On Subrecursive Representability of Irrational Numbers, Part II

Similar documents
On b-adic Representation of Irrational Numbers

arxiv: v1 [math.lo] 15 Apr 2018

On Subrecursive Representability of Irrational Numbers

Standard forms for writing numbers

Informal Statement Calculus

Introductory Analysis I Fall 2014 Homework #9 Due: Wednesday, November 19

Monotonically Computable Real Numbers

On the Hierarchy of 0 2-Real Numbers

DR.RUPNATHJI( DR.RUPAK NATH )

Connectedness. Proposition 2.2. The following are equivalent for a topological space (X, T ).

Limits and Continuity

This is logically equivalent to the conjunction of the positive assertion Minimal Arithmetic and Representability

Mathematical Reasoning & Proofs

The constructible universe

First-Order Logic. 1 Syntax. Domain of Discourse. FO Vocabulary. Terms

Notes for Math 290 using Introduction to Mathematical Proofs by Charles E. Roberts, Jr.

NOTES ON FINITE FIELDS

Arithmetic and Incompleteness. Will Gunther. Goals. Coding with Naturals. Logic and Incompleteness. Will Gunther. February 6, 2013

Hierarchy among Automata on Linear Orderings

Chapter One. The Real Number System

Supplementary Notes for W. Rudin: Principles of Mathematical Analysis

means is a subset of. So we say A B for sets A and B if x A we have x B holds. BY CONTRAST, a S means that a is a member of S.

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

CHAPTER 8: EXPLORING R

There are infinitely many set variables, X 0, X 1,..., each of which is

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

MATH 117 LECTURE NOTES

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

MATH 521, WEEK 2: Rational and Real Numbers, Ordered Sets, Countable Sets

Peano Arithmetic. CSC 438F/2404F Notes (S. Cook) Fall, Goals Now

M17 MAT25-21 HOMEWORK 6

Partial Collapses of the Σ 1 Complexity Hierarchy in Models for Fragments of Bounded Arithmetic

Seunghee Ye Ma 8: Week 2 Oct 6

The Completion of a Metric Space

Topological properties of Z p and Q p and Euclidean models

We are going to discuss what it means for a sequence to converge in three stages: First, we define what it means for a sequence to converge to zero

Arithmetical Hierarchy

What are the recursion theoretic properties of a set of axioms? Understanding a paper by William Craig Armando B. Matos

G δ ideals of compact sets

Arithmetical Hierarchy

KRIPKE S THEORY OF TRUTH 1. INTRODUCTION

CHAPTER 3. Sequences. 1. Basic Properties

s P = f(ξ n )(x i x i 1 ). i=1

3. Only sequences that were formed by using finitely many applications of rules 1 and 2, are propositional formulas.

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Laver Tables A Direct Approach

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

Foundations of Mathematics MATH 220 FALL 2017 Lecture Notes

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Chapter 1. Sets and Numbers

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

Lecture 2: Syntax. January 24, 2018

VC-DENSITY FOR TREES

Discrete Mathematics. Spring 2017

Math 324 Summer 2012 Elementary Number Theory Notes on Mathematical Induction

Department of Computer Science University at Albany, State University of New York Solutions to Sample Discrete Mathematics Examination II (Fall 2007)

Bounded Derivatives Which Are Not Riemann Integrable. Elliot M. Granath. A thesis submitted in partial fulfillment of the requirements

Topics in Logic and Proofs

The Real Number System

SEARCH FOR GOOD EXAMPLES OF HALL S CONJECTURE. 1. Introduction. x 3 y 2 = k

NOTES (1) FOR MATH 375, FALL 2012

On Aperiodic Subtraction Games with Bounded Nim Sequence

Boolean Algebras. Chapter 2

CS 125 Section #10 (Un)decidability and Probability November 1, 2016

The cardinal comparison of sets

A MODEL-THEORETIC PROOF OF HILBERT S NULLSTELLENSATZ

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.

Lecture Notes on DISCRETE MATHEMATICS. Eusebius Doedel

Proof Techniques (Review of Math 271)

* 8 Groups, with Appendix containing Rings and Fields.

Lecture 11: Minimal types

Preliminaries. Introduction to EF-games. Inexpressivity results for first-order logic. Normal forms for first-order logic

Mathematics 114L Spring 2018 D.A. Martin. Mathematical Logic

MA131 - Analysis 1. Workbook 6 Completeness II

MAT 417, Fall 2017, CRN: 1766 Real Analysis: A First Course

PEANO AXIOMS FOR THE NATURAL NUMBERS AND PROOFS BY INDUCTION. The Peano axioms

MATH1050 Greatest/least element, upper/lower bound

Gödel s Incompleteness Theorems

PRACTICE PROBLEMS: SET 1

Real Analysis - Notes and After Notes Fall 2008

MORE ON CONTINUOUS FUNCTIONS AND SETS

CONSTRUCTION OF THE REAL NUMBERS.

Short Introduction to Admissible Recursion Theory

Contribution of Problems

Basic counting techniques. Periklis A. Papakonstantinou Rutgers Business School

Notes on ordinals and cardinals

Math 210B. Artin Rees and completions


Practice Test III, Math 314, Spring 2016

A version of for which ZFC can not predict a single bit Robert M. Solovay May 16, Introduction In [2], Chaitin introd

Scalar multiplication in compressed coordinates in the trace-zero subgroup

Löwenheim-Skolem Theorems, Countable Approximations, and L ω. David W. Kueker (Lecture Notes, Fall 2007)

Writing proofs for MATH 51H Section 2: Set theory, proofs of existential statements, proofs of uniqueness statements, proof by cases

2. Metric Spaces. 2.1 Definitions etc.

Measure and integration

Set, functions and Euclidean space. Seungjin Han

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

Chapter 4: Computation tree logic

Optimization and Optimal Control in Banach Spaces

Constructions with ruler and compass.

Transcription:

On Subrecursive Representability of Irrational Numbers, Part II Lars Kristiansen 1,2 1 Department of Mathematics, University of Oslo, Norway 2 Department of Informatics, University of Oslo, Norway Abstract. We consider various ways to represent irrational numbers by subrecursive functions. An irrational number can be represented by its base-b expansion; by its base-b sum approximation from below; and by its base-b sum approximation from above. Let S be a class of subrecursive functions, e.g., the class the primitive recursive functions. The set of irrational numbers that can be obtained by functions from S depends on the representation and the base b. We compare the sets obtained by different representations and bases. We also discuss how representations by base-b expansions and sum approximations relates to representations by Cauchy sequences and Dedekind cuts. Keywords: computable analysis, subrecursive functions, irrational numbers. 1 Introduction The first n digits of a decimal expansion suffice to determine the first n digits of a binary expansion of the same number (we are talking about the digits after the period). On the other hand, any fixed number of digits of a decimal expansion are not sufficient to determine the first digit of the decimal expansion of the same number. For example, consider the following two binary expansions α = 0.0(0011) n 0010 β = 0.0(0011) n 0100. The decimal expansion of α and β start with 0.0 and 0.1, respectively. In other words, to determine the first digit after the period, we possibly need to read 4(n + 1) digits of a binary expansion where n can be arbitrary large. Unbounded search cannot occur in subrecursive algorithms. The example above shows that unbounded search is required to convert a binary representation into a decimal representation. In contrast, unbounded search is not required to convert a binary representation into a hexadecimal representation. We can compute the first fractional digit of the hexadecimal representation from the first four fractional digits of the binary representation. Then, we can can compute

2 Lars Kristiansen the next fractional digit of the hexadecimal representation from the next four fractional digits of the binary representation, and so on. We can represent the base-b expansion 3 of an irrational number α between 0 and 1 by a function Eb α where Eα b (n) yields the nth digit of the base-b expansion of α. Let S be a sufficiently large natural class of subrecursive functions, e.g., the class of elementary functions, the class of primitive recursive functions or the class of functions that we can prove is total in Peano Arithmetic, and let S be denote the set of irrational number between 0 and 1 that has a base-b expansion in S, that is, α S be if and only if Eb α S. The informal considerations above indicate that S be will depend on b: we should expect that S 2E S 16E, and we should expect that S 10E S 2E. For which bases a and b do we, or do we not, have S be S ae? It turns out that the inclusion S be S ae holds if and only if every prime factor of a is a prime factor of b. We will prove this equivalence for any subrecursive class S closed under elementary operations (a subrecursive class will be formally defined as an efficiently enumerable set of computable total functions). Sum approximations form below and above were introduced in Kristiansen [6]. Let α be an irrational number between 0 and 1. We can uniquely write α as an infinite sum of the form where α = 0 + D 1 b k1 + D 2 2 k2 + D 3 2 k3 +... b N \ {0, 1} and D i {1,..., b 1} (note that D i 0 for all i) k i N \ {0} and k i < k i+1. Let Âα b (i) = D ib ki when i > 0, and let Âα b (0) = 0. The rational number n Âα b (i) is an approximation of α that lies below α, and we will say that the function Âα b is the base-b sum approximation from below of α. The base-b sum approximation from above of α is a symmetric function Ǎα b such that 1 n Ǎα b (i) is an approximation of α that lies above α (and we have Âα b (i)+ Ǎα b (i) = 1). Let S b denote the set of irrational numbers that has a base-b sum approximation from below in a sufficiently large subrecursive class S, and let S b denote the set of irrational numbers that has a base-b sum approximation of from above in S. An interesting fact about sum approximations is that S b and S b are incomparable classes, that is, S b S b and S b S b (and thus it follows rather straightforwardly that S be S b and S be S b ). This is really what to expect from results already proven in Kristiansen [6], but we give detailed and neat proofs in this paper. This paper s main result on sum approximations is that S b S a if and only if S b S a if and only if every prime factor of a is a prime factor of b. We prove 3 Base-b expansions are often called b-adic expansions or b-adic representations in the literature.

On Subrecurive... Part II. 3 these equivalences for any S closed under primitive recursion (it is an open problem if it suffices to assume that S is closed under elementary operations). We will also discuss the relationship between sum approximations and Dedekind cuts, and we will prove, or at least sketch proofs of, a number of results conjectured in Kristiansen [6]. The research presented in this paper continues the research presented in Kristiansen [6]. Although this paper is meant to be self-contained, the reader may in several respects benefit from being familiar with [6] before reading this paper. In [6] we treat a number of notions, e.g., continued fractions, trace functions, general sum approximations and S-irrational numbers, that are closely related to base-b sum approximations. We also provide some intuition that might helpful when reading technical parts of this paper. The research presented in this paper is also related to research of Specker [19], Mostkowski [12], Lehman [14], Ko [4, 5], Labhalla and Lombardi [13] and a line of of research by Georgiev, Skordev and Weiermann, see [3, 17, 18]. For more on computable real numbers, see Aberth [1] or Weihrauch [21]. The author wants to thank one of the referees for valuable advice regarding the presentation. 2 Notation and Terminology Definition 1. A base is a natural number strictly greater than 1, and a base-b digit is a natural number in the set {0, 1,..., b 1}. Let M be an integer, let b be a base, and let D 1,..., D n be base-b digits. We will n use (M.D 1 D 2... D n ) b to denote the rational number M + D i b i. Let M be an integer, and let D 1, D 2,... be an infinite sequence of base-b digits. We say that (M.D 1 D 2...) b is the base-b expansion of the real number α if we have (M.D 1 D 2... D n ) b α < (M.D 1 D 2... D n ) b + b n for all n 1. Moreover, we say that the base-b expansion of α is finite if there exists k such that α = (M.D 1 D 2... D k ) b, and we say that the base-b expansion of α is infinite if no such k exists. We will use prim(b) denote the set of prime factors of the base b, that is, i=1 prim(b) = {p p is a prime and p b}. We will use D 1... D j (D j+1... D k ) ω to denote the infinite sequence D 1... D j D j+1... D k D j+1... D k D j+1... D k.... If D is a base-b digit, then D denotes the complement digit of D, that is, D = (b 1) D.

4 Lars Kristiansen It is easy to verify the following claims: Any real number has a unique base-b expansion: E.g., (0.0(9) ω ) 10 is not a base-10 expansion of 10 1 according to the definition above. The one and only base-10 expansion of 10 1 is (0.1(0) ω ) 10. When (M.D 1 D 2...) b is the base-b expansion of the real number α, we have α = lim n (M.D 1D 2... D n ) b moreover, if the base-b expansion of α is infinite, we have (M.D 1 D 2... D n ) b < α < (M.D 1 D 2... D n ) b + b n for all n 1. The base-b expansion of the real number α is finite iff the k th digit of the expansion is 0 for all sufficiently large k. For any M Z and any base-b digits D 1,... D n, there exists m Z such that (M.D 1... D n ) b = mb n. For any m Z, there exist M Z and base-b digits D 1,... D n such that mb n = (M.D 1... D n ) b. Assume that the base-b expansion of the rational number q is infinite. Then, the base-b expansion of q will be of the form (M.D 1 D 2... D j (D j+1... D s ) ω ) b where j < s and at least one of the digits D j+1... D s is different from 0, moreover, at least one of the digits D j+1... D s is different from 0, and if q = mn 1 where m, n N, then s < n. If (M.D 1 D 2... D n ) b α < (M.D 1 D 2... D n ) b + b n then (M.D 1 D 2... D j ) b α < (M.D 1 D 2... D j ) b + b j for all j {1,..., n 1}. For any base b and any base-b expansion (0.D 1 D 2...) b, we have (0.D 1 D 2 D 3...) b + (0.D 1 D 2 D 3...) b = 1. 3 The Base Transition Factor If we consider the first four fractional digits of the base-10 expansion of a real number, we have enough information to determine the first fractional digit of the base-16 expansion of the number. If the base-10 expansion starts with 0.0624, the base-16 expansion will start with 0.0, and if the base-10 expansions starts with 0.0625, the base-16 expansion will with by 0.1. Thus, we have to consider at least four fractional digits, but four will be enough. We never have

On Subrecurive... Part II. 5 to consider the fifth fractional digit to determine the first fractional digit of the base-16 expansion. If we want to determine the first two fractional digits of the base-16 expansion, we have to consider the first 8 fractional digits of the base-10 expansion, in general, if we want to determine the first k fractional digits of the base-16 expansion, we have to consider the first 4k fractional digits of the base-10 expansion. The base transition factor from base a to base b will be formally defined below. The factor tells us how many digits we have to consider when we want to convert a real from base b to base a. The base transition factor from base 16 to base 10 is 4. (It might sound a bit backwards that we are talking about the base transition factor from base 16 to base 10 and about converting reals from base 10 to base 16, but the terminology makes sense when you read on.) The base transition factor from base 2 to base 10 is 1. It is possible to determine k fractional digits of a base-2 expansion by considering k fractional digits of the base-10 expansion. The base transition factor from base 10 to base 2 is not defined. This coincides with the fact that we cannot convert an irrational number from base 2 to base 10 without carrying out an unbounded search, see the example at the start of Section 1. Definition 2. Let a and b be bases such that prim(a) prim(b). We will now define the base transition factor from a to b. Let b = p k1 1 pk2 2... pkn n, where p i is a prime and k i N \ {0} (for i = 1,..., n), be the prime factorization of b. Then, a can be written of the form a = p j1 1 pj2 2... pjn n where j i N (for i = 1,..., n). The base transition factor from a to b is the natural number k such that ji k = max{ 1 i n }. k i The base transition factor from a to b is not defined if prim(a) prim(b). When we assume that the base transition factor from a to b exists, it is understood that we have prim(a) prim(b). Clause (3) of the next lemma justifies our terminology. If k is the base transition factor from a to b, then a number of the form M.D 1... D n in base a can be written of the form M.D 1... D kn in base b. Lemma 3. Let k be the base transition factor from a to b. Then (1) there exists m N such that a 1 = mb k (2) for any m Z and any n N, there exists m Z such that ma n = mb kn (3) for any n N and any base-a digits D 1,..., D n, there exists m N such that (0.D 1... D n ) a = mb kn (and thus there exists base-b digits Ḋ 1,..., Ḋ kn such that (0.D 1... D n ) a = (0.Ḋ 1... Ḋ kn ) b ). Proof. Let b = p k1 1... pkn n and a = p j1 1... pjn n

6 Lars Kristiansen where p i is a prime and k i N \ {0} and j i N (for i = 1,..., n). It is easily seen from the definition of k that k ji k i, and thus we have k i k j i 0 (for i = 1,..., n). Furthermore, we have a 1 = (p j1 1... pjn n ) 1 = (p (k1k j1) 1... p (knk jn) n = (p (k1k j1) 1... p (knk jn) n = (p (k1k j1) 1... p (knk jn) n )(p j1+(k1k j1) 1... pn jn+(knk jn) )(p k1k = (p (k1k j1) 1... p (knk jn) n )b k. 1... p knk n ) 1 )(p k1 1... pkn n ) k ) 1 Hence, let m = p (k1k j1) 1... p (knk jn) n and (1) holds. We turn to the proof of (2). By (1), we have t N such that ma n = m(a 1 ) n = m(tb k ) n = mt n b kn. Thus, let m = mt n, and (2) holds. Now it is easy to see that (3) holds. By our definitions, we have (0.D 1... D n ) a = n i=1 D ia i. Thus, we have (0.D 1... D n ) a = ma n for some m N. By (2), we m N such that (0.D 1... D n ) a = mb kn. Theorem 4 (The Base Transition Theorem). Let k be the base transition factor from a to b, and let (M.D 1 D 2...) a and (M.Ḋ 1 Ḋ 2...) b be, respectively, the base-a and base-b expansion of an arbitrary real number α. Then, for all n N, we have (M.D 1... D n ) a (M.Ḋ 1... Ḋ kn ) b α < (M.Ḋ 1... Ḋ kn ) b + b kn (M.D 1... D n ) a + a n. Proof. We can w.l.o.g. assume 0 < α < 1. By Definition 1, we have and for all i N. (0.D 1... D i ) a α < (0.D 1... D i ) a + a i ( ) (0.Ḋ 1... Ḋ i ) b α < (0.Ḋ 1... Ḋ i ) b + b i ( ) (Claim I) For all n N, we have (0.D 1... D n ) a (0.Ḋ 1... Ḋ kn ) b.

On Subrecurive... Part II. 7 Assume that the claim does not hold. Then, for some n we have (0.Ḋ 1... Ḋ kn ) b < (0.D 1... D n ) a. By Lemma 3, we have m 1, m 2 N such that Thus, m 1 b kn = (0.Ḋ 1... Ḋ kn ) b < (0.D 1... D n ) a = m 2 b kn. (0.D 1... D n ) a (0.Ḋ 1... Ḋ kn ) b > b kn and then, by ( ), we have (M.D 1... D kn ) b + b kn < (M.D 1... D n ) a α. This contradicts ( ). This proves (Claim I). (Claim II) For all n N, we have (0.Ḋ 1... Ḋ kn ) b + b kn (0.D 1... D n ) a + a n. It follows straightforwardly from ( ) and ( ) that Thus, (0.Ḋ 1... Ḋ kn ) b < (0.D 1... D n ) a + a n. 0 < (0.D 1... D n ) a + a n (0.Ḋ 1... Ḋ kn ) b. By Lemma 3, we have m 1, m 2 N such that 0 < (0.D 1... D n ) a + a n (0.Ḋ 1... Ḋ kn ) b = m 1 b kn m 2 b kn (thus m 1 has to be strictly greater than m 2 ). It follows that Hence (0.D 1... D n ) a + a n (0.Ḋ 1... Ḋ kn ) b b kn. (0.D 1... D n ) a + a n (0.Ḋ 1... Ḋ kn ) b + b kn. This completes the proof of (Claim II). Now it is easy to see that our theorem holds. By (Claim I) and ( ), we have (0.D 1... D n ) a (0.Ḋ 1... Ḋ kn ) b α < (0.Ḋ 1... Ḋ kn ) b + b kn and then the theorem follows by (Claim II). The next corollary will be needed in the proof of one of our main results. Corollary 5. Let k be the the base transition factor from a to b, and let (M.D 1 D 2...) a and (M.Ḋ 1 Ḋ 2...) b be, respectively, the base-a and base-b expansion of an arbitrary real number α. Then, for all n, l N, we have (M.Ḋ 1... Ḋ kn ) b < (M.Ḋ 1... Ḋ l ) b (M.D 1... D n ) a < (M.D 1... D ml ) a where m = log a b.

8 Lars Kristiansen Proof. Assume (M.Ḋ 1... Ḋ kn ) b < (M.Ḋ 1... Ḋ l ) b. By the Base Transition Theorem, we have Hence (M.D 1... D n ) a (M.Ḋ 1... Ḋ kn ) b < (M.Ḋ 1... Ḋ l ) b < α. α (M.D 1... D n ) a > b l. (*) Assume for the sake of contradiction that (M.D 1... D n ) a (M.D 1... D ml ) a. Then we have α (M.D 1... D n ) a α (M.D 1... D ml ) a < a ml = (a m ) l = (a log a b ) l b l. This contradicts (*), and thus we conclude that (M.D 1... D n ) a < (M.D 1... D ml ) a. (Note that we by our definitions have α (M.D 1... D n ) a < a n for all n. See Definition 1.) We round off the section with another theorem on the base transition factor. The theorem is a kind of converse version of Theorem 4. We will not need this theorem later, but we will need the lemma leading up to the theorem. Lemma 6. Let a and b be bases, let p prim(a) \ prim(b), and let m Z and n N be such that mp n Z. Then, the rational number mp n has a finite base-a and an infinite base-b expansion. Proof. There is a base-a digit D 1 such that a = p D 1. Thus, we have p 1 = (0.D 1 ) a. The product of two numbers with a finite base-a expansion also have a finite base-a expansion. Thus, mp n has a finite base-a expansion as mp n can be written of the form m (0.D 1 ) a... (0.D 1 ) a. Assume that mp n has a finite base-b expansion (and recall that mp n is not an integer). Then there exist m Z and n N such that mp n = mb n. But p does not occur in the prime factorization of b. Thus, the equality mp n = mb n contradicts that every base has a unique prime factorization. Theorem 7. Assume that k is a natural number such that we for every α R and every n N \ {0} have (M.D 1... D n ) a (M.Ḋ 1... Ḋ kn ) b α < (M.Ḋ 1... Ḋ kn ) b + b kn (M.D 1... D n ) a + a n where (M.D 1 D 2...) a and (M.Ḋ 1 Ḋ 2...) b are, respectively, the base-a and base-b expansion of α. Then, prim(a) prim(b). Moreover, the base transition factor from a to b is the least k with this property. Proof. It follows from Lemma 6 that prim(a) prim(b). We leave to the reader to check that there cannot be a number less than the base transition factor from a to b that possesses the property.

On Subrecurive... Part II. 9 4 Subrecursion Theory 4.1 General Preliminaries We assume acquaintance with subrecursion theory and, in particular, with the elementary functions. An introduction to this subject can be found in [15] or [16]. Here we just state some important basic facts and definitions, see [15] and [16] for proofs. We will also assume that the reader is familiar with basic concepts of computability theory, e.g., Kleene s T -predicate and computable indexes. An introduction to elementary computability theory can be found in, e.g., [2] or [11]. The initial elementary functions are the projection functions (Ii n ), the constants 0 and 1, addition (+) and modified subtraction (. ). The elementary definition schemes are composition, that is, f( x) = h(g 1 ( x),..., g m ( x)) and bounded sum and bounded product, that is, respectively f( x, y) = i<y g( x, i) and f( x, y) = i<y g( x, i). A function is elementary if it can be generated from the initial elementary functions by the elementary definition schemes. A relation R( x) is elementary when there exists an elementary function f with range {0, 1} such that f( x) = 0 iff R( x) holds. Relations may also be called predicates, and we will use the two words interchangeably. A function f has elementary graph if the relation f( x) = y is elementary. The definition scheme (µz x)[...] is called the bounded µ-operator, and (µz y)[r( x, z)] denotes the least z y such that the relation R( x, z) holds. Let (µz y)[r( x, z)] = y + 1 if no such z exists. The class of elementary functions is closed under the bounded µ-operator. The definition scheme f( x, 0) = g( x) and f( x, y + 1) = h( x, y, g( x, y)) is called primitive recursion. If f is defined by a primitive recursion over g and h and f( x, y) j( x, y), then f is defined by bounded primitive recursion over g, h and j. The class of elementary functions is closed under bounded primitive recursion, but not under primitive recursion. Moreover, the the class of elementary relations is closed under the operations of the propositional calculus and under bounded quantification. Let 2 x 0 = x and 2 x n+1 = 2 2x n, and let s denote the successor function. The class of elementary functions equals the closure of {0, s, I n i, 2x, max} under composition and bounded primitive recursion. Given this characterization of the elementary functions, it is easy to see that for any elementary function f, we have f( x) 2 max( x) k for some fixed k. We will say that a class of functions is closed under the elementary operations when the class contains all the elementary functions and is closed under composition and bounded primitive recursion. We will say that a class of functions is closed under the primitive recursive operations when the class contains all the elementary functions and is closed under composition and (unbounded) primitive recursion.

10 Lars Kristiansen Uniform systems for coding finite sequences of natural numbers are available inside the class of elementary functions. Let f(x) be the code number for the sequence f(0), f(1),..., f(x). Then f belongs to the elementary functions if f does. We will indicate the use of coding functions with the notations... and (x) i where ( x 0,..., x i,..., x n ) i = x i. (So (x, i) (x) i is an elementary function.) Our coding system is monotone, that is, x 0,..., x n < x 0,..., x n, y holds for any y, and x 0,..., x i,..., x n < x 0,..., x i +1,..., x n. All the closure properties of the elementary functions can be proved by using Gödel numbering and standard coding techniques. We use f k to denote the k th iterate of the function f, that is, f 0 (x) = x and f k+1 (x) = f(f k (x)). 4.2 Coding of Rationals Subrecursive functions in general, and elementary functions in particular, are formally functions over natural numbers (N). We assume some coding of integers (Z) and rational numbers (Q) into the natural numbers. We consider such a coding to be trivial. Therefore we allow for subrecursive functions from rational numbers into natural numbers, from pairs of rational numbers into rational numbers, etc., with no further comment. As seen above, uniform systems for coding finite sequences of natural numbers are available inside the class of elementary functions. Hence, for any reasonable coding, basic operations on rational numbers like addition, subtraction and multiplication will obviously be elementary. We also consider the next lemma to be obvious, and we skip its proof. Lemma 8. Let (M.D q 1 D q 2...) b denote base-b expansion of the rational number q. There exists an elementary function digit : N Q N N such that for any rational number q, we have digit n (q, b) = D q n. 4.3 Honest Functions and Subrecursive Classes The proof of our main results are based on the theory of honest functions. In this subsection, we state and prove lemmas and theorems on honest functions that will be needed later. For more on honest functions, see [9] or [10]. Definition 9. A function f : N N is honest if it is monotone (f(x) f(x + 1)), dominates 2 x (f(x) 2 x ) and has elementary graph. From now on, we reserve the letters f, g, h,... to denote honest functions. Small Greek letter like φ, ψ, ξ,... will denote number-theoretic functions not necessarily being honest.

On Subrecurive... Part II. 11 Definition 10. A function φ is elementary in a function ψ, written φ E ψ, if φ can be generated from the initial functions ψ, 2 x, max, 0, s (successor), I n i (projections) by composition and bounded primitive recursion. Lemma 11. Let ψ E f where f is an honest function. Then there exists k N such that ψ(x 1,..., x n ) f k (max(x 1,..., x n )). Proof. The function ψ can be generated from the initial functions f, 2 x, max, 0, s, Ii n by composition and bounded primitive recursion. Use induction on such a generation of ψ to prove that the lemma holds. Use that f is monotone and dominates 2 x. Let T n denote the Kleene T -predicate, and let U denote the decoding function known from Kleene s Normal Form Theorem. We have φ(x 1,..., x n ) = {e}(x 1,..., x n ) = U(µt[T n (e, x 1,..., x n, t)]) when e is a computable index for φ. We will need the next theorem which is superficially proved in Kristiansen [7]. A more detailed proof can be found in Kristiansen [8]. Theorem 12 (Normal Form Theorem). Let f be an honest function. Let φ be any (Turing) computable function. Then, φ E f iff there exists a computable index e for φ and a fixed k N such that φ(x 1,..., x n ) = {e}(x 1,..., x n ) = U(µt f k (max(x 1,..., x n ))[T n (e, x 1,..., x n, t)]). Moreover, U is an elementary function, and T n is an elementary predicate. Definition 13. For any honest function f, we define the jump of f, written f, by f (x) = f x+1 (x). Lemma 14. Let f be an honest function. Then, f is an honest function. Proof. It is obvious that f is monotone and dominates 2 x. Let ψ(x, y) be an elementary function that places a bound on the code number for the sequence y, y,..., y of length x + 1. Then, f (x) = y is equivalent to ( s ψ(x, y))[(s) 0 = f(x) ( i < x)[(s) i+1 = f((s) i )] (s) x = y]. (*) Thus, the relation f (x) = y is elementary since all the functions, relations and operations involved in (*) are elementary. This proves that f has elementary graph.

12 Lars Kristiansen Lemma 15. Let f be an honest function, and let ψ be a unary function such that ψ E f. Then we have ψ(x) < f (x) for all sufficiently large x. Proof. By Lemma 11, we have k N such that ψ(x) f k (x). For x k, we have ψ(x) f k (x) < f x+1 (x) = f (x). Definition 16. Let σ : N N be a total function, and let [e] σ (x) = U( µt[t 1 (σ(e), x, t)] ) where T 1 and U are the elementary functions from Kleene s Normal Form Theorem (see Theorem 12). A set S of functions over the natural numbers is a subrecursive class when there exists a total computable function σ : N N such that the function [e] σ is total for every φ S there exists e N such that φ(x 1,..., x n ) = [e] σ ( x 1,..., x n ). We say that the function σ generates the class S. (So, a subrecursive class is a subset of an efficiently enumerable class of total functions.) Theorem 17. For any subrecursive class S, there exists an honest function f such that ψ S ψ E f. Proof. Assume that S is generated by the the total computable function σ. Let e σ be a computable index for σ, and let f(x) = µt[ t 2 x ( i x)( t 1 t)[ T 1 (e σ, i, t 1 ) ( j x)( t 2 t)[ T 1 (U(t 1 ), j, t 2 )] ] ]. Now, f is a total computable function as σ and [e] σ are total computable functions. The graph of f is elementary, moreover, f is monotone and dominates 2 x. Thus, f is honest. A proof of the claim below can be found in Section 8 of Kristiansen [6]. x e f(x) µt[t 1 (σ(e), x, t)]. (Claim) Now, let ψ be any function in S. Then, we have e such that ψ( x) = [e] σ ( x ). Let d = σ(e). By the claim, we have ψ( x) = [e] σ ( x ) = U(µt[T 1 (d, x, t)]) = U(µt f( x )[T 1 (d, x, t)]) whenever x e. Thus, we have ψ( x) = U ( (µt f( x + e))[t 1 (d, x, t)] ) for all x. This proves that ψ is elementary in f.

On Subrecurive... Part II. 13 Lemma 18. Let ψ be any function over the natural numbers. For any honest function g, there exists an honest function f such that ψ P R g ψ E f. Proof. Let S = {ψ ψ P R g}. It is easy to see that S is a subrecursive class in the sense of Definition 16. Assume ψ P R g. Then, ψ S. By Theorem 17, we have f such that ψ E f. 5 Base-b Expansions From now on we will restrict our attention to irrational numbers between 0 and 1. This entails no loss of generality. Definition 19. Let (0.D 1 D 2...) b be the base-b expansion of the real number α. We define the function Eb α : N {0,.., b 1} by Eα b (0) = 0 and Eα b (i) = D i (for i 1). For any class of functions S, let S be = { α E α b S }. We will occasionally identify the function Eb α with the the base-b expansion of α, and we may, e.g., say that S be is the set of irrational numbers with a base-b expansion in S. Theorem 20. Let a and b be bases such that prim(a) prim(b). For any real number α [0, 1), we have E α a E E α b. Proof. Let (0.D 1 D 2...) b be the base-b expansion of α and let k be the base transition factor from a to b. By the Base Transition Theorem, we have E α a (n) = digit n ((0.D 1... D kn ) b, a) where digit is the elementary function given by Lemma 8. Moreover (0.D 1... D kn ) b = kn i=1 E α b (i)b i. Thus, E α a is elementary in E α b. The preceding theorem says that we can compute Ea α elementarily in Eb α if prim(a) is a subset of prim(b). We cannot in general compute Ea α elementarily in Eb α if prim(a) is not a subset of prim(b). This is a consequence of the next theorem. In the proof of the theorem we construct a sequence of rationals q 0, q 1, q 2,... that converges to an irrational number α. The sequence is constructed such that

14 Lars Kristiansen α becomes different from every real whose base-a expansion is elementary in a given honest function f. Still, it turns out that α has an elementary base-b expansion. It is possible to construct the sequence q 0, q 1, q 2,... for any honest function f and any bases a and b where prim(a) prim(b). We will explain some of the ideas behind the construction such that it becomes easier for the reader to follow the technical proof. We start the construction by picking a sequence d 0, d 1, d 2,... of natural numbers. We set d 0 to some suitable number, and then we define d i+1 by d i+1 = f (d i ) where f is the jump of f. There are two reasons why we use f to determine the elements of the sequence. One reason is that d i and d i+1 must be very far apart from each other. For any fixed k, it must be the case f k (d i ) < d i+1 when i is large. If this is not the case, we will not be able to force the sequence q 0, q 1, q 2,... to converge to a desired limit, that is, a limit which base-a expansion is not elementary in f. When d i+1 = f (d i ), the distance between d i and d i+1 will be big enough. Another reason is that f has elementary graph (f is honest since f is honest, and thus the graph of f is elementary, see Lemma 14). This entails that we given x elementarily can decide if there is i such that d i = x. This will help us to pick q 0, q 1, q 2,... such that the base-b expansion of the limit becomes elementary. Now we are ready to explain the definition of q 0, q 1, q 2,.... The first element in the sequence q 0 is some suitable rational that has finite base-a expansion and infinite base-b expansion. In order to avoid confusing and annoying indexes, every second element of the sequence is just a copy of the preceding one, more precisely, q 2i+1 equals q 2i for all i N. Thus, q 2, q 4, q 6,... are the essential elements of the sequence. For any i N, we determine the value of q 2i+2 by the following scheme: Step 1. Pick a real number γ which base-a expansion is elementary in f. Comments to step 1. The number i tells us how to pick γ, more precisely, the number i yields a computable index that tells us how to compute the function E γ a. If the base-a expansion of a real is elementary in f, we will eventually come across an i that tells us to pick that real (we will indeed encounter infinitely many such i s). Step 2. Compute the rational number q γ such that q γ = d 2i+1 j=0 E γ a (j). Comments to step 2. The rational number q γ lies close to γ (it lies slightly below). In the next step, we use q γ to force the sequence q 0, q 1, q 2,... to converge to something else than γ.

On Subrecurive... Part II. 15 Step 3. Let q 2i+2 = { q 2i+1 ɛ 0 if q γ q 2i+1 q 2i+1 + ɛ 1 if q γ < q 2i+1 ( ) where ɛ 0 and ɛ 1 are suitable rational numbers. Comments to step 3. The rationals ɛ 0 and ɛ 1 are suitable when ɛ 0 and ɛ 1 are so small that the first d 2i+2 digits of the base-b expansion of q 2i+2 coincide with the first d 2i+2 digits of the base-b expansion of q 2i+1 (and thus with the first d 2i+2 digits of the base-b expansion of q 2i ). This will ensure that the sequence converges. Moreover, the first d 2i+2 digits of the base-b expansion of q 2i will coincide with the first d 2i+2 digits of the base-b expansion of the sequence s limit. ɛ 0 guarantees that lim i q i < γ. ɛ 1 guarantees that lim i q i > γ. ɛ 0 and ɛ 1 ensure that q 2i+2 has a finite base-a expansion and infinite base-b expansion. It is essential that all the rationals q 0, q 1, q 2... have finite base-a expansion and infinite base-b expansion. When we set q 2i+2 to something smaller than q 2i+1, we need a huge initial segment of the base-b expansion of that something smaller to coincide with a huge initial segment of the base-b expansion of q 2i+1. That would not be possible if the base-b expansion of q 2i+1 were finite (e.g., the first digit of the base-10 expansion of 10 1 is different from the first digit in the base-10 expansion of 10 1 ɛ for any small ɛ > 0). Moreover, we have to determine if γ is above or below q 2i+1 by examine a bounded segment of the base-a expansion of γ. This would not be possible if the base-a expansion of q 2i+1 were infinite. So far we have explained why the base-a expansion of the limit of q 0, q 1, q 2,... is not elementary in f. We have not yet explained why the base-b expansion of this limit is elementary. Well, this is the reason: Recall that we used a function with elementary graph to define the sequence d 0, d 1, d 2,.... This entails that we given n easily can find i such that d i n < d i+1. It is easy in the sense that i can be computed elementarily in n. Now, i is very small compared to n. Thus, we can compute q i elementarily in n. We cannot compute q i elementarily in i, but we can do it elementarily in n since n is very much bigger than i. Finally, when q i is available, we can elementarily compute the n th digit of the base-b expansion q i. But then we have elementarily computed the n th digit of the base-b expansion of α as the n th digit of the base-b expansion of α is the same as n th digit of the base-b expansion of q i. This concludes our intuitive explanation of the proof of Theorem 21. Theorem 21. Let a and b be bases such that prim(a) prim(b), and let f be any honest function. There exists an irrational number α such that (1) Eb α is elementary, and (2) Ea α E f.

16 Lars Kristiansen Proof. Let p prim(a) \ prim(b). We define the sequence of natural numbers d 0, d 1, d 2,... by d 0 = max(p, b) and d i+1 = f (d i ) where f is the jump of f. Recall that U and T 1 are the elementary function and the elementary predicate from Theorem 12, furthermore, digit is the elementary function given by Lemma 8. We define the sequence of rational numbers q 0, q 1, q 2,... by q 0 = p 1 and q 2i+1 = q 2i and { q 2i+1 p bk if U(µt d 2i+2 [T 1 ((i) 1, d 2i+1, t)]) q 2i+1 q 2i+2 = q 2i+1 + p bl ( ) otherwise where k is the least natural number greater than or equal to d 2i+2 such that digit k (q 2i+1, b) 0 l is the least natural number greater than or equal to d 2i+2 such that digit l (q 2i+1, b) 0. Let α = lim n q n. (Claim I) For any i N, there exists m N such that q 2i = mp d2i+1. We prove (Claim I) by induction on i. It is easy to see that the claim holds when i = 0. Assume by induction hypothesis that we have m such that q 2i = mp d2i+1 (we will prove that there is m such that q 2i+2 = mp d2i+3 ). We can w.l.o.g. assume that q 2i+2 = q 2i+1 p bk (the case when q 2i+2 = q 2i+1 + p bl is similar). Now, k is the least natural number greater than or equal to d 2i+2 such that digit k (q 2i+1, b) 0. We need to find an upper bound for k. By our induction hypothesis and the definition of q 2i+1 we have q 2i+1 = q 2i = mp d2i+1 where p prim(a)\prim(b). Thus, the base-b expansion of q 2i is infinite by Lemma 6. Since q 2i is rational, its base-b expansion will be of the form (0.D 1 D 2... D j (D j+1... D s ) ω ) b where j < s and at least one of the digits in the sequence D j+1... D s will be different from 0 (if they all were zeros, the base b expansion q 2i would be finite). Now, q 2i = mp d2i+1 where m and p d2i+1 are natural numbers. Thus, we have s < p d2i+1. Hence, we have k < d 2i+2 + p d2i+1. Now, since f is an honest function and d j max(p, b) 2 (for any j N), we have bk < b(d 2i+2 + p d2i+1 ) b(d 2i+2 + 2 log 2 p d2i+1 ) b(d 2i+2 + f( log 2 p d 2i+1 )) b(d 2i+2 + ff(d 2i+1 )) < b(d 2i+2 + f (d 2i+1 )) = b(d 2i+2 + d 2i+2 ) < f (d 2i+2 ) = d 2i+3.

On Subrecurive... Part II. 17 This proves that d 2i+3 is greater than bk. Now it is easy to see that there exists m N such that q 2i+2 = q 2i p bk = mp d2i+1 p bk = mp d2i+3. This completes the proof of (Claim I). (Claim II) For any j N, we have and q j+1 > q j lim n q n > q j q i+j < q j lim n q n < q j. For each i we have m 1,..., m i { 1, 1} and a very fast increasing sequence of natural numbers k 1, k 2,..., k i such that q 2i = q 2i+1 = p 1 + m 1 p bk1 + m 2 p bk2 +... + m i p bki. Thus, it is easy to see that (Claim II) holds. We are now prepared to prove clause (2) of the theorem. Let α n denote the sum n Eα a (n)a n. Then, we have lim q n = α = lim α n. (*) n n Assume for the sake of a contradiction that E α a E f. Thus, α n E f (view α n as a function of n). By Theorem 12, we have e, k N such that α n = U(µt f k (n)[t 1 (e, n, t)]). Pick i, j such that i = e, j and d 2i+1 k (there are infinitely many such i and j) and recall that f (x) = f x+1 (x). Then, we have α d2i+1 = U(µt f k (d 2i+1 )[T 1 (e, d 2i+1, t)]) = U(µt f k (d 2i+1 )[T 1 ((i) 1, d 2i+1, t)]) = U(µt f (d 2i+1 )[T 1 ((i) 1, d 2i+1, t)]) = U(µt d 2i+2 [T 1 ((i) 1, d 2i+1, t)]) (**) Now our proof splits into the the cases α d2i+1 q 2i+1 and α d2i+1 < q 2i+1. In both cases we will deduce a contradiction. Thus, we can conclude that E α a is not elementary in f. The case α d2i+1 q 2i+1. By (**) and ( ), we have α d2i+1 q 2i+1 > q 2i+2. By (Claim II), we have q 2i+1 > lim q n. Since α 0, α 1, α 2,... is an increasing n sequence, we have This contradicts (*). lim n α n α d2i+1 q 2i+1 > lim n q n.

18 Lars Kristiansen The case α d2i+1 < q 2i+1. By the definition of q 2i+1 and (Claim I), we have q 2i+1 = q 2i = mp d2i+1 for some m N. Since p prim(a), we also have q 2i+1 = m 0 a d2i+1 for some m 0 N. Furthermore, it is easy to see that we have α d2i+1 = m 1 a d2i+1 for some m 1 N. Thus, as α d2i+1 < q 2i+1, we have q 2i+1 α d2i+1 a d2i+1. Since α 0, α 1, α 2,... converges to an irrational number, we conclude that lim α n < q 2i+1. By (**) and ( ), we have q 2i+1 < q 2i+2. n By (Claim II), we have q 2i+1 < lim q n. Hence, lim α n < lim q n, and this n n n contradicts (*). This concludes the proof of clause (2) of the theorem. It remains to prove that clause (1) also holds. To this end we need the next claim. (Claim III) Let (0.D 1 D 2...) b be the base-b expansion of q 2i. Then, for any natural number j, we have 1 j < d 2i+2 (0.D 1 D 2 D 3... D j ) b < q 2i+2 < (0.D 1 D 2 D 3... D j ) b + b j. By (Claim I) and Lemma 6, q 2i has an infinite base b expansion. Thus, we have (0.D 1 D 2 D 3... D s ) b < q 2i < (0.D 1 D 2 D 3... D s ) b + b s for any s 1. We may w.l.o.g. assume that q 2i+2 = q 2i p bk where k is the least natural number k such that k d 2i+2 and D k 0 (the proof when q 2i+2 = q 2i + p bl is symmetric). Now, p bk = (p b ) k > b k. Hence, q 2i+2 = q 2i+1 p bk = q 2i p bk > q 2i b k. Now, since D k 0, we have (0.D 1 D 2 D 3... D k 1 ) b < q 2i+2 < q 2i < (0.D 1 D 2 D 3... D k 1 ) b + b (k 1). As k d 2i+2, we have (0.D 1 D 2 D 3... D j ) b < q 2i+2 < (0.D 1 D 2 D 3... D j ) b + b j for any j {1,..., d 2i+2 1}. This completes the proof of the (Claim III). The next claim follows easily from (Claim III). We have α = lim i q i, and then, by (Claim III), the first d 2i+2 digits of the base-b expansion of α will coincide with the first d 2i+2 digits of the base-b expansion of q 2i. Hence, (Claim IV) holds. (Claim IV) Let d i n < d i+1. Then, the n th digit of the base-b expansion of α is the same as the n th digit of the base-b expansion of q i, that is E α b (n) = digit n (q i, b).

On Subrecurive... Part II. 19 Now we ready to prove clause (1) of our theorem. We have d i+1 = f (d i ). The function f is the jump of f, and f is honest when f is, see Lemma 14. It follows that d z = y is an elementary relation. Let { d i if d i n ψ(i, n) = n otherwise It is not hard to see that ψ is an elementary function when we know that the relation d z = y is elementary. We will now define the function φ(i, n) by courseof-values recursion on i. Let φ(i, n) = p 1, let φ(2i + 1, n) = φ(2i, n), and let φ(2i + 1, n) p bk φ(2i + 2, n) = φ(2i + 1, n) + p bl where if U(µt ψ(2i + 2, n)[t 1 ((i) 1, ψ(2i + 1, n), t)]) φ(2i + 1, n) otherwise k is the least natural number greater than or equal to d 2i+2 such that digit k (q 2i+1, b) 0 l is the least natural number greater than or equal to d 2i+2 such that digit l (q 2i+1, b) 0. We observe that φ is defined by by course-of-values recursion over elementary functions. More careful, but very tedious, considerations will show that this course-of-values recursion over elementary functions can be reduced to bounded primitive recursion over elementary functions. Thus, as the elementary functions are closed under bounded primitive recursion, φ is an elementary function. It follows straightforwardly from the definition of the sequence q 0, q 1, q 2,... that we have φ(i, n) = q i when d i n. We can compute the i such that d i n < d i+1 elementarily in n. Thereafter, we can compute q i elementarily by computing φ(i, n). By (Claim IV) we have Eb α(n) = digit n(q i, b), and the function digit is elementary. This proves that the function Eb α is elementary. Corollary 22. Let S be any subrecursive class which is closed under elementary operations. For any bases a and b, we have prim(a) prim(b) S be S ae. Proof. Assume prim(a) prim(b). Let α S be. Thus, Eb α is in S. By Theorem 20, Ea α will also be in S. Thus, α S ae. Thus, S be S ae. Assume prim(a) prim(b). Let f be an honest function such that ψ E f implies ψ S. Such an f exists by Theorem 17. By Theorem 21, we have α such that Eb α is elementary and Eα a E f. Thus, Eb α is in S whereas Eα a is not. This shows that S be S ae.

20 Lars Kristiansen Mostowski [12] proves (a theorem obviously equivalent to) the left-right implication of Corollary 22 and poses the right-left implication as an open problem (Mostowski s S is the class of primitive recursive functions). Some of the results in Weihrauch [20] seem to be connected to the results proved above. Weihrauch proves that prim(a) prim(b) if and only if a base-b representation of a real can be continuously translated to a base-a representation of the same real, see Corollary 9 in [20]. 6 Base-b Expansions, Cauchy sequences and Dedekind Cuts Definition 23. A function C : N Q is a Cauchy sequence for the real number α when α C(n) < 1 2 n. A function D : Q {0, 1} is the Dedekind cut of the real number α when D(q) = 0 iff q < α. For any class of functions S, let S C denote the set of irrational numbers (between 0 and 1) that have Cauchy sequences in S; let S D denote the set of irrational numbers (between 0 and 1) that have Dedekind cut in S. Let S be a sufficiently large and natural subrecursive class. If (0.D 1 D 2...) b is the base-b expansion of the irrational number α, then C where C(n) = n+1 Eb α (i) is a Cauchy sequence for α. Thus, we have a Cauchy sequence for α in S if the base-b expansion of α is in S, and the inclusion S be S C holds for any base b. It is also easy to see that the inclusion S D S be holds for any base b. Assume that the Dedekind cut of α is in S. Then, the base-b expansion (0.D 1 D 2...) b of α will also be in S because we can compute D n+1 by the following procedure: First we compute (0.D 1... D n ) b. Then, we use (0.D 1... D n ) b and the Dedekind cut of α to determine D n+1 by searching for a base-b digit X such that (0.D 1... D n X) b < α < (0.D 1... D n X) b + b (n+1). No unbounded search is required. Thus, the base-b expansion of α is in S if the Dedekind cut of α is in S, and we have S D S be. Let b be an arbitrary base. Pick a base a such that prim(a) prim(b) and prim(b) prim(a). By Corollary 22, we have S be S ae and S ae S be. So, S be and S ae are incomparable sets, and by the considerations above both of them are subsets of S C and supersets of S D, that is S D S be S C and S D S ae S C.

On Subrecurive... Part II. 21 This implies that we have S D S be S C (*) for any base b. More careful considerations will show that (*) holds for any S which is closed under elementary operations. Kristiansen [6] proves the inclusion S D S C by a direct diagonalization argument, that is, without considering base-b expansions and the class S be. Specker [19] was the first to prove that we have S D S C for a subrecursive class S (Specker s S is the class of primitive recursive functions). He proves (*) with b = 10 by constructing α such that α S 10E and 3 α S 10E. Thus, S 10E is not closed under multiplication by natural numbers, but it is rather obvious that both S D and S C are. Hence, S 10E is different from both S D and S C, and since S 10E is a subset of S C and superset of S D, it follows that S 10E is a strict subset of S C and strict superset of S D. 7 Sum Approximations Sum approximations from below and above are explained at page 2. We will now give our formal definitions. Definition 24. Let (0.D 1 D 2...) b be the base-b expansion of an irrational number α. The base-b sum approximation from below of α is the function Âα b : N Q defined by Âα b (0) = 0 and Âα b (n + 1) = Eα b (m)b m where m is the least m such that n  α b (i) < (0.D 1... D m ) b. The base-b sum approximation from above of α is the function Ǎα b : N Q defined by Ǎα b (0) = 0 and Ǎα b (n + 1) = Eα b (m)b m where m is the least m such that n 1 Ǎ α b (n) > 1 (0.D 1... D m ) b (recall that D is the complement digit of D). For any class of functions S, let S b = { α  α b S } and S b = { α Ǎ α b S }. The functions Âα b and Ǎα are not defined if α is rational. When we use the notation it is understood that α is irrational.

22 Lars Kristiansen Lemma 25. Let (0.D 1 D 2...) b be the base-b expansion of the irrational number α. For any n N, there exists k, l n such that k  α b (i) = (0.D 1... D n ) b and l Ǎ α b (i) = (0.D 1... D n ) b. (*) Proof. We prove (*) by induction on n. We have (0.) b = 0 and Âα b (0) = Ǎα b (0) = 0. Thus, (*) holds with k = l = 0 when n = 0. Now, assume that (*) holds for n (we prove that (*) holds with n + 1 for n). Consider (0.D 1... D n D n+1 ) b. Now, D n+1 might be the digit 0, and D n+1 might be the digit 0, and D n+1 might be neither 0 nor 0. We split the proof into three cases. The case when D n+1 = 0. We have 1 l Ǎ α b (i) = 1 (0.D 1... D n ) b > 1 (0.D 1... D n D n+1 ) b. ( ) Thus, by the definition of Ǎα b (l + 1), we have l+1 Ǎ α b (i) = l Ǎ α b (i) + Ǎα b (l + 1) = (0.D 1... D n ) b + D n+1 b (n+1) = (0.D 1... D n+1 ) b. Furthermore, we have k  α b (i) = (0.D 1... D n ) b = (0.D 1... D n D n+1 ) b. Thus, (*) holds with n + 1 for n. The case when D n+1 = 0. Now we have k  α b (i) < (0.D 1... D n D n+1 ) b. ( ) By the definition of Âα b (k + 1), we have k+1  α b (i) = k  α b (i) + Âα b (k + 1) = (0.D 1... D n ) b + D n+1 b (n+1) = (0.D 1... D n+1 ) b.

On Subrecurive... Part II. 23 Furthermore, since the complement digit of 0 is the digit 0, we have l Ǎ α b (i) = (0.D 1... D n ) b = (0.D 1... D n D n+1 ) b. Thus, (*) holds with n + 1 for n. The case when D n+1 0 and D n+1 0. In this case both ( ) and ( ) hold, and we get k+1  α b (i) = (0.D 1... D n+1 ) b and l+1 Ǎ α b (i) = (0.D 1... D n+1 ) b. Theorem 26. For any irrational number α and any base b, we have Eb α (i)b i =  α b (i) = 1 Ǎ α b (i). Proof. The first equality follow straightforwardly from Lemma 25. It also follows from Lemma 25 that  α b (i) + Ǎ α b (i) = 1 and thus the second equality holds. Lemma 27. (i) Let α be an irrational number, and let p(x) be a polynomial such that i[ x i p(x) E α b (i) 0 ] for all x N. Then, Âα b E Eb α. (ii) Let α be an irrational number and let p(x) be a polynomial such that for all x N. Then, Ǎ α b E E α b. i[ x i p(x) E α b (i) 0 ] Proof. We prove (i). Assume Âα b (n) = Db m where D is some nonzero base-b digit. We know that Êα b (i) 0 for some i between m + 1 and p(m + 1). Hence, we can compute Âα b (n + 1) from the rational number p(m+1) i=1 E α b (i)b i.

24 Lars Kristiansen Such a computation of Âα b (n + 1) requires on application of primitive recursion, but more careful considerations will show that this application of primitive recursion can be reduced to bounded primitive recursion. The functions elementary in Êα b are closed under bounded primitive recursion. Hence, Âα b E Êα b. The proof of (ii) is similar. Lemma 28. Let f be an honest function, and let f be the jump of f. (i) Let α be an irrational number such that we have i[ x i f (x) 1 E α b (i) = 0 ] for infinitely many x N. Then,  α b such that we have E f. (ii) Let α be an irrational number i[ x i f (x) 1 E α b (i) = 0 ] for infinitely many x N. Then, Ǎ α b E f. Proof. We prove (i). Let x be such that we have Eb α (i) = 0 for all i between x and f (x) 1 (there are infinitely many such x). Then, there exists m < x such that Âα b (m + 1) (b 1)b f (x). Let ψ(z) = y if  α b (z + 1) = m 0b y for some m 0. Now, ψ is a total function. Moreover, ψ E  α b. We have ψ(m) f (x) for some m < x. Since ψ is strictly increasing, we have ψ(x) > f (x). Thus there are infinitely many x such that ψ(x) > f (x). Assume for the sake of a contradiction that Âα b E f. Then, ψ E f. This contradicts Lemma 15. Thus we conclude that Âα b E f. This completes the proof of (i). The proof of (ii) is symmetric. Theorem 29. For any honest function f there exist irrational numbers α and β such that such that (1) Ǎα b is elementary and Âα b E f, and (2) Âβ b is elementary and Ǎβ b E f. Moreover, (3) α and β have elementary Dedekind cuts. Proof. We prove (i). Let d 0 = 0 and d i+1 = f (d i ) where f is the jump of f. Let α be the irrational number given by the base-b expansion { Eb α 0 if there exists i such that d i = x (x) = 0 otherwise. Since f is honest, it is possible to check elementary in x if there is i such that d i = x. Hence, Eb α is elementary. By Lemma 27 (ii), Ǎα b is elementary. By Lemma 28 (i), we have Âα b (n) E f. This proves (i). The proof of (ii) is symmetric: Use Lemma 27 (i) in place of Lemma 27 (ii), use Lemma 28 (ii) in place of Lemma 28 (i), and let E β b (x) = 0 if there exists i such that d i = x, otherwise, let E β b (x) = 0. The proof of (3) is rather straightforward, and we omit the details. The reader that wants more details may consult the proof of Theorem 5.2 in Kristiansen [6].

On Subrecurive... Part II. 25 Corollary 30. Let S be a subrecursive class closed under elementary operations. For any base b, we have S b S b and S b S b. Proof. Pick an honest f such that Ǎα b S implies Ǎα b E f. Such an f exists by Theorem 17. By Theorem 29 (2), we have α such that Âα b S and Ǎα b f. Thus, S b S b. A symmetric argument yields S b S b. Use clause (1) of Theorem 29 in place of clause (2). (for any irrational α be- Theorem 31. We have Eb α E Âα b and Eα b E Ǎα b tween 0 and 1). Proof. Let (0.D 1 D 2...) b be the base-b expansion of α. It is trivial that we have n  α b (i) = (0.D 1... D m ) b for some m n. Let digit be the elementary function given by Lemma 8. Then we have ( n ) Eb α (n) = digit n  α b (i), b. Thus, Eb α E Âα b. Furthermore, by Lemma 25, we have 1 n+1 Ǎ α b (i) = 1 (0.D 1... D m D m+1 ) b = (0.D 1... D m (D m+1 + 1)) b for some m n. Thus, we have the equality ) n+1 Eb α (n) = digit n (1 Ǎ α b (i), b and we conclude that E α b E Ǎα b. Theorem 32. Let prim(a) prim(b). Then, we have Âα a P R  α b and Ǎα a P R (for any irrational α between 0 and 1). Ǎ α b Proof. We prove that Âα a P R  α b. Let (0.D 1D 2...) a and (0.Ḋ 1 Ḋ 2...) b be, respectively, the base-a and base-b expansion of α. Assume that we have computed Âα a (0),... Âα a (n) (the computation of  α a (0) is trivial). Then, we can compute Âα a (n + 1) by the following algorithm: