A CONSTRUCTION FOR ABSOLUTE VALUES IN POLYNOMIAL RINGS. than define a second approximation V 0

Similar documents
Chapter 8. P-adic numbers. 8.1 Absolute values

be any ring homomorphism and let s S be any element of S. Then there is a unique ring homomorphism

Rings. Chapter 1. Definition 1.2. A commutative ring R is a ring in which multiplication is commutative. That is, ab = ba for all a, b R.

(1) V (ab) = V (a) + V (b), V (a + b) min ( V (a), V (b) ) A CONSTRUCTION FOR PRIME IDEALS AS ABSOLUTE VALUES OF AN ALGEBRAIC FIELD

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Polynomial Rings. i=0. i=0. n+m. i=0. k=0

A connection between number theory and linear algebra

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

NOTES ON DIOPHANTINE APPROXIMATION

where c R and the content of f is one. 1

NOTES ON FINITE FIELDS

Section III.6. Factorization in Polynomial Rings

MA257: INTRODUCTION TO NUMBER THEORY LECTURE NOTES

Algebraic function fields

1 Adeles over Q. 1.1 Absolute values

Chapter 3. Rings. The basic commutative rings in mathematics are the integers Z, the. Examples

Local Fields. Chapter Absolute Values and Discrete Valuations Definitions and Comments

MATH 431 PART 2: POLYNOMIAL RINGS AND FACTORIZATION

School of Mathematics and Statistics. MT5836 Galois Theory. Handout 0: Course Information

Factorization in Integral Domains II

Algebra Review 2. 1 Fields. A field is an extension of the concept of a group.

Finite Fields. Sophie Huczynska. Semester 2, Academic Year

MT5836 Galois Theory MRQ

MTH310 EXAM 2 REVIEW

Lecture 7: Polynomial rings

= 1 2x. x 2 a ) 0 (mod p n ), (x 2 + 2a + a2. x a ) 2

RINGS: SUMMARY OF MATERIAL

φ(xy) = (xy) n = x n y n = φ(x)φ(y)

U + V = (U V ) (V U), UV = U V.

RUDIMENTARY GALOIS THEORY

g(x) = 1 1 x = 1 + x + x2 + x 3 + is not a polynomial, since it doesn t have finite degree. g(x) is an example of a power series.

55 Separable Extensions

1 Rings 1 RINGS 1. Theorem 1.1 (Substitution Principle). Let ϕ : R R be a ring homomorphism

How many units can a commutative ring have?

A field F is a set of numbers that includes the two numbers 0 and 1 and satisfies the properties:

ALGEBRA PH.D. QUALIFYING EXAM September 27, 2008

Math 201C Homework. Edward Burkard. g 1 (u) v + f 2(u) g 2 (u) v2 + + f n(u) a 2,k u k v a 1,k u k v + k=0. k=0 d

Extension fields II. Sergei Silvestrov. Spring term 2011, Lecture 13

Finite Fields: An introduction through exercises Jonathan Buss Spring 2014

MATH 326: RINGS AND MODULES STEFAN GILLE

Places of Number Fields and Function Fields MATH 681, Spring 2018

Moreover this binary operation satisfies the following properties

Finite Fields. Mike Reiter

Mathematics Course 111: Algebra I Part I: Algebraic Structures, Sets and Permutations

(1) A frac = b : a, b A, b 0. We can define addition and multiplication of fractions as we normally would. a b + c d

15. Polynomial rings Definition-Lemma Let R be a ring and let x be an indeterminate.

GEOMETRIC CONSTRUCTIONS AND ALGEBRAIC FIELD EXTENSIONS

Standard forms for writing numbers

Groups, Rings, and Finite Fields. Andreas Klappenecker. September 12, 2002

GALOIS GROUPS OF CUBICS AND QUARTICS (NOT IN CHARACTERISTIC 2)

Factorization in Polynomial Rings

50 Algebraic Extensions

ALGEBRA. 1. Some elementary number theory 1.1. Primes and divisibility. We denote the collection of integers

Public-key Cryptography: Theory and Practice

Polynomials. Chapter 4

x = π m (a 0 + a 1 π + a 2 π ) where a i R, a 0 = 0, m Z.

Module MA3411: Abstract Algebra Galois Theory Michaelmas Term 2013

Computations/Applications

THROUGH THE FIELDS AND FAR AWAY

p-adic fields Chapter 7

Ph.D. Qualifying Examination in Algebra Department of Mathematics University of Louisville January 2018

Abstract Algebra: Chapters 16 and 17

Quadratic reciprocity (after Weil) 1. Standard set-up and Poisson summation

Introduction to finite fields

1. Algebra 1.5. Polynomial Rings

Solutions to Homework for M351 Algebra I

Mathematical Olympiad Training Polynomials

12 16 = (12)(16) = 0.

Fields and Galois Theory. Below are some results dealing with fields, up to and including the fundamental theorem of Galois theory.

TC10 / 3. Finite fields S. Xambó

CYCLOTOMIC POLYNOMIALS

Formal power series rings, inverse limits, and I-adic completions of rings

Chapter 5. Number Theory. 5.1 Base b representations

Formal Modules. Elliptic Modules Learning Seminar. Andrew O Desky. October 6, 2017

COMMUTATIVE PRESEMIFIELDS AND SEMIFIELDS

Lecture Notes Math 371: Algebra (Fall 2006) by Nathanael Leedom Ackerman

EXAMPLES OF PROOFS BY INDUCTION

Finite Fields. Sophie Huczynska (with changes by Max Neunhöffer) Semester 2, Academic Year 2012/13

Basic Algebra. Final Version, August, 2006 For Publication by Birkhäuser Boston Along with a Companion Volume Advanced Algebra In the Series

An Approach to Hensel s Lemma

Quadratic reciprocity (after Weil) 1. Standard set-up and Poisson summation

AN INTRODUCTION TO THE THEORY OF FIELD EXTENSIONS

Handout #6 INTRODUCTION TO ALGEBRAIC STRUCTURES: Prof. Moseley AN ALGEBRAIC FIELD

Polynomial Rings. (Last Updated: December 8, 2017)

Course 311: Hilary Term 2006 Part IV: Introduction to Galois Theory

Math 2070BC Term 2 Weeks 1 13 Lecture Notes

Rings. EE 387, Notes 7, Handout #10

Algebra Homework, Edition 2 9 September 2010

CYCLOTOMIC POLYNOMIALS

Homework 8 Solutions to Selected Problems

Analysis I. Classroom Notes. H.-D. Alber

Math 429/581 (Advanced) Group Theory. Summary of Definitions, Examples, and Theorems by Stefan Gille

Chapter 4 Finite Fields

Algebraic structures I

Outline. MSRI-UP 2009 Coding Theory Seminar, Week 2. The definition. Link to polynomials

2 ALGEBRA II. Contents

Generator Matrix. Theorem 6: If the generator polynomial g(x) of C has degree n-k then C is an [n,k]-cyclic code. If g(x) = a 0. a 1 a n k 1.

CHAPTER I. Rings. Definition A ring R is a set with two binary operations, addition + and

Definitions. Notations. Injective, Surjective and Bijective. Divides. Cartesian Product. Relations. Equivalence Relations

Course 311: Abstract Algebra Academic year

Transcription:

A CONSTRUCTION FOR ABSOLUTE VALUES IN POLYNOMIAL RINGS by SAUNDERS MacLANE 1. Introduction. An absolute value of a ring is a function b which has some of the formal properties of the ordinary absolute value. More explicitly, for any b in the ring, b must be a real number with the properties bc = b c, b + c b + c. If the second inequality holds also in the stronger sense b + c max b, c then the value b is called non-archimedean. The thus delimited nonarchimedean values are of considerable arithmetic interest. They are useful in questions of divisibility and irreducibility and in fact often correspond exactly to the prime ideals of the given ring. This paper is devoted to the explicit construction of non-archimedean values. More specifically, given all such values for the field R of rational numbers, we construct all possible values of the ring R[x] of all polynomials in x with coefficients in R. In treating a non-archimedean value it is convenient to replace a by a related exponential value V a = log A, with corresponding forms 2 of the formal properties of V. The possible nonarchimedean values of the field of rational numbers were determined by Ostrowsi 1917: For every prime p there is a p-adic exponential value V 0 in which the value of any rational number is obtained by writing the number as p α u/v, where u and v are prime to p, and setting 1 V 0 p α u/v = α, where is any fixed positive constant. This value we denote by the symbol [ V 0 p = ]. The only other value V is a trivial one, in which V a is zero for a 0. On this basis we can determine all possible values in the ring of polynomials with rational coefficients. Any such value W gives a p-adic or trivial value V 0 a = W a for the rational numbers and a value µ = W x for the variable x. These facts alone give a first approximation V 1 to the value W, as follows: V 1 an x n + a n 1 x n 1 + + a 0 2 = min { V 0 a n + nµ, V 0 a n 1 + n 1µ,..., V 0 a 0 }. This V 1 is actually a value and is never larger than W. If V 1 is not equal to W, we choose a φx of smallest possible degree for which W φx > V 1 φx. We than define a second approximation V 0 fx by using the true value for φx. In this manner we construct successive approximations V 1, V 2, V 3,... which in the limit will give the arbitrary value W 8. The succession of values V 1, V 2,... is defined in Part I for polynomials with coefficients in any field K. This requires a general method 4 and 5 of constructing a value V from a previously obtained value V 1. The value given by the limit of such a sequence needs a special study 7. Here, as in 8 and 16, we assume that every value of the field K is discrete ; that is, that the real numbers used as values form an isolated point set, as in the case of p-adic values. Part II investigates the structure of the values which have been constructed. The central problem is the construction of the residue-class field which arises when polynomials which differ by a polynomial of positive value are put into the same residue-class. For the absolute values constructed in Part I this field is determined by an inductive construction of the homomorphism of polynomials to residueclasses 10 14. This homomorphism also yields a more specific description of how our values can be built up 9, 13. Since a given value W can be represented in many ways by a sequence of approximations V 1, V 2, V 3,..., we treat in 15 and 16 the questions as to when two such sequences can give the same ultimate value W, and how such a sequence can be put in a normal form. Among the applications of this construction of absolute values we mention the classification of irreducibility criteria of the Newton Polygon type. The theorem of Eisenstein states that a polynomial fx = x n + a n 1 x n 1 + a n 2 x n 2 + + a 0 with integral coefficients a i is irreducible if each coefficient a i is divisible by some fixed prime p, while the last term a 0 is not divisible by p 2. In terms of the value V 1 of 2 with µ = 1/n these hypotheses on fx become V 1 fx = V1 x n = V 1 a 0 < V 1 a i x i i = 1,..., n 1. In this form a simple proof of the theorem can be given. The theorems of Königsberger 1895, Dumas 1906, and Ore 1928 are liewise related to the values V 1. The second stage values V 2 can be similarly applied to interpret the irreducibility theorems of Schöneman 1846, Bauer 1905, Kürschá 1923, and Ore 1923. By using the general value V one can obtain a still more extensive irreducibility criterion which includes all these previous theorems MacLane Similar second-stage values V 2 appear implicitly in the irreducibility investigations of Ore 1923, Kürschá 1923, and Rella 1927. Absolute Values in Polynomial Rings 1 30-Oct-2014, 10:00 am

1935, and which asserts that certain polynomials fx with irreducible homomorphic images of sufficiently high degree are themselves irreducible. Our construction for absolute values can also be applied to give a new and complete treatment of the problem of constructing the prime ideal factors of a given rational prime in a given algebraic field. I. The construction of non-archimedean values 2. Elementary properties of values in rings. A ring S is said to have a non-archimedean value for short, a value V if to every element a 0 in S there is assigned a unique real number V a with the properties V ab = V a + V b, V a + b min V a, V b. These we call the product and triangle laws respectively. We assume also that 0 is assigned the value +, with the following conventions for any finite number γ: γ <, + γ = γ + = + =. Two simple consequences of the product law are 1 V 1 = V 1 = 0, V a = V a. More important is the strengthened form of the triangle law: 2 V a V b implies V a + b = min V a, V b. For suppose instead that V a > V b and V a + b > min V a, V b. Then a contradiction. V b = V a + b a min V a + b, V a > V b, Since we are using a value analogous to the negative logarithm of the ordinary absolute value, a small absolute value will correspond to a large value V. Hence we say that two ring elements a and b are of the same order of magnitude or equivalent in V denoted a V b if and only if V a b > V a. The product and strong triangle laws show that equivalent elements have the same value and that equivalence is a reflexive, symmetric, and transitive relation, provided the supplementary assumption that 0 V 0 be made. Here and in the sequel ring means commutative ring with unit element. Here and subsequently the element 0 plays an exceptional role. Two equivalences a V b and c V d can be multiplied to give 3 ac V bd. An element b is equivalence-divisible in V by a denoted a V b if and only if there exists a c in S such that b V ca. If this is true, it remains true when a or b is replaced by an equivalent element. The product law implies that a ring S with a value V must be an integral domain. The value V may be extended to the quotient field of S by defining, in accord with the product law, a 4 V = V a V b b for any elements a and b 0 in S. One then obtains the Theorem 2.1. Let S be an integral domain with quotient field K. If V is a value of S, then the function defined by 4 is a value of K. Conversely, every value of K can be obtained in this way from one and only one value of S. When S = K is a field, the set of all real numbers V a for a 0 is an additive group Γ, called the value-group of V. If the positive numbers of Γ have a positive minimum δ > 0, then the value V is said to be discrete. In this case the group Γ is cyclic and consists of all multiples of δ. If all elements not 0 have the value 0, then V is called trivial. Every ring has a trivial value, while the p-adic values for the field of rational numbers are examples of discrete values. Values of arithmetic interest are generally discrete. 3. The first stage values. Our problem is this: Given all values of a field K; to construct all values for the ring K[x] of all polynomials in x with coefficients in K. By Theorem 2.1, this is equivalent to determining all values in the field Kx of rational functions of x with coefficients in K. No gain in generality would result were a ring S used instead of the field K. As indicated in the introduction, the values for K[x] will be constructed in stages. For the first step, tae any value V 0 for the field K and any real number µ, and then define a corresponding first stage value V 1 for any polynomial by the equation 2 of 1. In particular, this gives V 1 x = µ, V 1 a = V 0 a any a K. Hence we use the symbol [ V 0, V 1 x = µ ] for the value V 1. Henceforth all polynomials considered are to have coefficients in K, unless otherwise noted. Absolute Values in Polynomial Rings 2 30-Oct-2014, 10:00 am

Theorem 3.1. If V 0 is a value of K and µ a real number, the function [ V 0, V 1 x = µ ] defined above is a value of K[x]. A particularly simple V 1 arises when µ = 0. On the other hand, if V 0 is trivial and µ < 0, then V 1 ax = µ deg ax. The symbol deg ax here and in the sequel denotes the degree in x of the polynomial ax. 4. Augmented values. Our construction now proceeds to build a second stage value on the basis of a first stage one; or, more generally, a th stage value from one at the stage 1. The process involved can be formulated once for all: Given a value W for K[x]; to construct an augmented value V by assigning larger values to a certain ey polynomial φx and to its equivalence-multiples. The ey polynomial φx must be suitably chosen. Definition 4.1. A ey polynomial φx 0 over a value W of K[x] is one which satisfies the following conditions: i Irreducibility. If a product is equivalence-divisible in W by φx, then one of the factors is equivalence-divisible by φx. ii Minimal degree. Any non-zero polynomial equivalence-divisible in W by φx has a degree in x not less than the degree of φx. iii The leading coefficient of φx is 1. This ey polynomial is to be assigned a new value 1 V φx = µ > W φx. To find the new values of other polynomials, we use expansions in φ; that is, expressions in powers of φx of the form 2 fx = f m xφ m + f m 1 xφ m 1 + + f 0 x, in which each coefficient polynomial f i x is either zero or of degree less than the degree of φx. Any polynomial has one and only one such expansion, which may be found by successive division by powers of φ. The new value V fx is computed from the expansion thus: 3 V f m xφ m + f m 1 xφ m 1 + + f 0 x = min i W fi x + iµ. Here min with subscript i means the smallest quantity of the form W f i x + iµ, for i = 0, 1,..., m. This assumption, although unnecessary, will simplify the subsequent wor. We use φ as an abbreviation for φx, and similarly for other polynomials. Theorem 4.2. If W is a value of K[x], φx is a ey polynomial over W and µ is a real number satisfying 1, then the function V defined in 3 is also a value of K[x]. V is called an augmented value, and is denoted by V = [ W, V φ = µ ]. Proof. The product and triangle laws for V must be verified. We first prove the triangle law for a sum fx + gx. Let f and g have the expansions 2 and 4 gx = g n xφ n + g n 1 xφ n 1 + + g 0 x respectively. By adjoining zero coefficients we can mae m = n. Hence f + g has the expansion n fx + gx = [f i x + g i x] φ i. i=0 By the definition of V and the triangle law for W, V f + g = min i W fi + g i + iµ min i min W fi, W g i + iµ min i W fi + iµ, W g i + iµ = min min i W fi + iµ, min i W gi + iµ, V f + g min V f, V g. To prove the product law we will use the quotient-remainder expression for a polynomial fx, 5 fx = qx φ + rx, where rx is zero or of degree less than that of φx. Lemma 4.3. If φ is a ey polynomial over a value W of K[x], and if fx 0 has the quotient-remainder expression 5, then 6 7 W rx W fx, W qxφ W fx. The inequality in 6 holds if and only if φ W f. Proof of Lemma. Were the first conclusion 6 false, then W fx > W rx in 5 and the definition of equivalence would give rx W qx φ. Absolute Values in Polynomial Rings 3 30-Oct-2014, 10:00 am

Hence φ W r and rx 0, a contradiction to the minimal property of φ and the restricted degree of rx. The second conclusion 7 now follows from 6 by the triangle law. The third conclusion gives a test for equivalence-divisibility in terms of ordinary division. When W rx > W fx, then 5 shows φ W f. Conversely, if φ W f then there exist polynomials hx and sx so that fx = hx φ + sx, W sx > W fx. If now the equality sign in 6 should hold, we would have with rx = fx qx φ = hx qx φ + sx, so that φ W r, again a contradiction. W sx > W fx = W rx, Return to Theorem 4.2 and consider the product law first for a product of two monomial expansions ax φ t and bx φ u. Because of the limited degrees of ax and bx, the product axbx has an expansion with not more than two terms, 8 ax bx = cx φ + dx. Were it the case that φ W ab then the equivalence-irreducibility of φ Definition 4.1 would require that either φ W a or φ W b, contrary to the minimal property. Hence φ W ab. Lemma 4.3 and the triangle axiom then yield W cx φ W ax bx = W dx. Since the new value of φ exceeds the old value, 9 W cx + µ > W ax bx = W dx. The product under consideration has by 8 the expansion ax φ t bx φ u = cx φ t+u+1 + dx φ t+u ; hence the definition of V and the conclusion 9 give V [ ax φ t bx φ u] = min { W cx + µ + t + uµ, W dx + t + uµ } = W dx + t + uµ = W ax + tµ + W bx + uµ = V ax φ t + V bx φ u. This is the product law for monomial expansions. The product law for polynomials fx and gx with arbitrary expansions 2 and 4 respectively is an immediate consequence. The product fx gx has an expansion obtained by adding expansions of monomial products; hence 10 V fx gx V fx + V gx. To show that the equality holds, choose t and u as the largest integers with V f t x φ t = V fx, V g u x φ u = V gx respectively. The monomial case then shows that the expansion of fx gx has a term rx φ t+u with the value V f + V g. The equality holds in 10, and Theorem 4.2 is established. 5. Properties of augmented values. An augmented value V is never less than the original value W. This characteristic property will now be established. As a consequence the method used to compute V can be extended Theorem 5.2 in a way subsequently useful in 12. Theorem 5.1 Monotonicity. The augmented value V = [ W, V φ = µ ] maes V fx W fx for all polynomials fx 0. The inequality sign holds if and only if φ W f. In particular, the equality sign holds whenever the degree of φx exceeds that of fx. Proof. The proof is by induction on the degree m of the expansion of fx in φ see 4, 2. If m = 0, the definition of V shows V fx and W fx equal. If m > 0, the quotient-remainder expression 1 fx = qx φ + rx indicates that qx has an expansion of degree m 1 in φ; hence the induction assumption will be V qx W qx. The value of the first term on the right of 1, by 4, 1, and the quotientremainder Lemma 4.3, is V qx φ W qx + V φ > W qx φ W fx. The details here omitted are given in Rella s proof. Absolute Values in Polynomial Rings 4 30-Oct-2014, 10:00 am

For the second term, the case m = 0 and Lemma 4.3 imply V rx = W rx W fx, where the inequality holds if and only if φ W f. The strong triangle law for V applied to 1 now gives the result see 2, 2. Theorem 5.2. If in the expression ax = a n x φ n + a n 1 x φ n 1 + + a 0 x the degrees of the a i x are not limited, but φ W a i for i = 0,..., n and a i 0, then the augmented value V = [ W, V φ = µ ] is V ax = min i [ W ai + iµ ] i = 0,..., n. Proof. A quotient-remainder expression for each coefficient polynomial gives a i x = q i x φ + r i x ax = n i=0 q ix φ i+1 + n i=0 r ix φ i. Lemma 4.3 shows that the second summation has a value V i r i φ i = min i [ W ri + iµ ] = min i [ W ai + iµ ] and that the first summation has a larger value i = 0,..., n, V i q i φ i+1 min i [ V qi + V φ + iµ ] > min i [ W qi + W φ + iµ ], because of the monotonicity. The strong triangle law for the sum of these two summations yields the desired conclusion. 6. Inductive and limit-values. This section classifies the values and valuegroups obtained by successive augmented values. Definition 6.1. A th stage inductive value V is any value of K[x] obtained by a sequence of values V 1, V 2,..., V, where V 1 = [ V 0, V 1 x = µ 1 ] is a first stage value 3 and where each V i is obtained by augmenting V i 1 : V i = [ V i 1, V i φ i = µ i ] Furthermore, for i = 2,...,, the ey polynomials φ i x must satisfy: 1 2 deg φ i x deg φ i 1 x; φ i x Vi 1 φ i 1 x. Here the first ey polynomial is understood to be φ 1 x = x. i = 2, 3,...,. These conditions involve no loss of generality, but simplify subsequent proofs see Theorem 6.7 and the end of 9. The value V may be conveniently symbolized thus: 3 V = [ V 0, V 1 φ1 = µ1, V 2 φ2 = µ2, V 3 φ3 = µ3,..., V φ = µ ]. Given an infinite sequence V 1, V 2,..., V,... of such values, we set 4 V fx = lim V fx. The monotonic character of V indicates that this limit, if not finite, is +. V satisfies the product law for values, as can be shown by taing limits in the product law for V. As for the sum fx + gx, note that the triangle law in V indicates that one of the inequalities V fx + gx V fx, V fx + gx V gx holds for infintely many. One of the conclusions V fx + gx V fx, V fx + gx V gx then results, and thence follows the triangle law for V. We have Theorem 6.2. Let {φ x} and {µ } be fixed infinite sequences such that all the functions V indicated in 3 are inductive values. Then the function V fx defined in 4 is a value of K[x], provided some polynomials not zero be allowed to have the value +. This function V will be called a limit-value. The case when several successive ey polynomials have the same degree will often require separate treatment, based on Lemma 6.3. If in the inductive value V in 3 the ey polynomials φ t+1 x, φ t+2 x,..., φ x all have the same degree, for t with 0 t 1, then i V t φ j+1 φ j = µ j j = t + 1, t + 2,..., 1, ii µ > µ 1 > > µ t+1, iii V t φ = V t φ 1 = = V t φ t+1 if t > 0. Proof. Let j range from t + 1 to 1, and set 5 s j x = φ j+1 x φ j x. Since both φ s have the first coefficient 1, the degree of s j x is less than that of φ j x. Therefore, by Theorem 5.1, V t sj x = V t+1 sj x = = V sj x. Absolute Values in Polynomial Rings 5 30-Oct-2014, 10:00 am

If the first conclusion were false for some j, we would have V t sj x = V j φ j+1 φ j > µ j = V j φ j, for the other inequality is impossible by Lemma 4.3. This would give φ j+1 Vj φ j, a contradiction of assumption 2. The conclusion i is thus established. Coupled with the monotonicity and the triangle axiom for 5, it gives the second conclusion, for µ j+1 = V j+1 φj+1 > Vj φj+1 min { Vj φ j, V t s j } = µ j. For similar reasons, assuming now that t > 0, V t sj x = µ j = V j φ j > V j 1 φ j V t φ j. The strong triangle axiom for V t in 5 then yields conclusion iii, V t φ j+1 = min { V t sj x, V t φ j } = V t φ j. An interesting consequence of this lemma is the invariance of the values assigned to the ey polynomials. Theorem 6.4. If the ith stage of the inductive value V in 3 uses a ey polynomial φ i with an assigned value µ i, then V φi x = V i φi x = µ i. For this conclusion follows directly from Theorem 5.1 if the degree of φ i+1 x, and hence that of every subsequent ey polynomial, exceeds the degree of φ i x. The only case remaining is that of Lemma 6.3, with t = i 1. But, by 5, φ i = φ s 1 x s 2 x s i x. The terms on the right have by the preceding lemma the V values µ, µ 1,..., µ i respectively, so that the conclusion follows by the strong triangle law. Both this theorem and Lemma 6.3 hold equally well for limit-values. The monotonic property of inductive values can be sharpened thus: Theorem 6.5. Let a limit or inductive value be built up by the inductive values V 1, V 2,.... Then, for any fixed polynomial fx 0, either V +1 fx > V fx = 1, 2,..., or else there is an i 1 such that V 1 fx < V2 fx < < Vi 1 fx = V i fx = Vi+1 fx = Vi+2 fx =. In the latter case there is an rx of degree less than that of φ i+1 with fx V rx = i + 1, i + 2,.... Suppose, contrary to the first alternative, that for some i Then the quotient-remainder expression V i+1 fx = Vi fx. fx = qx φ i+1 + rx must by Theorem 5.1 and Lemma 4.3 have V i r = V i f. Hence, for any i+1, V f r V i+1 f r = V i+1 qφ i+1 > V i qφ i+1 V i f = V i r = V r. Therefore fx V rx and V f = V r = V i r = V i f, so that V i fx is constant for i, which is the second alternative. An inductive value V of K[x] gives by Theorem 2.1 a value for the field Kx of rational functions. This value has by 2 a value-group Γ, which we call the value-group associated with V. It may be determined in the following way: Theorem 6.6. The value V in 3 has a value-group Γ consisting of all real numbers of the form ν + m 1 µ 1 + m 2 µ 2 + + m µ, where the m i are integers and ν is an element of the value-group of the original value V 0. That every number of Γ must be of this form follows by induction from the definition of the augmented value V. Conversely, any number of this form is by Theorem 6.4 the value in V of the rational function b x m1 φ m2 2 φ m, where b is a constant in K with the value ν. For a more precise description, designate a real number µ as commensurable with an additive group of numbers whenever some integral multiple of µ lies in the group. Then Absolute Values in Polynomial Rings 6 30-Oct-2014, 10:00 am

Theorem 6.7. In an inductive value V from 3 every assigned value µ i, except perhaps µ, is commensurable with the value-group Γ i 1 of the preceding value the case i = 1 included. Proof. Consider the expansion in φ i of the next ey, φ i+1 x = f m x φ m i + f m 1 x φ m 1 i + + f 0 x. If µ i is not commensurable with Γ i 1, no two terms here can have the same value in V i. Only one term, say the jth, has the minimum value, and φ i+1 x Vi f j x φ j i. By the irreducibility of φ i+1 at least one of the conditions 6 7 φ i+1 Vi f j x, φ i+1 Vi φ i, must hold. Because of the minimal property of φ i+1 the first possibility 6 contradicts the assumption 1 of Definition 6.1. For the same reasons the second possibility 7 implies that φ i+1 and φ i have the same degree, while sx = φ i+1 x φ i x has a smaller degree. Because of 7, Lemma 4.3 applied to V i and the ey polynomial φ i+1 shows V i sx > Vi φi+1 x. Hence φ i x Vi φ i+1 x, a contradiction of assumption 2. There can be no next ey φ i+1. 7. Constant degree limit-values. A limit-value V for polynomials does not give a value for all rational functions if some of the polynomials have the value +. Hence the problem: When is V finite; that is, when is V fx finite for all fx 0? We obtain an answer in the discrete case. If the ey polynomials φ x increase indefinitely in degree, then V fx is by Theorem 5.1 ultimately constant for fixed fx and V is finite. A different situation arises if the degrees of φ x are then all equal to some M for sufficiently large. For an example of such a constant degree limit-value, start with the p-adic value [ V 0 3 = 1 ] for the rational field see 1, 1 and set V 1 = [ V 0, V 1 x = 1 ], V = [ V 1, V x + 2p + p 2 + p 3 + + p 1 = ] = 2, 3,.... This gives a limit-value of constant degree 1. Since p 2 = 2p + p2 + p 3 + + p 1 p 2 holds by the usual methods for p-adic numbers, we find > 1, p = 3 V x + p = lim 2 V x + 2p + p 2 + p 3 + + p 1 p = lim 2 =. Hence this V is not finite. This use of p-adic numbers suggests the general notion of a perfect ring. In any ring S with a value V, a sequence {a n } is a Cauchy sequence if V a n a m approaches with n and m. If every Cauchy sequence has a V -limit b such that V a n b approaches with n, the ring S is said to be perfect. Any ring can be embedded in a perfect ring by the usual procedure of adjoining limits of Cauchy sequences. Theorem 7.1 Finiteness criterion. Let V be a limit-value with ey polynomials φ x of constant degree M for > t > 0. Extend the ring K[x] with the value V t to be a perfect ring S. Assume that all values of K are discrete. Then {φ } is a Cauchy sequence in V i and has a limit φ in S. Furthermore V is finite if and only if there is no gx 0 in K[x] divisible in S by the limit φ. For V the symbolism of Theorem 6.2 may be used. Since φ t+1, φ t+2,... all have the same degree M, the conclusions of Lemma 6.3 on constant degree values are applicable. Each number µ i is by Theorem 6.7 commensurable with the valuegroup Γ i 1 of V i 1. Our assumption shows the original value-group Γ 0 of V 0 to be discrete, hence, by Theorem 6.6 and by induction, the group Γ t is discrete. But Lemma 6.3 gives 1 µ i = V i φ i+1 φ i Γ i i > t; hence Γ i = Γ t for i > t. This lemma also shows the sequence {µ i } to be monotone increasing for i > t; it lies in the discrete set Γ t, hence 2 lim i µ i =. The strong triangle law combined with 1 then proves V t φ i+j φ i = V t i+j 1 =i φ +1 φ = min {µ } = µ i. Therefore, by 2, {φ i } is a Cauchy sequence with a limit φ in S. This φ need not be a polynomial, but, by conclusion iii of Lemma 6.3, φ 0. Absolute Values in Polynomial Rings 7 30-Oct-2014, 10:00 am

Now consider the necessary condition for finiteness. If gx 0 is divisible by φ in S, then gx = h φ, where h is the V i -limit of a Cauchy sequence {h i x} from K[x]. argument for the convergence of a product shows 3 lim i V t gx hi x φ i x =. By the triangle axiom and the monotonic property for i > t, The usual 4 V i g min { V i h i φ i, V i g h i φ i } min { V t h i + µ i, V t g h i φ i }. But {h i x} is a convergent sequence in V i with a limit not zero, so that, as is well nown, V t h i is ultimately constant. Consequently 2, 3, and 4 prove 5 V gx = lim i V i gx =, so that V is not a finite limit-value. Conversely, suppose that V is not finite. Then 5 holds for some gx 0. If gx has the quotient-remainder expressions q i xφ i +r i x, then, by Theorem 5.1 and by Lemma 4.3, V i g q i φ i = V i r i = V i 1 r i V i 1 gx i > t. Thus the sequence {q i φ i } converges in V i to the limit gx 0. Since {φ i } already converges to the limit φ 0, the standard argument for the limit of a quotient q i φ i /φ i shows that {q i } must converge in V i to some limit q in S, such that fx = qφ. Hence φ is a factor of fx in S, as asserted. 8. Completeness. We have the following theorem. Theorem 8.1. If every value of the field K is discrete, then every non-archimedean value W of the ring K[x] can be represented either as an inductive or as a limit-value. Given W, we shall construct by stages a corresponding inductive value V with the following three properties notation as in 6, 3: W fx 1 V fx for all fx, deg fx < deg φ implies W fx 2 = V fx, W φix 3 = V φi x = µ i i = 1, 2,...,. The initial value V 1 is defined by µ 1 = W x, V 0 a = W a any a K; the triangle axiom for W and the definition of V 1 in 1, 2, then show that conditions 1, 2, and 3 hold for = 1. Suppose now that an inductive value V with these three properties has already been constructed, and that the equality in 1 does not always hold. As a prospective ey polynomial, choose a ψx of smallest possible degree with the property 4 W ψx > V ψx. Multiplication with some constant gives ψx the first coefficient 1. Furthermore the two statements W fx 5 > V fx, 6 ψx fx, V are logically equivalent. For if 5 is given, and if fx has the quotient-remainder expression qxψ + rx, then V qψ f = W qψ f min { W qψ, W f } > min { V qψ, V f }, because of 2, the minimum degree choice of ψ and the induction assumption 1 for qx. Hence the strong triangle law shows f V qψ, which is the conclusion 6. Conversely, if 6 holds there exist polynomials hx and sx with fx = hxψ + sx, V sx > V fx = V hxψ. Then, because of the induction assumption 1, W f min { W hψ, W s } min { V h + W ψ, V s } > V h + V ψ = V f, which gives conclusion 5. The equivalence of 5 and 6 is established. From the equivalence one readily shows that ψx satisfies the Definition 4.1 of a ey polynomial over the value V. Finally we can assign ψx = φ +1 the new value 7 µ +1 = W ψ > V ψ, satisfying the proper inequality, and then construct the augmented value V +1 = [ V, V +1 φ +1 = µ +1 ]. This will be an inductive value if only conditions 1 Absolute Values in Polynomial Rings 8 30-Oct-2014, 10:00 am

and 2 of Definition 6.1 hold. By the choice of φ +1 = ψ and the induction assumption 2, φ x cannot exceed φ +1 x in degree, therefore condition 1 of 6 is true. Condition 2 of 6 could only be false if φ +1 V φ ; in other words, only if V φ φ +1 > V φ = V φ +1. By 2, 3, and the choice of ψ in 4 this would entail W φ min { W φ +1, W φ φ +1 } > min { V φ +1, V φ } = V φ = W φ, a contradiction which establishes the desired condition. The inductive value V +1 thus constructed satisfies the analogues of the desired conditions 1, 2, and 3. The latter two are consequences of the definitions in 4 and 7, while 1 follows from the definition see 4, 3 of the augmented value V +1 by the triangle axiom for W : W m i=0 f ixψ i min i W fi x + iµ +1 = V+1 m i=0 f ixψ i. The inductive construction of the value V associated with W is complete. This process either will ultimately yield an inductive value V equal to W or will give an infinite sequence of inductive values with a limit-value V such that W fx V fx = lim V fx for all fx. In the discrete case the first inequality sign never occurs. For suppose instead that it did hold for some fx; then since {V f} is monotone non-decreasing, W fx > V fx = 1, 2,.... The equivalence of 5 and 6 then implies that φ +1 x V fx. Hence the monotonicity Theorem 5.1 shows V +1 fx > V fx = 1, 2,.... This cannot hold if the degrees of the ey polynomials φ x increase indefinitely, so that we have the case where φ x has the fixed degree M for > t, as in Theorem 7.1. The monotonic increasing sequence {V fx } consists of numbers all from the discrete group Γ t, with the result W fx V fx = lim V fx =. This can occur only for fx = 0, a trivial case. Accordingly, W = V, and the completeness theorem is established. II. The structure of inductive values 9. Properties of ey polynomials. To apply the preceding construction of values to any particular case it is necessary to now what polynomials can be used as ey polynomials. This question is not constructively answered by the definition in 4. Part of this question will be answered at once Theorem 9.4; the rest after the structure of the inductive values V has been more explicitly formulated. We first show that certain polynomials act lie equivalence-units : Lemma 9.1. If V is an inductive value with > 1, then for every polynomial bx with V bx = V 1 bx there is a polynomial b x with 1 b x bx V 1, V b x = V 1 b x. The hypothesis on bx implies that bx is not divisible by the last ey polynomial φ x. Since φ is certainly irreducible in the ordinary sense, there are polynomials b x and cx with b x bx + cx φ x = 1, deg b x < deg φ x. By Theorem 5.1, V b = V 1 b. The transition from V 1 to V increases the value of cφ, but leaves unchanged the values of b b and 1 in this equation. Hence b b V 1, as in 1. Lemma 9.2. In any inductive V, the last ey polynomial φ is equivalenceirreducible in V ; a polynomial gx with φ V gx has a value V g in Γ 1. Proof. If a polynomial fx has the expansion 2 fx = f n x φ n + f n 1 x φ n 1 + + f 0 x, deg f i x < deg φ x, then φ V fx if and only if V f 0 > V f. For if V f 0 > V f, then f f 0 is a polynomial equivalent to f with a factor φ. Conversely, if f V hxφ, then the last term f 0 of the expansion for f is obtained from f hφ, where V f hφ > V f, so that V f 0 > V f. In particular, if φ V f then V f = V f 0 = V 1 f 0 Γ 1, as asserted. This criterion shows φ equivalence-irreducible in V. For suppose instead that φ V f g, although neither factor is so divisible. Then the criterion gives V f 0 = V f, V g 0 = V g, where g 0 x is the last term in the expansion for g. The last term in the expansion for f g is the remainder r 0 x obtained by dividing f 0 g 0 by φ ; but since φ V 1 f 0 g 0, Lemma 4.3 shows V r 0 = V 1 r 0 = V 1 f 0 g 0 = V f 0 + V g 0 = V f g. This means that φ V f g, a contradiction proving the lemma. Absolute Values in Polynomial Rings 9 30-Oct-2014, 10:00 am

An inductive value V will be called commensurable if the value µ assigned the last ey polynomial is commensurable with the previous value-group Γ 1 cf. Theorem 6.7. There is then a smallest positive integer τ such that τ µ is in Γ 1. For each t there is a similar τ t : 3 τ t is the smallest integer such that τ t µ t Γ t 1. We will subsequently need polynomials with any given values: Lemma 9.3. If V is a commensurable inductive value, then for any real number λ in the value-group Γ of V there is a polynomial R λ = R λ x with value λ in V and in every value V +1 obtained by augmenting V. Proof. As in Theorem 6.6, λ has the form λ = ν + m 1 µ 1 + m 2 µ 2 + + m µ, ν Γ 0. Each integer m may be made non-negative by adding to m i µ i and subtracting from ν a sufficiently large term qµ i, so chosen that qµ i Γ 0 e.g., choose q 0 mod τ 1, τ 2,..., τ i. If then a is a constant of value ν, R λ = R λ x = a x m1 φ m2 2 φ m3 3 φ m, V R λ = λ, m 0, is the required polynomial. Theorem 6.4. In any augmented value V +1, R λ has value λ by Theorem 9.4. A polynomial fx is a ey polynomial for an inductive value over V if and only if the following conditions hold: i the expansion 2 has a last term with V f = V f 0 ; ii the expansion has a first term f n xφ n with f nx = 1, V φ n = V f, and n 0 mod τ ; iii fx is equivalence-irreducible in V. Proof. Condition i means, as in the proof of Lemma 9.2, that φ x V fx. Assume first that fx is a ey. Condition iii is necessary by definition. Were i false, then V f < V f 0, so that f V f f 0, while 2 shows f f 0 = qxφ for a qx of degree less than fx. Thus f V qφ. Since f is a ey, this leads to a contradiction much as in the proof of Theorem 6.7. The assumption V f 0 V f is false. Since fx is minimal Definition 4.1 it has no equivalence-multiples of degree less than itself. Hence f n x is a constant in K, for otherwise > 1, and Lemma 9.1 supplies a b x with b xf n x V 1. The product b xfx formed from 2 and modified by replacing the first coefficient by 1 and by reducing the other coefficients modulo φ is then an equivalence-multiple of fx. Its degree is n deg φ, and is less than that of fx unless f n x K. As the leading coefficient of f must be 1, f n x = 1 follows, as in ii. Certainly V φ n = V f is necessary, for otherwise f φ n is an equivalent polynomial of smaller degree. Thus V φ n = V f = V f 0 = V 1 f 0 Γ 1, so that n 0 mod τ by 3. This establishes the necessity of ii. Conversely, if fx satisfies i, ii, and iii, it has first coefficient 1 and is minimal, because any equivalence-multiple of fx must be of degree at least n in φ cf. the proof of Theorem 4.2. The remaining restrictions of Definition 6.1 are readily verified, so that fx is in fact a ey polynomial. 10. Residue-class fields. The structure of a ring S with a value V involves the corresponding value-ring S +, which consists of all elements a of S with V a 0 these elements are the so-called integers of S. A congruence for integers can be defined thus 1 a b mod V if and only if V a b > 0. All elements of S + congruent to a given b form a residue-class; these classes together yield as usual the residue-class ring of V in S. This ring can also be considered as the residue-class ring S + /P, where P, the set of all elements of S + with positive value, is a prime ideal in S +. If S is a field, then S + /P is also a field, the residue-class field of V in S. The structure of V depends essentially on this residue-class field. For the p-adic value V 0 of the rational numbers see 1, 1 this field is simply the field of integers modulo p. Our problem is the determination of the residue-class field for any discrete inductive value. If the residue-class of each integer a be denoted by [a ], then H : a [a ] is a homomorphism of S + to the residue-class ring = S + /P, so that H has the following properties: I. H is a many-one correspondence between S + and ; II. H leaves sums and products unchanged; i.e., for V a 0 and V b 0: 2 [a + b ] = [a ] + [b ]; [ab ] = [a ][[b ]. III. If V a 0, then [a ] = [0]] if and only if V a > 0. By II, the last condition means that H carries congruent elements and only congruent elements into the same residue-class. Absolute Values in Polynomial Rings 10 30-Oct-2014, 10:00 am

For an inductive value V we denote the residue-class rings thus, for t = 1, 2,..., : 3 4 5 Λ t is the residue-class field of V t in Kx; H t is the homomorphism from Kx + to Λ t ; t is the residue-class ring of V t in K[x]. But fx and gx are congruent as polynomials mod V t if and only if they are congruent as rational functions mod V t. Hence each residue-class of t is contained in a residue-class of Λ t, and no two residue-classes of t are contained in the same class of Λ t. Addition and multiplication of classes are defined as addition and multiplication on elements in the classes, and hence are the same in t as in Λ t. Therefore t is isomorphic to a subring of Λ t. Since isomorphism does not alter the structure of a ring, we will replace t henceforth by the isomorphic subring of Λ t. Then the H t of 4 is also the homomorphism from K[x] + to t. The correspondence H t for rational functions is usually determined by the H t for polynomials. For if fx/gx 0 is a rational function with non-negative value and if V t is commensurable, there is by Lemma 9.3 a polynomial g t x with V t g t = Vt g, and by 2 6 [ ] [ ] fx g tf = = [g t f ]t. gx t g t g t [g t g ]t Both [g t f ]t and [g t g ]t are residue-classes of polynomials, while [[g t g ]t [0 ] t by Property III. We have proved Lemma 10.1. For a commensurable V, the residue-class field Λ of Kx is the quotient-field of the residue-class ring of K[x]. Theorem 10.2. For a commensurable first stage inductive value V 1 = [ V 0, V 1 x = µ 1 ], the residue-class ring 1 is isomorphic to the ring F 0 [y] of all polynomials in a variable y with coefficients in F 0, the residue-class field of the value V 0 for K. Proof. There is given a homomorphism H 0 from the ring K + of all V 0 -integers b in K to the residue-class field F 0. Each residue-class [b ] 1 of 1 contains the residue-class [b ] 0 of F 0, and this correspondence [b ] 1 [b ] 0 is an isomorphism between F 0 and the set of those classes of 1 containing elements of K. We will identify F 0 with this isomorphic subfield of 1 ; then 1 is an extension of F 0 and [b ] 1 = [b ] 0 for all b in K +. Any monomial bx n of value zero has V 0 b = nv 1 x = nµ 1, so that the exponent n is a multiple of the integer τ 1, defined in 9, 3. Any fx with V 1 f 0 thus has the form fx b m x mτ1 + b m 1 x m 1τ1 + + b 1 x τ1 + b 0 mod V 1 after terms of positive value are omitted. If e is a constant in K of value V 0 e = τ 1 µ 1, each term b j x jτ1 may be rewritten as a product b j e j e j x jτ1 of two factors of value 0. The application of the homomorphism H 1 then yields 7 [fx ] 1 = m j=0 [b j e j ] 0 y j ; y = [e 1 x τ1 ] 1. With y so defined, any [f ] 1 in 1 becomes a polynomial in y with coefficients [b j e j ] 0 in F 0, so that the residue-class ring 1 is contained in F 0 [y]. Since y 1 and 1 is a ring, 1 = F 0 [y]. The element y is transcendental i.e., a variable over F 0, for otherwise it would satisfy an algebraic relationship αy = [0 ] 1, where αy = α m y m + α m 1 y m 1 + + α 0 ; α m [0]] 1, α j F 0. Then the residue-class αy contains the polynomial fx = a m e m x mτ1 + a m 1 e m+1 x m 1τ1 + + a 0, where each a j is a constant with [a j ] 0 = α j. Then V 1 f 0 and [[f ] 1 = αy = [[0 ] 1, so that, by Property III of H 1, V 1 f > 0. But α m [0 ] 1, so that V 0 a m = 0 and V 1 a m e m x mτ1 = V 1 f = 0, a contradiction. The theorem is established. We note also that 7 enables us to calculate the residue-class of any given fx. 11. Conditions for equivalence-irreducibility. If φ +1 is a ey polynomial over a value V then φ V φ +1 Theorem 9.4, condition i. For any fx with this property, questions of equivalence-divisibility can be handled as follows: Lemma 11.1. In a commensurable V let fx be a polynomial with φ V f, and choose a polynomial f x so that V 1 f = V f = V f. Then a polynomial gx with V g = 0 satisfies f V g if and only if [g ] is divisible by [f f ] in the residue-class ring. By Lemma 9.2, V f is in Γ 1, so Lemma 9.3 yields the f desired, and [f f ] [0 ]. Suppose first that [g ] which is not [0 ] by Property III of H is divisible by [f f ]. Then [g ] = α [f f ] for some residue-class α = [hx ] [0]] in, and [g ] = α [f f ] = [h ] [f f ] = [hf f ]. Absolute Values in Polynomial Rings 11 30-Oct-2014, 10:00 am

Thus g and hf f have the same residue-class, V g hf f > 0 = V g, so g V hf f and f V g, as asserted. Conversely, if f V g, then g V hf V hf f f, where f x is a polynomial chosen as in Lemma 9.1 so that f f V 1. But f f, g, and hence hf have value 0, so that g V hf f f; [g ] = [hf ] [f f ], which shows [g ] divisible by [f f ]. Lemma 11.2. For f and f as in Lemma 11.1, fx is equivalence-irreducible in V if and only if every product in divisible by [f f ] has a factor divisible by [f f ] in. Proof. Suppose first that f is equivalence-irreducible and that [[g ] [h ] = [gh ] is a multiple of [f f ]. As we can assume V g = V h = 0, the previous lemma shows the product gh equivalence-divisible by the equivalence-irreducible f, so that one of the factors is so divisible. By Lemma 11.1 this means that [[g ] or [[h ] is a multiple of [[f f ], as asserted in the lemma. Conversely, suppose that every product [g ] [h ] divisible by [f f ] has a factor so divisible, and consider a product ax bx with f V ab, so that ab V cxf for some c. Write ax V gxφ d and bx V hxφ e, where the powers d and e are chosen so large that φ V g and φ V h. Then V g and V h are by Lemma 9.2 in Γ 1, so that there are polynomials g x and h x with V g g = V h h = 0. Then g h a b V g g h h φ d+e V g h c f. But φ V f, while φ is equivalence-irreducible Lemma 9.2, so φ d+e V g h c. Removal of the factor φ d+e gives f V g g h h, so that as in the previous lemma [g g ] [h h ] is divisible by [f f ]. One of the factors, say [g g ], is then divisible by [f f ], and Lemma 11.1 f V g g. But a V g g g φ d, where g is chosen so that g g V 1. Hence f V ax, and f is equivalenceirreducible. That is, the principal ideal [[f f ]] is a prime ideal in. 12. Residue-class rings for commensurable values. We have Theorem 12.1. If V is a commensurable inductive value of K[x], given as in 6, 3, and if the original value V 0 of K has a residue-class field F 0, then there is a sequence of fields F 1 = F 0, F 2, F 3,..., F, each an algebraic extension of the preceding, such that for any t = 1,..., the V t -residue-class ring of K[x] is isomorphic to the ring F t [y] of polynomials in a variable y with coefficients in F t. For t > 1 the degree m t of F t is determined by cf. 9, 3 m t τ t 1 deg φ t 1 = deg φ t ; m t = deg[ F t : F t 1 ]. By Lemma 10.1 we can then conclude at once Corollary 12.2. F t y is the V t -residue-class field of Kx. Proof of Theorem. The case t = 1 of this theorem is nown Theorem 10.2; hence we use induction, and assume the theorem true for V t. It is convenient to omit the subscript t + 1 but not the subscript t and to write V, φ, H, τ, etc., for V t+1, φ t+1, H t+1, τ t+1, etc. By the monotonic character of V Theorem 5.1 polynomials fx and gx with V t f 0 and V t g 0 are congruent mod V t only if they are congruent mod V. Each residue-class mod V t is thus contained in a residue-class mod V, and this gives a homomorphism between t = F t [y] and a subring F of the residue-class ring cf. 10, 5, where F = F t+1 is composed of all residue-classes mod V containing an fx with V t f 0. Polynomials f and g incongruent mod V t become congruent mod V if and only if φ t+1 Vt f g Theorem 5.1. This means that [f ] t [g ] t is divisible by the polynomial 1 ψ t+1 y = ψy = [φ t φ ] t ; Vt 1 φ t = V t φ t = V t φ, φ t = φ t t+1 x, constructed as in Lemma 11.1. Since not all polynomials are equivalence-divisible by φ in V t, ψy is not a constant in F t, while Lemma 11.2 shows ψy an irreducible polynomial in F t [y]. In the above homomorphism between F t [y] and F the mutiples of ψy in F t [y] are the elements corresponding to [0]], so that F is isomorphic to the ring of polynomials F t [y] modulo ψy, or, alternatively, to the field obtained by adjoining to F t a root θ of ψy. We identify F with this isomorphic field: 2 F = F t+1 = F t θ; ψθ = 0 θ = θ t+1. Then the residue-class [f ] t, when reduced modulo ψy, will be identical to the residue-class [[f ]; that is, 3 V t fx 0 implies [f ] = [f ]t y=θ. Absolute Values in Polynomial Rings 12 30-Oct-2014, 10:00 am

A monomial expansion axφ n of value 0 must have n a multiple of τ cf. 9, 3. Hence any fx with V f 0 has the form 4 fx f m xφ mτ + f m 1 xφ m 1τ + + f 0 x mod V, deg f i < deg φ. Since V φ τ Γ t, there are by Lemmas 9.1 and 9.3 polynomials φ τ x and φ τ x such that 5 V t φ τ = V φ τ = V φ τ, φ τ φ τ 1 mod V, V t φ τ = V φ τ = V φ τ. The terms f i φ iτ in 4 can be rewritten as products f j φ τ j φ jτ φ τ j, where V t f j φ τ j 0 and V φ jτ φ τ j = 0, the former because V f 0. The application of H, with 3, then proves 6 [fx ] = m j=0 [f j x φ τ j ] t y=θ y j 1 ; y 1 = [φ τ φ τ 1 ]. This shows that every [[f ] in the residue-class ring is also in F [y 1 ], while by 5, y 1 = [φ τ φ τ 1 ] = [φ τ φ τ ] is a residue-class of a polynomial, hence is in. Consequently, F [y 1 ] =, as asserted in the theorem. The element y 1 is transcendental over F ; for suppose instead that y 1 satisfied an algebraic relation αy 1 = [0 ], with αy 1 = α m y m 1 + α m 1 y m 1 1 + + α 0 ; α m [0]], α j F. By the original italicized definition of F each residue-class α j of F contains a polynomial b j x with V t b j 0, so that [b j ] = α j. Then fx = m j=0 b jx φ τ j φ jτ m j=0 b jx φ τ j φ jτ in V is a polynomial of non-negative value which has [f ] = αy 1 = [0 ]. By Property III of H, V f > 0. On the other hand V f must equal 0, for [b m ] = α m [0]] gives V b m = 0 and V f V b m φ τ m φ mτ = 0, by Theorem 5.2. This contradiction shows y 1 a variable over F. The formula 6 enables us to calculate [[fx ] effectively for any fx given in 4, provided only that V t f j φ τ j 0 for all j. It remains to determine the degree of the field F over F t, which by 2 is the degree of ψy. The ey φ has by Theorem 9.4 an expansion of the form 7 φ = φ mτt t + mτ t 1 a i xφ i t, i=0 V t φ = V t φ mτt t = V t a 0. If t > 1, ψ = [φ t φ ] t can be computed by the analog of 6 for the preceding stage with t in 6 replaced by t 1, for the coefficients φ t a i must by the choice of φ t have V t 1 φ t a i 0. This calculation shows ψy to be a polynomial in y with a first term [[φ t φ τ t t m t ] t y m arising from the first term of 7. But V t φ t φ τ t t m t = V t φ t + V t φ τ t t m t = V t φ + V t φ mτt t = 0, so that the coefficient of y m is not [0 ]. The polynomial ψ has degree m, and by 7 mτ t deg φ t = deg φ, m = deg ψ = deg[ F : F t ], as asserted in Theorem 12.1. This theorem has now been demonstrated. 13. Conditions for ey polynomials. In the criterion of Theorem 9.4 for a ey polynomial the condition iii of equivalence-irreducibility can now be replaced by the condition of Lemma 11.2, iiia, [f f ] is an irreducible polynomial in F [y]. This yields a final explicit description of ey polynomials. A partial converse is possible: Theorem 13.1. In a given V, let ψy y be a polynomial of degree m > 0, irreducible in F [y] and with first coefficient 1. Then there is one and, except for equivalent polynomials in V, only one φx which is a ey polynomial and which has [φ τ m φ ] = ψy. Proof. There is a polynomial fx with the residue-class ψ, so that [f ] = ψ and V f = 0. If we multiply f by φ τ m chosen as in 12, 5 and in the expansion of the resulting product drop all terms not of minimum value and then replace the leading coefficient of φ by 1, we obtain a polynomial φx with the value V φ τ m. Then [φ τ m φ ] = [φ τ m φ τ m f ] = [f ] = ψy. Furthermore φ can be shown to satisfy the remaining conditions i and ii of Theorem 9.4, hence φ is a ey polynomial. The uniqueness is readily established. The proof given holds for t > 1, but may be simplified for the case t = 1. The assumption ψy y is needed, for the condition V f = V f 0 in Theorem 9.4 implies [[f f ]] y. Absolute Values in Polynomial Rings 13 30-Oct-2014, 10:00 am

Since [[f f ] can be effectively constructed by 12, 6, the problem of testing whether a given fx is a ey polynomial is reduced to that of testing the image [f f ] of fx for irreducibility in F [y]. If K is the field of rationals, then F is a finite field and the latter problem is completely solvable. This result can be used to construct examples for inductive values of any stage and for limitvalues of both constant degree and increasing degree types. The construction of constant degree values may be simplified by deducing from Theorem 9.4 the following partial converse of Lemma 6.3: Corollary 13.2. In V let sx be a polynomial of degree less than that of φ x and with V sx = V φ. Then φ x + sx is a ey polynomial for an inductive value over V. 14. Special cases of homomorphism. The residue-class fields can be similarly found for finite discrete limit-values and for inductive values where the value for K is trivial 2 or where the last assigned value µ is incommensurable 6. Theorem 14.1. Let V be a limit-value constructed as in Theorem 6.1 from a sequence of values V 0, V 1, V 2,..., and satisfying one of the conditions: a the degrees of the eys φ are not bounded as ; b V is finite and discrete, and deg φ = M for all > t. The fields F of Theorem 12.1 yield a possibly infinite extension field F = F 0 + F 1 + F 2 + which is isomorphic to the V residue-class field both for K[x] and for Kx. In case b, F = F t+1. Let H be the homomorphism of K[x] to the V residue-class ring F [y] and H the homomorphism of K[x] to the V residue-class ring. Then 1 V 1 fx 0 implies that [[f ]+j = [f ] F j = 1, 2,.... For, according to 12, 3, [f ] = [f ] 1 y=θ, [f ] +1 = [f ] y=θ+1, which indicates that [[f ] is a constant free of y in F, and that [f ] +1 must equal [[f ], and so on. In case a there is for every fx a so large that deg φ > deg f, so that V f = V 1 f, as in Theorem 5.1. If V f 0, then, by 1, [f ] +j = [f ] is a constant in F independent of j. By using a transcendental p-adic number the finiteness condition of 7 can be satisfied. The correspondence 2 [f ] [f ], for with [[f ] = [f ] +1 = [f ] +2 =, carries each element of into an element of F. Every element α of F is used, for, by the definition of F, α is in some F so that α has the form [f ], and [f ] +j = [f ] as in 1, whence α corresponds in 2 to [f ]. The correspondence 2 is one-to-one, for elements are congruent mod V if and only if they are congruent modulo some V. Finally, 2 is an isomorphism, maing F =, as asserted. The residue-class field of Kx is, by the argument of Lemmas 9.3 and 10.1, just the quotient field of F, and must then be F itself. In the case b, the degrees of the extensions F +1 : F as determined in Theorem 12.1 are all 1 for > t. Hence F = F t+1. Because V is finite 7 and discrete, Theorem 6.5 yields for any fx with V f 0 an i t so large that V f = V i f 0. Then [f ] is again ultimately constant, and 2 gives the isomorphism as before. Theorem 14.2. For an incommensurable inductive value V of Kx the field F determined from F 1 and φ exactly as in 12 is the V residue-class field of both K[x] and Kx. Proof. Since no non-zero multiple of µ = V φ lies in Γ 1, no two terms in a φ -expansion can have the same value in V. Hence any polynomial is equivalent to a monomial expansion, and every rational function has by Lemma 9.1 the form fx/gx V cxφ m, V c = V 1 c. If f/g has value 0, then m = 0, V 1 c 0, and [[f/g ] = [c ]. But F is defined in 12, italics or, for = 1, in 10 as all residue-classes [h ] with V 1 h 0. In this case every residue-class has this form, so that F is as asserted the whole residue-class field, either for K[x] or for Kx. In particular, over the trivial value V 0 2 of K the only non-trivial inductive values are V 1 = [ V 0, V 1 x = µ 1 ], µ 1 0; V 2 = [ V 0, V 1 x = 0, V 2 φ = µ 2 ], µ 2 > 0, φx irreducible. Both are incommensurable no multiple of µ 2 lies in the group Γ 0, which contains only 0. Furthermore, the residue-class field of K for the trivial V 0 is K itself. Hence the residue-class field for V 1 is K and for V 2 is Kθ, where θ is a root of φx. Absolute Values in Polynomial Rings 14 30-Oct-2014, 10:00 am