(t) ⊥ (ej − 1) × 0
r
∼
= ⊥
j =1
ej g(αj ) ⊥ (ej − 1) × 0 ,
r
s
f (t) = (t − βj )mj qj (t)nj ,
j =1 j =1
r s
∼
= ⊥ Syl(t − β ; g) ⊥ ⊥ Syl(q ; g) ⊥ 0, . . . , 0
(c) j =1
j
j =1
j
s
∼
= g(β1 ), . . . , g(βr ) ⊥
(e)
⊥ Syl(q ; g) ⊥ 0, . . . , 0.
j =1
j
The forms√ Syl(qj ; g) are all hyperbolic (or zero in case qj | g) since R[t]/(qj ) and
C = R( −1) are isomorphic as R-algebras, and [C; α]R ∼ = 1, −1 = H for all
α ∈ C ∗ . (Choosing β ∈ C with 2αβ 2 = i, the form [C; α]R is represented by the
matrix 01 10 with respect to the basis β, −iβ.)
It follows that
Syl(f ; g) ∼
= g(β1 ), . . . , g(βr ) ⊥ t × H ⊥ (n − r − 2t) × 0
S(f ) ∼
= Syl(f ) = Syl(f ; 1) ∼
= r × 1 ⊥ s × H ⊥ (n − r − 2s) × 0
30 1 Ordered Fields and Their Real Closures
and
Syl(f ; λ − t) ∼
= λ − β1 , . . . , λ − βr ⊥ s × H ⊥ (n − r − 2s) × 0,
from which the statements of the theorems of Sylvester and Hermite can be read off
directly.
In addition to the works [65] and [39], mentioned before, the reader may want
to compare the current topic and related themes with the article [10] that contains
interesting historical remarks about Sylvester.
This section, and the next one, will not be used in the remainder of this book and can
therefore be skipped during a first reading. Having said that, its content is a classical
part of real algebra, and is essential for understanding the historical development
of mathematics in the nineteenth century. Furthermore, it has become important in
many applications such as the stability of motion or control theory.
We will follow the paper [65] of Krein and Naimark, but can only cover a small
part for lack of space. The interested reader may therefore want to consult the
original source.
R will always denote a real closed field.
Definition 1.9.1 Let ϕ(t) = g(t)/f (t) ∈ R(t) be a rational function with g and f
relatively prime polynomials over R.
(a) Let α ∈ R be a pole of ϕ (i.e., a zero of f ). The Cauchy index indα (ϕ) of ϕ at
α is defined as follows:
⎧
⎪
⎨+1 if the value of ϕ(t) jumps from −∞ to +∞ at t = α
⎪
indα (ϕ) := −1 if the value of ϕ(t) jumps from +∞ to −∞ at t = α .
⎪
⎪
⎩ 0 otherwise
(b) The Cauchy index of ϕ in an open interval ]a, b[ ⊆ R (a and b are allowed to be
±∞), denoted Iab (ϕ), is defined as the sum of the Cauchy indices of ϕ at those
poles of ϕ that are in ]a, b[. If a = −∞ and b = +∞, we speak of the global
Cauchy index of ϕ. (Note that if ϕ has a pole at ∞, this pole is ignored for the
Cauchy index, no matter what a and b are.)
1.9 Cauchy Index of a Rational Function, Bézoutian and Hankel Forms 31
r
mi
ϕ(t) = ψ(t) + ,
t − αi
i=1
for some rational function ψ without real poles. Thus ϕ = f /f has precisely
r real poles α1 , . . . , αr and its Cauchy index at each one of them is +1.
(2) More generally: let f, h ∈ R[t] with f = 0. For ϕ = hf /f we have
+∞
I−∞ (ϕ) = # α ∈ R : f (α)= 0 and h(α) > 0 − # α ∈ R : f (α)= 0 and h(α)< 0 ,
f (x)g(y) − f (y)g(x)
n−1
= cij x i y j ,
x−y
i,j =0
32 1 Ordered Fields and Their Real Closures
where
i
cij = dk,i+j +1−k with dij := an−j bn−i − an−i bn−j .
k=0
Independently from these explicit formulas (for which we will have no further use),
it is clear that cij = cj i for all i, j = 0, . . . , n − 1.
Definition 1.9.3 The symmetric n × n-matrix (cj i )0≤i,j ≤n−1 is called the Bézout
matrix of f and g, and is denoted B(f, g). The associated quadratic form over K is
called the Bézoutian of f , also denoted B(f, g). Thus,
n−1
B(f, g) (x0 , . . . , xn−1 ) = cij xi xj .
i,j =0
Since deg g ≤ deg f we can write the rational function ϕ(t) = g(t)/f (t) as a
formal power series in t −1 :
ϕ(t) = s−1 + s0 t −1 + s1 t −2 + · · · .
Definition 1.9.4 The symmetric n × n-matrix (si+j )0≤i,j ≤n−1 is called the Hankel
matrix of f and g and is denoted H (f, g). (In general, any quadratic matrix (aij )
whose entry aij only depends on i +j is called a Hankel matrix.) Again we interpret
H (f, g) as a quadratic form,
n−1
H (f, g) (x0 , . . . , xn−1 ) = si+j xi xj ,
i,j =0
and then refer to it as the Hankel form of f and g. For 1 ≤ p ≤ n the truncated
matrix (si+j )0≤i,j ≤p−1 is denoted Hp (f, g).
Remark 1.9.5 H (f, g) clearly only depends on n and the rational function ϕ =
g/f . More generally, a Hankel matrix Hn (ϕ) as in Definition 1.9.4 can be associated
to any formal power series ϕ(t) ∈ K[[t −1 ]] in t −1 and any n ≥ 0. A well-known
(elementary) theorem of Frobenius says that ϕ(t) is rational (i.e., ϕ(t) ∈ K(t)) if
and only if there exists n ≥ 0 such that det Hm (ϕ) = 0 for all m ≥ n. Furthermore,
in this case the smallest such n is the degree of the denominator of ϕ, written as
an irreducible fraction, and every matrix Hm (ϕ) with m > n has rank n. See, for
example, [40, §16.10].
1.9 Cauchy Index of a Rational Function, Bézoutian and Hankel Forms 33
Example 1.9.6 We will compute the Hankel form H (f, f ), where f ∈ K[t] is a
non-constant polynomial. Let f (t) = a(t − α1 ) · · · (t − αn ) be the factorization of
f over the algebraic closure of K, then (geometric series!)
∞
f (t) 1
n n
sj t −j −1
j
= = with sj = αk .
f (t) t − αi
i=1 j =0 k=1
Proposition 1.9.7 The Bézoutian B(f, g) is isometric to the Hankel form H (f, g).
Proof As usual, let ϕ(t) = g(t)/f (t). Then
n−1
ϕ(y) − ϕ(x)
cij x i y i = f (x)f (y)
x−y
i,j =0
∞
y −i − x −i
= f (x)f (y) si−1
x−y
i=0
∞
i−1
= f (x)f (y) si−1 x j −i y −j −1
i=1 j =0
∞
= f (x)f (y) sk+l x −(k+1) y −(l+1)
k,l=0
∞
= sk+l (a0 x n−k−1 + · · · + an x −k−1 ) (a0 y n−l−1 + · · · + an y −l−1 ).
k,l=0
Since there are no negative powers of x or y on the left hand side, we may omit all
such terms on the right hand side and obtain
n−1
n−1
cij x i y j = sk+l (a0 x n−k−1 + · · · + an−k−1 ) (a0 y n−l−1 + · · · + an−l−1 ).
i,j =0 k,l=0
34 1 Ordered Fields and Their Real Closures
we then get
of the Bézout matrix B(f, g) (the “principal minors from the bottom right up”).
Then
2p
Bp = a0 det Hp (f, g). (1.3)
for all i ≥ 0. These can be used to transform the identity (1.3) as follows:
1 0 · · · 0 0 · · · 0
0 1 · · · 0 0 · · · 0
. . . ..
.. .. . . ... ..
. .
p(p−1) 2p+1 0 0 · · ·
1 0 · · · 0
a0 Bp = (−1) 2 a0
(1.3) 0 s−1 · · · sp−2 sp−1 · · · s2p−2
0 0 · · · sp−3 sp−2 · · · s2p−3
.. .. .. .. ..
. . . . .
0 0 · · · s−1 s0 · · · sp−1
1 0 0 0 ··· 0
0 s ··· s2p−2
−1 s0 s1 a 0 a 1 a 2 · · · a2p
0 1 0 0 ··· 0
0 a0 a1 · · · a2p−1
0 0 s−1 s0 ··· s2p−3
= (−1)p · . . . . .. · 0 0 a0 · · · a2p−2 .
.. .. .. .. . .. .. .. .. .
. . . . ..
0 0 0 0 · · · 1 0 · · · 0
0 0 0 0 · · · 0 s−1 · · · sp−1 0 0 0
· · · a0
0 0 0 0 · · · 0 1 · · · 0
a 0 a1 · · · an−1 an 0 0 ··· 0
0 a0 · · · 0 ··· 0
an−2 an−1 an
. .. . . .. .. .. .. ..
..
. . . . . . .
0 0 ··· a0 a1 a2 a3 · · · an
.
b 0 b1 · · · bn−1 bn 0 0 ··· 0
0 b0 · · · bn−2 bn−1 bn 0 ··· 0
.. .. . . .. .. .. .. ..
. . . . . . . .
0 0 · · · b0 b1 b2 b3 · · · bn
If b0 = 0, this is the resultant of f and g in the usual algebra sense (cf. [55, 73, 113]).
In contrast, if deg(g) = m < n = deg(f ), then R(f, g) is obtained from the usual
resultant via multiplication with the (unimportant) factor a0n−m .
In case n = p, Theorem 1.9.9 says
n(n−1)
det B(f, g) = (−1) 2 R(f, g). (1.6)
Since the resultant of two polynomials vanishes if and only if they have a common
divisor (cf. [113], . . . ), we obtain
Corollary 1.9.11 The Bézoutian of f and g is nondegenerate if and only if f and
g have no common divisor.
Denoting the quadratic space (K n , B(f, g)) by V (f, g), we have more generally
Proposition 1.9.12 Let D = t q + d1 t q−1 + · · · + dq be the greatest common divisor
of f and g, and let f = f 0 · D, g = g 0 · D. Then the radical Rad V (f, g) of V (f, g)
(see Sect. 1.2) has dimension q, and the quadratic space V (f, g)/ Rad V (f, g) is
isometric to V (f 0 , g 0 ).
Proof Let (cij0 ) := B(f 0 , g 0 ). Then
n−1
f (x)g(y) − f (y)g(x)
cij x i y j =
x−y
i,j =0
n−q−1
= cij0 (d0 x q+i + · · · + dq x i ) (d0 y q+j + · · · + dq y j ),
i,j =0
1.9 Cauchy Index of a Rational Function, Bézoutian and Hankel Forms 37
u0 = dq x0 + dq−1 x1 + · · · + d0 xq
u1 = dq x1 + · · · + d1 xq + d0 xq+1
..
.
un−q−1 = dq xn−q−1 + ··· + d0 xn−1
(d0 := 1). From the above identity we then read off that
From now on we assume again that K = R is real closed. As before, let f and g
be polynomials of degree n and degree at most n, respectively, and let ϕ := g/f .
+∞
Theorem 1.9.13 The global Cauchy index I−∞ (ϕ) is equal to the signature of the
Hankel form H (f, g), and thus also equal to the signature of the Bézoutian B(f, g)
of f and g.
Proof By Proposition 1.9.12 we may assume without loss of generality that f and
g are relatively prime and thus that H (f, g) is nondegenerate.
We use complex analysis to prove the case R = R. Specifically, we will compute
the residues of the differential form ω = ϕ(z) θ (z)2 dz, where z is a complex
variable,
and x0 , . . . , xn−1 are independent real parameters. In order to compute the residue
of ω at z = ∞, we introduce the local coordinate ζ = z−1 at ∞. Then we have
and so − Res∞ (ω) is the coefficient of ζ in the Laurent expansion of ϕ(z) θ (z)2 in
terms of ζ . We see from
n−1
ϕ(z) θ (z)2 = (s−1 + s0 ζ + s1 ζ 2 + · · · ) · xj xk ζ −j −k
j,k=0
38 1 Ordered Fields and Their Real Closures
that
n−1
− Res∞ (ω) = sj +k xj xk = H (f, g) (x0 , . . . , xn−1 ) . (1.7)
j,k=0
By the Residue Theorem this expression equals the sum of the residues of ω at the
simple poles of ω, in other words at the zeroes of f .
Assume thus that z = α is a (complex) zero of f (z) of multiplicity m. There is a
Laurent expansion
θ (z + α) = x0 + x1 (z + α) + · · · + xn−1 (z + α)n−1
= u0 + u1 z + · · · + un−1 zn−1 ,
r−1
r−1
Resα (ω) + Resα (ω) = (uj vj + uj v j ) = 2 (pj qj + pj qj ), (C)
j =0 j =0
and if m = 2r + 1 we obtain
r−1
Resα (ω) + Resα (ω) = 2 (pj qj + pj qj ) + 2c0 (pr2 − pr2 ). (D)
j =0
In total precisely n real linear forms occur in the identities (A), (B) (for real
α) and (C), (D) (for nonreal α) since f has precisely n zeroes, counted with
multiplicities. Using the Residue Theorem, (1.7) yields a representation of H (f, g)
as a quadratic form in those n linear forms. These must be linearly independent since
otherwise H (f, g) is degenerate. Furthermore, H (f, g) is the orthogonal sum of the
quadratic forms determined by (A), (B) for α ∈ R and (C), (D) for Im(α) > 0.
The forms coming from (A), (C) and (D) are clearly hyperbolic, while the form
coming from (B) is the orthogonal sum of a hyperbolic form and the form c0 u2r ,
and so its signature is the sign of c0 . Observe now that for α ∈ R the Cauchy index
indα (ϕ) = 0 if m is even, but indα (ϕ) = sgn c0 if m is odd. We conclude that
+∞
sign H (f, g) = indα (ϕ) = I−∞ (ϕ).
α∈R,f (α)=0
There are now two different options for proving Theorem 1.9.13 for an arbitrary
real closed field R. The first option√is to use a purely algebraic residue theorem for
rational differential forms over R( −1). Indeed, for the form ω introduced above
one can determine formal Laurent series at all poles over any algebraically closed
field, use them to define the residues of ω, and then show that their sum is zero. The
proof above then remains valid word for word. See also [109, p. 31].
The other option consists of applying Tarski’s Principle (cf. [90, §5] or [91,
§4.2]). This theorem from model theory states among other things that every
“elementary sentence” in the language of ordered fields that is valid over R is also
valid over every real closed field. Even if one has had only very little exposure to
model theory, one readily recognizes that Theorem 1.9.13 is such an elementary
sentence. Tarski’s Principle plays an important role in higher real algebra (as
mentioned in the Preface), but will not be considered any further in this book.
In this way we declare the theorem proven.
40 1 Ordered Fields and Their Real Closures
From this result we immediately obtain a method for expressing Cauchy indices
Iab (ϕ) on finite open intervals (a, b) as signatures of “modified” Bézoutians. Without
loss of generality we may assume deg g < deg f and define, for each λ ∈ R, the
quadratic form
by Theorem 1.9.13. Thus, if a, b are at most poles of order one of ϕ and if a < b,
then
ρ(b) = I−∞
a
(ϕ) + inda (ϕ) + Iab (ϕ) − Ib∞ (ϕ),
ρ(a) = I−∞
a
(ϕ) − Iab (ϕ) − indb (ϕ) − Ib∞ (ϕ).
This proves
Corollary 1.9.14 If ϕ has at most simple poles at a and b and if a < b, then
Moreover, Hurwitz’ Theorem 1.9.9 gives us tight control over the signature of
B(f, g). As an illustration we mention
Theorem 1.9.15 (Hurwitz) Let f and g be polynomials over R as above (deg g ≤
deg f = n). The following statements are equivalent:
(i) All the roots of f are real and simple, and between any two such roots there is
a root of g;
(ii) The entries in the sequence of determinants
a0 b0 0 0 · · · 0 0
a b a1 b1 a0 b0 · · · 0 0
1, 0 0 ,..., . .. .. .. .. ..
a1 b1 .. . . . . .
a b2n−1 a2n−2 b2n−2 ··· a b
2n−1 n n
In this section R denotes a real closed field and, in contrast to previous practice, all
zeroes will be counted with multiplicities.
Let f ∈ R(t) be a non-constant rational function without pole in the interval
[a, b] (a, b ∈ R, a < b). We fix n ≥ 1 such that f (n) (a)f (n) (b) = 0 and denote by
Ni the number of real zeroes of the i-th derivative f (i) in ]a, b] (i = 0, 1, . . . , n); in
particular we let N := N0 denote the number of zeroes of f in ]a, b]. For x ∈ [a, b]
we let V (x) := Var f (x), f (x), . . . , f (n) (x) (cf. Sect. 1.7).
Theorem 1.10.1 (Hurwitz) There exists v ∈ N0 such that
Corollary 1.10.3
(a) (Rule of Descartes) There exist v, v ∈ N0 such that p = V − 2v and p =
V − 2v .
(b) If cn = 0 and if all roots of f are real, then p = V and p = V .
(c) If cn = 0 and if ci−1 = ci = 0 for some i with 1 < i < n, then f has non-real
roots.
Proof (a) V (0) = Var(cn , 1! cn−1 , . . . , n! c0 ) = V and V (b) = 0 for b 0,
together with Corollary 1.10.2, give the first formula. For the second one, simply
replace f (t) by f (−t).
(b) For x ∈ R, let
f (a + t) = 1 (m)
m! f (a) t m + t m+1 g(t).
1.10 An Upper Bound for the Number of Real Zeroes (with Multiplicities) 43
The Leibniz rule gives f (m) (a) = m! h(0). There exists g ∈ R(t) such that g(0) =
∞ and h(t) = h(0) + t g(t), from which the statement follows.
Proposition 1.10.5 Let a, b ∈ R, a < b, and f ∈ R(t) without a pole in [a, b].
(a) If f has no zeroes in [a, b[, f (b) = 0 and f (a) = 0, then the number of zeroes
of f in ]a, b[ is
even if f (a)f (a) < 0
.
odd if f (a)f (a) > 0
(b) If f has no zeroes in ]a, b], f (a) = 0 and f (b) = 0, then the number of zeroes
of f in ]a, b[ is
odd if f (b)f (b) < 0
.
even if f (b)f (b) > 0
(r − 1) + 2u + (m1 − 1) + · · · + (mr − 1) = N + 2u − 1
i.e.,
f (k) (x)
f (x + h) = · hk−1 + hk g(h).
(k − 1)!
It follows that sign f (x + h) = sign f (k) (x) for a small h > 0, i.e., w(x + h) =
(Wf )(x) for a possibly even smaller h > 0.
3) If f (x) f (x) = 0, then (Wf )(x) = w(x) = w(x + h) for a small h > 0.
Lemma 1.10.8 There exists v0 ∈ N0 such that N −N1 = (Wf )(a)−(Wf )(b)−2v0 .
Proof Let h > 0 be sufficiently small such that f · f has no zeroes nor poles
in ]a, a + h] ∪ ]b, b + h] and thus such that N, resp. N1 , is the number of zeroes
of f , resp. f , in [a + h, b + h]. From Lemma 1.10.6 it follows that N − N1 =
w(a + h) − w(b + h) − 2v0 for some v0 ≥ 0, and the statement follows from
Lemma 1.10.7.
1.10 An Upper Bound for the Number of Real Zeroes (with Multiplicities) 45
This result can now be iterated. Since the assumptions made about f are also
valid for f , . . . , f (n−1) , we obtain
Corollary 1.11.3 Let R/K and R /K be real closures of (K, P ). Then there exists
a unique K-isomorphism R → R .
Because of Corollary 1.11.3 we will say the real closure of (K, P ) from now on.
Proposition 1.11.4 Let L/K be a finite field extension such that K ⊆ L ⊆ R. Then
ϕ has an order preserving extension ψL : L → S. (Here “order preserving” is with
respect to the ordering of L induced by R.)
Proof Let α ∈ L be such that L = K(α), and let f ∈ K[t] be the minimal
polynomial of α. By Hermite’s Theorem (Theorem 1.7.15) the Sylvester form S(f )
of f has positive signature over R, and thus also over S. Again by this theorem,
f thus also has a zero in S and so there is at least one extension χ : L → S of ϕ.
Let χ1 , . . . , χr be all such extensions (r ≥ 1). Suppose that none of the χi is order
preserving. Then there exist a1 , . . . , ar ∈ L with ai > 0 (in R), but χi (ai ) < 0
√ √
(in S). Consider L := L( a1 , . . . , ar ) ⊆ R. Then L /K is finite, but there are
actually no extensions χ : L → S of ϕ to L (since χ |L must be one of the χi , i.e.,
√
χ ( ai )2 < 0 for some i), contradicting Hermite’s Theorem.
Proof of Theorem 1.11.2 By Zorn’s Lemma there exists an intermediate field
K /K of R/K and a maximal order preserving extension ψ : K → S of ϕ.
However, it follows from Proposition 1.11.4 that K = R, which establishes the
existence of ψ. Concerning the uniqueness see the next remark. This finishes the
proof of Theorem 1.11.2.
Corollary 1.11.7 Let (K, P ) be an ordered field with real closure R, and let L/K
be a finite algebraic extension. If L = K(α) and f ∈ K[t] is the minimal
polynomial of α over K, then the extensions of P to L correspond bijectively to
the zeroes of f in R. In particular, P has at most [L : K] distinct extensions to L.
Example 1.11.8 Let L = Q(α) be a number field and f ∈ Q[t] the minimal
polynomial of α over Q. The zeroes of f correspond to the field embeddings
L → C. The real zeroes of f thus correspond to the embeddings L → R, and
therefore also to the orderings of L.
The following theorem ties in with Corollary 1.6.3:
Theorem 1.11.9 (Orderings and Involutions in the Galois Group) Let K be a
field (of arbitrary characteristic) with absolute Galois group = Gal (Ks /K).
There exists a natural bijection from the set of all conjugacy classes [τ ] of
involutions in to the set of orderings of K. More precisely: ([τ ]) is √
the ordering
√
of K defined by the inclusion K → Fix(τ ), i.e., ([τ ]) = {a ∈ K : τ ( a) = a}.
(Note that Fix(τ ) is real closed by Sect. 1.6.)
Proof (Sketch) is well-defined since Fix(σ τ σ −1 ) = σ (Fix τ ) (σ, τ ∈ ). The
injectivity of ϕ follows from Theorem 1.11.2, and the surjectivity from the existence
of the real closure and the findings in Sect. 1.6. We leave the (easy) details to the
reader.
If τ is an involution in , then the centralizer of τ in is equal to {1, τ }; this
is just a reformulation of Aut(Fix(τ )/K) = {1} (Proposition 1.5.6(b)). Therefore,
the conjugates of τ are in bijective correspondence with the elements of the
homogeneous space /{1, τ }.
To conclude this section, we express several of our previous results in the
language of Witt rings and signatures. Let K be a field of char K = 2. By
48 1 Ordered Fields and Their Real Closures
of K (Exercise 1.2.17), and in this way one obtains all signatures in case R/ϕ(K)
is algebraic (existence of the real closure). If ϕ : K → R, ϕ : K → R are
homomorphisms into real closed fields R, R , and if R/ϕ(K) is algebraic, then
W (ϕ) = W (ϕ ) if and only if there exists a K-homomorphism R → R (uniqueness
of the real closure).
Theorem 1.11.6 can be formulated as follows: If L/K is an algebraic field
extension, ϕ : K → R a homomorphism into a real closed field R, and σ = W (ϕ)
the induced signature of K, then the extensions ψ : L → R of ϕ correspond
bijectively to the signatures τ of L that extends σ (i.e., that satisfy σ = τ ◦ iL/K )
via τ = W (ψ).
b(v, W ) = 0 ⇔ (s ◦ b)(v, W ) = 0.
Proof One direction is trivial. For the other direction we note that the K-bilinear
form
β : L × L → K, β(a, a ) := s(aa )
1.12 Transfer of Quadratic Forms 49
Remark 1.12.6 The additive subgroup s∗ W (L) of W (K) does not depend on s.
Indeed, if s : L → K is another K-linear map, s = 0, then there exists a ∈ L∗
with s (b) = s(ab) for all b ∈ L. It follows that s∗ (η) = s∗ (a · η) for η ∈ W (L).
The following theorem shows in particular that s∗ W (L) is an ideal of W (K).
s∗ iL/K (ξ ) · η = ξ · s∗ (η).
s∗ (W (L)) · W (L/K) = 0.
On the other hand, tr∗ (η) is precisely the Sylvester form SylK (f ; g) (Sect. 1.8).
Then, by Lemma 1.8.5(a),
where the identity () follows from the explicit computation of the Sylvester form
over real closed fields (see the end of Sect. 1.8).
x ≤ z ≤ y and x, y ∈ X ⇒ z ∈ X.
Arbitrary intersections and upward directed unions of convex subsets are again
convex. In particular, for every subset Y ⊆ M there exists a smallest convex superset
X of Y in M, the convex hull of Y in M.
Definition 2.1.2
(a) An ordered abelian group is a pair (, ≤), where is an abelian group (usually
written additively) and ≤ is a total order relation on the set such that for all
α, β, γ ∈ ,
α ≤ β ⇒ α + γ ≤ β + γ.
α ≥ 0 ⇒ ϕ(α) ≥ 0.
Remarks 2.1.4
(1) An ordered abelian group can also be defined as a pair (, ), where is an
abelian group and ⊆ is a subset such that
(2) Let (K, ≤) be an ordered field and denote the set of positive elements of K by
K+ ∗ . Then (K, +, ≤) and (K ∗ , ·, ≤) are ordered abelian groups.
+
(3) Ordered abelian groups are torsion free: nα = 0 for all 0 = α ∈ and 0 = n ∈
Z.
(4) Let be an ordered abelian group. The convex hulls of subgroups of are
subgroups again. A subgroup is convex if and only if γ ∈ , δ ∈ and
0 ≤ γ ≤ δ imply that γ ∈ . The convex subgroups of form a chain: every
two convex subgroups are comparable with respect to inclusion.
(5) If ϕ : → is an order preserving homomorphism of ordered abelian groups,
then Ker(ϕ) is convex in . Conversely, for every convex subgroup of
there is precisely one ordering of = / that makes into an ordered
abelian group and π : → order preserving, namely = π(), where
:= {α ∈ : α ≥ 0} (Remark (1)). The quotient group = / will
always be equipped with this ordering. There exists an obvious homomorphism
theorem for ordered abelian groups.
(6) Convex subgroups are sometimes called isolated subgroups in the literature (see
for instance [13, Chap. VI]).
2.1 Convex Subrings of Ordered Fields 55
Examples 2.1.5
(1) Let 1 , . . . , r be ordered abelian groups. We equip := 1 × · · · × r with
the lexicographic order: (γ1 , . . . , γr ) > 0 if and only if
Then (, >) is an ordered abelian group which we denote by (1 × · · · × r )lex .
The projections → (1 × · · · × i )lex where i = 0, . . . , r, are order
preserving, and so their kernels i are convex subgroups of . Clearly we have
= 0 ⊇ 1 ⊇ · · · ⊇ r = 0. If the i are archimedean, then 0 , . . . , r
are the only convex subgroups of .
(2) In the special case 1 = · · · = r = Z we call (e1 , . . . , er ) the lexicographic
basis of Zrlex , where ei denotes i-th unit vector.
Exercise 2.1.6 Let G be a torsion free abelian group and H ⊆ G a subgroup. Let
an ordering ≤ of H be given. Show that ≤ can be extended to an ordering of G, and
that this extension is uniquely determined if G/H is a torsion group.
In the rest of this section (K, P ) will be an ordered field. Whenever it is clear
that convexity, the ≤ sign, etc. refer to P we will not mention this explicitly.
Proposition 2.1.7 Let A be a subring of K.
(a) The convex hull of A in K is a subring of K.
(b) A is convex in K if and only if [0, 1] ⊆ A. In particular, if A is convex in K,
then every overring of A is also convex in K.
Proof (a) This is clear. (b) Obviously [0, 1] ⊆ A is necessary for the convexity of
A. Conversely, if [0, 1] ⊆ A and if a ∈ A, b ∈ K with 0 < b < a, then ba −1 ∈ A
and so b = ba −1 · a ∈ A.
for all f ∈ F (t) with f (0) ∈ {0, ∞}. Now let h(t) = t r f (t) ∈ F (t)∗ , where r ∈ Z
and f ∈ F (t) with f (0) ∈ {0, ∞}. By what we have just established, there exist
a, b ∈ F with 0 < a < b and at r < |h(t)| < bt r . It follows that
The following classical theorem of Hölder gives (in principle) a summary of all
archimedean ordered abelian groups and fields:
Theorem 2.1.10 (O. Hölder [52], 1901)
(a) Let be an archimedean ordered abelian group and let 0 < γ ∈ . Then there
exists precisely one order preserving injective group homomorphism ϕ : →
R such that ϕ(γ ) = 1.
(b) Let (K, P ) be an archimedean ordered field with underlying archimedean
ordered abelian group (K, +, P ). The order preserving embedding ϕ : K → R
such that ϕ(1) = 1, furnished by (a), is a ring homomorphism.
Corollary 2.1.12 (von Staudt’s Theorem) The identity is the only endomorphism
of the field R of real numbers.
Corollary 2.1.13 A proper overfield of R does not have any archimedean order-
ings.
Proof of Theorem 2.1.10 With some heuristics one sees quickly how the proof
should proceed. Namely, if there exists a map ϕ as in (a), then for all α ∈ , m ∈ Z,
and n ∈ N we have
m
n ≤ ϕ(α) ⇔ mϕ(γ ) ≤ nϕ(α) ⇔ mγ ≤ nα.
2.2 Valuation Rings 57
U (α) := { m
n : m ∈ Z, n ∈ N, mγ ≤ nα}, O(α) := { n : m ∈ Z, n ∈ N, mγ > nα}.
m
The pair U (α), O(α) constitutes a proper Dedekind cut of Q. In other words (cf.
Definition 2.9.3):
(1) U (α) ∪ O(α) = Q;
(2) U (α), O(α) are nonempty;
(3) for all a ∈ U (α) and b ∈ O(α) we have a < b.
Here (1) is trivial, (2) uses the archimedean property, and (3) uses the torsion
freeness of .
As is well-known, the real numbers are axiomatically defined via such cuts. Thus,
for every α ∈ there exists precisely one ϕ(α) ∈ R such that
One verifies that for all α, β ∈ , U (α) + U (β) ⊆ U (α + β) and O(α) + O(β) ⊆
O(α + β), from which it follows that ϕ(α + β) = ϕ(α) + ϕ(β). Then ϕ is order
preserving since γ > 0 and is archimedean. Also, ϕ is injective since Ker(ϕ) =
{α ∈ : − γ ≤ nα ≤ γ for all n ∈ Z}.
For (b), let = K and γ = 1. The map ϕ is then also multiplicative: since
Q ⊆ K we have
The theory of Krull valuations revolves around three fundamental concepts: valua-
tions, valuation rings, and places. These are more or less equivalent to each other,
58 2 Convex Valuation Rings and Real Places
but it may be more convenient to work with one concept and not the other two,
depending on the situation. Therefore it is important to be familiar with all three of
them and to be able to translate from one to the other, as and when required. In this
section we investigate the first part of this triptych: valuation rings. In Sect. 2.4 we
will consider valuations. Finally, places will be investigated in Sect. 2.8.
Definition 2.2.1 A subring A of a field K is called a valuation ring of K if for
every a ∈ K ∗ we have a ∈ A or a −1 ∈ A. An arbitrary ring A is called a valuation
ring if it has no zero divisors and is a valuation ring of its field of fractions.
Definition 2.2.2 A ring A is called local if A = 0 and A has just one maximal ideal,
which is usually denoted by mA or simply m. We call the field κ(A) := A/mA the
residue field of A.
Definition 2.2.5 A valuation ring A is called residually real if the field κ(A) =
A/mA is real.
Remarks 2.2.6
(1) Let (K, P ) be an ordered field, A ⊆ K a convex subring of K and π : A →
κ(A) = A/mA the residue homomorphism. We see immediately that
P := π(A ∩ P )
Corollary 2.2.10 There exists an inclusion reversing bijection between the set of
prime ideals p of A and the set of overrings B of A in K, given by p → Ap =: B.
The inverse is given by B → mB =: p. Both sets are totally ordered (by inclusion).
A ⊆ B ⇔ mA ⊇ mB .
60 2 Convex Valuation Rings and Real Places
bn + a1 bn−1 + · · · + an = 0.
where δij denotes the Kronecker symbol. Let f (t) ∈ A[t] be the characteristic
polynomial of the matrix (aij ). Since for every n × n-matrix S we have S S =
SS = det(S) · 1 (where S denotes the adjugate of S), it follows that f (b)ui = 0 for
i = 1, . . . , n, and so f (b)M = 0. Since M is faithful, we have f (b) = 0, which
shows that b is integral over A by Definition 2.3.1.
Proof of Proposition 2.3.2 If b, b ∈ B are integral over A, then A[b, b ] is a
finitely generated A[b]-module and A[b] is a finitely generated A-module. Hence,
A[b, b ] is also a finitely generated A-module. It is a faithful A[c]-module for every
c ∈ A[b, b ], and the statement follows from Lemma 2.3.3.
Definition 2.3.4
(a) Let B be a ring and A ⊆ B a subring. The subring of elements of B that are
integral over A is called the integral closure of A in B. If the integral closure of
A in B is equal to A, then A is called integrally closed in B.
(b) A ring A is called integrally closed if it has no zero divisors and coincides with
its integral closure in Quot A.
If {A
α } is a family of subrings of a ring B and if all Aα are integrally closed in
B, then Aα is also integrally closed in B.
α
Theorem 2.3.5 (Cohen–Seidenberg) Let A ⊆ B be an integral ring extension.
(a) For every prime ideal p of A there exists a prime ideal q of B such that p = A∩q.
(b) For every prime ideal q of B we have: q is maximal in B ⇔ A ∩ q is maximal
in A.
Proof We first prove (b). It suffices to show that for every integral extension K ⊆ L
of rings K and L without zero divisors, K is a field if and only if L is a field
(consider the integral extension A/A ∩ q ⊆ B/q). If K is a field and 0 = b ∈ L,
then there exists a monic irreducible polynomial f ∈ K[t] such that f (b) = 0,
and it follows that b−1 ∈ L. If L is a field and 0 = a ∈ K, then there exist
a1 , . . . , an ∈ K such that a −n + a1 a 1−n + · · · + an = 0, and multiplying by a n−1
shows that a −1 ∈ K.
Next we prove (a). Let p be a prime ideal of A and let S := A \ p. The natural
homomorphism Ap = S −1 A → S −1 B is injective, and Ap ⊆ S −1 B is also an
integral ring extension. Choosing a maximal ideal q of S −1 B gives q ∩ Ap = pAp
by (b) (Ap is local), and it follows that q ∩ A = p for the preimage q of q in B.
bn = a1 bn−1 + · · · + an .
over κ(B) by the residue class of y and y satisfies an integrality equation. It follows
again that B = C, and so y ∈ B.
Corollary 2.3.9 The valuation rings of K are precisely those local subrings that
are not properly dominated by any other local subring.
Proof It follows from Theorem 2.3.7 that the integral closure of a local subring
A ⊆ K is equal to the intersection of all valuation rings of K that dominate A. Thus,
if A is not properly dominated by any local subring, then A must be a valuation ring.
The converse direction follows from Proposition 2.2.9.
Corollary 2.3.10 For every subring A of a field K and every prime ideal p of A
there exists a valuation ring B of K such that A ⊆ B, p = A ∩ mB , and κ(B) is
algebraic over κ(p) = Quot A/p.
Proof Choose B such that B dominates the local ring Ap and κ(B) is algebraic over
κ(Ap ) = κ(p) (Theorem 2.3.7). Then A ∩ mB = A ∩ Ap ∩ mB = A ∩ p Ap = p.
Corollary 2.3.11 The fields K that have no proper (i.e., different from K) valuation
rings are precisely the algebraic field extensions of finite fields.
Proof Let K be algebraic over Fp . Since every subring of K contains the prime field
Fp , K is the only subring of K that is integrally closed in K. Conversely, let K be
a field without any proper valuation rings. Then char K =: p > 0, for otherwise K
would be integral over Z by Theorem 2.3.7, contradicting that Z is integrally closed
in Q. If there was a t ∈ K, transcendental over Fp , then t −1 would not be integral
over Fp [t], a contradiction.
Exercise 2.4.3 Let a, b ∈ K with v(a) = v(b). Show that equality holds in (V3),
i.e., that v satisfies the following stronger property:
ov := {a ∈ K : v(a) ≥ 0}
a o∗v ≤ b o∗v :⇔ a −1 b ∈ ov .
(The multiplicative notation for this ordered abelian group should not be a source of
confusion for the reader!)
Now, let A be an arbitrary valuation ring of K. For all a, b ∈ K ∗ we define
analogously,
aA∗ ≤ bA∗ :⇔ a −1 b ∈ A.
It is immediately clear that this definition turns K ∗ /A∗ into an ordered abelian
group.
Definition 2.4.6 Let A be a valuation ring of K. The value group of A is the
ordered abelian group A := K ∗ /A∗ with ordering as above. The canonical
epimorphism vA : K ∗ → A is called the canonical valuation of K associated
to A.
The second definition is justified by:
Proposition 2.4.7 Let A be a valuation ring of K. Then vA : K → A ∪ ∞ is a
valuation of K and the associated valuation ring satisfies ovA = A.
Proof The map vA |K ∗ is a homomorphism, so it remains to show that (V3) holds.
Let a, b ∈ K ∗ with a + b = 0 and vA (a) ≤ vA (b). Then a −1 b ∈ A and it follows
that vA (a + b) = (a + b)A∗ = a(1 + a −1 b)A∗ ≥ aA∗ = vA (a). Clearly, ovA = A.
Example 2.4.11 Every ordered abelian group occurs as the value group of a
surjective valuation. Indeed, let k be a field and A := k[+ ] the semigroup algebra
of + := {α ∈ : α ≥ 0}. This means that A has a k-vector space basis
{xα : α ∈ + } with xα = xβ for α = β, and multiplication in A is given by
xα xβ = xα+β for all α, β ∈ + .
For 0 = a = α∈+ aα xα ∈ A (with aα ∈ k and almost all aα = 0) let
v (a) := min{α ∈ + : aα = 0}.
Example 2.4.13 We continue with the set-up of the previous example and assume
in addition that k has an ordering P . Then P induces an ordering Q of K as follows:
for 0 = a = α aα xα ∈ A we define the “leading coefficient” of a as
L(a) := av(a) (∈ k ∗ ),
We claim that the valuation ring ov is convex in K with respect to Q. To see this,
let 0 = x ∈ mv . There exist a, b ∈ A with v(a) = v(b) = 0, and 0 < α ∈
such that x = xα · a/b. It follows that 1 + x = b−2 (b2 + abxα ) ∈ Q since
L(b2 + abxα ) = L(b)2 ∈ P . This shows that 1 + mv ⊆ Q, from which the claim
follows by Proposition 2.2.7.
x ≤ y ⇒ y ∈ X.
a → v(a \ {0})
M → v −1 (M ∪ {∞}).
defines an inclusion reversing bijection from the prime ideals of A to the convex
subgroups of . The inverse map is given by
(c) The ideals of A form a chain (i.e., are totally ordered by inclusion).
(d) A is a Bézout ring, i.e., A has no zero divisors and every finitely generated ideal
is principal. More precisely: for a1 , . . . , an ∈ A and a = Aa1 + · · · + Aan , if
v(a1 ) ≤ v(ai ) for i = 1, . . . , n, then√
a = Aa1 .
(e) If a = A is a radical ideal (i.e., a = a), then a is a prime ideal.
Proof The proofs of (a) and (b) are elementary, (c) follows immediately from (a),
and (d) is clear.
(e) Let a, b ∈ A with ab ∈ a. We may assume that a | b in A, i.e., b = ac for
some c ∈ A. Then b2 = abc ∈ a, and so b ∈ a.
Recall that a sequence
ϕi−1 ϕi
· · · → Gi−1 −−→ Gi −
→ Gi+1 → · · ·
of abelian groups and homomorphisms is called exact if Im(ϕi−1 ) = Ker(ϕi ) for all
i. Sequences of the form
0→G →G→G →0
0 → C → A → B → 0
B → B/A := B ∗ /A∗ = vA (B ∗ ).
∼ .
A / B/A −→ B
Proof This follows from Corollary 2.2.10 and Propositions 2.4.17 and 2.4.18
Definition 2.4.20
(a) Let be an ordered abelian group. The rank of , denoted rank , is the number
of convex subgroups of that are different from (a natural number or ∞).
(b) Let A be a valuation ring. The rank of A, denoted rank A, is defined as the rank
of the value group A = K ∗ /A∗ , where K = Quot A.
Remarks 2.4.21
(1) Many authors (Bourbaki, for example) refer to the rank of as the height of .
(2) By Corollaries 2.4.19 and 2.2.10, the rank of a valuation ring A is also equal to
the number of prime ideals of A that are different from {0} (the Krull dimension
of A), as well as the number of overrings of A in K that are different from K.
(3) The valuation rings of rank 0 are the fields. The ordered abelian groups of rank 1
are precisely the nontrivial subgroups of (R, +) (Theorem 2.1.10). The only
“discrete” ones among these are the infinite cyclic groups, which justifies the
following definition:
Definition 2.4.22 A valuation ring A is called discrete of rank one if the value
group A is infinite cyclic.
Proposition 2.4.23 Let A be a valuation ring which is not a field. The following
are equivalent:
(i) A is discrete of rank one;
(ii) A is noetherian;
(iii) A is a principal ideal domain.
Proof (i) ⇒ (ii): This follows from Proposition 2.4.17(a). (ii) ⇒ (iii): This follows
from Proposition 2.4.17(d). (iii) ⇒ (i): Let mA = Aπ. In particular, π is a prime
70 2 Convex Valuation Rings and Real Places
Exercise 2.4.24 Show that a field K has a non-trivial valuation ring with real
residue field if, and only if, K has a non-archimedean ordering.
Exercise 2.4.25 Let k be a field, let t be a variable, and let V be the set of all non-
trivial valuation rings of k(t) that contain k. Show that there is a natural bijection
between V and P ∪ {∞},where P is the set of monic irreducible polynomials in k[t]
and ∞ is an additional symbol. Every B ∈ V is a discrete valuation ring of rank
one.
Proof Let f ∈ A[t] be a monic polynomial. If f has a root in K, this root is already
in A since A is integrally closed. In particular, the mod mA reduced polynomial has
a root in κ(A), from which (a) follows immediately.
2.5 Residue Fields and Subfields of Convex Valuation Rings 71
Assume now that K is real closed. It follows from the previous argument that
κ(A) has no proper field extensions of odd degree and that P := κ(A)2 satisfies
the ordering axioms (O1) and (O2) (Sect. 1.1). Therefore, κ(A) is real closed if and
only if −1 ∈ κ(A)2 (Proposition 1.5.2).
If A is convex in K, then mA ∩ [1, ∞[ = ∅ (Proposition 2.2.7). Hence, 1 + a 2 ∈
mA for a ∈ A, and so −1 ∈ κ(A)2. Conversely, assume that A is not convex. Then
there exists a ∈ K such that 1 + a 2 ∈ mA (Proposition 2.2.7). It follows that a ∈ A
for if not, we have a −1 ∈ mA and thus 1 = a −2 (1 + a 2 ) − a −2 ∈ mA . We conclude
that −1 ∈ κ(A)2.
Let A be a valuation ring of K. If K is ordered and A is convex in K, then A
contains subfields of K (e.g., Q). More generally, one sees easily that A contains a
subfield if and only if char K = char κ(A) (the so-called “equal characteristic case”;
the unequal characteristic case only occurs when char K = 0 and char κ(A) > 0).
Every subfield F of A is contained in a maximal subfield of A (Zorn’s Lemma).
Also, via F → A → κ(A) we always consider F as a subfield of κ(A).
Proposition 2.5.2 Let A be a valuation ring of K and F a maximal subfield of A.
Then F is algebraically closed in K, and the field extension κ(A)/F is algebraic.
Proof Since A is integrally closed in K, it contains the algebraic closure of F in
K. Let π : A → κ(A) be the residue homomorphism. If there were an a ∈ A
with π(a) transcendental over π(F ), we would have F [a] ∩ mA = {0}, and thus
F F (a) ⊆ A, contradicting the maximality of F .
F ⊆ F0 F1 · · · Fr = K
and
F ⊆ F0 F1 · · · Fs = K
Example 2.5.8 (E. Artin) It is essential that K is real closed for the isomorphism
Fi ∼= Fi !
Let R0 be the real closure of Q (i.e., the relative algebraic closure of Q in R) and
F the relative algebraic closure of Q(e) in R (note that e = 2.71828 . . . is transcen-
dental). Consider K = F (t), equipped with the ordering P0,+ (Example 1.1.13(3)).
Then 0 < t < a for all 0 < a ∈ F . Finally, let F := R0 (e + t) ⊆ K. Since F
is not real closed, F and F are not isomorphic. We claim that not only F , but also
F , is archimedean saturated in K. This is clear for F . To see this for F , we need
to cite a result from field theory (cf. e.g., [54, Vol. 3, p. 199]):
If E/k is a field extension such that k is (relatively) algebraically closed in
E, and if E(x)/E is a simple transcendental extension, then k(x) is also
algebraically closed in E(x).
In the above situation, it suffices to verify that F is algebraically closed in K
since tr.deg.(K/F ) = 1 and o(K/F ) = o(K/Z) = K. This, however, is an
immediate application of the cited theorem (R0 is algebraically closed in F and
K = F (e + t)).
2.6 The Topology of Ordered and Valued Fields 73
In this example, one should also observe that R0 (e) is a subfield of K which
is order isomorphic to F , but which is not archimedean saturated in K (F is
archimedean over R0 (e)).
we have that v −1 (α) is also closed. All sets Bα (a) in the basis that defines Tv are
thus clopen. In particular, (K, Tv ) is totally disconnected. Furthermore, the obvious
assumption that the closure and boundary of Bα (a) are given by {x : v(x − a) ≥ α}
and {x : v(x − a) = α}, respectively, is false.
Proposition 2.6.2 Let (K, P ) be an ordered field and v : K → ∪ ∞ a nontrivial
surjective valuation, compatible with P . Then the order and valuation topologies
on K coincide: TP = Tv .
Proof For every 0 < a ∈ K there exists α ∈ such that Bα (0) ⊆ ]−a, a[, for
instance α = v(a). Indeed, if x ∈ K and v(x) > α = v(a), then x/a ∈ mv ⊆
]−1, 1[, hence |x| < a. Conversely, for every α ∈ there exists 0 < a ∈ K such
that ]−a, a[ ⊆ Bα (0). Indeed, choose a such that v(a) > α. Then if |x| < a, it
follows that |x/a| < 1, hence v (x/a) ≥ 0 and so v(x) ≥ v(a) > α.
The converse of this theorem is false: there may exist valuations that are not
compatible with P , yet induce the same topology as P . This follows easily from the
following:
Exercise 2.6.3 Let v and v be surjective nontrivial valuations of K. Show that if
v is a coarsening of v, then Tv = Tv .
In addition to Krull valuations we introduce absolute values. In number theory
especially they are indispensable.
Definition 2.6.4 Let K be a field. An absolute value on K is a map f : K → R
such that for all a, b ∈ K:
(A1) f (a) ≥ 0, and f (a) = 0 ⇔ a = 0;
(A2) f (ab) = f (a) f (b);
(A3) f (a + b) ≤ f (a) + f (b).
If instead of (A3) the ultrametric inequality
(A4) f (a + b) ≤ max{f (a), f (b)} for all a, b ∈ K,
is satisfied, then f is called an ultrametric absolute value.
In the literature the term “valuation” is often used instead of absolute value.
Depending on whether it is ultrametric or not, this “valuation” is then called “non-
archimedean” or “archimedean”. Clearly this easily causes confusion.
2.7 The Baer–Krull Theorem 75
Remarks 2.6.5
(1) If ϕ : K → C is a field embedding, then a → |ϕ(a)| is a non-ultrametric
absolute value on K. The converse is essentially also true, as the following
classical theorem of A. Ostrowski ([86], 1916) shows:
If f : K → R is a non-ultrametric absolute value on a field K, then there
exists an embedding ϕ : K → C and a real number s with 0 < s ≤ 1,
such that f (a) = |ϕ(a)|s for all a ∈ K. (See for instance also [55, Vol. II,
§9.5].)
(2) If v : K → ∪ ∞ is a surjective valuation with archimedean value group ,
then one obtains an ultrametric absolute value fv on K as follows: choose an
order preserving embedding i : → R (cf. Sect. 2.1) as well as a real number
c with 0 < c < 1 and let fv (a) := ci(v(a)) (a = 0), fv (0) := 0. Conversely,
every ultrametric absolute value can be “logarithmified” to a valuation of rank
≤ 1. Upon defining an equivalence of absolute values in the obvious way,
one then obtains a bijective correspondence between the equivalence classes
of valuations of rank ≤ 1 and the equivalence classes of ultrametric absolute
values. Together with Remark (1) we thus obtain a description of all absolute
values on a field.
Every absolute value f on K defines a metric on K via (a, b) → f (a − b) and
makes K into a metric topological space. We denote the associated topology by Tf .
Proposition 2.6.6 Let (K, P ) be an ordered field and assume that K is either
archimedean, or that K contains a proper convex subring of finite rank. Then
there exists an absolute value f on K such that TP = Tf . (In particular, TP is
metrizable).
Proof If (K, P ) is archimedean with order preserving embedding ϕ : K → R,
then one obtains f as in Remark 2.6.5(1). In the other case, K contains a maximal
proper convex subring A. The valuation vA = v has rank 1, and one obtains f as
in Remark 2.6.5(2). It is clear that Tf = Tv , and Proposition 2.6.2 gives Tf = TP .
We have seen that for every convex subring A of an ordered field (K, P ) there is an
induced ordering P of the residue field κ(A) (Remark 2.2.6(1)). Conversely, given
a valuation ring A of a field K and an ordering Q of κ(A), one can wonder if there
exists an ordering P of K that makes A convex and for which P = Q.
The Baer–Krull Theorem provides a (positive) answer to this question. In fact,
the theorem is considerably more precise in that it determines all such orderings Q.
Consider a surjective valuation v : K ∗ → and let A := ov . Then v induces a
surjective homomorphism v2 : K ∗ /K ∗2 /2. Let {γi : i ∈ I } ⊆ be a subset
such that {γi + 2 : i ∈ I } is a basis of the F2 -vector space /2 and γi > 0 for all
76 2 Convex Valuation Rings and Real Places
2
be two representations of the form (2.1). Then cc ∈ A∗ and, since u = cc u, it
follows that signQ (u ) = signQ (u). Hence σ : a → signQ (u) is a well-defined map
σ : K ∗ → {±1}. This is a homomorphism since, if
a = u c2 πj and b = v d 2 π
j ∈J ∈L
i.e., a + b ∈ P .
Thus we have shown that P is an ordering of K. From the definition of P the
convexity of A with respect to P follows immediately (since 1 + mA ⊆ P ), as does
P = Q.
As a result, the map P → P from Y to Xk is invariant under this action of /2
and takes different values on different orbits. Thus we obtain:
Theorem 2.7.4 (Baer–Krull Theorem, “Invariant” Formulation) Let v : K ∗
be a surjective valuation with valuation ring A and residue field k = A/mA . Let
Y be the set of all orderings of K compatible with v, and Xk the set of all orderings
on Y and a canonical bijection
of k. Then there is a natural free action of /2
→ Xk .
Y /2
2.8 Places
The operations
Let K and L be fields. We denote the disjoint union K ∪ {∞} by K.
+ and · are partially extended to K as follows:
a + ∞ = ∞ + a = ∞ (for all a ∈ K)
a·∞=∞·a =∞ a = 0).
(for all a ∈ K,
Proof (a) If λ(∞) = ∞, then λ(∞) + λ(∞) is defined, and thus also ∞ + ∞,
contradicting (P1). If λ(0) = 0, then λ(0) · λ(∞) is defined, and thus also 0 · ∞,
contradicting (P2).
(b) This is clear for a ∈ {0, ∞} by (a). Thus let a ∈ K ∗ . If λ(a) = λ(−a) = ∞,
then λ(−a) = −λ(a). Otherwise λ(a)+λ(−a) is defined and it follows from (a) that
λ(a) + λ(−a) = λ(0) = 0, hence λ(−a) = −λ(a). Similarly, λ(a −1 ) = λ(a)−1 .
(c) This is trivial.
Notation 2.8.6
(a) A place λ : K→L will from now on be written as λ : K → L ∪ ∞. If ϕ : K →
L is a homomorphism, we denote its associated trivial place K → L ∪ ∞ by ϕ
as well (instead of by
ϕ ).
(b) Let λ : K → L ∪ ∞ be a place. We denote the valuation ring λ−1 (L) of K
by oλ , its maximal ideal by mλ , its residue field oλ /mλ by κλ , and the induced
homomorphism κλ → L by λ. We also denote the value group of oλ by λ :=
K ∗ /o∗λ and the associated canonical valuation by vλ : K ∗ → λ .
Places provide an alternative way of looking at valuations and valuation rings.
We illustrate this in the results below.
Proposition 2.8.7 Let A ⊆ K be a subring. An element a ∈ K is integral over A if
and only if λ(a) = ∞ for every place λ of K that is finite on A.
Proof By Theorem 2.3.7(a).
Proposition 2.8.10 Let (K, P ) and (L, Q) be ordered fields and let λ : K → L∪∞
be a place. The following statements are equivalent:
(i) λ is compatible with P and Q;
(ii) the subring oλ of K is convex with respect to P and the homomorphism
λ : κλ → L is order preserving with respect to P and Q.
Proof (i) ⇒ (ii): It follows from λ(1 + mλ ) = {1} that 1 + mλ ⊆ P , proving the
convexity of oλ with respect to P . It is clear that λ is order preserving.
(ii) ⇒ (i): Let a ∈ P ∩ oλ and let a be the image of a in κλ . Then a ∈ P and so
λ(a) = λ(a) ∈ Q.
Example 2.8.12 If K is a real closed field and (L, Q) an ordered field, then every
place K → L ∪ ∞ is order preserving with respect to Q.
We finish this section with an extension theorem for real places.
Lemma 2.8.13 Let K ⊆ L be a field extension and B ⊆ L a valuation ring of L.
Then A := K ∩ B is a valuation ring of K and mA = K ∩ mB . If L is algebraic
over K, then κ(B) is algebraic over κ(A).
Proof It is clear that A is a valuation ring of K. If a ∈ K ∗ , then
a ∈ mA ⇔ a −1 ∈ A ⇔ a −1 ∈ B ⇔ a ∈ mB .
in κ(B), where the cj = aj +m /am are in κ(A). It follows that b is algebraic over
κ(A).
Let R be a real closed field. In this section we will explicitly determine the orderings
of R(t), R((t)) and Quot R{t}.
We start with the rational function field in one variable R(t). Let c ∈ R. Then c
determines a place λc : R(t) → R ∪ ∞, namely λc (f ) = f (c) (setting f (c) = ∞
if f has a pole at c, as usual). The valuation ring associated to λc is R[t](t −c) . It is
discrete of rank one and t − c is a generator of its maximal ideal. By the Baer–Krull
Theorem, R(t) has precisely two λc -compatible orderings: Pc,+ and Pc,− . These are
characterized by
Its maximal ideal is generated by t −1 , and as above there are precisely two λ∞ -
compatible orderings P∞,+ and P∞,− , characterized by t −1 ∈ P∞,+ and −t −1 ∈
P∞,− . (In a certain sense this notation would be more consistent the other way
around. See the following paragraph however.)
The following geometric interpretation of these orderings may be helpful. If 0 =
f ∈ R(t), then f has only finitely many zeroes and poles. Given any c ∈ R it
therefore makes sense to speak of the sign of f immediately to the right or to the
left of c. Here “f > 0 immediately to the right of c” means that there exists an
ε > 0 such that f (a) > 0 for all a ∈ ]c, c + ε[. Similarly, f (c) has a well-defined
sign for c → +∞ and for c → −∞. Thus we obtain
Proposition 2.9.1 The orderings Pc,+ and Pc,− , for all c ∈ R ∪ {∞}, are precisely
the orderings of R(t) that are non-archimedean over R. If R = R, these are all the
orderings of R(t).
Proof Let P be an ordering of R(t) that is non-archimedean over R and let A :=
oP R(t)/R be the convex hull of R.
Let t ∈ A. Then p := mA ∩ R[t] is a prime ideal of R[t] (the centre of A in
R[t]). Since R[t]p ⊆ A and A = R(t) we have p = (0). Thus p = R[t] · f for some
monic irreducible polynomial f ∈ R[t]. Since κ(A) is a real field and R[t]/(f ) is a
subfield of κ(A) there exists an element c ∈ R such that f = t − c. Since R[t](t −c)
is a valuation ring of R(t) (by Proposition 2.2.8) and is dominated by A, we have
A = R[t](t −c) (by Corollary 2.3.9) and it follows that P = Pc,+ or P = Pc,− .
If t ∈ A, we let u := t −1 and obtain as above that A = R[u](u) . Hence P =
P∞,+ or P = P∞,− .
The statement about R follows from Corollary 2.1.13.
As a consequence of this proof we obtain:
Corollary 2.9.2 Let k be an arbitrary field. Then the nontrivial valuation rings of
k(t) that contain k are precisely the following:
(1) A = k[t](f ) with f ∈ k[t] monic and irreducible;
(2) A = k[t −1 ](t −1) .
In each case the value group is Z and the residue field is a finite simple field
extension of k.
In order to determine the remaining orderings of R(t) we consider for every
ordering P of R(t) the sets
UP := {a ∈ R : t − a ∈ P } and OP := {a ∈ R : a − t ∈ P }.
84 2 Convex Valuation Rings and Real Places
respectively, cf. Proposition 1.7.2. There is precisely one i ∈ {0, . . . , r} such that
Ii ∩ U = ∅ = Ii ∩ O and so we say that f has sign εi at ξ . Thus, if we let
given by P → ηP and η → Pη .
Proof Let P be archimedean over R and η := ηP = (UP , OP ). Then UP = ∅ =
OP and for every a ∈ UP there exists 0 < b ∈ R such that b < t − a (otherwise
2.9 The Orderings of R(t), R((t)) and Quot R{t} 85
r
s
f (t) = α (t − ai ) (t − bj )2 + cj2
i=1 j =1
Corollary 2.9.6 There exists a natural bijection between the orderings of R(t) and
the generalized Dedekind cuts of R.
Next on our list is the field R((t)) of formal Laurent series i≥n ai t i (with n ∈ Z
and ai ∈ R). Recall that R((t)) is the field of fractions of the ring A = R[[t]] of
formal power series in one variable over R. Let 0 = f (t) = i≥n ai t i with an = 0.
Then v(f ) := n defines a valuation v of R((t)) with valuation ring A, maximal ideal
mA = tA, value group Z and residue field A/mA = R. Since for every f ∈ mA the
element
2
1+f = 1/2
i
f i
i≥0
and
0 → ν → ν◦λ → λ → 0 (2.2)
ϕ : α → (πα, α − sπα)
∼ ( × ) = (Z × )
ν◦λ → λ ν lex ν lex
for all f ∈ K ∗ . This solves the problem of expressing the valuation vν◦λ in terms
of vλ , vν and λ. (The argument can be repeated without any assumptions on λ , but
then the description of the section s will be more complicated in general.)
Notation 2.10.4 We denote by Zrantilex the group Zr equipped with the antilex-
icographic order, i.e., the positive elements are the tuples (a1 , . . . , as , 0, . . . , 0)
(1 ≤ s ≤ r) with as > 0. Of course, Zrantilex ∼ = Zrlex as ordered abelian groups;
we use the antilexicographic order just to simplify our notation below.
Let us consider the case of rational function fields (in several variables). Let
k be an arbitrary field, t1 , . . . , tr algebraically independent variables over k and
Ki := k(t1 , . . . , ti ) for 0 ≤ i ≤ r. Given a tuple c = (c1 , . . . , cr ) ∈ k r we construct
a place λ : Kr → k ∪ ∞ such that λ(f ) = f (c) for all f ∈ k[t1 , . . . , tr ] as follows:
For every i = 1, . . . , r there is a unique place λi : Ki → Ki−1 ∪ ∞ over Ki−1
such that λi (ti ) = ci (cf. Corollary 2.9.2). Let λ := λ1 ◦ · · · ◦ λr . Since λi = Z
for i = 1, . . . , r, it follows by induction from Proposition 2.10.2 and Lemma 2.10.3
that λ ∼
= Zrlex . We wish to determine vλ explicitly. For α = (α1 , . . . , αr ) ∈ Zr we
88 2 Convex Valuation Rings and Real Places
(Let be an ordered abelian group. Then ⊗Q has a unique ordering that makes
the embedding → ⊗Q order preserving. In the statement of the theorem, λ ⊗Q
is equipped with this ordering.)
Proof The inclusion K ∗ ⊆ L∗ induces an order preserving embedding λ ⊆ μ .
We claim that the group μ / λ is torsion. Indeed, let b ∈ L∗ and let a0 , . . . , an ∈ K
be such that an bn + · · · + a1 b + a0 = 0 and an = 0. Then there exist indices i and
j such that 0 ≤ i < j ≤ n and vμ (ai bi ) = vμ (aj bj ) < ∞ (for otherwise,
vμ ( ak bk ) = min v(ak bk ) < ∞)), and it follows that (j − i)vμ (b) = vμ aaji ∈
k
λ .
Since L is real closed or algebraically closed, μ is divisible (every positive
element, resp. every element of L has arbitrary roots) and so μ = Q·λ = λ ⊗Q.
the field Kr = k((t1 )) · · · ((tr ))!) Let λi : Ki → Ki−1 ∪ ∞ be the place over Ki−1 ,
associated to Ki−1 [[ti ]], and let λ = λ1 ◦· · ·◦λr : Kr → k ∪∞. Then λ ∼ = Zrlex with
lexicographic basis vλ (tr ), . . . , vλ (t1 ), as in Proposition 2.10.5. If k ⊆ K ⊆ Kr is
an intermediate field with t1 , . . . , tr ∈ K, then we also have λ|K = Zrlex .
Using some more commutative algebra we obtain an interesting application
(which will play no further role however):
Let A be a regular local k-algebra with κ(A) = k, A = lim A/mn its completion
←− A
and (f1 , . . . , fr ) a minimal set of generators of mA . It is well-known that there
exists a k-isomorphism A → ∼ k[[t , . . . , t ]] such that f → t for all i = 1, . . . , r,
1 r i i
cf. [81, p. 206] or [13, Ch. IX, § 3, no. 3]. Let K := Quot A. It follows from
K ⊆ Quot A ∼ = k((t1 , . . . , tr )) ⊆ Kr that the place λ|K : K − → k ∪ ∞ also has
k
value group Zrlex , with lexicographic basis v(fr ), . . . , v(f1 ) (where v := vλ|K ).
Furthermore, its valuation ring dominates A.
We briefly recall some facts from field theory. Let L ⊇ K be a field extension.
Then all maximal families of elements of L that are algebraically independent over
K have the same cardinality. Any such family is called a transcendence basis of L
over K and its cardinality is called the transcendence degree of L over K, denoted
tr.deg.(L/K). The field L is called a d-dimensional function field over K if L is
finitely generated over K and tr.deg.(L/K) = d. (The terminology comes from
the fact that up to K-isomorphism such fields are precisely the fields of rational
functions on irreducible algebraic K-varieties of dimension d, cf. [97, I, §3] or [45].)
The purpose of this section is to provide a proof of:
Theorem 2.11.1 Let R be a real closed field, K an r-dimensional real function
field over R and (t1 , . . . , tr ) a fixed transcendence basis of K over R. Then there
exist elements ai , bi ∈ R with ai < bi for i = 1, . . . , r such that for every real
closed overfield S ⊇ R the following property is satisfied:
(Tr ) For every tuple c = (c1 , . . . , cr ) ∈ S r such that ai < ci < bi for i = 1, . . . , r,
there exists a place λ : K → S ∪ ∞ over R that satisfies λ(ti ) = ci for
i = 1, . . . , r.
Note that such a place λ must be trivial (i.e., a homomorphism) if c1 , . . . , cr
are algebraically independent over R. (Indeed, in this case we have R[t1 , . . . , tr ] ∩
λ−1 (0) = {0}, which implies R(t1 , . . . , tr ) ⊆ oλ . It follows that oλ = K since oλ is
integrally closed.)
Corollary 2.11.2 (Lang’s Embedding Theorem) Let R be a real closed field and
K a real function field over R. For every real closed overfield S of R such that
tr.deg.(S/R) ≥ tr.deg.(K/R) there exists an R-homomorphism K → S.
90 2 Convex Valuation Rings and Real Places
Proof The statement follows from Theorem 2.11.1 and the remark following it.
Indeed, there are elements ci ∈ S with ai < ci < bi for i = 1, . . . , r such that
c1 , . . . , cr are algebraically independent over R; this is clear for r = 1 and follows
by induction for all r.
We first establish two auxiliary results.
Lemma 2.11.3 Let L ⊇ K be a finite field extension, B a valuation ring of L and
A = K ∩ B. Then the ramification index e = [B : A ] and the residue degree
f = [κ(B) : κ(A)] are finite and ef ≤ [L : K].
Proof Write v = vB . Consider elements b1 , . . . , br ∈ B\{0} such that the images of
the v(bi ) in B / A are distinct, and elements c1 , . . . , cs ∈ B such that the images
cj of the cj in κ(B) are linearly independent over κ(A). We will show that for all
aij ∈ K (1 ≤ i ≤ r, 1 ≤ j ≤ s),
v aij bi cj = min v(aij bi ).
i,j
i,j
where aij := aij /aik(i) ∈ A. Since the cj are linearly independent over κ(A) (and
aik(i) = 1) we have aij cj ∈ B ∗ and thus
j
v aij bi cj = v(aik(i) ) + v(bi ) (i = 1, . . . , r).
j
Proposition 2.11.6 (
Tr ) implies (Tr ) for all r ≥ 1.
Proof Let c ∈ S r with ai < ci < bi and c1 , . . . , cr not necessarily algebraically
independent over R. Let u1 , . . . , ur be independent variables over S and T :=
S(u1 , . . . , ur ) the r-dimensional rational function field over S. Consider the place
μ : T → S ∪ ∞ over S with μ (ui ) = 0 for i = 1, . . . , r, described in Sect. 2.10.
By Corollary 2.8.16 there exists a real closure T of T and an extension μ : T → S ∪
∞ of μ . Since the ui are infinitely small with respect to S we have ai < ci +ui < bi
for i = 1, . . . , r. Since the ci + ui are algebraically independent over R, there exists
a homomorphism ϕ : K → T such that ϕ(ti ) = ci + ui by property ( Tr ). The place
λ := μ ◦ ϕ : K → S ∪ ∞ then satisfies λ(ti ) = ci for i = 1, . . . , r.
Let us now prove ( T1 ). Assume thus that r = 1, that R and K are as in
Theorem 2.11.1 and that F := R(t), where t := t1 . Let tr : K → F be the trace
of the finite extension K ⊇ F and let g1 , . . . , gn ∈ F ∗ be such that tr∗ 1K ∼ =
g1 , . . . , gn . After multiplying the gi with the square of their denominators we may
assume without loss of generality that all gi are in R[t]. Let d1 < · · · < dN be the
distinct zeroes of g1 · · · gn in R, d0 := −∞ and dN+1 := +∞. Every gi has a
92 2 Convex Valuation Rings and Real Places
n
εij = signP tr∗ 1K
i=1
n
n
signPc tr∗ 1K = signPc (gi ) = εij ≥ 1,
i=1 i=1
tr.deg.(K/F ) = 1 (and S/F is algebraic) this place is nontrivial and satisfies μ(z) =
∞. Let A := oμ and λA : K → κ(A) ∪ ∞ the canonical place associated to A. By
Lemma 2.11.4 we have [κ(A) : F ] < ∞ and A ∼ = Z. Furthermore, since κ(A) is
real and tr.deg. κ(A)/R = tr.deg.(F /R) = r − 1 it follows by induction that there
exists a place ν : κ(A) → R ∪ ∞ over R such that ν λA (z) = ∞ and ν ∼ = Zr−1
lex .
Let λ := ν ◦ λA : K − → R ∪ ∞. Then λ(z) = ∞ and λ ∼ = Zlex by Sect. 2.10.
r
R
n
n
n
1 ≤ signP tr∗ 1K = signP (gj ) = sign μ(gj ) = sign gj (p),
j =1 j =1 j =1
where the final equality follows from the fact that μ(f ) = f (p) for all f ∈
R[t1 , . . . , tr ].
In order to finish the proof of Theorem 2.11.1 we need the following proposition
that we will prove at the end of this section:
Proposition 2.11.10 Let g ∈ R[x1 , . . . , xr ] be a polynomial and p ∈ R r such that
g(p) = 0. Then there exists an element ε ∈ R, ε > 0, such that for every real closed
overfield S of R and every q ∈ S r with pi − ε < qi < pi + ε for i = 1, . . . , r we
have sign g(p) = sign g(q).
For g1 , . . . , gn and p as above, the lemma provides us with an 0 < ε ∈ R. Let
S ⊇ R be a real closed overfield and consider a tuple c = (c1 , . . . , cr ) ∈ S r such
that c1 , . . . , cr are algebraically independent over R and pi − ε < ci < pi + ε for
i = 1, . . . , r. Let ψ : F → S be the embedding determined by ψ(ti ) = ci and Pc
94 2 Convex Valuation Rings and Real Places
n
n
n
signPc tr∗ 1K = signPc (gj ) = sign gj (c) = sign gj (p) ≥ 1,
j =1 j =1 j =1
where the third equality follows from Proposition 2.11.10. It follows from
Sects. 1.11 and 1.12 that ψ has an extension ϕ : K → S as required. This concludes
the proof of Theorem 2.11.1
Proof of Proposition 2.11.10 Let x = (x1 , . . . , xr ) and let
g(p + x) − g(p) = cα x α
α
be the Taylor expansion of g at p (where the sum is over all 0 = α ∈ Nr0 , and
almost all cα are 0). If a ∈ S r is such that |ai | < ε ≤ 1 for i = 1, . . . , r, then
|g(p +a)−g(p)| ≤ ε |cα |. Hence, it suffices to choose ε > 0 such that ε |cα | <
|g(p)| and ε ≤ 1. α α
f = f (t , tn ) = f0 (t ) · tnd + · · · + fd (t ),
Zk (f ) = {a ∈ k n : f (a) = 0}.
Dk (f ) = {a ∈ k n : h(a) = 0}.
Remark 2.12.5 Let (k, P ) be an ordered field with real closure R. It follows from
Proposition 2.12.4(b) that every function of the form a1 f12 +· · ·+ar fr2 with ai ∈ P
and fi ∈ k(t) is positive semidefinite over R. A rational function which is positive
semidefinite over k is in general no longer positive semidefinite over R. This is
96 2 Convex Valuation Rings and Real Places
because k is usually not dense in R, and f (considered over R) may become negative
in the “gaps”. (The reader may want to construct an example, cf. Sect. 2.1.)
In 1900 David Hilbert gave his famous address to the International Congress of
Mathematicians in Paris in which he presented his list of 23 unsolved mathematics
problems. Since then many of these problems have been solved, often involving the
development of completely new methods. This is especially true of the 17th problem
about the representation of forms by sums of squares and its solution by Emil Artin
in 1927, cf. [3, 4]. Artin discovered the notion of ordering of a field, putting it to
elegant and successful use. His ideas sparked new directions in the development of
real algebra in the twentieth century (see the Preface).
Hilbert’s 17th Problem can be formulated as follows:
Hilbert’s 17th Problem Given a positive semidefinite polynomial f =
f (t1 , . . . , tn ) ∈ R[t1 , . . . , tn ], do there exist an integer r ≥ 1 and rational functions
g1 , . . . , gr ∈ R(t1 , . . . , tn ) such that f = g12 + · · · + gr2 ?
Supplementary question: If k ⊆ R is a subfield that contains the coefficients of f ,
can the gi already be found in k(t1 , . . . , tn )?
Remarks 2.12.6
(1) In the formulation of the supplementary question (which is also due to Hilbert)
it would be better to consider representations in the sense of Remark 2.12.5
since otherwise it is immediately clear that there is no solution in general.
(2) An obvious follow-on question is whether there exists a representation f =
g12 +· · ·+gr2 where g1 , . . . , gr are actually polynomials. Hilbert already showed
in 1888 that this is indeed the case if n = 1, but that for n ≥ 2 this is false
in general. (Hilbert’s proof actually provided more precise details.) The first
example of a positive semidefinite polynomial that cannot be written as a sum
of squares of polynomials was not published until 1967. Indeed, in [84] Motzkin
considered the polynomial (here presented in homogeneous form)
f (x, y, z) = x 4 y 2 + x 2 y 4 + z6 − 3x 2 y 2 z2 .
1 4 2
(x y + x 2 y 4 + z6 ) ≥
3
x 6 y 6 z6 .
3
A further example is given by
g(x, y, z) = x 4 y 2 + y 4 z2 + z4 x 2 − 3x 2y 2 z2 .
In [26] one can find a simple method for proving that f and g are not sums of
squares of polynomials, as well as further examples.
2.12 Artin’s Solution of Hilbert’s 17th Problem and the Sign Change Criterion 97
f = a1 g12 + · · · + ar gr2 .
The use of the word “criterion” is motivated by (ii) which signifies that f changes
sign “along” the hypersurface H := {f = 0}. Algebraic geometry tells us that K is
the field of rational functions on H .
The proof makes use of the following proposition, which we establish first:
Proposition 2.12.9 Let A be a unique factorization domain, E = Quot A and
f ∈ A[u]\A a polynomial in one variable u, irreducible in A[u]. Then A[u]/(f ) →
E[u]/(f ) is injective and E[u]/(f ) is a field (and thus the quotient field of
A[u]/(f )).
Proof The statement follows immediately from the fact that since A is a UFD, A[u]
is also a UFD (a consequence of Gauss’s Lemma, cf. [73, IV, Theorem 2.1] or [55,
Vol. I, §2.16]): if g ∈ A[u] and h ∈ E[u] such that g = f h in E[u], then there exist
0 = a ∈ A and h ∈ A[u] such that ag = f h (in A[u]). Since f is not a divisor
of a, it follows that f | g in A[u]. A similar straightforward argument shows that f
remains irreducible in E[u].
Proof of Theorem 2.12.8 (i) ⇒ (ii): Assume that f is semidefinite. Without loss of
generality we may assume that f ≥ 0 on R n . By Artin’s theorem there exists an
equation of the form
r
f h2 = ai gi2 ,
i=1
(2.3)
We start by recalling some definitions. Let A be a ring and a an ideal of A. The set
√
a := {a ∈ A : a n ∈ a for some n ∈ N}
√
√ the radical of a. If a = a, then a is called a radical ideal.
is an ideal of A, called
The ideal Nil A := (0) is the nilradical of A, and A is reduced if Nil A = (0). As
usual, we also let Ared := A/ Nil A.
Definition 3.1.1 The (Zariski) spectrum Spec A of A is the set of all prime ideals
of A. Every p ∈ Spec A has an associated residue field
Corollary 3.1.8 The closed points of Spec A are precisely the maximal ideals of
A.
Recall that a topological space X is called a T0 space if for every two distinct
points x, y ∈ X, we have x ∈ {y} or y ∈ {x}. If we replace “or” by “and”, then X is
a T1 space.
104 3 The Real Spectrum
The Zariski topology is thus very weak as a rule and not even a T1 topology.
However:
Exercise 3.1.9 Show that the Zariski topology is a T0 topology.
√
Corollary
√ 3.1.12 For every two ideals a, b of A we have V(a) ⊆ V(b) ⇔ a ⊇
b.
√
Proof of Proposition 3.1.11 We must show that a = {p ∈ Spec A : a√⊆ p}.
The inclusion ⊆ is clear. For the reverse inclusion, let f ∈ A with f ∈ a. By
Zorn’s Lemma, there exists and ideal p that contains a and that is maximal for the
property p ∩ {1, f, f 2 , . . . } = ∅. Then p is a prime ideal by Exercise 3.1.10, and
f ∈ p.
For a = 0, Proposition 3.1.11 says that Nil A is the intersection of all the prime
ideals of A. In particular, A and Ared have “the same” Zariski spectrum (more
precisely: the map A → Ared induces a homeomorphism Spec Ared → Spec A).
Definition 3.1.13 A topological space X is called irreducible if it is not the union of
two proper closed subsets. A point x ∈ X is called a generic point of X if X = {x}.
If X contains a generic point, then X is irreducible. One is often interested in
the converse. Note that in case X is a T0 space (for example a subspace of some
Spec A), then X has at most one generic point.
Proposition 3.1.14 Every nonempty closed irreducible subspace of Spec A con-
tains a (unique) generic point. Hence, the nonempty closed irreducible subsets of
Spec A are precisely the sets V(p) where p ∈ Spec A.
Proof Let ∅ = Y ⊆ Spec A be closed and irreducible, and let a := I(Y ). By
Proposition 3.1.7 we must show that a is a prime ideal. Since Y = ∅, we have
a = A. If f, g ∈ A are such that fg ∈ a, then Y ⊆ V(f ) ∪ V(g). Since Y is
irreducible, we obtain without loss of generality that Y ⊆ V(f ), and in particular
that f ∈ a.
Recall that a topological space is quasi-compact if each of its open coverings has
a finite subcovering, and compact if it is quasi-compact and Hausdorff.
3.1 The Zariski Spectrum. Affine Varieties 105
Remark 3.1.16 We use K̊(Spec A) to denote the set of all finite unions D(f1 ) ∪
· · · ∪ D(fr ), where r ≥ 1 and fi ∈ A. By Proposition 3.1.15, the elements of
K̊(Spec A) are precisely the open quasi-compact subsets of Spec A. It follows in
particular that K̊(Spec A) is completely determined by the topological space Spec A.
This is generally not the case for the open basis {D(f ) : f ∈ A} of Spec A, cf.
Sect. 3.4.
Let us look at the functorial behaviour of the Zariski spectrum. Let ϕ : A → B
be a ring homomorphism. If q ∈ Spec B, then ϕ −1 (q) is a prime ideal of A. Hence
ϕ induces a map of spectra in the opposite direction,
is commutative.
106 3 The Real Spectrum
Thus, interpreting the rings A and B as rings of functions on their spectra (as
explained in Remark 3.1.3), the map ϕ : A → B just pulls back functions on Spec A
via ϕ ∗ : Spec B → Spec A and via the embeddings ϕq of the residue fields.
Two special cases are worthy of investigation in more detail. For the first case, let
S be a multiplicative subset of A and iS : A → S −1 A the canonical homomorphism
a → a/1.
Proposition 3.1.17 The map
is inverse to iS∗ . If a
s ∈ S −1 A, then
which shows that the map iS∗ is open, and thus a homeomorphism. For the second
statement, let q ∈ Spec S −1 A and p := iS∗ (q) = iS−1 (q) (i.e., q = S −1 p).
It is immediately clear that the embedding A/p → (S −1 A)/(S −1 p) induces an
isomorphism of the quotient fields.
By Proposition 3.1.17 we can identify Spec S −1 A with the subspace DA (S) =
{p ∈ Spec A : p ∩ S = ∅} of Spec A. In particular we can interpret every basic open
set DA (f ) as the Zariski spectrum of a ring, namely the ring f −∞ A := S −1 A,
where S := {1, f, f 2 , . . . }.
For the second special case, let a ⊆ A be an ideal and ϕ : A → A/a the residue
homomorphism.
Proposition 3.1.18 The map
is a homeomorphism from Spec A/a to the closed subspace VA (a) of Spec A. The
residue field embeddings ϕq with q ∈ Spec A/a are all isomorphisms.
Proof Analogous to the proof of Proposition 3.1.17, where this time ϕ ∗ DA/a (f +
a) = DA (f ) ∩ VA (a) for all f ∈ A.
The observation we made above for S −1 A goes through, mutatis mutandis, for
the ring A/a. The closed subspaces of Spec A are thus also spectra of rings.
3.1 The Zariski Spectrum. Affine Varieties 107
In the final part of this section we will explain what we mean by affine varieties
over a field. Our point of view is in essence Weil’s in the sence that all varieties
are reduced. We will show that by considering the associated function algebras we
obtain a coordinate-free description.
Let k be a field with algebraic closure k. If K ⊇ k is an overfield, T a subset of
k[t1 , . . . , tn ] and V a subset of K n , we write
and
n
Definition 3.1.19 An affine k-variety is a subset V of k of the form V = Zk (a),
where a is an ideal of k[t1 , . . . , tn ]. If k ⊆ K ⊆ k is an intermediate field, then the
m
elements of V (K) := V ∩ K n are called the K-rational points of V . If V ⊆ k
n
and W ⊆ k are affine k-varieties, then a map ϕ : V → W is called a k-morphism
from V to W if there exist polynomials f1 , . . . , fn ∈ k[t1 , . . . , tm ] such that ϕ(x) =
f1 (x), . . . , fn (x) for all x ∈ V .
Remark 3.1.20 It is not sufficient to just consider the k-rational points V (k) of V
as these usually do not contain enough information about V . For example, V (k) can
be empty even though V is not the empty variety (concrete example: the R-variety
in the plane defined by 1 + t12 + t22 = 0).
n
Remark 3.1.21 If V ⊆ k is a k-variety, then Ik (V ) is clearly the largest ideal a of
k[t1 , . . . , tn ] such that V = Zk (a).
the affine (k-)algebra of V . It is clear that k[V ] is an affine (i.e., a finitely generated)
reduced k-algebra, which can be interpreted as the algebra of k-morphisms V →
k. If ϕ : V → W is a k-morphism of affine k-varieties, then ϕ induces a
homomorphism ϕ ∗ : k[W ] → k[V ] of k-algebras by “pulling back”.
The point is now that up to isomorphism, V and k[V ] are the “same”. More
precisely, V is the same as k[V ] plus a choice of coordinates. Let us elaborate this.
Consider any overfield K of k, and denote the set of k-algebra homomorphisms
from k[V ] to K by Homk (k[V ], K). Writing ti := ti + Ik (V ) ∈ k[V ], there is a
108 3 The Real Spectrum
canonical bijection
Homk k[V ], K −∼
→ ZK Ik (V ) , x → x(t1 ), . . . , x(tn ) .
follows easily via localization that the closed points of Spec A are dense in Spec A
and the theorem then follows by the Weak Nullstellensatz (Proposition 3.1.11). For
a complete proof of Hilbert’s Nullstellensatz we refer to [68] or [13, Ch. V, §3.3].
Assume now that A is an affine reduced k-algebra. Then Hilbert’s Nullstellensatz
says that for every 0 = f ∈ A there exists x ∈ VA such that f (x)= 0, which in
turn signifies that the canonical homomorphism A → k[VA ] = A/ Ker(x) is an
x∈VA
isomorphism.
n
Summary We have seen that an affine k-variety V ⊆ k is “the same” as an affine
reduced k-algebra A together with a system of generators of A. Henceforth, when
talking about an affine k-variety we will thus usually mean an affine reduced k-
algebra A together with the set VA = Homk (A, k), which offers the advantage of a
coordinate-free approach.
is a bijection from VA (k) = Homk (A, k) to the set of those maximal ideals m of A
for which the map k → A → A/m is an isomorphism (i.e., A = m + k · 1). The
topology induced by Spec A on VA (k) is also called the Zariski topology on VA (k).
However, if k = k, then the variety VA = VA (k) associated to A cannot in general
be squeezed into Spec A: if m ∈ Spec A is a closed point there exist in general
several k-embeddings A/m → k and accordingly several points in VA associated to
m.
On the other hand, if k = k is algebraically closed, then x → Ker x is a bijection
from VA to the space of closed points in Spec A. Therefore, the topology induced
on VA by VA → Spec A is also called the Zariski topology on the variety VA . Of
course, historically speaking the situation was reversed since O. Zariski studied the
n
topology on k-varieties V ⊆ k that was named after him long before the spectrum
was conceived in the 1950s.
110 3 The Real Spectrum
Remarks 3.2.2
(1) The trivial ring is real reduced, but not real. Every nontrivial real reduced ring is
also real. The ring R[x, y]/(x 2 +y 2 ) is real (cf. (5) below), but not real reduced.
(2) If A is real reduced, then A is reduced, and every subring of A is also real
reduced.
(3) An integral domain is real reduced if and only if its field of fractions is real. A
field K is real (reduced) if and only if it has an ordering (cf. Sect. 1.1).
(4) A valuation ring A is real reduced if and only if it is real. (Proof: Real reduced
implies real is clear from (1). Let A be real and let ai ∈ A, ai = 0 such that
a12 +· · ·+an2 = 0. We may assume that v(a1 ) ≤ · · · ≤ v(an ) < ∞. Letting bi :=
ai /a1 ∈ A for i = 1, . . . , n, we then have 1 + b22 + · · · + bn2 = 0, contradiction.)
For example, if A is a residually real valuation ring, then A is real (reduced).
This follows from Corollary 2.7.2 and (5) below. A real (reduced) valuation ring
is in general not residually real though.
(5) If ϕ : A → B is a homomorphism and B is real, then A is also real.
(6) If ϕ : A → B is a homomorphism and b ⊆ B is a real (resp. real reduced) ideal,
then ϕ −1 (b) is also real (resp. real reduced). If ϕ is surjective, then b is real
(resp. real reduced) if and only if ϕ −1 (b) is real (resp. real reduced).
(7) Let A = A1 × · · · × Ar be a direct product of rings with Ai = 0. Then A is real
reduced if and only if the factors A1 , . . . , Ar are all real reduced, and A is real
if and only if at least one of the factors A1 , . . . , Ar is real.
1 In Einführung in die reelle Algebra we followed a suggestion of T.Y. Lam [71], and used
“semireal” and “real” instead of “real” and “real reduced”, respectively. In the meantime the
terminology we use here has become standard.
3.2 Reality for Commutative Rings 111
but their converses do not hold in general. (Proof of the first implication: If
−1
i (ai /si ) = 0 in S A, then i (ai ti ) = 0 in A for certain ti ∈ S. If A is
2 2
Notation 3.2.3 We denote the set of all real reduced prime ideals of A by
(Spec A)re , i.e.,
√
Definition 3.2.11 If a ⊆ A is an ideal, its real radical re
a is defined as the
intersection of all real reduced prime ideals p ⊇ a.
Remarks 3.2.12
√
(1) A ring is real reduced if and only if re
(0) = (0), cf. Proposition 3.2.4.
(2) By Proposition 3.2.10 we have
√
re
a = {f ∈ A : ∃ N ∈ N, ∃ a1 , . . . , ar ∈ A such that f 2N + a12 + · · · + ar2 ∈ a}.
When giving a direct proof that the right hand side is an ideal, showing closure
under addition requires some effort: assume f 2M + a12 + · · · + ar2 ∈ a and
g 2N + b12 + · · · + bs2 ∈ a. We may take M = N. Then (f + g)4N + (f − g)4N is
a sum of elements of the form f 2m g 2n with m + n = 2N. Therefore, m ≥ N or
n ≥ N in every such term. It follows that −f 2m g 2n is a sum of squares modulo
a, which yields (f + g)4N + c12 + · · · + ct2 ∈ a.
√
Let R be a real closed field, C = R( −1), A an affine R-algebra, and
VA = HomR (A, C) the variety associated to A. As before, we interpret VA (R) =
HomR (A, R) as a subspace of Spec A (cf. Sect. 3.1), as a matter of fact even of
(Spec A)re and present new formulations of the theorems of Artin and Lang that
were established in Chap. 2:
Theorem 3.2.13 (Stellensatz) An affine R-algebra A is real if and only if its
associated variety VA has real points, i.e., if and only if VA (R) = ∅.
Proof The sufficient direction is clear. For the necessary direction, assume that A is
real. By Corollary 3.2.8 there exists p ∈ (Spec A)re , and K := κ(p) is a real function
field over R. Let f1 , . . . , fr be generators of A, and f1 , . . . , fr their images in K.
By Theorem 2.11.8 there exists a place λ : K → R ∪∞ over R such that λ(fi ) = ∞
for i = 1, . . . , r. The composition λ ◦ ρp is an element of VA (R).
In fact, the following stronger statement is true:
Corollary 3.2.14 Let A be an affine R-algebra. Then VA (R) is Zariski dense in
(Spec A)re .
Proof Let f ∈ A such that DA (f ) ∩ (Spec A)re = ∅. Then f −∞ A is a real affine
R-algebra, and so HomR (f −∞ A, R) = ∅ by Theorem 3.2.13. Every element of
this set yields an element of DA (f ) ∩ VA (R) (via composition with A → f −∞ A).
Let A be a ring.
Definition 3.3.1 (M. Coste, M.-F. Roy, 1979)
(a) The real spectrum Sper A of A is the set of all pairs α = (p, T ) with p ∈ Spec A
and T an ordering of the field κ(p) = Quot A/p. (Other notations in use are
Specr A and SpecR A. In Sect. 3.6 we will motivate our notation Sper A.) The
prime ideal p is called the support of α, denoted p = supp α.
(b) For α = (p, T ) ∈ Sper A, k(α) denotes the real closure of κ(p) with respect to
T . There is a canonical homomorphism
ρp
rα : A −−→ κ(p) → k(α).
Remarks 3.3.2
(1) A difference with the Zariski spectrum is that the Harrison subbasis is usually
not an open basis. An example of an open basis for the Harrison topology is
given by
Proposition 3.3.3 The map supp : Sper A → Spec A is continuous. Its image is the
subspace (Spec A)re of real reduced prime ideals.
Proof We have supp−1 DA (f ) = H̊A (f 2 ). The second statement is clear.
An ordering T of κ(p) is determined by T ∩(A/p), i.e., by ρp−1 (T ). Consequently
the elements of Sper A can also be interpreted as certain subsets of A. Given α =
(p, T ) ∈ Sper A, let Pα := ρp−1 (T ) = {f ∈ A : f (α) ≥ 0}. Then P = Pα satisfies
the following properties:
(RO1) P + P ⊆ P, PP ⊆ P;
(RO2) P ∪ (−P ) = A;
(RO3) supp P := P ∩ (−P ) is a prime ideal of A.
Since (RO1) and (RO2) already imply that supp P is an ideal of A, we may replace
(RO3) by
(RO3’) −1 ∈ P , and: a ∈ P , b ∈ P ⇒ −ab ∈ P (a, b ∈ A).
It is easy to see that every subset P ⊆ A that satisfies properties (RO1) to (RO3)
defines an element αP ∈ Sper A, namely αP = (p, P ), where p := supp P and P
is the ordering on κ(p) induced by P (for a, b ∈ A, b ∈ p we have: ρp (a)/ρp (b) ∈
P ⇔ ab ∈ P ). This motivates:
Definition 3.3.4 An ordering of A is a subset P of A that satisfies properties (RO1)
to (RO3) above. The terms cone and prime cone are also used.
116 3 The Real Spectrum
which is continuous since (Sper ϕ)−1 H̊A (f ) = H̊B ϕ(f ) . Thus Sper, like Spec,
is a functor from commutative rings to topological spaces. (Moreover, supp is a
morphism Sper → Spec of functors since suppA ◦(Sper ϕ) = (Spec ϕ) ◦ suppB .)
The map ϕ also yields embeddings of real closed fields as follows: let β ∈ Sper B
and α = ϕ ∗ (β) ∈ Sper A, then ϕ induces an order preserving (with respect to α, β)
3.3 Definition of the Real Spectrum 117
embedding ϕsupp β : κ(supp α) → κ(supp β), and hence ϕβ : k(α) → k(β) (cf.
Sect. 1.11). Letting p := supp α and q := supp β we thus obtain the commutative
diagram
Continuing the analogy with the Zariski spectrum we can think of ϕ as pulling back
the “functions” g ∈ B that are defined on Sper B via Sper ϕ (and via the embeddings
ϕβ ).
Proposition 3.3.10 Consider a multiplicative subset S ⊆ A and its associated
canonical homomorphism iS : A → S −1 A. Then Sper(iS ) is a homeomorphism
from Sper S −1 A to the subspace {α ∈ Sper A : S ∩ supp α = ∅} = s∈S H̊A (s 2 )
of Sper A. If β ∈ Sper S −1 A and α = (Sper iS )(β), then (iS )β : k(α) → k(β) is an
isomorphism.
Finally, (iS )β is an isomorphism since (iS )supp β is an isomorphism (cf. Sect. 3.1)
and order preserving.
induced by ρp : A → κ(p).
Example 3.3.13 Let R be a real closed field and A = R[[t]] the ring of formal
power series over R. The only prime ideals of A are (0) and mA = (t), and κ(A) =
118 3 The Real Spectrum
A/mA ∼ = R. By Sect. 2.9, Sper A has precisely three elements, namely the orderings
P0 , P− , P+ of A, where supp P0 = mA and supp P− = supp P+ = (0). From
P− ∪ P+ ⊆ P0 it follows that P0 is closed and has P− and P+ as generalizations.
The situation can be illustrated by means of the following symbolic diagram, where
the arrows represent specialization:
Example 3.3.14 Assume again that R is a real closed field and let A = R[t]. The
real reduced prime ideals of R[t] are (0) and all (t − c), where c ∈ R. The orderings
of R(t) were determined in Sect. 2.9. The following list therefore gives a complete
description of all points of Sper R[t]:
1. for every c ∈ R, a closed point Pc and also two generalizations Pc,− and Pc,+ of
Pc (simply written as c, c− , c+ );
2. the closed points P−∞ , P+∞ with support (0) (simply written as −∞, +∞);
3. (only if R = R) for every free Dedekind cut ξ of R, a closed point Pξ with
support (0) (simply written as ξ ).
As in the previous example, we illustrate the situation with a symbolic diagram:
The diagram shows that Sper R[t] possesses a natural linear order relation “≤” that
extends the ordering of R.
An open basis of the topology of Sper R[t] is given by all open intervals
and all [−∞, b[, ]a, ∞] where a, b ∈ R, defined similarly. Note that ]a, b[ contains
the points a+ and b− , but not the points a− , a, b, b+ , and similarly for the other
intervals.
In this concrete example it is not difficult to see that Sper R[t] is quasi-compact
and that the subspace topology induced on R ⊂ Sper R[t] is precisely the strong
topology. In fact, this observation is true in general, as we will show now.
√
Let R be a real closed field, C = R( −1) and A an affine R-algebra with
associated variety V := VA = HomR (A, C) and set of real points V (R) := VA (R).
3.3 Definition of the Real Spectrum 119
Remark 3.3.16 The real points of V , equipped with the strong topology, therefore
form a subspace of Sper A. In other words, we do not loose any information about
V (R) after the transition to Sper A. Furthermore, Sper A has a great advantage
compared to V (R) as V (R) has bad topological properties when R = R (totally
disconnected, almost never (locally) compact), whereas Sper A is quasi-compact
and only has finitely many connected components (cf. [11, §7.5]), which facilitates
the definition of a non-trivial concept of connectedness even when R = R. (A
different approach that does not use the real spectrum can be found in [31].) Thus,
the addition of “ideal” points results in a much more geometric space. The example
of the affine line A = R[t] above (Example 3.3.14) already illustrates nicely how
the additional points in Sper A “close the gaps” (if R = R) and make the space
quasi-compact.
An immediate consequence of the Artin–Lang theorem is that V (R) is dense as
a subspace of Sper A!
(b) A Boolean lattice is a lattice (L, ≤) that satisfies the following properties:
(BL1) Distributivity: x ∧ (y ∨ z) = (x ∧ y) ∨ (x ∧ z) for all x, y, z ∈ L;
(BL2) Complement: for every x ∈ L, there exists x ∈ L such that x ∧ x = 0
and x ∨ x = 1.
x ∨ (y ∧ z) = (x ∨ y) ∧ (x ∨ z),
x + y := (x ∧ y ) ∨ (x ∧ y) and xy := x ∧ y
3.4 Constructible Subsets and Spectral Spaces 121
turns it into a Boolean algebra, and conversely every Boolean algebra becomes
a Boolean lattice via
x ∧ y := xy , x ∨ y := x + y + xy , x := 1 + x.
Definition 3.4.3
(a) K(Sper A) denotes the Boolean sublattice of the power set 2Sper A generated by
the Harrison subbasis H̊A of Sper A. The elements of K(Sper A) are called the
constructible subsets of Sper A. We let
Remarks 3.4.4
(1) K(Sper A) is the smallest subset K of 2Sper A such that H̊A ⊆ K and which
is stable under taking finite intersections and complements. The elements of
K(Sper A) are precisely the finite unions of sets of the form
Proof Let us prove the second statement first. Consider the constructible topology
on Sper A. Then all Y ∈ K(Sper A) are clopen by definition. Conversely, let Y ⊆
Sper A be clopen. Since Sper A is compact (see below), Y is the union of finitely
many constructible sets, and thus constructible
itself.
To prove compactness, consider Z = f ∈A {0, 1}, endowed with the product
topology. Then Z is totally disconnected and also compact by Tykhonov’s theorem.
We identify an element (εf )f ∈A with the subset {f ∈ A : εf = 1} of A, in
other words we identify Z with the power set of A. It follows that j (α) :=
{f ∈ A : f (α) ≥ 0} = Pα defines an injection j : Sper A → Z, and the
topology induced by Z (via j ) on Sper A is precisely the constructible topology, cf.
Remark 3.4.4(1). It remains to show that j (Sper A) is closed in Z, or equivalently
that Z \ j (Sper A) is open. To this end, let S ⊆ A be a subset such that S ∈
j (Sper A), i.e., such that S is not and ordering of A. Then one of the following
axioms (cf. Sect. 3.3) is violated by S:
(1) S + S ⊆ S,
(2) SS ⊆ S,
(3) S ∪ (−S) = A,
(4) −1 ∈ S,
(5) a ∈ S, b ∈ S ⇒ −ab ∈ S (a, b ∈ A).
But then this particular axiom is also violated in a neighbourhood of S in Z! For
example, if S violates axiom (5), then there exist a, b ∈ A \ S such that −ab ∈ S.
Then U := {T ⊆ A : a ∈ T , b ∈ T , −ab ∈ T } is a neighbourhood of S in Z, and
each T ∈ U violates axiom (5) for the same reason that S violates axiom (5).
Since the Harrison topology is coarser than the constructible topology, we obtain:
Corollary 3.4.6 The Harrison topology on Sper A is quasi-compact.
Let ϕ : A → B be a ring homomorphism. Then Sper ϕ : Sper B → Sper A is
continuous in the constructible topology since (Sper ϕ)−1 H̊A (f ) = H̊B ϕ(f )
for all f ∈ A and taking inverse images commutes with the Boolean operations.
Nevertheless, the image of a constructible set is in general not constructible anymore
(consider for example the map R[t] → R(t) and the set Sper R(t)). Consequently
it is often necessary to consider more general subspaces.
Definition 3.4.7 A subset Y of Sper A is called proconstructible if Y is the
intersection of (arbitrarily many) constructible subsets of Sper A or, equivalently,
if Y is closed in the constructible topology.
Remarks 3.4.8
(1) The equivalence of the conditions in the definition follows from the fact that the
complement of a subset Y which is closed in the constructible topology can be
written as the union of constructible subsets of Sper A.
(2) Some examples of proconstructible subspaces of Sper A are given by
Sper S −1 A and Sper A/a. If S, resp. a, is finitely generated as semigroup,
3.4 Constructible Subsets and Spectral Spaces 123
Remark 3.4.14 If A is a noetherian ring, for instance an affine algebra over a field,
then X = Spec A has only finitely many irreducible components (i.e., A has only
finitely many minimal prime ideals). These play an important role in algebraic
geometry. In contrast, the spaces X = Sper A almost always have infinitely many
minimal points, also in the “geometric setting”, which is why the concept of
irreducible components is less useful in this context.
We take a temporary leave of real algebra and embark on an excursion into point-
set topology. Specifically we will have a closer look at spectral spaces. These satisfy
all the topological properties that we already observed for the Zariski spectrum and
the real spectrum. Spectral spaces elucidate the boundary between real algebraic
and purely topological arguments. Moreover, in one swoop we obtain the properties
of many new spaces that occur naturally in the study of spectra (for example,
proconstructible subspaces are spectral). The main examples of spectral spaces are
the real spectra and especially the Zariski spectra. In fact, Hochster has shown that
these are actually all the examples, cf. [51].
Definition 3.4.15
(a) A spectral space is a topological space X that satisfies the following
properties:
(1) X is a quasi-compact T0 space;
(2) the intersection of any two open quasi-compact subsets of X is again quasi-
compact;
(3) the open quasi-compact sets form a basis of the topology on X;
(4) every nonempty closed irreducible subspace of X has a (unique) generic
point.
(b) Let X and Y be spectral spaces. A spectral map is a map f : X → Y such that
for all V ⊆ Y , if V is open and quasi-compact, then so is f −1 (V ). (In particular,
f is continuous.)
Examples 3.4.17
(1) The spectral maps are precisely those continuous maps X → Y for which
Xcon → Ycon is also continuous.
(2) Sper A is a spectral space, and the current definition of constructible subsets is
the same as the original one. If ϕ is a ring homomorphism, the map Sper ϕ is
spectral. In a similar fashion, Spec A is a spectral space and Spec ϕ is a spectral
map.
(3) Every Boolean space (i.e., a compact and totally disconnected topological
space) is spectral. Its constructible sets are precisely the clopen sets, which
means that in this case the spectral and constructible topologies coincide. The
Boolean spaces are precisely the spectral spaces that are Hausdorff.
Proposition 3.4.18 Consider a spectral space X. Then Xcon is compact and totally
disconnected. The identity map Xcon → X is spectral, and (Xcon )con = Xcon .
Proof The only nontrivial statement that needs a proof is the quasi-compactness
of Xcon , i.e., the finite intersection property for K(X). We start by showing the
following
Claim: If Y ={Yi : i ∈ I } is a family of either open
quasi-compact or closed sets in
X such that Yj = ∅ for all finite J ⊆ I , then Yi = ∅.
j ∈J i∈I
Proof of the claim: Without loss of generality we may assume that Y is maximal
with respect to the indicated property (by Zorn’s Lemma). Let Z be the intersection
of the closed sets in Y. Then Z = ∅ (since X is quasi-compact), hence also Z ∈ Y.
Assume for the sake of contradiction that Z is reducible. Then Z = Z1 ∪Z2 for some
closed Zi Z. Since Zi ∈ Y, there exist open U1 , U2 ∈ Y such that Ui ∩ Zi = ∅
for i = 1, 2. In particular, (U1 ∩ U2 ) ∩ Z = ∅. But U1 ∩ U2 ∈ Y since Y is maximal,
a contradiction.Hence Z is irreducible. It is now clear that the generic point of Z is
an element of Yi . This proves the claim.
i
The remainder of the proof is a modification of an argument provided by
Marcus Tressl. Consider the set Ultra K(X) of all ultrafilters in the lattice K(X),
cf. Definition 3.5.10. For x ∈ X, let
Fx := {U ∈ K(X) : x ∈ U }
Transferring the Zariski topology from Spec K(X) to the set X via the bijection (3.2)
yields the constructible topology on X, which can be seen by direct verification.
Hence the map (3.2) is a homeomorphism Xcon ≈ Spec K(X) of topological spaces.
Since every Zariski spectrum is quasi-compact (cf. Proposition 3.1.15) we conclude
that Xcon is also quasi-compact.
Corollary 3.4.19 The elements of K̊(X) are precisely the open quasi-compact
subsets of X.
The “geometric setting” refers to the case of an affine variety (or rather an affine
algebra) over a real closed field. In this context there exists quite a concrete
geometric interpretation of Sper A and its constructible subspaces via semialgebraic
subsets of VA (R). √
We assume throughout this section that R is a real closed field, C = R( −1),
A is an affine R-algebra, V = VA = HomR (A, C) is its associated variety and
V (R) = VA (R) = HomR (A, R) is the set of real points of the variety. We view
V (R), endowed with the strong topology, as a subspace of Sper A, as explained in
Sect. 3.3. Given any subset Y of Sper A we may then consider the subset Y ∩ V (R)
of V (R).
Definition 3.5.1
(1) We let
S V (R) := V (R) ∩ K(Sper A) = Y ∩ V (R) : Y ∈ K(Sper A)
and call the sets M ∈ S V (R) the semialgebraic subsets of V (R). More
generally, for M ∈ S V (R) we denote the set of semialgebraic subsets of
M by
S(M) := M ∩ K(Sper A) = M ∩ S V (R) = N ∈ S V (R) : N ⊆ M .
Remark 3.5.2 The semialgebraic subsets of V (R) are thus precisely those subsets
that can be described by finitely many (equalities and) inequalities. In other words,
they are the finite unions of sets of the form
Definition 3.5.6 We use the tilde to denote the map S V (R) → K(Sper A)
which is the inverse of Y → Y ∩V (R). Thus, for a semialgebraic subset M of V (R),
denotes the (uniquely determined) constructible subset of Sper A that satisfies
M
3.5 The Geometric Setting: Semialgebraic Sets and Filter Theorems 129
M = M ∩ V (R). In fact, M
is the closure of M in Sper A with respect to the
constructible topology.
In this context the following fact is very important but beyond the scope of this
book:
Theorem 3.5.7 Let M and N be semialgebraic subsets of V (R) such that N ⊆ M.
is also open in M
If N is open in M, then N (the reverse is trivially true). Hence the
following equalities also hold:
where r ∈ N and f1 , . . . , fr ∈ A.
Example 3.5.9 Let A = R[t], hence V (R) = R. We refer to Sect. 3.3 for an explicit
description of Sper R[t] and use the natural total order relation ≤ on Sper R[t]. The
images of the intervals M ⊆ R under the map : S V (R) → K Sper R[t] are
as follows:
A ∪ B ∈ F ⇒ A ∈ F or B ∈ F.
Remarks 3.5.11
(1) If E is a subset of L such that A∈E A = ∅ for all finite E ⊆ E, then E is
contained in a filter of L. The smallest such filter is
!
B ∈ L : there exists a finite subset E ⊆ E such that A⊆B .
A∈E
Lemma 3.5.14 Let X be a set and L a well-behaved sublattice of 2X . Then the map
!
Filt(L) → (pro-L) \ {∅}, F → A
A∈F
3.5 The Geometric Setting: Semialgebraic Sets and Filter Theorems 131
Y → FY := {B ∈ L : Y ⊆ B}.
A, B ∈ L, Y ⊆ A ∪ B ⇒ Y ⊆ A or Y ⊆ B;
These are thus the sets of maximal, resp. minimal points with respect to the
specialization relation. We observe that Xmax is precisely the subset of closed points
of X, and that Xmin = (X∗ )max and Xmax = (X∗ )min , where X∗ is the inverse
spectral space of X, cf. Sect. 3.4.
Now we are ready to describe X in terms of filters of lattices of constructible
subsets.
Proposition 3.5.17 Let X be a spectral space.
(a) There exists a canonical inclusion reversing bijection
Sper A → Ultra S V (R) , x → Fx = {M : x ∈ M}
with inverse F → M∈F M.
(b) There exists a bijection from Filt S V (R) to the set of nonempty procon-
structible subsets of Sper A, namely F → M∈F M.
Let A be an affine algebra over a real closed field R with associated variety
V = VA . We illustrate the correspondence between the real spectrum and ultrafilters
in several examples below. For x ∈ Sper A, we consider the corresponding ultrafilter
Fx = {M ∈ S V (R) : x ∈ M}.
= {F : M ∈ F }
M
Example 3.5.21 Let A = R[t], hence V (R) = R (cf. Example 3.3.14). The
correspondence Sper A → Ultra S(R) is as follows:
where the intersection is over all closed subvarieties W of V such that W (R) ∈ Fx .
then x ∈ M
If M ⊆ V (R) is semialgebraic and x ∈ M, max if and only if for every
N ∈ S(M) with N ∈ Fx there exists an N ∈ Fx which is closed in M and such
that N ⊆ N.
134 3 The Real Spectrum
ϕ0 : A → R, f →
f (0, 0)
ϕ1 : A → R(x), f → f (x, 0)
ϕ2 : A → R(x, y), f → f.
Assume that the ordering of R(x, y) is such that x compared to R and y compared
to R(x) are infinitely small and positive, and assume that R(x) is endowed with the
induced ordering. Let αi ∈ Sper A be the point defined by ϕi (and the indicated
orderings) for i = 0, 1, 2 (cf. Remark 3.3.2(3)). We have
⇔ α0 = (0, 0) ∈ M;
α0 ∈ M
⇔ ∃ε > 0 such that ]0, ε[ × {0} ⊆ M;
α1 ∈ M
⇔ ∃ε > 0, ∃g ∈ R[t] with g(t) > 0 for t ∈ ]0, ε[ ,
α2 ∈ M
such that {(a, b) ∈ R 2 : 0 < a < ε, 0 < b < g(a)} ⊆ M.
(The verification only requires some effort in the case of α2 . Since α2 ∈ (Sper A)min ,
we have: α2 ∈ M ⇔ there exist f1 , . . . , fn ∈ A such that α2 ∈ H̊A (f1 , . . . , fn ) and
i {x ∈ R : fi (x) > 0} ⊆ M. Then one needs to consider that for every f ∈ A
2
with f (α2 ) > 0, the set {x ∈ R 2 : f (x) > 0} contains a set of the form described
above.)
We can picture α1 as the bud of the positive branch of the x-axis at the origin,
and α2 as the tip of the upper shoreline of the half-plane to the right of the x-axis.
Thus we see that α0 = (0, 0) has a chain of generalizations of (maximal possible)
length 2. It is a straightforward exercise (cf. Example 3.7.7) to show that every
point x ∈ R n has a chain of generalizations of length n in R #n . (In more generality
the analogous statement holds for smooth R-varieties.)
3.6 The Space of Closed Points 135
Example 3.5.25 Assume that A is an integral domain and let K = Quot A. The
orderings of K can be visualized in V (R) as follows.
We call a semialgebraic subset M ⊆ V (R) thin if M is not Zariski dense in
Spec A. (If we define the dimension of a semialgebraic subset as the dimension
of its Zariski closure in Spec A, then M is thin if and only if dim M < dim A =
tr.deg.(K/R) = dim V .)
Let x ∈ Sper A. Then supp(x) = 0 (i.e., x ∈ Sper K ⊆ Sper A) if and only if Fx
contains no thin sets. Expressed in a different way: there is an equivalence relation
on S V (R) defined by
M ∼ N :⇔ M ) N := (M \ N) ∪ (N \ M) is thin,
and the quotient set S V (R) / ∼ carries a lattice structure induced by S V (R) .
The ultrafilters of S V (R) / ∼ then correspond precisely to those ultrafilters of
S V (R) that do not contain any thin sets. This proves:
Remark 3.5.27 We have shown that for every spectral space X there exists a natural
bijection X → Prime K̊(X). In this way it is possible to “recover” the set X
from the lattice K̊(X). In fact this relationship goes much further. Indeed, since
we can determine the topology of X from K̊(X), (the topological space) X and (the
distributive lattice) K̊(X) are essentially the same. We give a short sketch (without
proof) of this duality.
If L is a distributive lattice (with 0 and 1), then the set Prime L carries a natural
topology in which the sets Ux := {F ∈ Prime L : x ∈ F } (x ∈ L) form an open
basis. This topological space is called the Stone space St(L) of L. It is a spectral
space and L is (via x → Ux ) canonically isomorphic to K̊ St(L) . Conversely, if
X is a spectral space, then X is canonically homeomorphic to St K̊(X) (we did
describe the bijection). The following holds (see for instance [43, p. 103] or [56, p.
65 ff.]):
This section deals with properties of real spectra that are quite false for general
spectral spaces. As usual, A denotes an arbitrary commutative unitary ring.
136 3 The Real Spectrum
Remarks 3.6.2
(1) If X is spectral, the above statement already follows from Proposition 3.5.17.
After passing to the inverse spectral space X∗ , we see that X = {x} (but
Xmin is in general not quasi-compact!). x∈X min
(2) By Corollary 3.4.21(c) the subspace Xmax of a spectral space X is procon-
structible if and only if it is closed. Usually this is not the case however. For
example, if R is a real closed field and A an affine R-algebra with associated
variety V , then we know from Sect. 3.5 that V (R) is already dense in Sper A,
hence a fortiori in (Sper A)max .
(3) Let R be a real closed field and X = Sper R[t]. Then Xmax consists of R ∪
{−∞, +∞} together with all free Dedekind cuts of R. In the special case that
R = R, the space Xmax = R ∪ {±∞} is the natural compactification (with two
end points) of R.
We have reached the point mentioned at the start of the chapter, where the real
spectra of rings exhibit special features that general spectral spaces do not. Despite
its simplicity, the consequences of the next lemma are far reaching!
Lemma 3.6.3 Let P and Q be orderings of A. The following statements are
equivalent:
(i) P ⊆ Q and Q ⊆ P (i.e., P and Q are incomparable with respect to
specialization);
(ii) there exists f ∈ A such that P ∈ H̊A (f ) and Q ∈ H̊A (−f );
(iii) there exist open U, V ⊆ Sper A such that U ∩ V = ∅ and P ∈ U, Q ∈ V .
Proof (i) ⇒ (ii): Choose a ∈ P \ Q and b ∈ Q \ P . Then f := a − b yields
the desired property. Indeed, arguing by contradiction, if −f ∈ P , then also b =
a − f ∈ P , and if f ∈ Q, then also a = b + f ∈ Q. Either conclusion is false.
The implications (ii) ⇒ (iii) ⇒ (i) are trivial.
Remark 3.6.5 Theorem 3.6.4 is in general completely false for the Zariski spec-
trum. For example, if A is an affine C-algebra with associated variety V =
V (C), then (Spec A)max = V (C) carries the (induced) Zariski topology, which is
profoundly non-Hausdorff whenever dim A > 0.
Remark 3.6.6 Let R be a real closed field and A an affine R-algebra with associated
variety V = VA . It follows from the density of V (R) in (Sper A)max that
(Sper A)max is a natural compactification of V (R).
If x1 , . . . , xn is a system of generators of A and α ∈ (Sper A)max , then k(α)
is archimedean over R if and only if there exists c ∈ R such that x1 (α)2 + · · · +
xn (α)2 < c, as we will see later. (The condition is clearly equivalent to A/ supp(α)
being archimedean over R; we will see in Sect. 3.7 that k(α) is archimedean over
A/ supp(α) for all α ∈ (Sper A)max .)
We call those α ∈ (Sper A)max for which k(α)/R is archimedean the finite points
and all the remaining α ∈ (Sper A)max the infinitely distant points of (Sper A)max .
The set of finitepoints thus forms an open dense subspace of (Sper A)max (equal to
(Sper A)max ∩ c∈R H̊A (c − x12 − · · · − xn2 )). This is interesting in the particular
case R = R, since the set of finite points is already equal to V (R) in this case (cf.
Sect. 2.1). Thus we see that V (R) is contained as an open dense subspace in its
compactification (Sper A)max . This fact has been used by G.W. Brumfiel, who gave
an interpretation of the Thurston compactification of Teichmüller spaces by means
of the real spectrum, cf. [20].
A further and equally fundamental consequence is
Theorem 3.6.7 Let x ∈ Sper A. Then {x} forms a chain with respect to specializa-
tion. In other words, for all y, z ∈ Sper A we have
x % y and x % z ⇒ y % z or z % y.
Remark 3.6.9 For every x ∈ Sper A we can visualize {x} as a “spear”, with the
closed specialization of x as its tip. The notation Sper for real spectrum alludes to
this picture and also to the French term “spectre réel”.
Remark 3.6.11 As mentioned in Sect. 3.5 we will show in Sect. 3.7 that every point
x0 ∈ R n is the tip of a spear
also ρ(Y1 ) ∩ ρ(Y2 ) = ∅ and ρ(Y1 ), ρ(Y2 ) are closed in Xmax by Proposition 3.6.16,
from which the assertion follows.
We recognize now that the spaces Xmax (with X ⊆ Sper A proconstructible)
are topologically pleasing, not only because they are compact, but also because they
retain some information about the space X. For example, one can use ρ to reduce the
cohomology of X to the cohomology of Xmax ; both have isomorphic cohomology,
cf. [23]. In the geometric setting it seems important to treat M max as the natural
compactification of M ⊆ V (R). In spite of everything however, Sper A should be
considered as the central object, and (Sper A)max rather as a useful auxiliary space.
f + g ∈ a ⇒ f, g ∈ a.
Lemma 3.7.2 The map a → π(a) defines a bijection from the set of P -convex
ideals of A to the set of ideals of A/p that are convex with respect to P . The inverse
is given by a → π −1 (a).
Proof We endow A/p with the total order relation induced by P . Since we have
f ∈ P and −f ∈ P for every f ∈ p, it follows that p is contained in every P -
convex ideal of A. This shows that the two described maps are inverse to each other.
If a ⊆ A is a P -convex ideal, then π(a) is convex in A/p. Indeed, if a ∈ a and
f ∈ A such that 0 ≤ π(f ) ≤ π(a), then f ∈ P and a − f ∈ P , hence f ∈ a by
assumption. Conversely, if a ⊆ A/p is a convex ideal and if f, g ∈ P are such that
f + g ∈ a := π −1 (a), then 0 ≤ π(f ) ≤ π(f + g) ∈ a, and thus also π(f ) ∈ a.
Corollary 3.7.3
(a) The P -convex ideals form a chain, and p = supp(P ) is the smallest ideal in it.
(b) For every P -convex ideal a of A we have P + a = P ∪ a.
Proof (a) See Remark 2.1.4(4).
(b) Let p ∈ P and a ∈ a. If p+a ∈ P , then −(p+a) ∈ P , and −a = p−(p+a) ∈
a implies p + a ∈ a.
3.7 Specializations and Convex Ideals 141
Definition 3.7.6 Let B be a valuation ring and A a subring of B. The prime ideal
A ∩ mB of A is called the centre of B on A. Similarly, given a place λ : K → L ∪ ∞
and a subring A of K on which λ is finite, we call the kernel of the homomorphism
λ|A : A → L the centre of λ on A.
Sper L, λ∗ (z) has a chain of generalizations with the pi as supports, i.e., there exist
y0 , . . . , yn ∈ Sper A with supp(yi ) = pi such that y0 % y1 % · · · % yn = λ∗ (z).
This shows by way of example that for a real closed field R every point a ∈ R n
has chains of generalizations of length n in Sper R[t1 , . . . , tn ]: namely, in Sect. 2.10
we constructed a place λ = λn ◦ · · · ◦ λ1 : R(t1 , . . . , tn ) → R ∪ ∞ over R for
which λi ◦ · · · ◦ λ1 has centre (tn − an , . . . , tn−i+1 − an−i+1 ) on R[t1 , . . . , tn ] for
i = 0, . . . , n.
Definition 3.7.8 A valuation ring B is called real closed if Quot B is a real closed
field and B is residually real. (The second condition is equivalent to: B is convex in
Quot B, cf. Sect. 2.5.)
is surjective.
√
Proof If p is a P0 -convex prime ideal √
of A, then pc is a prime ideal of B by
√
Proposition 2.4.17(e), and we have A ∩ pc = p = p.
3.7 Specializations and Convex Ideals 143
Then r ∗ (Sper C) = {β} and ψ ∗ (Sper C) = {α} by Proposition 3.7.12, from which
ϕ ∗ {β} = {α} follows.
We want to make the statements of Lemma 3.7.10 and Proposition 3.7.12 a bit
more precise. Lemma 3.7.10 can also be formulated as follows: for every P0 -convex
prime ideal p of A there exists a convex overring C of A with centre p on A. More
accurately:
Proposition 3.7.15 Let R be a real closed field, A ⊆ R a subring and P = A ∩
[0, ∞[R . If p ⊆√A is a P -convex prime ideal, then C := (Ap )c is the smallest and
Cc , where c := pc , the largest convex overring of A in R with centre p on A.
Proof Let B ⊆ R be a convex overring of A with centre p on A. Then Ap ⊆ B, thus
also C = (Ap )c ⊆ B, and B = Cq for some √ prime ideal q of C. Since p ⊆ mB = q
it must also be the case that pc ⊆ q and c = pc ⊆ q, i.e., B ⊆ Cc (note that c is a
prime ideal of C, cf. Lemma 3.7.10).
{γ ∈ Sper A : α % γ % β}.
Proof Let q := supp β, A := rα (A) and q := rα (q). It suffices to take for B the
convex hull of Aq in k(α) (and for ϕ the map induced by rα : A → k(α)).
Corollary 3.7.18 Let α ∈ Sper A. Then α is closed in Sper A if and only if k(α) is
archimedean over rα (A).
Proof We may assume without loss of generality that supp α = (0), i.e., that
R := k(α) is the real closure of K := Quot A. Since R is archimedean over K
(cf. Proposition 1.7.1), we have C = R for p = (0) in Proposition 3.7.15, i.e., every
convex overring B of A which is distinct from R induces a proper specialization of
α. Hence, {α} = {α} if and only if there are no such overrings.
3.8 The Real Spectrum and the Reduced Witt Ring of a Field 145
Since convex valuation rings of a real closed field R and real places R → S ∪ ∞
are essentially the same thing, the previous statements can also be formulated in the
language of places. Let us give an example.
Let R and S be real closed fields, ϕ : A → R a homomorphism and λ : R →
S ∪ ∞ a place which is finite on ϕ(A). Then the element αλ◦ϕ is a specialization
of αϕ in Sper A, and for every specialization of αϕ there is such a place λ. More
precisely:
Definition 3.7.19 A homomorphism ϕ : A → R into a real closed field is called
tight if R is archimedean over Quot ϕ(A).
3.8 The Real Spectrum and the Reduced Witt Ring of a Field
In the previous section we made a detailed study of the specialization chains (the
“spears”) in the real spectrum of a ring. Visually speaking these are transversal to
the fibres of the support map supp : Sper A → Spec A since every spear has at most
one point in common with such a fibre. The fibres are precisely the real spectra
of the residue fields of A (cf. Corollary 3.3.12). Thus, in order to gain a better
understanding of the spectral space Sper A, a deeper study of the real spectrum of
fields suggests itself. In this section we will initiate this study.
Throughout, F denotes a field, without loss of generality assumed to be a real.
For the sake of brevity we write XF instead of Sper F . The spectral space XF is
Hausdorff, thus compact and totally disconnected, its (Harrison) topology coincides
with the constructible topology, and the constructible subsets of XF are precisely the
clopen subsets (cf. Remark 3.3.2(4) and Sect. 3.4). If a1 , . . . , an are elements of F ∗ ,
we simply write HF (a1 , . . . , an ) or H (a1, . . . , an ) instead of H̊F (a1 , . . . , an ) =
H F (a1 , . . . , an ).
146 3 The Real Spectrum
(the total signature of F ) is a ring homomorphism with kernel equal to the nilradical
of W (F ).
Proof Since all signP are homomorphisms, the map ϕ → sign ϕ is a homomor-
phism from W (F ) to the ring of all maps XF → Z. Given ϕ = a with a ∈ F ∗ , the
continuity of sign ϕ is clear. It follows that sign ϕ is continuous for all ϕ ∈ W (F ).
The kernel of sign is the intersection of the kernels of all the signP , and these are
precisely the minimal prime ideals of W (F ) (cf. Sect. 1.4). The statement follows.
W (F ) := W (F )red = W (F )/ Nil W (F )
the reduced Witt ring of F and denote the image of the fundamental ideal I (F ) in
W (F ) by I (F ) = I (F )/ Nil W (F ).
Given ϕ ∈ W (F ), we denote its image in W (F ) by ϕ. By Proposition 3.8.2 we
may identify W (F ) canonically with a subring of C(XF , Z), and usually this is what
we will do. Under this identification we have ϕ = sign ϕ.
We denote the characteristic function of a subset Y ⊆ XF by χY . Given a 2-
dimensional form ϕ = 1, a with a ∈ F ∗ , we then have ϕ = 2 · χH (a) . Since I (F )
3.8 The Real Spectrum and the Reduced Witt Ring of a Field 147
Proposition 3.8.6 The abelian group C(XF , Z)/W (F ), i.e., the cokernel of the
total signature of F , is a 2-primary torsion group.
Proof Every element of C(XF , Z) can be written in the form m1 χY1 + · · · + mr χYr
with r ≥ 1, mi ∈ Z and Yi ⊆ XF constructible. It follows from Lemma 3.8.5 that
C(XF , Z) is generated as an additive group by the f = χH (a1 ,...,an ) with n ≥ 1 and
ai ∈ F ∗ . For such an f we have 2n f ∈ W (F ). The statement follows.
Definition 3.8.9 A real field F is called a SAP field if, given any two disjoint closed
subsets Y and Y of XF , there exists an a ∈ F ∗ such that a|Y > 0 and a|Y < 0.
(SAP stands for “strong approximation property”, which is unrelated to the number
theoretic concept of strong approximation.)
Theorem 3.8.12 (Special case of [9, Satz 15]) Let F be a real field and n ∈ N
such that n ≥ st(F ). For every (n + 1)-fold Pfister form ϕ over F there exists an
(n-fold) Pfister form ψ over F such that ϕ = 2ψ.
It is not difficult to deduce new characterizations of st(F ) from Theorem 3.8.12.
Since we do not prove the theorem, nor have any use for its corollary below in the
remainder of this book, we leave the proof of the corollary as an exercise to the
reader.
Corollary 3.8.13 Let F be a real field and s ∈ N. The following statements are
equivalent:
(i) st(F ) ≤ s;
(ii) every set H (a1, . . . , aN ) with ai ∈ F ∗ and N ≥ 1 can be written in the form
H (b1, . . . , bs ) with bj ∈ F ∗ ;
(iii) for every (s + 1)-fold Pfister form ϕ over F there exists an (s-fold) Pfister form
ψ over F such that ϕ = 2ψ;
s+1 s
(iv) I (F ) = 2 I (F ).
3.8 The Real Spectrum and the Reduced Witt Ring of a Field 149
The term stability index for st(F ) comes from characterization (iv): st(F ) is the
n
smallest index from which point onwards the sequence I (F ) : n = 1, 2, . . . is
stable in the sense of (iv). (This is a K-theoretic notion of stability, cf. [83].)
In the final part of this section we turn our attention to finitely generated field
extensions. If F ⊆ K is a field extension we write rK/F for the restriction map from
XK := Sper K to XF = Sper F . We consider finite extensions first:
Proposition 3.8.14 Let K ⊇ F be a field extension of degree n < ∞. Then:
(a) r := rK/F : XK → XF is an open (and closed) map, i.e., the images of
constructible sets are constructible.
(b) For each x ∈ XF , the fibre r −1 (x) is finite and its cardinality t satisfies t ≤ n
and t ≡ n mod 2.
(c) Let Ut := {x ∈ XF : #r −1 (x) = t} for t ≥ 0. Then every Ut is constructible
and r is topologically trivial over every Ut (i.e., there exists a decomposition
of r −1 (Ut ) into t disjoint open subsets W1 , . . . , Wt such that r|Wi is a
homeomorphism from Wi to Ut for every i = 1, . . . , t).
(d) The map f : XF → Z, x → #r −1 (x) is an element of W (F ).
Proof Consider the trace form ϕ := tr∗ 1K associated to the extension K/F .
Then f = ϕ by Corollary 1.12.10, which proves (d) as well as the constructibility
of the Ut . Furthermore, (b) follows immediately since dim ϕ = n.
Let us prove (a) next: r is closed since it is a continuous map between compact
spaces. In addition, r(XK ) = U1 ∪· · ·∪Un is constructible in XF . Let a1 , . . . , am ∈
√ √
K ∗ . Then HK (a1 , . . . , am ) = rL/K (XL ) for L := K( a1 , . . . , am ) (cf.
Proposition 1.3.3). Consequently, rK/F HK (a1 , . . . , am ) = rK/F rL/K (XL ) =
rL/F (XL ) is also constructible in XF . Since these HK (a1 , . . . , am ) form a basis of
the topology of XK , (a) follows.
It remains to prove the second statement of (c). Let t ≥ 0 and fix an x ∈ Ut . If
r −1 (x) = {y1 , . . . , yt }, there exist pairwise disjoint constructible neighbourhoods
Wi of the yi in r −1 (Ut ) for i = 1, . . . , t. Then V (x) := r(W1 ) ∩ · · · ∩ r(Wt ) is a
constructible neighbourhood of x in Ut and the restriction of r to every Wi (x) :=
Wi ∩ r −1 V (x) is a homeomorphism to V (x). Then we cover Ut with finitely
many V (xi ) (i = 1, . . . , N) and make this covering disjoint by shrinking the V (xi )
if necessary. For every 1 ≤ i ≤ N we have r −1 V (xi ) = tj =1 Wj (xi ) (disjoint
union), and r|Wj (xi ) is a homeomorphism from Wj (xi ) to V (xi ). Finally, letting
Wj := N i=1 Wj (xi ) (j = 1, . . . , t), the sets W1 , . . . , Wt yield the desired result.
With help from the sign change criterion (Theorem 2.12.8), we can generalize
statement (a) from Proposition 3.8.14 to finitely generated field extensions:
Proposition 3.8.15 (R. Elman, T.Y. Lam, A. Wadsworth [38]) Let K ⊇ F be a
finitely generated field extension. Then rK/F : XK → XF is an open (and closed)
map.
150 3 The Real Spectrum
Proof The argument used in the proof of Proposition 3.8.14(a) shows that it suffices
to prove that rK/F (XK ) is open in XF . Let α1 , . . . , αd be a transcendence basis of
K over F , and let β ∈ K be such that K = F (α1 , . . . , αd , β) (Primitive Element
Theorem). If p ∈ F [t1 , . . . , td , td+1 ] = F [t] is an irreducible polynomial such that
p(α1 , . . . , αd , β) = 0, then the field of fractions of F [t]/(p) is isomorphic over F
to K (cf. Proposition 2.12.9).
Consider x ∈ rK/F (XK ) and let k(x) be the associated real closure of K.
By the sign change criterion (Theorem 2.12.8) there exist a, b ∈ k(x)d+1 such
that p(a) < 0 < p(b). Let L = F (a, b) be the finite extension of F that is
generated over F by the coordinates of a and b in k(x). Let x be the restriction
of the ordering of k(x) to L. Then x ∈ HL (−p(a), p(b)). By Proposition 3.8.14,
Y := rL/F HL (−p(a), p(b)) is a constructible neighbourhood of x in XF . The
sign change criterion then yields Y ⊆ rK/F (XK ). Indeed, if y ∈ HL (−p(a), p(b))
and y = rL/F (y ) ∈ Y , then k(y) = k(y ) is an overfield of L, and p is indefinite
over k(y) by the choice of y . This implies that y has an extension to K, as required.
Let A be ring. The purpose of this section is to investigate the following question:
let X ⊆ Sper A be an intersection of sets of the form {f > 0}, {f ≥ 0} and {f = 0}
(where f ∈ A), and explicitly represented in this manner. Determine those functions
in A that are positive (resp. are nonnegative, resp. vanish) on X.
Definition 3.9.1 A preordering of A is a subset T ⊆ A that satisfies the following
properties:
(RP1) T + T ⊂ T , T T ⊂ T (i.e., T is a subsemiring of A);
(RP2) a 2 ∈ T for all a ∈ A.
If in addition
(RP3) −1 ∈ T
holds, then T is called a proper preordering of A. Otherwise T is an improper
preordering of A.
3.9 Preorderings of Rings and Positivstellensätze 151
H A (F ) = H A (P [F ])
(indeed, the inclusion ⊇ is trivial and the inclusion ⊆ follows from the
description).
(6) For fields k with char k = 2 (in particular real fields), k 2 is the intersection of
all orderings of k (cf. Sect. 1.1). For rings this is false in general. For example,
A = R[t1 , . . . , tn ] is a real reduced ring and if n ≥ 2, then A2 is not
equal to the intersection of all orderings. Indeed, letting K = Quot A, we
can identify (Sper A)min with Sper K (we did not fully prove this, but compare
with Example 3.7.7) and the intersection of all orderings of A is thus equal to
A ∩ K 2 , the positive semidefinite polynomials. But these are not always sums
of squares in A when n ≥ 2 (cf. Sect. 2.12).
As for fields we nevertheless have:
Proposition 3.9.3 Every proper preordering of A is contained in an ordering.
We first prove:
Lemma 3.9.4 Let T ⊆ A be a proper preordering and a ∈ A. Then:
(a) (aT ) ∩ (1 + T ) = ∅ or (−aT ) ∩ (1 + T ) = ∅;
(b) if (aT ) ∩ (1 + T ) = ∅, then T − aT is also a proper preordering of A.
Proof (a) Assume that as = 1 + s and −at = 1 + t for certain s, s , t, t ∈ T .
Then
−a 2 st = 1 + s + t + s t ,
hence −1 = a 2 st + s + t + s t ∈ T , a contradiction.
(b) The statements −1 ∈ T − aT and aT ∩ (1 + T ) = ∅ are equivalent.
152 3 The Real Spectrum
as = 1 + s and bt = 1 + t
abst = 1 + s + t + s t ,
t t
a· =1+ in B
a 2m a 2n
by Proposition 3.9.5. We may assume without loss of generality that m = n, and it
follows that
at = a 2n + t in B,
i.e.,
a · a 2N t = a 2(N+n) + a 2N t in A
of Sper A, i.e.,
X = α ∈ Sper A : f (α) > 0 ∀f ∈ F, g(α) ≥ 0 ∀g ∈ G, e(α) = 0 ∀e ∈ E .
154 3 The Real Spectrum
t t
a· =1+ 2 in B.
s2 s
3.9 Preorderings of Rings and Positivstellensätze 155
in other words, an identity of the form (ii). The implication (ii) ⇒ (iii) follows using
the same trick as in the proof of Proposition 3.9.5 and the implication (iii) ⇒ (i) is
clear.
Proof of Theorem 3.9.10. For the implication (i) ⇒ (ii) we let a ≥ 0 on X. By
Proposition 3.9.8 there exist s, s ∈ S, t, t ∈ T and n ≥ 0 such that
t t
a· 2
= a 2n + 2 in B.
s s
an identity of the form (ii). The implication (ii) ⇒ (iii) follows as in the proof of
Proposition 3.9.8 and the implication (iii) ⇒ (i) is clear.
Proof of Theorem 3.9.11. For the implication (i) ⇒ (ii) we let a = 0 on X. By
Theorem 3.9.10 there exist s1 , s2 ∈ S, t1 , t1 , t2 , t2 ∈ T and m, n ≥ 0 such that
−a 2 t1 t2 = a 2(m+n) (s1 s2 )2 + t
for some t ∈ T , an identity of the form (ii). The converse implication is clear.
Remark 3.9.13 Consider a real closed field R and an affine R-algebra A with
associated variety V . Letting F, G, E ⊆ A and keeping everything else as it was,
we obtain geometric Stellensätze as long as F and G are assumed to be finite. In
this situation X is namely constructible (since A is noetherian, the cardinality of E
does not matter) and in (i) of Theorems 3.9.9, 3.9.10 and 3.9.11 we may replace X
by X(R) := X ∩ V (R).
Example 3.9.14 The solution of Hilbert’s 17th Problem can be deduced from
Theorem 3.9.10: If A = R[t1 , . . . , tn ] and a ∈ A is positive semidefinite on R n ,
there exist f, g ∈ A2 and m ≥ 0 such that
(a 2m + f )(a 2m + g)
a= ,
(a 2m + g)2
for all x ∈ X.
Conversely, if such an inequality is satisfied for all x ∈ X, then of course g
vanishes at the zeroes of f in X.
Proof By assumption we have X ∩ ZA (f ) ⊆ ZA (g). Since X is an intersection
of closed constructible subsets of Sper A and ZA (g) is constructible, there exists
a constructible closed overset Y of X such that Y ∩ ZA (f ) ⊆ ZA (g) (cf.
Corollary 3.4.10). We may replace X by Y . In other words, we may assume without
loss of generality that X is constructible. Then there exist (finitely generated)
preorderings T1 , . . . , Tr of A such that
X = H A (T1 ) ∪ · · · ∪ H A (Tr ).
(To see this, write the complement of X as a finite union of sets of the form
H̊A (f1 , . . . , fN ).) An application of the Nullstellensatz (Theorem 3.9.11) yields
equations
g 2ni + ti = ai f 2
158 3 The Real Spectrum
Let A be a ring. In Sect. 3.7 we studied the ideals of A that are convex with
respect to an ordering P . In this section we will more generally consider the notion
of convexity with respect to a preordering T . Furthermore, the Stellensätze from
Sect. 3.9 will provide a geometric insight in the case of T -convex radical ideals.
Almost all the results concerning T -convex radical ideals proved in this section
go back to G. Brumfiel [17, 18], who also conducted fruitful research on more
general T -convex ideals (“completely convex ideals”) with his student R. Robson,
cf. [18, 95].
3.10 The Convex Radical Ideals Associated to a Preordering 159
a ≤T b :⇔ b − a ∈ T .
Then ≤T is reflexive and transitive, but only antisymmetric (and thus a partial order
relation on A) if and only if T ∩ (−T ) = 0. (If T ∪ (−T ) = A, then ≤T can be
interpreted as an ordering of the abelian group A/T ∩ (−T ).) For every subgroup
G of the additive group (A, +) the following properties are equivalent:
(i) From t, t ∈ T and t + t ∈ G it follows that t ∈ G (and t ∈ G);
(ii) from a, b ∈ G, c ∈ A and a ≤T c ≤T b it follows that c ∈ G;
(iii) from a ∈ G, c ∈ A and 0 ≤T c ≤T a it follows that c ∈ G.
(For (i) ⇒ (ii), take t := c − a and t := b − c. For (iii) ⇒ (i), take a := t + t and
c := t. The multiplicative structure of A does not play a role here. Thus T need not
be a preordering either.)
Definition 3.10.1 Let T be a preordering of A. A subgroup G of A (considered as
additive group) is called T -convex (or convex with respect to T ) if the equivalent
conditions (i)–(iii) are satisfied.
This generalizes Definition 3.7.1, see also Definition 2.1.1. In this section we
only consider T -convex ideals, while in the next section we focus on T -convex
subrings. Since every intersection of T -convex subgroups of A is again T -convex,
every subgroup (resp. every ideal, every subring) of A yields a smallest T -convex
subgroup (resp. a smallest T -convex ideal, a smallest T -convex subring) of A that
contains the given object.
If T = A is the trivial (improper) preordering, then A is the only T -convex
subgroup of A, an uninteresting case.
Consider the smallest preordering T0 = A2 of A and an ideal a of A. If a is
real reduced (i.e., if the ring A/a is real reduced), then a is T0 -convex. The converse
√ take A = k[t]/(t ) and a = (0) for any real
is false in general (counterexample: 2
2ba = (1 + b)2 a − b 2a − a.
Since the right hand side is in a, we also have ba ∈ a. Hence a is an ideal. The T -
convexity of a follows immediately (if t, t ∈ T such that t +t ∈ a, then t +t ∈ −T ,
hence t = (t + t ) − t ∈ −T ) and it is clear that a is contained in every T -convex
ideal of A.
0 ≤T u ≤T a and 0 ≤T v ≤T b.
0 ≤T u2 v 2 ≤T a 2 v 2 ≤T a 2 v 2 + b2 u2 and 0 ≤T u2 v 2 ≤T u2 b2 ≤T a 2 v 2 + u2 b2 ,
= {a ∈ A : −a 2n ∈ T − aT for some n ≥ 0}
T
= a ∈ A : there exist t, t ∈ T and n ≥ 0 such that a(a 2n + t) = t
between the power set of A and Sper A. These maps are such that for every subset
F ⊆ A we have
P H (F ) = P
[F ],
H P (X) = X,
convex in x, etc.) Proposition 3.10.10 thus says that for radical ideals a and arbitrary
preorderings T , a is T -convex if and only if a is H (T )-convex. A subgroup of A is
X-convex if and only if it is X-convex.
Since for subsets Y ⊆ X ⊆ Sper A, every Y -convex ideal is trivially also X-
convex, it follows in particular that for every x ∈ X the ideal supp(x) is X-convex.
We will show next that the converse also holds if X is closed.
Lemma 3.10.12 If X is a proconstructible subset of Sper A and p is a prime ideal
of A which is not convex in any x ∈ X, then there exists a closed constructible set
Y ⊇ X such that p is not Y -convex. In particular, p is not X-convex.
Proof For every x ∈ X there exist elements ax , bx ∈ A \ p such that ax (x) ≥ 0,
bx (x) ≥ 0 and ax + bx ∈ p. The sets Y (x) := H (ax , bx ) (x ∈ X) then constitute
a covering of X by closed constructible sets. Since the constructible topology is
compact (cf. Corollary 3.4.10(a)) there exist finitely many x1 , . . . , xr ∈ X such
that X ⊆ Y1 ∪ · · · ∪ Yr =: Y , where Yi := Y (xi ). There is no i ∈ {1, . . . , r} for
which p is Yi -convex. It follows from Lemma 3.10.8 that p is not Y -convex since
P (Y ) = P (Yi ).
i
Theorem 3.10.13 Let X be a closed subset of Sper A. Then the X-convex prime
ideals of A are precisely the supports supp(x) of all points x ∈ X.
Proof For x ∈ X, the ideal supp(x) is X-convex, as already remarked. Conversely,
consider an X-convex prime ideal p. By Lemma 3.10.12, p is convex in some x ∈ X,
i.e., p = supp(y) for some specialization y of x (cf. Theorem 3.7.4). Then y ∈ X
since X is closed.
From Theorem 3.10.13 we can deduce a very concrete geometric interpretation
of the X-convex radical ideals. If X ⊆ Sper A is a subset, we call a subset Z of X
algebraic in X if there exists and ideal a ⊆ A such that
Writing
!
I (Y ) := supp (y) = {a ∈ A : a(y) = 0 for all y ∈ Y },
y∈Y
we have:
Theorem 3.10.14 Let X be closed in Sper A. Then the X-convex radical ideals a
of A correspond bijectively to the subsets Z of X that are algebraic in X via
What does Theorem 3.10.14 mean in the “geometric setting”? Let V be an affine
variety over a real closed field R and let A := R[V ] be the associated R-algebra.
Consider a closed semialgebraic subset M ⊆ V (R) of the form
r
M= x ∈ V (R) : fi1 (x) ≥ 0, . . . , fis(i) (x) ≥ 0 (3.4)
i=1
for finitely many fij ∈ A. (By the—as yet unproven—Finiteness Theorem 3.5.8,
every closed semialgebraic M is of the form (3.4). We only require this assumption
in order to ensure that M is closed in V(R).) The (closed) constructible subset
M of Sper A that corresponds to M is the closure of M in Sper A (with respect
to the constructible or Harrison topology). It follows that P (M) = P (M) for the
associated (saturated) preorderings of A.
Let a be a radical ideal of A and W the associated closed subvariety of V . By
Theorem 3.10.14, a is M-convex if and only if a = I ZM (a) = I M ∩ ZA (a) .
However, ZA (a) = W (R), and so M ∩ ZA (a) = M ∩ W (R). Since I (N) = I (N)
for semialgebraic N ⊆ V (R) we can formulate Theorem 3.10.14 in the current
situation as follows:
Theorem 3.10.15 Let V be an affine R-variety, M a closed semialgebraic subset
of V (R) (of the form (3.4)), a a radical ideal of R[V ] and W the subvariety of V
associated to a. Then a is M-convex if and only if M ∩ W (R) is Zariski dense in W .
We return to our arbitrary commutative ring A and deduce two consequences
from Theorems 3.10.13 and 3.10.14. Let T be a preordering and a an ideal of A.
Definition 3.10.16 We denote the intersection of all T -convex ideals b that contain
a (i.e., the smallest T -convex ideal that contains a) by ciT (a). (Here “ci” stands for
“convex ideal”.)
2n
(b + b )2 + (b − b )2 = (2b2 + 2b 2 )2n = b 2n u + b 2n u
with u, u ∈ T . Thus,
2n
(b + b )2 + (b − b )2 + tu + t u = au + a u ∈ a.
Since the left hand side is of the form (b + b )4n + v with v ∈ T , it follows that
b + b ∈ b.
The next theorem concerns the relationship between a subset X of Sper A and
that it generates:
the probasic closed set X
have the same image
Proposition 3.10.18 If X ⊆ Sper A is closed, then X and X
under the support map supp : Sper A → Spec A.
Proof This follows immediately from Theorem 3.10.13 since P (X) = P (X).
We finish this section with the interesting and arguably difficult problem of
determining the relationships between X and X in reasonable situations. For
instance, if
r
X= H (fi1 , . . . , fis(i) )
i=1
3.11 Boundedness
In Sect. 2.1 we touched upon the notion of an archimedean field extension with
respect to a fixed ordering. We will now generalize this idea in two respects: instead
of field extensions we allow arbitrary ring extensions A ⊇ , and instead of a
fixed ordering we consider arbitrary subsets X of Sper A. This leads to the notion of
archimedean closure of in A on X, which we will characterize in several ways in
this Section. In fact, the archimedean closure can also be viewed as the convex hull
of a subring with respect to a preordering, as we will explain below.
In this section ϕ : → A will always denote a ring homomorphism.
Definition 3.11.1 Consider a subset X ⊆ Sper A.
(a) An element a ∈ A is called bounded on X over (or with respect to ϕ) if there
exists a λ ∈ such that
|a(x)| ≤ (ϕλ)(x)
for all x ∈ X. Clearly the elements of A that are bounded on X over form
a subring of A that contains ϕ( ). We denote this subring by oX (A/ ) or
oX (A, ϕ). If X = {x}, we simply write ox (A/ ) := o{x} (A/ ).
(b) We call A archimedean on X over (or with respect to ϕ) if oX (A/ ) =
A. If is a subring of A and ϕ the inclusion map, then we call oX (A/ )
the archimedean closure of in A on X. In case = oX (A/ ), we call
archimedean closed in A on X.
If we replace ϕ by the inclusion of the subring ϕ( ) of A, then oX (A/ ) does
not change. Hence, one can always think of as a subring of A (and of ϕ as the
inclusion map).
An element a ∈ A is (already) bounded on X with respect to ϕ if and only if
there exists a μ ∈ such that
|a(x)| ≤ |(ϕμ)(x)|
oX (A/ ) = oX (A/ ) = oX
(A/ ),
where X denotes the probasic closed subspace generated by X (cf. Sect. 3.10).
Indeed, by definition an element a ∈ A is in oX (A/ ) if and only if there exists
a λ ∈ such that X ⊆ H ϕ(λ)2 − a 2 . (Since the occurring rings are convex
hulls with respect to preorderings, the identities follow from P (X) = P (X) =
P (X).)
particular, if F is a real field, then all elements of the form a1 (1 + a12 + · · · + ar2 )−1
with a1 , . . . , ar ∈ F are absolutely bounded. More generally for any preordering T
of A, every element b ∈ A that satisfies an equation of the form b(1 + a 2 + t) = a
with a ∈ A and t ∈ T is absolutely bounded on H (T ).
Examples of archimedean ring extensions can be obtained from
Proposition 3.11.5 Every integral extension of rings A ⊇ is archimedean (on
the whole of Sper A).
Proof Every a ∈ A satisfies an equation of the form
a n + λ1 a n−1 + · · · + λn = 0
for every x ∈ Sper A. Since the inequality |λ| ≤ 1 + λ2 holds in every ordered field
we conclude that
We now prove a result that goes substantially beyond Proposition 3.11.5 and
which characterizes the archimedean closure of a subring on a subset of the
real spectrum. We may assume that char A = 0, i.e., that Z ⊆ A, since otherwise
Sper A = ∅ and the concepts become meaningless.
Definition 3.11.6 Let be a ring with Z ⊆ . We call a polynomial f (t) ∈ [t]
almost monic if its leading coefficient is a positive integer.
Theorem 3.11.7 (G.W. Brumfiel [17, Ch. VI]) Let A be a ring with Z ⊆ A, a
subring of A and X a subset of Sper A. For every a ∈ A the following statements
are equivalent:
(i) a ∈ oX (A/ ), i.e., a is bounded on X over ;
(ii) there exists an almost monic non-constant polynomial f (t) ∈ [t] such that
f (a 2 ) ≤ 0 on X;
(iii) there exists an almost monic non-constant polynomial f (t) ∈ [t] of even
degree such that f (a) ≤ 0 on X.
(For b ∈ A, the expression “b ≤ 0 on X” of course means that b(x) ≤ 0 for all
x ∈ X.)
Proof If a ∈ oX (A/ ) and λ ∈ such that |a(x)| ≤ λ(x) for all x ∈ X, then the
polynomial f (t) = t − λ2 satisfies (ii). The implication (ii) ⇒ (iii) is trivial (just
replace f (t) by f (t 2 )). It remains to show (iii) ⇒ (i). For this implication we appeal
to the remarkable
3.11 Boundedness 169
k
t 2n + u1 t 2n−1 + · · · + u2n = b(u) + hi (u, t)2 .
i=1
β(x) ≤ sa(x) ≤ β(x)
u1 2 u21
t 2 + u1 t + u2 = t + 2 + u2 − 4 .
u1
If n > 1, we apply the “Tschirnhausen transformation” s := t − 2n and obtain
we can apply the induction hypothesis to the second summand, which proves the
claim for any n.
Now we come to a different description of the archimedean closure of in A on
X, in which a more prominent role is played by the preordering P (X) of elements
of A that are non-negative on X. If T is an arbitrary preordering of A, we will
also write oT (A/ ) instead of oH (T ) (A/ ) (and oT (A) instead of oT (A/Z)). Thus
oT (A/ ) is the T-convex hull of in A. Because of Example 3.11.2(2), every ring
oX (A/ ) is of the form oT (A/ ), for instance with T = P (X).
170 3 The Real Spectrum
Proof From the existence of λ (resp. n) and t in (a) (resp. (b)), it follows that
−λ2 ≤ a ≤ λ2 (resp. −n ≤ a ≤ n) on X := H (T ). Hence it suffices to verify
the converse direction in (a). Given a ∈ oT (A/ ), we choose μ ∈ such that
|a(x)| ≤ |μ(x)| for all x ∈ X, and set λ := 1 + μ2. Since |a| ≤ |μ| < λ ≤ λ2 on X,
λ2 ± a are positive on X and the statement follows from Prestel’s Positivstellensatz
(Proposition 3.9.5, the (i) ⇒ (ii) implication).
and
o o(A/ )/ = o(A/ ).
λ4 − b 2 = a12 + · · · + ar2 .
Every |ai | is bounded by λ2 on the whole of Sper A, which shows that the ai are in
B. We conclude that |b| is bounded by λ2 on the whole of Sper B.
3.11 Boundedness 171
λa
1 + a2 + t
2λa
z= ,
1 + a2 + t
as required.
2a (a + 1)2 − (a − 1)2
= .
1+a +t
2 (a + 1)2 + (a − 1)2 + 2t
of a over is closed in Sper A and for each point in it, its generalizations are also
in it.
Proof We have
!
!(a/ ) = {x ∈ Sper A : ∀λ ∈ |a(x)| ≥ |λ(x)|} = H (a 2 − λ2 ),
λ∈
Example 3.11.18 Consider again a real closed field R and let V be an affine R-
variety with associated coordinate algebra A = R[V ]. Let u1 , . . . , un be generators
of R[V ] over R and a := u21 + · · · + u2n ∈ A. Then !(a/R) consists of all elements
x ∈ Sper A = V (R) whose “distance to the origin” (with respect to the embedding
of V into the affine space An given by u1 , . . . , un ) is infinitely large compared to R.
By Proposition 3.11.16, !(a/R) consists precisely of all the generalizations of the
points of
N
X⊆ {y ∈ Sper A : |a(y)| < λxi (y)} =: U.
i=1
where we may restrict the intersection on the right hand side to the points x of Xmax
or Xmin .
Proof This follows from Proposition 3.11.19 and Corollary 3.11.17.
Remark 3.11.21 Let V be an affine variety over a real closed field R and A = R[V ]
its coordinate algebra. Given a semialgebraic subset M of V (R) with associated
constructible subset M of Sper A, one might ask if oM (A/ ) is already the
intersection of the rings ox (A/ ) with x ∈ M, for instance for = R. In general
the answer to this question is negative. Trivially, ox (A/R) = A for all x ∈ V (R),
whereas only in special cases the R-valued function x → a(x) is bounded on M for
every a ∈ A. In all other cases, oM (A/R) is a proper subring of A. In connection
with Corollary 3.11.20, the usefulness of “ideal points” in M becomes clear at this
point once more: For every polynomial function f ∈ A that is unbounded on M,
and in this point f does indeed attain an infinitely
there exists such a point in M,
large function value compared to R.
174 3 The Real Spectrum
We finish Chap. 3 with a detailed study of the holomorphy ring o(F ) of a (real) field
F . The well-developed theory of the holomorphy ring—and more generally the ring
oT (F )—as well as interesting applications to power sums in fields is mainly due to
E. Becker and his student H.-W. Schülting (cf. [7, 8, 106], . . . ). A key notion in this
context is the concept of a Prüfer ring. We will only explain the rudiments of this
theory, which for the most part already go back to the work of D.W. Dubois [37].
Our presentation (strongly) emphasizes the use of the real spectrum and differs in
this respect from these earlier works. Once again the real spectrum often plays a
helpful and clarifying role.
Definition 3.12.1
(a) An integral domain A is called a Prüfer ring if every localization Ap of A at a
prime ideal p is a valuation ring. If in addition A is a subring of a field F and if
F is the field of fractions of A, then A is called a Prüfer ring of F .
(b) The Prüfer ring A is called residually real if for all p ∈ Spec A the residue field
κ(p) = Ap /pAp is real (i.e., if all valuation rings Ap are residually real in the
sense of Sect. 2.2). The notion of residually real Prüfer ring of F is defined
analogously.
Remarks 3.12.2
(1) Every valuation ring B is a Prüfer ring. If in addition κ(B) = B/mB is real,
then B is convex in F := Quot(B) with respect to an ordering P of F by
Corollary 2.7.2. The same is true for every overring of B in F , and it follows
that Bp /pBp is real for every p ∈ Spec B. This shows that our two definitions
of “residually real” (Definitions 3.12.1(b) and 2.2.5) are equivalent.
(2) If A is a Prüfer ring of F , then every overring B of A in F is also a Prüfer ring
since Bq is an overring of Aq∩A in F for every q ∈ Spec B. Similarly, every
overring (in F ) of a residually real Prüfer ring of F is again a residually real
Prüfer ring (cf. Remark 3.12.2(1)).
(3) Let A be an integral domain. If all localizations Am at maximal ideals m of A
are valuation rings (resp. residually real valuation rings), then A is already a
Prüfer ring (resp. a residually real Prüfer ring) since for every p ∈ Spec A such
that p ⊆ m, Ap is an overring of Am in the field of fractions of A.
(4) If A is an integral domain and F = Quot A, then A = m Am where the
intersectionis taken over all maximal ideals m of A in F . (This is easy to see:
Given λ ∈ m Am , the ideal of all elements a of A such that aλ ∈ A is not
contained in any maximal ideal of A, hence must be the unit ideal.) It follows
that every Prüfer ring of F is the intersection of valuation rings of F and thus
in particular integrally closed.
(5) If A is a Prüfer ring of F and K a further field, then there exists a natural
bijection from the set of places λ : F → K ∪ ∞ that are finite on A to
Hom(A, K), namely λ → λ|A . (Here and later Hom(·, ·) denotes the set of
3.12 Prüfer Rings and the Real Holomorphy Ring of a Field 175
1
(a ∈ F, a 2 = −1)
1 + a2
is a Prüfer ring of F .
Proof Let p be a prime ideal of A and B := Ap . Consider x ∈ F ∗ . We must show
that x ∈ B or x −1 ∈ B. First of all, x 2 ∈ B or x −2 ∈ B. This is clear if x 2 = −1.
Otherwise this follows from the identity
1 1
+ = 1,
1 + x2 1 + x −2
since the summands on the left hand side are not both in p, i.e., since one of those
summands is invertible in B. We may assume without loss of generality that x 2 ∈ B.
Let y := 1 + x. If y 2 ∈ B, then 2x = y 2 − 1 − x 2 ∈ B, hence x ∈ B. Otherwise,
y = 0 and y −2 ∈ B. Since (y − 1)2 = x 2 ∈ B we also have
y −2 (y − 1)2 = 1 − 2y −1 + y −2 ∈ B,
In particular, the real holomorphy ring o(F ) of F is the smallest residually real
Prüfer ring of F .
Proof (i) ⇒ (ii): Since o(F ) contains all elements of the form (1 + a 2 )−1 with a ∈
F , o(F ) is a Prüfer ring of F by Theorem 3.12.4. Consider a valuation ring B of F
such that o(F ) ⊆ B. If B is not residually real, there exist b1 , . . . , br ∈ B such that
1+b12 +· · ·+br2 ∈ mB , which is impossible since (1+b12 +· · ·+br2)−1 ∈ o(F ) ⊆ B.
Hence A is a residually real Prüfer ring of F .
(ii) ⇒ (iii): Since intersections of archimedean closed subrings of F are again
archimedean closed in F , it suffices to prove that residually real valuation rings of F
are archimedean closed. If B is a residually real valuation ring of F , then B is convex
in F with respect to some x ∈ Sper F (cf. Sect. 2.7). It follows that ox (F /B) = B,
hence a fortiori o(F /B) = oSper F (F /B) = B.
The implications (iii) ⇒ (iv) ⇒ (v) ⇒ (i) are trivial.
Corollary 3.12.6 The real holomorphy ring o(F ) is the intersection of all residu-
ally real valuation rings of F .
Corollary 3.12.10 If F contains R, then the support map supp : Sper o → Spec o
yields a bijection from (Sper o)max to the space (Spec o)max of maximal ideals of o.
Note that this bijection is not a homeomorphism in general.
Proof Since the field R does not possess any proper endomorphisms by Corol-
lary 2.1.12, every homomorphism ϕ : o → R is surjective, and ϕ is determined
by its kernel. Thus supp(x) is a maximal ideal for every x ∈ (Sper o)max
by Proposition 3.12.9. Since all prime ideals of o are real reduced, the map
supp : (Sper o)max → (Spec o)max is surjective. In the commutative diagram
the map Ker is injective. Thus the map supp is also injective by Proposition 3.12.9.
where a ∈ A, because S(a) = Hom(o, R) ∩ H̊o (a). Since all S(a) are T -open, T
is finer than T. A subbasis of open sets of T is given by all the sets
τ : o → C(Y, R), a →
a,
where C(Y, R) denotes the ring of continuous R-valued functions on Y and for all
a ∈ o, a is defined by
a (ϕ) := ϕ(a) for all ϕ ∈ Y .
It follows immediately from the definition that the functions
a separate the points
of Y . Furthermore, Q is contained in o. The Stone–Weierstrass Theorem [14, Ch. 10,
178 3 The Real Spectrum
§4, Th. 3] then implies that τ (o) is a dense subring of C(Y, R) with respect to the
uniform convergence topology.
The subset C + (Y, R) of nonnegative functions is a preordering of the ring
C(Y, R) (identical to the set of squares). Its preimage under τ satisfies
Proposition 3.12.12 Let a ∈ o. Then a ∈ C + (Y, R) if and only if a + 1
n is a sum of
squares (in F or o) for every n ∈ N.
Proof First of all, it is clear that o2 = o ∩ F 2 , since if a12 + · · · + ar2 is bounded
a + n1 ∈ C + (Y, R), and
over Z (on Sper F ), then so is each ai . If a + n1 ∈ F 2 , then
if this holds for all n ∈ N it follows that +
a ∈ C (Y, R). Conversely, a ∈ C + (Y, R)
implies that a is nonnegative as a function on (Sper o) max (cf. Proposition 3.12.9).
For all n ∈ N, the intersection of
H o −a − 1
n = {x ∈ Sper o : 1
n + a(x) ≤ 0}
Remark 3.12.16 Let ρ : Sper o → (Sper o)max = M(F ) denote the (continuous)
canonical retraction, which maps every x ∈ Sper o to the tip of the spear
{x} in Sper o (i.e., the closed specialization of x in Sper o), cf. Sect. 3.6. Given
x ∈ Sper F = (Sper o)min, it is easy to describe ρ(x) without reference to the
holomorphy ring o. Namely, ρ(x) is the “canonical” place F → R ∪ ∞ associated
to the ordering x of F , in other words the uniquely determined x-compatible place
whose valuation ring is the convex hull ox (F ) of Z in F with respect to x. The
restriction Sper F → M(F ) of ρ then makes M(F ) into a topological quotient
space of Sper F (since this map is closed).
A deeper geometric understanding of the space of places M(F ) (for instance in
the case of function fields over real closed base fields) is beyond our current abilities,
as this would require further knowledge of algebraic geometry beyond what we
can cover in this book (e.g., blow-ups). Nevertheless we present two illustrative
examples. These are quite special since they are “one-dimensional”. Therefore we
must emphasize that the holomorphy ring o(F ) and the space M(F ) are also useful
for higher-dimensional geometry.
Example 3.12.17 (Schülting, [106, p. 11 ff.]) Consider a one-dimensional real-an-
alytic manifold N and let M be a nonempty compact connected subset of N. Let
O(M) := lim O(U ), where U runs through the open neighbourhoods of M and
−→
O(U ) denotes the ring of real-analytic functions U → R. Then O(M) is an integral
domain. Let F denote the field of fractions of O(M), i.e., the field of “meromorphic
functions on M”. We assume that the functions of O(M) separate the points of
M (in fact, this is always the case, cf. [106, p. 13]). Then O(M) = o(F ) and the
canonical evaluation map M → Hom(o(F ), R) is a homeomorphism. Thus we have
rediscovered M with its topology as the space of places M(F ), and the holomorphy
ring o(F ) as the holomorphic functions on it.
Proof Every point x ∈ M gives rise to a discrete valuation vx of F with residue
field R, namely the order (of zeroes) at x. We have O(M) = {f ∈ F : vx (f ) ≥
0 for all x ∈ M}, from which o(F ) ⊆ O(M) follows. Conversely, every f ∈ O(M)
is bounded on M, hence bounded on an open neighbourhood U of √ M since M
is compact. If n ∈ N is chosen such that |f (x)| < n on U , then n ± f are
180 3 The Real Spectrum
Remarks 3.12.20
(1) The place λ is compatible with the preordering F 2 of F if and only if λ is real
(i.e., K is a real field). In this case we have λ(F 2 ) = K 2 .
3.12 Prüfer Rings and the Real Holomorphy Ring of a Field 181
s 2 + (c − s)s = cs ∈ p
b(v + v ) = uv + u v ∈ U,
(iii) follows by applying the proven implication (i) ⇒ (iii) to μ. All this establishes
the equivalence of statements (i) to (vii).
Assume now that λ is real, i.e., o ⊆ oλ (cf. Theorem 3.12.5). The equality
oλ = op was already established in Exercise 3.12.3, and the equivalence of (iv)
and (viii) follows from Lemma 3.12.5. The equivalence of (viii) and (ix) follows
from Proposition 3.10.10 and Theorem 3.10.13, and (ii) ⇒ (x) is trivial. Conversely,
assume (x). If q is an ideal of o that is maximal with respect to the property
q ∩ (1 + T ∩ o) = ∅, then q is prime and convex in o with respect to T ∩ o by
Proposition 3.10.5. Hence, oq is T -convex in F by the proven implication (viii) ⇒
(iv). From oq ⊆ op and Lemma 3.12.4 then follows the T -convexity of op = oλ , in
other words, (iv). This proves the theorem.
The first claim then follows since λ(T ) = λ(T ∩ oλ ), while the second claim is the
special case T = F 2 (i.e., λ(T ) = K 2 ), cf. Remark 3.12.20(1).
Since the prime ideals p of o(F ) correspond to the residually real valuation rings
of F (via p → o(F )p , cf. Exercise 3.12.3 and Theorem 3.12.5), the second claim of
Proposition 3.12.25 can also be formulated as follows:
If p is a prime ideal of o(F ), then o(F )/p is the holomorphy ring of the field
of fractions κ(p) of o(F )/p.
only if there exists an ordering P ⊇ T of F that is compatible with λ and such that
Q = λ(P ).
Proof For the nontrivial direction, let Q ⊇ λ(T ) and write Q∗ := Q\{0}. If P is an
ordering of F such that T ∪ λ−1 (Q∗ ) ⊆ P , then P clearly satisfies the claim. Hence
we must show that T ∪ λ−1 (Q∗ ) generates a proper preordering of F . Assume for
the sake of contradiction that this is not the case. Then there exist si ∈ λ−1 (Q∗ ),
ti ∈ T (i = 1, . . . , N) such that
1 + s1 t1 + · · · + sN tN = 0.
Hence there must be at least one i for which λ(ti ) = ∞. Let v denote the valuation
associated to λ and assume without loss of generality that v(t1 ) ≤ · · · ≤ v(tN ).
Then we obtain
t1−1 + s1 + s2 t2 + · · · + sN tN = 0,
where t1−1 and all ti = ti /t1 are in T ∩ oλ . Applying λ to this identity gives a
contradiction.
Proof Let p be a prime ideal of A. From Exercise 3.12.3 it follows that op∩o = Ap
(both valuation rings have the same centre o). This implies the commutativity of the
diagram
where the horizontal arrows are the bijections of Exercise 3.12.3 (localization). It
follows that the restriction map Spec A → Spec o is injective and that it preserves
the residue fields, from which already follows that rA is injective and that Y (A) =
{x ∈ Sper o : A ⊆ osupp(x)}. For this reason (c) is clear, as is the stability of Y (A)
under generalization. A priori Y (A) is proconstructible. If a = b/c ∈ A with 0 = b
and c ∈ o, then rA (H̊A (a)) = H̊o (bc)∩Y (A). It follows that rA is a homeomorphism
to Y (A). It remains to show the second claim of (b). If q ∈ Spec o and A ⊆ oq , then
Aq ⊆ A ∩ q oq = A. Conversely, if Aq ⊆ p with p ∈ Spec A, then A ⊆ Ap =
op∩o ⊆ oq .
Corollary 3.12.30 If A and A are residually real Prüfer rings of F with Y (A) =
Y (A ), then A = A .
Proof From Proposition 3.12.29 it follows that given a prime ideal q of o, oq
contains A if and only if oq contains A , and that A resp. A is the intersection
of all such oq .
By Proposition 3.12.29 we may identify Sper A with the subspace Y (A) of
Sper o. This space can be described quite explicitly as follows. Let y ∈ Sper F =
(Sper o)min . The specializations x of y in Sper o correspond uniquely to the subrings
B of F that are convex with respect to y, via the map x → osupp(x). (This
follows from Proposition 3.12.29 and Exercise 3.12.3.) This correspondence is
order reversing. Hence y corresponds to the largest such subring (namely B = F )
and the closed specialization y of y in Sper o corresponds to the smallest such
subring (namely B = oy (F )). Given any specialization x of y, it follows from
Proposition 3.12.29(b) that x is in Y (A) = Sper A if and only if A is contained in
osupp(x). In particular, the support ideal of the closed specialization of y in Sper A is
precisely the centre of A · oy (F ) (the y-convex hull of A) in o.
Results that are similar to Propositions 3.12.9–3.12.15 for the holomorphy ring
o(F ) can also be established for the ring A = oT (F ), where T is any proper
preordering of F . We will just indicate these results and leave the proofs to the
reader.
The canonical map Hom(A, R) → Sper A is a bijection onto the subset
of (Sper A)max . This subset contains the tips of all the spears {x} in Sper o for x ∈
H F (T ), but is larger in general. There is a natural bijective correspondence between
Hom(A, R) and the closed (hence compact) subspace
of M(F ). A result similar to Proposition 3.12.11 then also holds for A, and
Proposition 3.12.12 can be generalized as follows:
If a ∈ A and
a : Hom (A, R) → R denotes the evaluation map, then
a ≥ 0 if
and only if
a+ 1
n ∈T for all n ∈ N.
Chapter 4
Recent Developments
The results from Sects. 1.7–1.10 on counting real zeroes of univariate polynomials
are just the classical tip of a modern iceberg, that has emerged in the past decades.
We can merely sketch the general idea, for more details we have to refer to the
literature. The support supp(f ) of a polynomial f in n variables is the finite set
of n-tuples (α1 , . . . , αn ) ∈ Zn for which the monomial x1α1 · · · xnαn appears in f
with nonzero coefficient. Let a system of n real polynomials fi (x) (i = 1, . . . , n)
in n variables x = (x1 , . . . , xn ) be given. The number of nondegenerate
complex
solutions of the system f1 (x) = · · · = fn (x) = 0 is at most ni=1 deg(fi ) (Bézout),
and for the number of solutions in (C∗ )n there is the Bernstein-Khovanskiı̆-
Kushnirenko (BKK) upper bound in terms of the mixed volume of the convex
hulls of the supports supp(fi ). When it comes to real zeroes of real polynomials,
Khovanskiı̆ discovered around 1980 that what matters for their count is just the
number of different monomials in the fi , not their degrees or the volume of their
convex hull. More precisely, he gave an explicit number ϕ(n, N) with the following
property: Given any real polynomials f1 , . . . , fn in n variables that together have
at most N monomials, the system f1 (x) = · · · = fn (x) has at most ϕ(n, N)
nondegenerate real solutions in the open positive orthant. Note how in the univariate
case such a statement follows from Descartes’ rule of sign (cf. Corollary 1.10.3(a)).
In more than one variable, proofs for results of this kind become a lot more difficult.
This discovery gave rise to the term fewnomials, and it has meanwhile grown
into a large theory with many ramifications. The bound ϕ(n, N) itself is huge and
is generally believed to be far from best possible. For small values of N − n, much
better or even sharp bounds have been proved. For example, it has been shown for
n = 2 that two real trinomials f1 (x, y), f2 (x, y) can have at most 5 nondegenerate
positive common zeroes, and that this bound is sharp.
For more details we refer to the books by Khovanskiı̆ [58] and Sottile [110].
very hard. For the coordinate rings A = R[V ] of affine algebraic varieties V over
R, Mahé [79] proved an explicit (although very large) upper bound: There exists
a function w : N → N such that the cokernel of the global signature of R[V ] is
annihilated by 2w(d) whenever dim(V ) ≤ d.
In the case when A is a field, or more generally a semilocal ring, the integer s ≥ 0
(or s = ∞) for which 2s is the exponent of the cokernel of sign is called the stability
index st(A) of A (see Definition 3.8.7). Via the theory of fans, the stability index is
closely related to real valuations of A, and can be understood via powerful local–
global principles. Marshall’s abstract framework of spaces of orderings provides
strong tools for analyzing the stability index and related invariants.
The stability index has found attractive applications in geometry, and Corol-
lary 3.8.13 (the equivalence of (i) and (ii)) already points in this direction. Let V be
an affine R-variety. If M ⊂ V (R) is a basic open semialgebraic set, let s(M) denote
the minimal number r such that M = {x ∈ V (R) : f1 (x) > 0, . . . , fr (x) > 0} for
suitable fi ∈ R[V ], and put s(V ) = sup{s(M) : M ⊂ V (R) basic open}. For basic
closed sets M, define s(M) similarly using nonstrict inequalities fi (x) ≥ 0, and let
s(V ) = sup{s(M) : M ⊂ V (R) basic closed}.
Theorem 4.2.1 Let V be an affine R-variety of dimension d ≥ 1.
(a) (Bröcker–Scheiderer) s(V ) ≤ d,
(b) s(V ) ≤ 12 d(d + 1).
In fact, equality holds in (a) and (b) if V (R) is Zariski-dense in V .
See [16, 99] where much more general versions are proved. Basic closed sets
that actually require d2 (d + 1) inequalities are somewhat pathological; Averkov–
Bröcker [6] proved later that s(P ) ≤ d holds for any d-dimensional polyhedron P
in Rn . Note however that it remains very difficult to produce an explicit description
of minimal (or just small) length, for a general basic (open or closed) set M. Not
surprisingly, the price one has to pay for a short description is that the inequalities
tend to have very large degrees. See e.g., Burési–Mahé [21] for some explicit degree
bounds.
To allow for more flexible applications to geometrical questions, Marshall’s
spaces of orderings were later generalized to abstract real spectra [80] and to spaces
of signs [2]. The book [2] presents an extensive account of such applications,
at varying levels of abstractness (description of semialgebraic, semianalytic and
constructible sets by few inequalities).
4.3 Stellensätze
In Sect. 3.9 we discussed various Stellensätze. This is another topic where remark-
able progress was made since the German edition appeared. The results in Propo-
sitions 3.9.5 and 3.9.8 and Theorem 3.9.11 are of a general nature and hold for
arbitrary preorderings in arbitrary rings. Their common feature is that if a ring
190 4 Recent Developments
as deg(si ) ≤ 2d, then finding the supremum f(d) of all numbers c for which
(∗) exists becomes a standard semidefinite program. Semidefinite programs (SDP)
are solved very efficiently by interior path methods, due to ground-breaking ideas
of Karmarkar, Nesterov and Nemirenko in the 1980s and 1990s. Observe how
the Schmüdgen and Putinar Positivstellensätze imply that the restricted optima
converge to the true optimum when the degree d is taken to infinity. In other words,
f∗ = min f (K) = limd→∞ f(d) . In practice the semidefinite programs become
more and more complex when d is increased. Still, these ideas have become very
important and are now used in a large variety of situations.
192 4 Recent Developments
where the (finitely many) xi ∈ A are symmetric and α1 , . . . , αn are the elements of
F appearing in a diagonalization of the form x → TrdA (x ∗ x).
They showed that if deg A = 2, i.e., dimK A = 4, i.e., A is a quaternion algebra,
the weights are superfluous ([94, Cor. 5.5]). In general they are not, cf. [59], [60].
See [5] for an approach via signatures of hermitian forms. The weights are also
superfluous if every ordering of F is a ∗-ordering, cf. [94, Prop. 5.3]. This happens
for example in the “split” case where A = Mn (F ) and ∗ = T , as already observed
in 1974 by Gondard and Ribenboim, cf. [41].
In the context of real multivariate matrix polynomials (or polynomial matrices)
A = Mn (R[x1, . . . , xm ]), analogues have been obtained of the Krivine strict
Positivstellensatz by Cimprič [27] in 2011 and the Schmüdgen and Putinar strict
4.4 Noncommutative Stellensätze 193
Positivstellensätze by Scherer and Hol [102] in 2006. These are the equivalences
(i) ⇔ (ii), (i) ⇔ (ii’) and (i) ⇔ (ii”), respectively, in [27, Thm. 2]: Consider a
finite set of polynomials S = {g1 , . . . , gd } in R[x] = R[x1 , . . . , xm ] and let
S = {g1α1 · · · gdαd : α ∈ {0, 1}d }. Consider the basic closed set KS = {x ∈
Rm : g1 (x) ≥ 0, . . . , gd (x) ≥ 0} and the quadratic modules
d
MSn = C0 + Ci gi : C0 , . . . , Cd ∈ Mn (R[x])2 , TSn = M
n
S
, and TS = M
1
S
.
i=1
Let M ∈ Mn (R[x]) be any symmetric matrix. Then: (i) M(x) is strictly positive
definite for every x ∈ KS if and only if (ii) there exist t ∈ TS and V ∈ TSn such that
(1 + t)M = In + V .
Furthermore, if KS is compact, then (i) and (ii) are equivalent to: (ii’) there exists
ε > 0 such that M − εIn ∈ TSn . Finally, if one of the sets K{gi } = {x ∈ Rm : gi (x) ≥
0}, i = 1, . . . , d, is compact, then (i), (ii) and (ii’) are equivalent to: (ii”) there exists
ε > 0 such that M − εIn ∈ MSn .
A generalization is obtained by dropping the requirements that the variables
x1 , . . . , xm commute and that the matrix size is fixed. For example, in his seminal
2002 paper [48], Helton considered the free ∗-algebra A = Rx, x ∗ generated
by m noncommuting variables x = (x1 , . . . , xm ) and their “transposes” x ∗ =
(x1∗ , . . . , xm
∗ ). It is common practice to refer to the elements of A as noncommutative
(or NC) polynomials (or simply polynomials). When substituting matrices for the
variables of a noncommutative polynomial, the involution ∗ is replaced by matrix
transposition T . A symmetric noncommutative polynomial p ∈ A is matrix-
positive provided the matrix p(X, XT ) is positive semidefinite for all matrix tuples
X ∈ Mr (R)m and all r ∈ N. With this definition of positivity, Helton proved the
following noncommutative analogue of Artin’s Theorem [48, Thm. 1.1]: Let p be
any symmetric noncommutative polynomial. Then p is matrix-positive if and only
if it is a sum of hermitian squares.
In [82] McCullough and Putinar pointed out that Helton’s Theorem remains valid
for Cx, x ∗ and that Helton’s proof can easily be adapted to show this. They also
presented a new proof, based upon Carathéodory’s Theorem from convex geometry
and a Hahn–Banach separation argument.
In 2004, Helton and McCullough [50] established a strict Positivstellensatz
for three classes of matrix-valued noncommutative polynomials. To simplify the
exposition we only describe the case of symmetric noncommutative polynomials.
Given a collection P of symmetric polynomials in noncommuting variables x =
(x1 , . . . , xm ), its positivity domain DP is defined as follows: let H be a separable
real Hilbert space and let DP (H ) denote the collection of tuples X = (X1 , . . . , Xm )
such that each Xj is a symmetric operator on H and p(X) is positive semidefinite
for each p ∈ P. Then DP is defined to be the collection of tuples X such that
X ∈ DP (H ) for some H . The positivity domain DP is bounded if there exists a
194 4 Recent Developments
and involution given by ak∗ = a−k for k = 1, . . . , d. The Weyl algebra W(d)
has a natural filtration with corresponding graded algebra the polynomial algebra
C[z, z] in 2d complex variables z = (z1 , . . . , zd ), z = (z1 , . . . , zd ), where zj and
zj correspond to aj and aj∗ , respectively. If c ∈ W(d) is an element of degree n,
the polynomial associated to the n-th component of c is denoted cn (z, z). Let α be
a fixed positive number which is not an integer, and let N denote the set of all finite
products of the elements a1∗ a1 + · · · + ad∗ ad + (α + n)1, where n ∈ Z. Finally,
consider the positive cone
1. Andradas, C., Bröcker, L., Ruiz, J.M.: Minimal generation of basic open semianalytic sets.
Invent. Math. 92(2), 409–430 (1988). https://doi.org/10.1007/BF01404461
2. Andradas, C., Bröcker, L., Ruiz, J.M.: Constructible sets in real geometry, Ergebnisse der
Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)],
vol. 33. Springer, Berlin (1996). https://doi.org/10.1007/978-3-642-80024-5
3. Artin, E.: Über die Zerlegung definiter Funktionen in Quadrate. Abh. Math. Sem. Univ.
Hamburg 5(1), 100–115 (1927). https://doi.org/10.1007/BF02952513
4. Artin, E., Schreier, O.: Algebraische Konstruktion reeller Körper. Abh. Math. Sem. Univ.
Hamburg 5(1), 85–99 (1927). https://doi.org/10.1007/BF02952512
5. Astier, V., Unger, T.: Signatures of Hermitian forms, positivity, and an answer to a question
of Procesi and Schacher. J. Algebra 508, 339–363 (2018). https://doi.org/10.1016/j.jalgebra.
2018.05.004
6. Averkov, G., Bröcker, L.: Minimal polynomial descriptions of polyhedra and special semi-
algebraic sets. Adv. Geom. 12(3), 447–459 (2012). https://doi.org/10.1515/advgeom-2011-
059
7. Becker, E.: The real holomorphy ring and sums of 2nth powers. In: Real Algebraic Geometry
and Quadratic Forms (Rennes, 1981). Lecture Notes in Math., vol. 959, pp. 139–181.
Springer, Berlin-New York (1982)
8. Becker, E.: Valuations and real places in the theory of formally real fields. In: Real Algebraic
Geometry and Quadratic Forms (Rennes, 1981). Lecture Notes in Math., vol. 959, pp. 1–40.
Springer, Berlin-New York (1982)
9. Becker, E., Köpping, E.: Reduzierte quadratische Formen und Semiordnungen reeller Körper.
Abh. Math. Sem. Univ. Hamburg 46, 143–177 (1977). https://doi.org/10.1007/BF02993018
10. Becker, E., Spitzlay, K.J.: Zum Satz von Artin-Schreier über die Eindeutigkeit des reellen
Abschlusses eines angeordneten Körpers. Comment. Math. Helv. 50, 81–87 (1975). https://
doi.org/10.1007/BF02565735
11. Bochnak, J., Coste, M., Roy, M.F.: Géométrie algébrique réelle. Ergebnisse der Mathematik
und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 12. Springer,
Berlin (1987)
12. Bourbaki, N.: Algèbre. Masson, Paris
13. Bourbaki, N.: Algèbre Commutative. Masson, Paris
14. Bourbaki, N.: Topologie Générale. Masson, Paris
15. Bröcker, L.: Zur Theorie der quadratischen Formen über formal reellen Körpern. Math. Ann.
210, 233–256 (1974). https://doi.org/10.1007/BF01350587
16. Bröcker, L.: On the stability index of Noetherian rings. In: Real Analytic and Algebraic
Geometry (Trento, 1988). Lecture Notes in Math., vol. 1420, pp. 72–80. Springer, Berlin
(1990). https://doi.org/10.1007/BFb0083912
17. Brumfiel, G.W.: Partially ordered rings and semi-algebraic geometry. London Mathematical
Society Lecture Note Series, vol. 37. Cambridge University Press, Cambridge-New York
(1979)
18. Brumfiel, G.W.: Real valuation rings and ideals. In: Real Algebraic Geometry and Quadratic
Forms (Rennes, 1981). Lecture Notes in Math., vol. 959, pp. 55–97. Springer, Berlin-New
York (1982)
19. Brumfiel, G.W.: Witt Rings and K-Theory. pp. 733–765 (1984). https://doi.org/10.1216/RMJ-
1984-14-4-733. Ordered fields and real algebraic geometry (Boulder, Colo., 1983)
20. Brumfiel, G.W.: The real spectrum compactification of Teichmüller space. In: Geometry of
Group Representations (Boulder, CO, 1987). Contemp. Math., vol. 74, pp. 51–75. Amer.
Math. Soc., Providence (1988). https://doi.org/10.1090/conm/074/957511
21. Burési, J., Mahé, L.: Reducing inequalities with bounds. Math. Z. 227(2), 231–243 (1998).
https://doi.org/10.1007/PL00004371
22. Burgdorf, S., Scheiderer, C., Schweighofer, M.: Pure states, nonnegative polynomials and
sums of squares. Comment. Math. Helv. 87(1), 113–140 (2012). https://doi.org/10.4171/
CMH/250
23. Carral, M., Coste, M.: Normal spectral spaces and their dimensions. J. Pure Appl. Algebra
30(3), 227–235 (1983). https://doi.org/10.1016/0022-4049(83)90058-0
24. Cassels, J.W.S., Ellison, W.J., Pfister, A.: On sums of squares and on elliptic curves
over function fields. J. Number Theory 3, 125–149 (1971).https://doi.org/10.1016/0022-
314X(71)90030-8
25. Chaillou, J.: Hyperbolic differential polynomials and their singular perturbations. Mathe-
matics and its Applications, vol. 3. D. Reidel Publishing Co., Dordrecht-Boston (1979).
Translated from the French by J. W. Nienhuys
26. Choi, M.D., Lam, T.Y.: An old question of Hilbert. In: Conference on Quadratic Forms—
1976 (Proc. Conf., Queen’s Univ., Kingston, Ont., 1976), pp. 385–405. Queen’s Papers in
Pure and Appl. Math., No. 46 (1977)
27. Cimprič, J.: Strict Positivstellensätze for matrix polynomials with scalar constraints. Linear
Algebra Appl. 434(8), 1879–1883 (2011). https://doi.org/10.1016/j.laa.2010.11.046
28. Coste, M., Roy, M.F.: La topologie du spectre réel. In: Ordered Fields and Real Algebraic
Geometry (San Francisco, Calif., 1981). Contemp. Math., vol. 8, pp. 27–59. Amer. Math.
Soc., Providence (1982)
29. Coste-Roy, M.F.: Faisceau structural sur le spectre réel et fonctions de Nash. In: Real
Algebraic Geometry and Quadratic Forms (Rennes, 1981). Lecture Notes in Math., vol. 959,
pp. 406–432. Springer, Berlin-New York (1982)
30. de Oliveira, M.C., Helton, J.W., McCullough, S.A., Putinar, M.: Engineering systems and free
semi-algebraic geometry. In: Emerging Applications of Algebraic Geometry. IMA Vol. Math.
Appl., vol. 149, pp. 17–61. Springer, New York (2009). https://doi.org/10.1007/978-0-387-
09686-5_2
31. Delfs, H., Knebusch, M.: Semialgebraic topology over a real closed field. I. Paths and
components in the set of rational points of an algebraic variety. Math. Z. 177(1), 107–129
(1981). https://doi.org/10.1007/BF01214342
32. Delfs, H., Knebusch, M.: Semialgebraic topology over a real closed field. II. Basic theory of
semialgebraic spaces. Math. Z. 178(2), 175–213 (1981). https://doi.org/10.1007/BF01262039
33. Delfs, H., Knebusch, M.: Locally semialgebraic spaces. Lecture Notes in Mathematics,
vol. 1173. Springer-Verlag, Berlin (1985). https://doi.org/10.1007/BFb0074551
34. Delzell, C.N., Madden, J.J.: A completely normal spectral space that is not a real spectrum. J.
Algebra 169(1), 71–77 (1994). https://doi.org/10.1006/jabr.1994.1272
35. Dickmann, M., Schwartz, N., Tressl, M.: Spectral spaces. New Mathematical Mono-
graphs, vol. 35. Cambridge University Press, Cambridge (2019). https://doi.org/10.1017/
9781316543870
References 197
36. Dress, A.: Lotschnittebenen mit halbierbarem rechtem Winkel. Arch. Math. (Basel) 16, 388–
392 (1965). https://doi.org/10.1007/BF01220047
37. Dubois, D.W.: Infinite primes and ordered fields. Dissertationes Math. (Rozprawy Mat.) 69,
43 (1970)
38. Elman, R., Lam, T.Y., Wadsworth, A.R.: Orderings under field extensions. J. Reine Angew.
Math. 306, 7–27 (1979)
39. Fuller, A.T.: Aperiodicity determinants expressed in terms of roots. Int. J. Control 47(6),
1571–1593 (1988). https://doi.org/10.1080/00207178808906122
40. Gantmacher, F.R.: Matrizentheorie. Springer-Verlag, Berlin (1986). https://doi.org/10.1007/
978-3-642-71243-2. With an appendix by V. B. Lidskij, With a preface by D. P. Želobenko,
Translated from the second Russian edition by Helmut Boseck, Dietmar Soyka and Klaus
Stengert
41. Gondard, D., Ribenboim, P.: Le 17e problème de Hilbert pour les matrices. Bull. Sci. Math.
(2) 98(1), 49–56 (1974)
42. Gorin, E.A.: Asymptotic properties of polynomials and algebraic functions of several
variables. Uspehi Mat. Nauk 16(1 (97)), 91–118 (1961)
43. Grätzer, G.: General Lattice Theory. Birkhäuser Verlag, Basel-Stuttgart (1978). Lehrbücher
und Monographien aus dem Gebiete der Exakten Wissenschaften, Mathematische Reihe,
Band 52
44. Harrison, D.K.: Witt Rings. Lecture Notes. Univ. of Kentucky, Lexington (1970)
45. Hartshorne, R.: Algebraic Geometry. Springer-Verlag, New York-Heidelberg (1977). Gradu-
ate Texts in Mathematics, No. 52
46. Helmke, U.: Rational functions and Bezout forms: a functorial correspondence. Linear
Algebra Appl. 122/123/124, 623–640 (1989). https://doi.org/10.1016/0024-3795(89)90669-
1
47. Helmke, U., Fuhrmann, P.A.: Bezoutians. Linear Algebra Appl. 122/123/124, 1039–1097
(1989). https://doi.org/10.1016/0024-3795(89)90684-8
48. Helton, J.W.: “Positive” noncommutative polynomials are sums of squares. Ann. Math. (2)
156(2), 675–694 (2002). https://doi.org/10.2307/3597203
49. Helton, J.W., Klep, I., McCullough, S.: The convex Positivstellensatz in a free algebra. Adv.
Math. 231(1), 516–534 (2012). https://doi.org/10.1016/j.aim.2012.04.028
50. Helton, J.W., McCullough, S.A.: A Positivstellensatz for non-commutative polynomials.
Trans. Am. Math. Soc. 356(9), 3721–3737 (2004). https://doi.org/10.1090/S0002-9947-04-
03433-6
51. Hochster, M.: Prime ideal structure in commutative rings. Trans. Am. Math. Soc. 142, 43–60
(1969). https://doi.org/10.2307/1995344
52. Hölder, O.: Die Axiome der Quantität und die Lehre vom Maß. Leipz. Ber. 53, 1–64 (1901)
53. Hörmander, L.: On the division of distributions by polynomials. Ark. Mat. 3, 555–568 (1958).
https://doi.org/10.1007/BF02589517
54. Jacobson, N.: Lectures in Abstract Algebra (3 volumes). D. Van Nostrand Co., Inc., Toronto,
New York, London (1951–1954)
55. Jacobson, N.: Basic Algebra (2 volumes). W. H. Freeman and Co., San Francisco, Calif. (1974
and 1980)
56. Johnstone, P.T.: Stone Spaces. Cambridge Studies in Advanced Mathematics, vol. 3. Cam-
bridge University Press, Cambridge (1982)
57. Kato, K.: A Hasse principle for two-dimensional global fields. J. Reine Angew. Math. 366,
142–183 (1986). https://doi.org/10.1515/crll.1986.366.142. With an appendix by Jean-Louis
Colliot-Thélène
58. Khovanskiı̆, A.G.: Fewnomials. Translations of Mathematical Monographs, vol. 88. American
Mathematical Society, Providence (1991). https://doi.org/10.1090/mmono/088. Translated
from the Russian by Smilka Zdravkovska
59. Klep, I., Unger, T.: The Procesi-Schacher conjecture and Hilbert’s 17th problem for algebras
with involution. J. Algebra 324(2), 256–268 (2010). https://doi.org/10.1016/j.jalgebra.2010.
03.022
198 References
60. Klep, I., Špenko, v., Volčič, J.: Positive trace polynomials and the universal Procesi-Schacher
conjecture. Proc. Lond. Math. Soc. (3) 117(6), 1101–1134 (2018). https://doi.org/10.1112/
plms.12156
61. Knebusch, M.: Symmetric bilinear forms over algebraic varieties. In: Conference on
Quadratic Forms—1976 (Proc. Conf., Queen’s Univ., Kingston, Ont., 1976), pp. 103–283.
Queen’s Papers in Pure and Appl. Math., No. 46 (1977)
62. Knebusch, M., Kolster, M.: Wittrings. Aspects of Mathematics, vol. 2. Friedr. Vieweg & Sohn,
Braunschweig (1982)
63. Knus, M.A.: Quadratic and Hermitian Forms Over Rings. Grundlehren der mathematischen
Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 294. Springer-
Verlag, Berlin (1991). https://doi.org/10.1007/978-3-642-75401-2. With a foreword by I.
Bertuccioni
64. Knus, M.A., Merkurjev, A., Rost, M., Tignol, J.P.: The Book of Involutions. American
Mathematical Society Colloquium Publications, vol. 44. American Mathematical Society,
Providence (1998). https://doi.org/10.1090/coll/044. With a preface in French by J. Tits
65. Kreı̆n, M.G., Naı̆mark, M.A.: The method of symmetric and Hermitian forms in the theory
of the separation of the roots of algebraic equations. Linear and Multilinear Algebra 10(4),
265–308 (1981). https://doi.org/10.1080/03081088108817420. Translated from the Russian
by O. Boshko and J. L. Howland
66. Krivine, J.L.: Anneaux préordonnés. J. Anal. Math. 12, 307–326 (1964). https://doi.org/10.
1007/BF02807438
67. Krull, W.: Allgemeine Bewertungstheorie. J. Reine Angew. Math. 167, 160–196 (1932).
https://doi.org/10.1515/crll.1932.167.160
68. Kunz, E.: Einführung in die kommutative Algebra und algebraische Geometrie. Vieweg
Studium: Aufbaukurs Mathematik [Vieweg Studies: Mathematics Course], vol. 46. Friedr.
Vieweg & Sohn, Braunschweig (1980). With an English preface by David Mumford
69. Lam, T.Y.: The algebraic theory of quadratic forms. W. A. Benjamin, Inc., Reading (1973).
Mathematics Lecture Note Series
70. Lam, T.Y.: Orderings, Valuations and Quadratic Forms. CBMS Regional Conference Series
in Mathematics, vol. 52. Published for the Conference Board of the Mathematical Sciences,
Washington, DC; by the American Mathematical Society, Providence (1983). https://doi.org/
10.1090/cbms/052
71. Lam, T.Y.: An introduction to real algebra. Rocky Mount. J. Math. 14(4), 767–814 (1984).
https://doi.org/10.1216/RMJ-1984-14-4-767. Ordered fields and real algebraic geometry
(Boulder, Colo., 1983)
72. Lam, T.Y.: Introduction to Quadratic Forms Over Fields. Graduate Studies in Mathematics,
vol. 67. American Mathematical Society, Providence (2005)
73. Lang, S.: Algebra. Graduate Texts in Mathematics, vol. 211, 3rd edn. Springer, New York
(2002). https://doi.org/10.1007/978-1-4613-0041-0
74. Lasserre, J.B.: An introduction to polynomial and semi-algebraic optimization. Cambridge
Texts in Applied Mathematics. Cambridge University Press, Cambridge (2015). https://doi.
org/10.1017/CBO9781107447226
75. Leicht, J.B.: Zur Charakterisierung reell abgeschlossener Körper. Monatsh. Math. 70, 452–
453 (1966). https://doi.org/10.1007/BF01300449
76. Łojasiewicz, S.: Sur le problème de la division. Stud. Math. 18, 87–136 (1959). https://doi.
org/10.4064/sm-18-1-87-136
77. Lorenz, F., Leicht, J.: Die Primideale des Wittschen Ringes. Invent. Math. 10, 82–88 (1970).
https://doi.org/10.1007/BF01402972
78. Mahé, L.: Signatures et composantes connexes. Math. Ann. 260(2), 191–210 (1982). https://
doi.org/10.1007/BF01457236
79. Mahé, L.: Théorème de Pfister pour les variétés et anneaux de Witt réduits. Invent. Math.
85(1), 53–72 (1986). https://doi.org/10.1007/BF01388792
80. Marshall, M.A.: Spaces of Orderings and Abstract Real Spectra. Lecture Notes in Mathemat-
ics, vol. 1636. Springer, Berlin (1996). https://doi.org/10.1007/BFb0092696
References 199
81. Matsumura, H.: Commutative Algebra. Mathematics Lecture Note Series, vol. 56, 2nd edn.
Benjamin/Cummings Publishing Co., Inc., Reading (1980)
82. McCullough, S., Putinar, M.: Noncommutative sums of squares. Pac. J. Math. 218(1), 167–
171 (2005). https://doi.org/10.2140/pjm.2005.218.167
83. Milnor, J.: Algebraic K-theory and quadratic forms. Invent. Math. 9, 318–344 (1969/1970).
https://doi.org/10.1007/BF01425486
84. Motzkin, T.S.: The arithmetic-geometric inequality. In: Inequalities (Proc. Sympos. Wright-
Patterson Air Force Base, Ohio, 1965), pp. 205–224. Academic Press, New York (1967)
85. Obreschkoff, N.: Verteilung und Berechnung der Nullstellen reeller Polynome. VEB
Deutscher Verlag der Wissenschaften, Berlin (1963)
86. Ostrowski, A.: Über einige Lösungen der Funktionalgleichung ϕ(x) · ϕ(x) = ϕ(xy). Acta
Math. 41(1), 271–284 (1916). https://doi.org/10.1007/BF02422947
87. Pascoe, J.E.: Positivstellensätze for noncommutative rational expressions. Proc. Am. Math.
Soc. 146(3), 933–937 (2018). https://doi.org/10.1090/proc/13773
88. Pfister, A.: Zur Darstellung definiter Funktionen als Summe von Quadraten. Invent. Math. 4,
229–237 (1967). https://doi.org/10.1007/BF01425382
89. Pourchet, Y.: Sur la représentation en somme de carrés des polynômes à une indéterminée sur
un corps de nombres algébriques. Acta Arith. 19, 89–104 (1971). https://doi.org/10.4064/aa-
19-1-89-104
90. Prestel, A.: Lectures on Formally Real Fields. Lecture Notes in Mathematics, vol. 1093.
Springer, Berlin (1984). https://doi.org/10.1007/BFb0101548
91. Prestel, A.: Einführung in die Mathematische Logik und Modelltheorie. Vieweg Studium:
Aufbaukurs Mathematik [Vieweg Studies: Mathematics Course], vol. 60. Friedr. Vieweg &
Sohn, Braunschweig (1986). https://doi.org/10.1007/978-3-663-07641-4
92. Prestel, A., Delzell, C.N.: Positive Polynomials. Springer Monographs in Mathematics.
Springer, Berlin (2001). https://doi.org/10.1007/978-3-662-04648-7. From Hilbert’s 17th
problem to real algebra
93. Prieß-Crampe, S.: Angeordnete Strukturen. Ergebnisse der Mathematik und ihrer Grenzge-
biete [Results in Mathematics and Related Areas], vol. 98. Springer, Berlin (1983). https://
doi.org/10.1007/978-3-642-68628-3. Gruppen, Körper, projektive Ebenen. [Groups, fields,
projective planes]
94. Procesi, C., Schacher, M.: A non-commutative real Nullstellensatz and Hilbert’s 17th
problem. Ann. Math. (2) 104(3), 395–406 (1976). https://doi.org/10.2307/1970962
95. Robson, R.O.: The Ideal Theory of Real Algebraic Curves and Affine Embeddings of Semi-
Algebraic Spaces and Manifolds. Ph.D. Thesis, Stanford University (1981)
96. Roman, S.: Field Theory. Graduate Texts in Mathematics, vol. 158, 2nd edn. Springer, New
York (2006)
97. Schafarewitsch, I.R.: Grundzüge der algebraischen Geometrie. Friedr. Vieweg+Sohn, Braun-
schweig (1972). Übersetzung aus dem Russischen von Rudolf Fragel, Logik und Grundlagen
der Mathematik, Band 12
98. Scharlau, W.: Quadratic and Hermitian Forms. Grundlehren der Mathematischen Wis-
senschaften [Fundamental Principles of Mathematical Sciences], vol. 270. Springer, Berlin
(1985). https://doi.org/10.1007/978-3-642-69971-9
99. Scheiderer, C.: Stability index of real varieties. Invent. Math. 97(3), 467–483 (1989). https://
doi.org/10.1007/BF01388887
100. Scheiderer, C.: Sums of squares on real algebraic surfaces. Manuscr. Math. 119(4), 395–410
(2006). https://doi.org/10.1007/s00229-006-0630-5
101. Scheiderer, C.: Positivity and sums of squares: a guide to recent results. In: Emerging
Applications of Algebraic Geometry. IMA Vol. Math. Appl., vol. 149, pp. 271–324. Springer,
New York (2009). https://doi.org/10.1007/978-0-387-09686-5_8
102. Scherer, C.W., Hol, C.W.J.: Matrix sum-of-squares relaxations for robust semi-definite
programs. Math. Program. 107(1-2, Ser. B), 189–211 (2006). https://doi.org/10.1007/s10107-
005-0684-2
200 References
103. Schmüdgen, K.: The K-moment problem for compact semi-algebraic sets. Math. Ann. 289(2),
203–206 (1991). https://doi.org/10.1007/BF01446568
104. Schmüdgen, K.: A strict Positivstellensatz for the Weyl algebra. Math. Ann. 331(4), 779–794
(2005). https://doi.org/10.1007/s00208-004-0604-4
105. Schmüdgen, K.: Noncommutative real algebraic geometry—some basic concepts and first
ideas. In: Emerging Applications of Algebraic Geometry. IMA Vol. Math. Appl., vol. 149,
pp. 325–350. Springer, New York (2009). https://doi.org/10.1007/978-0-387-09686-5_9
106. Schülting, H.W.: Über reelle Stellen eines Körpers und ihren Holomorphiering. Ph.D. Thesis,
Univ. Dortmund (1979)
107. Schwartz, N.: The Basic Theory of Real Closed Spaces. Regensburger Mathematische
Schriften [Regensburg Mathematical Publications], vol. 15. Universität Regensburg, Fach-
bereich Mathematik, Regensburg (1987)
108. Schwartz, N.: The basic theory of real closed spaces. Mem. Am. Math. Soc. 77(397), viii+122
(1989). https://doi.org/10.1090/memo/0397
109. Serre, J.P.: Groupes algébriques et corps de classes. Hermann, Paris (1975). Deuxième édition,
Publication de l’Institut de Mathématique de l’Université de Nancago, No. VII, Actualités
Scientifiques et Industrielles, No. 1264
110. Sottile, F.: Real Solutions to Equations from Geometry. University Lecture Series, vol. 57.
American Mathematical Society, Providence (2011). https://doi.org/10.1090/ulect/057
111. Stengle, G.: A nullstellensatz and a positivstellensatz in semialgebraic geometry. Math. Ann.
207, 87–97 (1974). https://doi.org/10.1007/BF01362149
112. Swan, R.G.: Topological examples of projective modules. Trans. Am. Math. Soc. 230, 201–
234 (1977). https://doi.org/10.2307/1997717
113. van der Waerden, B.L.: Algebra (2 volumes). Heidelberger Taschenbücher. Springer, Berlin-
Heidelberg-New York (1971 and 1967)
114. Weber, H.: Lehrbuch der Algebra (3 Volumes). Reprint: Chelsea Publ. Comp., New York
(1963)
Symbol Index
K 2 , 2 Var(c0 , . . . , cr ), 21
signP a, 4 V (x), 21
a1 , . . . , an , 5 sr (f ), 23
H, 5 S(f ), 24
ϕ ⊥ϕ ,ϕ⊗ϕ ,5 Sλ (f ), 25
n · ϕ, ϕ ⊗n , 6 trA/K , 26
U ⊥, 6 [A, a]K , 27
Rad (ϕ), 6 SylK (f ; g), 27
rank (ϕ), 6 indα (ϕ), 30
det ϕ, 6 Iab (ϕ), 30
dim ϕ, 6 B(f, g), 32
ϕ ∼ϕ,7 H (f, g), 32
[ϕ], 7 R(f, g), 35
W (K), 7 s∗ ϕ, 48
signP ϕ, signP , sign, 8 (1 × · · · × r )lex , 55
e : W (K) → Z/2Z, 10 oP (K/A), o(K/A), 55
I (K), 10 oP (K), o(K), 55
ϕL , 10 mA , 58
iL/K , 10 κ(A), 58
τ | σ , 11 ov , mv , 64
A , vA , 65 x % y, 116
rank , rank A, 69 Sper ϕ, ϕ ∗ , 116
TP , Tv , 74 ϕβ , 117
λA , 79 K(Sper A), 121
oλ , mλ , κλ , 80 K̊(Sper A), K(Sper A), 121
λ , vλ , 80 K(X), K̊(X), K(X), 124
R{t},
√ 85 Xcon , 124
a, 101 X ∗ , 126
Ared , 101 S(M), S̊(M), S(M), 128
Spec A, 102 128
M,
κA (p), κ(p), 102 Filt(L), Prime(L), Ultra(L), 130
ρp : A → κ(p), 102 pro-L, 130
f (p), 102 X max , X min , 131
DA (f ), DA (T ), 103 ρX , ρ, 139
VA (f ), VA (T ), 103 M c , Ac , ac , 142
I(Y ), 103 HF (a1 , . . . , an ), 145
K̊(Spec A), 105 sign ϕ, sign, 146
Spec ϕ, ϕ ∗ , 105 W (F ), I (F ), 146
ϕq , 105 ϕ, 146
f −∞ A, 106 χY , 146
n
V (K), 107 I n (F ), I (F ), 147
k[V ], 107 st(F ), 147
VA , VA (K), 108 rK/F , 149
ϕ ∗ , 108 P [F ], 151
A2 , 110 a ≤T b, 159
(Spec
√ A)re , 111 T, 161
re
a, 113 P (X), 162
Sper A, 114 162
X,
supp α, 114 ciT (a), 164
k(α), 114 oX (A/ ), ox (A/ ), 166
f (α), 114 oX (A), o(A), 167
k(P ), 116 oT (A/ ), oT (A), 169
f (P ), 116 !(a/ ), 172
H̊A (T ), H̊A (f1 , . . . , fr ), 114 M(F ), 178
H A (T ), H A (f1 , . . . , fr ), 114
a , 178
ZA (T ), ZA (f1 , . . . , fr ), 114 λ(T ), 180
H̊A , 114 M(F /T ), 186
Index
A in Sper A, 135
Absolutely bounded, 167 Constructible
Absolute value sets, 121, 124
on a field, 74 topology, 121, 124
ultrametric, 74 Convex
Affine algebra of an affine variety, 107 hull, 53
Affine k-variety, 107 subgroup, 54
Almost monic polynomial, 168 subring, 58
Anisotropic quadratic form, 6 subset, 53
Archimedean
closed subring, 166
over a subgroup, subring, . . . , 54, 55, 166 D
saturated subfield, 71 Dedekind cut (generalized), 84
Archimedean closure of a subring, 166 free, 84
Archimedes, axiom of, 55 proper, 57, 84
Artin–Lang Stellensatz, 92, 93, 113 Definite (positive, negative), 8, 95
Descartes, rule of, 42
Determinant of a quadratic form, 6
B Dimension index, 10
Bézout matrix, 32 Discrete valuation ring of rank one, 69
Bézoutian, 32 Distributive lattice, 135
Bilinear space, 5 Domination of local rings, 62
Boolean algebra, 120
Boolean lattice, 120
Boolean space, 115, 125 E
Bounded (over a subring), 166 Extension of
orderings, 11
places (theorem of Chevalley), 80
C real places, 82
Cauchy index, 30 signatures, 11
Centre
of a place, 141
of a valuation ring, 141 F
Closed points Field
in Spec A, 103 ordered, 1
M
H Maximally
Hankel form, 32 real ideal, 111
Hankel matrix, 32 real reduced ideal, 111
Harrison Maximal points, 131, 136
subbasis, 114 Minimal points, 131
topology, 114
Hilbert
Nullstellensatz, 108 N
17th problem, 96 Nichtnegativstellensatz, 154
solution of 17th problem, 97 Nilradical, 101
Holomorphy ring of a field, 167 Nonreal ring, 110
Hyperbolic Nullstellensatz
form, 6 Hilbert’s, 108
plane, 5 real, 113, 154
weak, 104
weak real, 112
I
Ideal
P -convex (P an ordering), 140 O
T -convex (T a preordering), 159 Ordered abelian group, 53
X-convex (X a subspace of Sper A), 162 rank, 69
Indefinite, 95 Ordered field, 1
inequality of Łojasiewicz, 158 Ordering
Inertia theorem of Sylvester, 8 of a field, 1
Infinitely compatible with a valuation, 66
distant points, 137, 173 of a ring, 115
large, 54, 55 Order preserving
small, 54, 55 homomorphism, 54
Integral closure, 61 place, 80
Integral element, 60 Order toplogy, 73
Integrally closed ring, 61
Integral ring extension, 60
Intermediate value theorem, 20 P
Inverse spectral space, 126 P -convex ideal, 140
Isotropic quadratic form, 6 Pfister form, 147
Place (of a field), 78
associated with a valuation ring, 79
K canonical, 79
k-morphism (of affine k-varieties), 107 compatible with a preordering, 180
K-rational points, 107, 109 compatible with an ordering, 80
Index 205
Budan–Fourier, 41 coarsening, 65
Chevalley (extension of places), 80 compatible with an ordering, 66
Cohen–Seidenberg, 61 equivalent, 65
Dress, 175 non-trivial, 64
Hölder, 56 trivial, 64
Hörmander, 156 value group, 64
Hermite, 25, 29 Valuation ring, 58
Hurwitz (different ones), 34, 41 associated to a place, 79
Rolle, 20 associated to a valuation, 64
Sturm, 21 discrete of rank 1, 59
Sylvester, 25, 29 maximal ideal, 58
Sylvester (inertia theorem), 8 prime ideals, 59
von Staudt, 56 rank, 69
Thin semialgebraic set, 135 real closed, 142
Tilde (), 128 residually real, 58
Topological space value group, 65
Boolean, 125 Variety (abstract) associated to an affine
irreducible, 104 algebra, 108
quasi-compact, 104
spectral, 124
Topology W
associated to a valuation, 74 Well-behaved sublattice, 130
associated to an ordering, 73 Witt
strong, 73, 119 decomposition, 6
Zariski, 103, 109 equivalence, 6
Trace (linear form), 26 index, 6
Trace formula, 50 ring, 7
Transfer of a bilinear space, 48 prime ideals, 13
reduced, 146
Witt’s cancellation theorem, 6
U
Ultrafilter, 130
Ultrafilter theorem, 132 X
Ultrametric, 74 X-convex (X a subspace of Sper A), 162
Unbounded locus, 172
Universal quadratic form, 6
Z
Zariski, O.
V spectrum, 102
Valuation, 63 topology, 103, 109
canonically associated to a valuation ring,
65