An example of a computable



Similar documents
CHAPTER II THE LIMIT OF A SEQUENCE OF NUMBERS DEFINITION OF THE NUMBER e.

SHARP BOUNDS FOR THE SUM OF THE SQUARES OF THE DEGREES OF A GRAPH

Continued Fractions and the Euclidean Algorithm

Mathematics for Econometrics, Fourth Edition

1. Prove that the empty set is a subset of every set.

CONTINUED FRACTIONS AND FACTORING. Niels Lauritzen

MASSACHUSETTS INSTITUTE OF TECHNOLOGY 6.436J/15.085J Fall 2008 Lecture 5 9/17/2008 RANDOM VARIABLES

Homework until Test #2

Degrees that are not degrees of categoricity

MA651 Topology. Lecture 6. Separation Axioms.

Undergraduate Notes in Mathematics. Arkansas Tech University Department of Mathematics

The Classes P and NP

F. ABTAHI and M. ZARRIN. (Communicated by J. Goldstein)

CONTINUED FRACTIONS AND PELL S EQUATION. Contents 1. Continued Fractions 1 2. Solution to Pell s Equation 9 References 12

GENERIC COMPUTABILITY, TURING DEGREES, AND ASYMPTOTIC DENSITY

Math 55: Discrete Mathematics

3. Mathematical Induction

U.C. Berkeley CS276: Cryptography Handout 0.1 Luca Trevisan January, Notes on Algebra

Stationary random graphs on Z with prescribed iid degrees and finite mean connections

1 if 1 x 0 1 if 0 x 1

11 Ideals Revisiting Z

THE TURING DEGREES AND THEIR LACK OF LINEAR ORDER

COFINAL MAXIMAL CHAINS IN THE TURING DEGREES

The sum of digits of polynomial values in arithmetic progressions

CONTRIBUTIONS TO ZERO SUM PROBLEMS

CHAPTER 5. Number Theory. 1. Integers and Division. Discussion

9 More on differentiation

THE SEARCH FOR NATURAL DEFINABILITY IN THE TURING DEGREES

Basic Concepts of Point Set Topology Notes for OU course Math 4853 Spring 2011

Prime Numbers and Irreducible Polynomials

HOMEWORK 5 SOLUTIONS. n!f n (1) lim. ln x n! + xn x. 1 = G n 1 (x). (2) k + 1 n. (n 1)!

Lecture 2: Universality

About the inverse football pool problem for 9 games 1

Low upper bound of ideals, coding into rich Π 0 1 classes

EMBEDDING COUNTABLE PARTIAL ORDERINGS IN THE DEGREES

WHAT ARE MATHEMATICAL PROOFS AND WHY THEY ARE IMPORTANT?

Gambling Systems and Multiplication-Invariant Measures

Some Polynomial Theorems. John Kennedy Mathematics Department Santa Monica College 1900 Pico Blvd. Santa Monica, CA

Sign changes of Hecke eigenvalues of Siegel cusp forms of degree 2

Lectures 5-6: Taylor Series

How To Prove The Dirichlet Unit Theorem

A simple criterion on degree sequences of graphs

SOLUTIONS TO ASSIGNMENT 1 MATH 576

How To Find Out How To Calculate A Premeasure On A Set Of Two-Dimensional Algebra

Notes on Factoring. MA 206 Kurt Bryan

Extension of measure

arxiv: v1 [math.pr] 5 Dec 2011

I. GROUPS: BASIC DEFINITIONS AND EXAMPLES

Collatz Sequence. Fibbonacci Sequence. n is even; Recurrence Relation: a n+1 = a n + a n 1.

THE DEGREES OF BI-HYPERHYPERIMMUNE SETS

Mathematical Induction. Lecture 10-11

There is no degree invariant half-jump

DEGREES OF ORDERS ON TORSION-FREE ABELIAN GROUPS

Mathematical Methods of Engineering Analysis

Turing Degrees and Definability of the Jump. Theodore A. Slaman. University of California, Berkeley. CJuly, 2005

FACTORING CERTAIN INFINITE ABELIAN GROUPS BY DISTORTED CYCLIC SUBSETS

Metric Spaces. Chapter 1

The Ideal Class Group

Lecture 13: Factoring Integers

Diagonalization. Ahto Buldas. Lecture 3 of Complexity Theory October 8, Slides based on S.Aurora, B.Barak. Complexity Theory: A Modern Approach.

Further Study on Strong Lagrangian Duality Property for Invex Programs via Penalty Functions 1

SOLUTIONS TO EXERCISES FOR. MATHEMATICS 205A Part 3. Spaces with special properties

THE FUNDAMENTAL THEOREM OF ALGEBRA VIA PROPER MAPS

This chapter is all about cardinality of sets. At first this looks like a

Gambling and Data Compression

Mathematics for Computer Science/Software Engineering. Notes for the course MSM1F3 Dr. R. A. Wilson

INTEGRAL OPERATORS ON THE PRODUCT OF C(K) SPACES

God created the integers and the rest is the work of man. (Leopold Kronecker, in an after-dinner speech at a conference, Berlin, 1886)

Global Properties of the Turing Degrees and the Turing Jump

1 Approximating Set Cover

Information Theory and Coding Prof. S. N. Merchant Department of Electrical Engineering Indian Institute of Technology, Bombay

Metric Spaces. Chapter Metrics

INDISTINGUISHABILITY OF ABSOLUTELY CONTINUOUS AND SINGULAR DISTRIBUTIONS

Prime power degree representations of the symmetric and alternating groups

A single minimal complement for the c.e. degrees

Linear Risk Management and Optimal Selection of a Limited Number

Lecture Notes on Measure Theory and Functional Analysis

A REMARK ON ALMOST MOORE DIGRAPHS OF DEGREE THREE. 1. Introduction and Preliminaries

MOP 2007 Black Group Integer Polynomials Yufei Zhao. Integer Polynomials. June 29, 2007 Yufei Zhao

CHAPTER 7 GENERAL PROOF SYSTEMS

International Journal of Information Technology, Modeling and Computing (IJITMC) Vol.1, No.3,August 2013

ON SEQUENTIAL CONTINUITY OF COMPOSITION MAPPING. 0. Introduction

FIRST YEAR CALCULUS. Chapter 7 CONTINUITY. It is a parabola, and we can draw this parabola without lifting our pencil from the paper.

The Settling Time Reducibility Ordering and 0 2 Setting

Reading 13 : Finite State Automata and Regular Expressions

Kolmogorov Complexity and the Incompressibility Method

Formal Languages and Automata Theory - Regular Expressions and Finite Automata -

A BASIS THEOREM FOR Π 0 1 CLASSES OF POSITIVE MEASURE AND JUMP INVERSION FOR RANDOM REALS ROD DOWNEY AND JOSEPH S. MILLER

Tiers, Preference Similarity, and the Limits on Stable Partners

FACTORING POLYNOMIALS IN THE RING OF FORMAL POWER SERIES OVER Z

Basic Probability Concepts

Lecture 3: Finding integer solutions to systems of linear equations

University of Miskolc

Mathematics 31 Pre-calculus and Limits

The degree, size and chromatic index of a uniform hypergraph

A Turán Type Problem Concerning the Powers of the Degrees of a Graph

CS 3719 (Theory of Computation and Algorithms) Lecture 4

MATH10212 Linear Algebra. Systems of Linear Equations. Definition. An n-dimensional vector is a row or a column of n numbers (or letters): a 1.

The Henstock-Kurzweil-Stieltjes type integral for real functions on a fractal subset of the real line

THE CENTRAL LIMIT THEOREM TORONTO

Transcription:

An example of a computable absolutely normal number Verónica Becher Santiago Figueira Abstract The first example of an absolutely normal number was given by Sierpinski in 96, twenty years before the concept of computability was formalized. In this note we give a recursive reformulation of Sierpinski s construction which produces a computable absolutely normal number. Introduction A number which is normal in any scale is called absolutely normal. The existence of absolutely normal numbers was proved by E. Borel. His proof is based on the measure theory and, being purely existential, it does not provide any method for constructing such a number. The first effective example of an absolutely normal number was given by me in the year 96. As was proved by Borel almost all in the sense of measure theory real numbers are absolutely normal. However, as regards most of the commonly used numbers, we either know them not to be normal or we are unable to decide whether they are normal or not. For example we do not know whether the numbers, π, e are normal in the scale of 0. Therefore, though according to the theorem of Borel almost all numbers are absolutely normal, it was by no means easy to construct an example of an absolutely normal number. Examples of such numbers are fairly complicated. M. W. Sierpinski. Elementary Theory of Numbers, Warszawa, 964, p. 77. A number is normal to base q if every sequence of n consecutive digits in its q-base expansion appears with limiting probability q n. A number is absolutely normal if it is normal to every base q. For example, the rational number 0.00000... is not normal to base because although the probability to find is and so is the probability to find 0, the probability to find is not. There are also irrational numbers that Department of Computer Science, FCEyN, University of Buenos Aires, Argentina

are not normal in some base, as 0.000000000000000..., which is not normal to base. Another example is Champernowne s number 0.34567890345... which has all natural numbers in their natural order, written in base 0. It can be proved that Champernowne s number is normal to base 0, but not in some other bases. Let us notice that no rational is absolutely normal: a with a < b is written in base b as 0.a000000000.... b Moreover, every rational r is not normal to any base q [4]: the fractional expansion of r in base q will eventually repeat, say with a period of k, in which case the number r is about as far as being normal to the base q k as it can be. The first example of an absolutely normal number was given by Sierpinski in 96 [5], twenty years before the concept of computability was formalized. Sierpinski determines such a number using a construction of infinite sets of intervals and using the minimum of an uncountable set. Thus, it is a priori unclear whether his number is computable or not. In this note we give a recursive reformulation of Sierpinski s construction which produces a computable absolutely normal number. We actually give an ridiculously exponential algorithm to compute this number. The present work can be related to that of A. Turing [7] where he attempts to show how absolutely normal numbers may be constructed. However, the strategy he used is different from Sierpinski s and it is unclear whether the theorems announced in his paper actually hold. Another example of an absolutely normal but not computable number is Chaitin s random number Ω, the halting probability of a universal machine [3]. Based on his theory of program size Chaitin formalizes the notion of lack of structure and unpredictability in the fractional expansion of a real number, obtaining a definition of randomness stronger than statistical properties of randomness. Although the definition of Ω is known there is no algorithm to exhibit its fractional digits. That is, Ω is not computable. The fundamental constants, like π, and e, are computable and it is widely conjectured [6, ] that they are absolutely normal. However, none of these has even been proved to be normal to base 0, much less to all bases. The same has been conjectured of the irrational algebraic numbers []. In general, we lack an algorithm that decides on absolute normality. Let us recall that a real number is computable if there is a recursive function that calculates each of its fractional digits. Namely, there exists a total recursive f : N N such that for every n, fn is the n th fractional digit of the number in some base. We will use Lebesgue s definition of measure. The measure of the interval I = a, b, with a < b, is denoted by µ I = b a, and the measure of a set of intervals J, denoted by µ J is the measure of J. We will use several properties of Lebesgue s measure, e.g., countable additivity and subadditivity. After presenting Sierpinski s original construction we introduce our algorithmic version of his construction. Then, we discuss Sierpinski s number and we consider other variants defining absolutely normal numbers.

Sierpinski s result of 96 Sierpinski [5] achieves an elementary proof of an important proposition proved by E. Borel that states that almost all real numbers are absolutely normal. At the same time he gives a way to effectively determine one such number. He defines ε as a set of certain open intervals with rational end points. Although the set ε contains countably many intervals, they do not cover the whole of the 0, segment. Sierpinski proves that every real number in 0, that is external to ε is absolutely normal. ε is defined as the union of infinitely many sets of intervals q,m,n,p. The parameter ε is a number in 0, ] used to bound the measure of ε. ε = q q= m= n=n m,qε p=0 q,m,n,p where q ranges over all possible bases, n ranges over the lengths of fractional expansions, p ranges between 0 and q, m allows for arbitrarily small differences in the rate of appearance of the digit p in the fractional expansions, and q,m,n,p is the set of all open intervals of b the form q + b +... + bn, b q q n q n q + b +... + bn + such that q q n q cpb,b,...,b n n n q m where 0 b i q for i n and c p b, b,..., b n represents the number of times that the digit p appears amongst b, b,..., b n. The idea is that q,m,n,p contains all numbers that are not normal to base q. If a number is normal in base q we expect that the rate of appearance of the digit p in a prefix of length n to be as close as possible to. Each interval in q q,m,n,p contains all numbers that written in base q start with 0.b b...b n and the digit p appears in 0.b b...b n at a rate different from. Let us observe that the right end of the intervals in q q,m,n,p add : added q n only would leave the number 0.b q n b...b n... outside the open interval. Each interval in q,m,n,p has measure 3 and for fixed q, m, n, q n q,m,n,p is a finite set. From Sierpinski s proof follows that n m,q ε must be large enough as to imply µ ε < 4m ε; n m,q ε = 6 q + suffices. In order to bound the measure of ε, Sierpinski works ε with the sum of the measures of each interval of q,m,n,p : s ε = q q= m= n=n m,qε p=0 µ I and he proves that µ ε s ε < ε for every ε 0, ]. Sierpinski defines E ε as the set of all real numbers in 0, external to every interval of ε and he proves that for every ε 0, ], every real in E ε is absolutely normal. Since µe ε is greater than or equal to ε, for every ε in 0, ], every real in 0, is absolutely normal with probability. Although the measure of ε is less than ε, no segment c, d with c < d can be completely included in Eε. If this happened there would be infinitely many rationals 3

belonging to Eε, contradicting absolute normality. Thus, Eε is a set of infinitely many isolated irrational points. Sierpinski defines ξ = mine, and in this way he gives an effective determination an absolutely normal number. 3 An algorithm to construct an absolutely normal number Our work is based on one essential observation: we can give a recursive enumeration of Sierpinski s set ε, and we can bound the measure of error in each step. To simplify notation, we will fix a rational ε 0, ] in fact, ε can be any computable real in 0, ] and we rename = ε ; s = s ε ; n m,q = n m,q ε. We define the computable sequence k : k = k+ k k n m,q q q= m= n=n m,q p=0 q,m,n,p. We also define the bounds to the measure of each term in the sequence: s k = k+ k k n m,q q q= m= n=n m,q p=0 µ I It is clear that lim k s k = s and lim k k =. Since s k is the sum of the measures of all intervals belonging to k, we have µ k s k and similarly we have µ s. Besides, for every natural k, s k s. Let us observe that for every pair of natural numbers k and l such that k l, we have k l, and for any k, k. Finally, we define the error of approximating s by s k, r k = s s k. We can give a bound on r k, a result that makes our construction computable. Theorem. For every natural number k, r k < 5ε k. Proof. Let us define S qmn = q s = k+ k k n m,q q= m= n=n m,q S qmn + k+ p=0 k q= m= n=k n m,q+ µ I. By splitting the sums in the definition of s, S qmn + But the first term of is s k, so the rest is r k. r k = k+ k q= m= n=k n m,q+ S qmn + k+ q= m=k+ k+ q= m=k+ n=n m,q n=n m,q S qmn + S qmn + q=k+ m= q=k+ m= n=n m,q S qmn. n=n m,q S qmn. 4

We will bound each of the three terms that appear in this equation. From Sierpinski s proof we know that S qmn < m4. The third term of equation can be bounded by n S qmn < m 4. 3 n m= n=n m,q m= n=n m,q q=k+ Let us now find a bound for the definition of n m,q we have, n m,q = Applying this last result to 3 we have q=k+ m= q=k+. For any i, it holds that n n=n m,q n=n m,q S qmn < ε q=k+ 4m 6 q ε + > 4m6 q. Then, ε n=i+ n <, and by i n=n m,q n < ε 4m 6 q. < ε q m k + < ε k. 4 m= Similarly, the second term of equation can be bounded by k+ S qmn < ε < ε q m k. 5 q= m=k+ n=n m,q q= m=k+ Finally, the first term of equation can be bounded by k+ k S qmn < m 4 q= m= n=k n m,q+ q= m= n n=k n m,q+ < ε k. 6 Replacing in the bounds found in 4,5 and 6 we conclude r k < ε k + ε k + ε k = 5ε k. We now give an overview of our construction of the binary number ν = 0.b b b 3... To determine the first digit of ν we divide the [0, ] interval in two halves, c 0 = [ 0, ] and c = [, ], each of measure. Thinking in base, in c 0 there are only numbers whose first fractional digit is 0 while in c there are only numbers whose first fractional digit is. By Sierpinski s result we know that neither nor any of the k cover the whole segment [0, ]. Of course, all points external to must either be in c 0 or in c. The idea now is to determine a subset of, p, big enough i.e. sufficiently similar to as to ensure that, whenever p does not cover completely a given interval, then does not either. We can guarantee this because we have an upper bound on the error of approximating at every step. We pick the interval c 0 or c, the least covered by p. If we select c 0 then there will be real numbers external to every interval of whose first digit in the binary expansion is 0; therefore, we define b = 0. Similarly, if we select c we define b =. To define the rest of the digits we proceed recursively. We will divide c n b n in two halves defining the intervals c n 0 and c n, each of measure n. At least one of the two will not be 5

completely covered by. The n th digit of ν will be determined by comparing the measure of a suitable set pn restricted to the intervals c n 0 and c n, where the index p n is obtained computably from n. If we select c n 0, then we will define b n to be 0, otherwise b n will be. Since these measures are computable we have obtained an algorithm to define a real number ν, digit by digit, such that ν is external to every interval of. By Sierpinski s result ν is absolutely normal. Before proceeding we shall prove some results. The following proposition gives us a bound on the measure of the sets that have not been enumerated in the step k. Proposition. For every k, µ k r k. Proof. Since k is included in, the measure of k is less than or equal to the sum of the measures of those intervals in but not in k. Hence µ k q q µ I s µ I. q= m= n=n m,q p=0 q= m= n=n m,q p=0 I k But = k+ q q= m= n=n m,q p=0 k k n m,q q q= m= n=n m,q p=0 I k I / k µ I k+ k k n m,q q q= m= n=n m,q p=0 µ I = I k µ I = s k. Hence, µ k s s k = r k. We are also able to bound the measure of the difference between two sets enumerated in different steps. Proposition 3. For any natural numbers k and l such that k l, µ l k r k r l. Proof. µ l k l+ and as k l, l+ l l l n m,q q q= m= n=n m,q p=0 l n m,q q q= m= n=n m,q p=0 I k I / k µ I l+ µ I = s l k+ µ l k s l s k = s s k s s l = r k r l. k k n m,q q= m= n=n m,q p=0 l l n m,q q q= m= n=n m,q p=0 q I k µ I µ I = s k. Thus, Let J be a set of intervals and let c be an interval. We will will denote with J c the restriction of J to c, J c = {x IR : x c j J : x j}. Lemma 4. For any interval c and any natural number k, µ c µ k c + r k. Proof. Since k, for any natural k we have that c k c k. Taking measure we obtain µ c µ k c + µ k. By Proposition we have µ c µ k c + r k. Lemma 5. For any interval c any natural numbers k and l such that k l, µ l c µ k c + r k r l. 6

Proof. Obvious, from Proposition 3 and l c k c l k. Lemma 6. µ k is computable for any k, and µ k c is computable for any k and for any interval c = a, b where a and b are rationals. Proof. k is a finite set of known intervals with rationals end points. An algorithm for the measure of k and for the measure of k c can be easily given. 3. Determination of the first digit We will compute b. We divide the interval [0, ] in two halves c 0 = [ 0, ] and c = [, ]. We will have to determine a suitable index p. It is clear that µ p c 0 + µ p c = µ p s p. Adding r p + r p in each side of this inequality and using the definition of r p gives µ p c 0 + rp + µ p c + rp sp + r p + r p = s + r p < ε + r p. It is impossible that both terms µ p c 0+r p and µ p c +r p be greater or equal to ε+rp. If they were, we would have ε + r p = ε + r p + ε + r p µ p c 0 + rp + µ p c + rp < ε + rp a contradiction. Thus,the following proposition is true µ p c ε + r p 0 < r p µ p c ε + r p < r p Now we determine the value of p. It has to be large enough so that the error r p is sufficiently small to guarantee that even if all the remaining intervals that have not yet been enumerated in step p fall in c b, the whole c b will not be completely covered by. We need µ p c b + rp <. We know that the measure µ p c b + rp < ε+rp. Theorem states that r p < 5ε p. So, for p = 5 we obtain r p < ε. Then, µ p c b +rp < ε+rp = ε + ε < ε. Using Lemma 4 we obtain µ c 4 b < = µ c b. This means that the union of all the intervals belonging to will never cover the whole interval c b, whose measure is. Thus, there exist real numbers belonging to no interval of that fall in the interval c b. These have their first digit equal to b. We define { 0 if µ p c b = 0 µ p c otherwise 7

3. Determination of the n th digit, for n > Let us assume that we have already computed b, b,..., b n and that at each step m, m < n, µ pm c m b m + rpm < m ε + j r m pj and we have chosen p m = 5 m. We have to prove that this condition holds for m = n and p n = 5 n, and that we can computably determine b n. We split the interval c n b n in two halves of measure, c n n 0 = [0.b b... b n, 0.b b... b n ] and c n = [0.b b... b n, 0.b b... b n...]. As they cover the interval c n µ pn c n 0 + µ pn c n = µ pn c n b n. Since p n p n and using Lemma 5 we obtain µ pn c n 0 + µ pn c n µ pn c n b n + r pn r pn. b n, we have Adding r pn + r pn to both sides of this inequality we obtain µ pn c n 0 + r pn + µ pn c n + r pn µ pn c n p n + rpn + r pn and by the previous equation µ pn c n 0 + r pn + µ pn c n + r pn < n ε + n j r pj. Hence, one of the two terms, µ pn c n 0 + r pn or µ pn c n + r pn, must be less than ε + n j r pj. This means that the following proposition is true: n µ p n c n 0 < ε + n j r pj r n pn µ p n c n < ε + n j r pj r n pn. We define b n as the first index i such that the interval c n i is less covered by pn, { 0 if µ pn c b n = n 0 µ pn c n otherwise By Theorem we have n n j j n r pj < ε = ε j < ε. j 8

From the last inequality and from the definition of b n we obtain µ pn c n b n + rpn < n ε + j r n pj < ε n n and using Lemma 4 we deduce µ c n b n < = µ c n n b n. Hence, the set does not cover the interval c n b n. There must be real numbers in the interval c n b n that belong to no interval of. Theorem 7. The number ν is computable and absolutely normal. Proof. In our construction we need only to compute the measure of the sets pn c n b n. Then, by Lemma 6, ν is computable. Let us prove that ν is external to every interval of. Suppose not. Then, there must be an open interval I such that ν I. Consider the intervals c b, c b, c b 3,... By our construction, ν belongs to every c n b n. Let us call c the first interval c n b n of the sequence such that c n b n I. Such an interval exists because the measure of c n b n goes to 0 as n increases. But then the interval c is covered by. This contradicts that in our construction at each step n we choose an interval c n b n not fully covered by. Thus, ν belongs to no interval of, so by Sierpinski s result, ν is absolutely normal. Finally, it follows from Theorem that the bound s on µ is also computable. Corollary 8. The real number s is computable. Proof. Let us define the sequence of rationals in base, a n = s 5ε n. By Theorem we have s a n = s s 5ε n = r 5ε n < n. Then, it is possible to approximate s by a computable sequence of rationals a n such that the first n digits of a n coincide with the first n digits of s. Therefore, s is computable. 3.3 About Sierpinski s ξ number We finish this section with some observations about the absolutely normal number defined by Sierpinski, ξ the first real number external to, that is, ξ = mine. As we see, Sierpinski defines ξ fixing ε =. Since our construction requires ε 0, ], we do not obtain the number ξ. Under a slight modification of our construction we can allow ε to be any computable number in the interval 0,, by defining p n = 5 n ε ε +. Now can speak of the family of numbers that are definable using Sierpinski s notion for different values of ε. Fix ε to be any computable real in 0, and let ξ = mineε. Then ξ is definable in our construction, in binary notation, in the following way: b n = 0 if µ pn c n 0 < n otherwise 9 ε + n j r pj r pn

For each n, ξ either falls in c n 0 or in c n ; therefore, we get a determination of each digit of its binary expansion. If we could prove that ε + n j r n pj r pn is irrational and computable, then we could assert that ξ is computable. Without this assumption we can assert a weaker property: ξ is computably enumerable. A real number is computably enumerable if there is a computable non decreasing sequence of rationals which converges to that number. Every computable number is computably enumerable, but the converse is not true. It is possible that a real number r be approximated from below by a computable non decreasing sequence of rationals but that there is no function which effectively gives each of the fractional digits of r. This happens when it is not possible to bound the error of approximating the number by any computable sequence of rationals. Let us sketch the proof that ξ is computably enumerable, for any computable real ε 0, ]. Since ε is a recursively enumerable set of intervals, we can scan them one by one. At each step we single out the first rational in [0, ] which is external to all the intervals scanned up to the moment. This procedure is computable and determines a non decreasing sequence of rationals which converges to ξ. This proves that ξ is computably enumerable. However, because of the shape of the intervals in ε, it doesn t seem easy to bound the error of this method of approximating ξ. 4 Other computable absolutely normal numbers The construction we gave defines ν, a computable absolutely normal number in base. We can adapt the construction to define numbers in any other bases: To compute a number in a base q, at each step we should divide the interval selected in the previous step in q parts. In the n th step we determine the n th digit defining the intervals c n 0, c n,..., c n q of measure q n, where c n i = [ n i q n b j, i+ q j q n + n b j q j ] for 0 i q. We will choose p n = 5 q n and following the same steps as in the construction of a number in base, there must be an index i such that µ pn c n i < n ε + q j r q n pj r pn. As before, we define b n as the first index corresponding to the interval least covered by pn, { b n = min i : j : 0 j q : µ pn c n i µ } pn c n j. 0 i q In principle, for different bases the numbers will be distinct they will not be ν expressed in different bases, while they will all be examples of computable absolutely normal numbers. The definition of absolute normality is asymptotic, that is, it states a property that has to be true in the limit. Thus, given an absolutely normal number, we can alter it 0

by adding or removing a finite number of digits of its fractional expansion to obtain an absolutely normal number. For example, we could fix an arbitrary number of digits of the fractional expansion and complete the rest with the digits of ν. However, we wonder whether it is possible to obtain an absolutely normal number by fixing a priori infinitely many digits and filling in the free slots. Obviously with only finitely many free slots we may not obtain an absolutely normal number for example, fix all the digits to be 0 except finitely many. Likewise we may not obtain an absolutely normal number by fixing infinitely many digits and leaving free also infinitely many slots: if we fix 0 in the even positions we will never obtain an absolutely normal number because we will never find the string in the fractional expansion. The same thing happens if we fix the digits in the positions which are multiple of 3, 4, etc. It s clear that the possibility to obtain an absolutely normal number depends on the values we assign to the fixed positions if we set the even positions with the digits of the even positions of ν it would be trivial to complete the rest to obtain an absolutely normal number. Finally, we wonder what happens if the digits we fix are progressively far apart, for example in the positions which are powers of or in the positions which are Fibonacci numbers. Let us also note that absolute normality is invariant under permutation of digits. Acknowledgements. The present work was originated in a discussion with Guillermo Martinez, who pointed out that Sierpinski gave the first example of an absolutely normal number in 96. The authors are also indebted to Cristian Calude and Max Dickmann for their valuable comments. References [] D. H. Bailey and R. E. Crandall. On the random character of fundamental constant expansions, http://www.perfsci.com/free/techpapers/freepapers.html, 000. [] E. Borel. Les probabilités dénombrables et leurs applications arithmétiques. Rend. Circ. Mat. Palermo, 7:47 7, 909. [3] G. J. Chaitin. A theory of program size formally identical to information theory. ACM, :39 340, 975. [4] G. Martin. Absolutely Abnormal Numbers, Amer. Math. Monthly, to appear. [5] M. W. Sierpinski. Démonstration élémentaire du théorème de M. Borel sur les nombres absolument normaux et détermination effective d un tel nombre. Bull. Soc. Math. France, 45:7 3, 97. [6] M. W. Sierpinski. Elementary Theory of Numbers, Warszawa, 964. [7] A. Turing. A Note on Normal Numbers. Collected Works of A. M. Turing, Pure Mathematics, edited by J. L. Britton, pp. 7-9. North Holland, 99.