K-regular sequence

Last updated

In mathematics and theoretical computer science, a k-regular sequence is a sequence satisfying linear recurrence equations that reflect the base-k representations of the integers. The class of k-regular sequences generalizes the class of k-automatic sequences to alphabets of infinite size.

Contents

Definition

There exist several characterizations of k-regular sequences, all of which are equivalent. Some common characterizations are as follows. For each, we take R′ to be a commutative Noetherian ring and we take R to be a ring containing R′.

k-kernel

Let k  2. The k-kernel of the sequence is the set of subsequences

The sequence is (R′, k)-regular (often shortened to just "k-regular") if the -module generated by Kk(s) is a finitely-generated R′-module. [1]

In the special case when , the sequence is -regular if is contained in a finite-dimensional vector space over .

Linear combinations

A sequence s(n) is k-regular if there exists an integer E such that, for all ej > E and 0 ≤ rjkej  1, every subsequence of s of the form s(kejn + rj) is expressible as an R′-linear combination , where cij is an integer, fijE, and 0 ≤ bijkfij  1. [2]

Alternatively, a sequence s(n) is k-regular if there exist an integer r and subsequences s1(n), ..., sr(n) such that, for all 1 ≤ ir and 0 ≤ ak  1, every sequence si(kn + a) in the k-kernel Kk(s) is an R′-linear combination of the subsequences si(n). [2]

Formal series

Let x0, ..., xk  1 be a set of k non-commuting variables and let τ be a map sending some natural number n to the string xa0 ... xae  1, where the base-k representation of x is the string ae  1...a0. Then a sequence s(n) is k-regular if and only if the formal series is -rational. [3]

Automata-theoretic

The formal series definition of a k-regular sequence leads to an automaton characterization similar to Schützenberger's matrix machine. [4] [5]

History

The notion of k-regular sequences was first investigated in a pair of papers by Allouche and Shallit. [6] Prior to this, Berstel and Reutenauer studied the theory of rational series, which is closely related to k-regular sequences. [7]

Examples

Ruler sequence

Let be the -adic valuation of . The ruler sequence ( OEIS:  A007814 ) is -regular, and the -kernel

is contained in the two-dimensional vector space generated by and the constant sequence . These basis elements lead to the recurrence relations

which, along with the initial conditions and , uniquely determine the sequence. [8]

Thue–Morse sequence

The Thue–Morse sequence t(n) ( OEIS:  A010060 ) is the fixed point of the morphism 0 → 01, 1 → 10. It is known that the Thue–Morse sequence is 2-automatic. Thus, it is also 2-regular, and its 2-kernel

consists of the subsequences and .

Cantor numbers

The sequence of Cantor numbers c(n) ( OEIS:  A005823 ) consists of numbers whose ternary expansions contain no 1s. It is straightforward to show that

and therefore the sequence of Cantor numbers is 2-regular. Similarly the Stanley sequence

0, 1, 3, 4, 9, 10, 12, 13, 27, 28, 30, 31, 36, 37, 39, 40, ... (sequence A005836 in the OEIS )

of numbers whose ternary expansions contain no 2s is also 2-regular. [9]

Sorting numbers

A somewhat interesting application of the notion of k-regularity to the broader study of algorithms is found in the analysis of the merge sort algorithm. Given a list of n values, the number of comparisons made by the merge sort algorithm are the sorting numbers, governed by the recurrence relation

As a result, the sequence defined by the recurrence relation for merge sort, T(n), constitutes a 2-regular sequence. [10]

Other sequences

If is an integer-valued polynomial, then is k-regular for every .

The Glaisher–Gould sequence is 2-regular. The Stern–Brocot sequence is 2-regular.

Allouche and Shallit give a number of additional examples of k-regular sequences in their papers. [6]

Properties

k-regular sequences exhibit a number of interesting properties.

Proving and disproving k-regularity

Given a candidate sequence that is not known to be k-regular, k-regularity can typically be proved directly from the definition by calculating elements of the kernel of and proving that all elements of the form with sufficiently large and can be written as linear combinations of kernel elements with smaller exponents in the place of . This is usually computationally straightforward.

On the other hand, disproving k-regularity of the candidate sequence usually requires one to produce a -linearly independent subset in the kernel of , which is typically trickier. Here is one example of such a proof.

Let denote the number of 's in the binary expansion of . Let denote the number of 's in the binary expansion of . The sequence can be shown to be 2-regular. The sequence is, however, not 2-regular, by the following argument. Suppose is 2-regular. We claim that the elements for and of the 2-kernel of are linearly independent over . The function is surjective onto the integers, so let be the least integer such that . By 2-regularity of , there exist and constants such that for each ,

Let be the least value for which . Then for every ,

Evaluating this expression at , where and so on in succession, we obtain, on the left-hand side

and on the right-hand side,

It follows that for every integer ,

But for , the right-hand side of the equation is monotone because it is of the form for some constants , whereas the left-hand side is not, as can be checked by successively plugging in , , and . Therefore, is not 2-regular. [22]

Notes

  1. Allouche and Shallit (1992), Definition 2.1.
  2. 1 2 Allouche & Shallit (1992), Theorem 2.2.
  3. Allouche & Shallit (1992), Theorem 4.3.
  4. Allouche & Shallit (1992), Theorem 4.4.
  5. Schützenberger, M.-P. (1961), "On the definition of a family of automata", Information and Control, 4 (2–3): 245–270, doi: 10.1016/S0019-9958(61)80020-X .
  6. 1 2 Allouche & Shallit (1992, 2003).
  7. Berstel, Jean; Reutenauer, Christophe (1988). Rational Series and Their Languages. EATCS Monographs on Theoretical Computer Science. 12. Springer-Verlag. ISBN   978-3-642-73237-9.
  8. Allouche & Shallit (1992), Example 8.
  9. Allouche & Shallit (1992), Examples 3 and 26.
  10. Allouche & Shallit (1992), Example 28.
  11. Allouche & Shallit (1992), Theorem 2.3.
  12. 1 2 Allouche & Shallit (2003) p. 441.
  13. Allouche & Shallit (1992), Theorem 2.5.
  14. Allouche & Shallit (1992), Theorem 3.1.
  15. Allouche & Shallit (2003) p. 445.
  16. Allouche and Shallit (2003) p. 446.
  17. Allouche and Shallit (2003) p. 441.
  18. Bell, J. (2006). "A generalization of Cobham's theorem for regular sequences". Séminaire Lotharingien de Combinatoire. 54A.
  19. Cobham, A. (1969). "On the base-dependence of sets of numbers recognizable by finite automata". Math. Systems Theory. 3 (2): 186–192. doi:10.1007/BF01746527.
  20. Allouche & Shallit (1992) Theorem 2.10.
  21. Allouche and Shallit (2003) p. 444.
  22. Allouche and Shallit (1993) p. 168–169.

Related Research Articles

Real analysis Mathematics of real numbers and real functions

In mathematics, real analysis is the branch of mathematical analysis that studies the behavior of real numbers, sequences and series of real numbers, and real functions. Some particular properties of real-valued sequences and functions that real analysis studies include convergence, limits, continuity, smoothness, differentiability and integrability.

Sequence Finite or infinite ordered list of elements

In mathematics, a sequence is an enumerated collection of objects in which repetitions are allowed and order matters. Like a set, it contains members. The number of elements is called the length of the sequence. Unlike a set, the same elements can appear multiple times at different positions in a sequence, and unlike a set, the order does matter. Formally, a sequence can be defined as a function from natural numbers to the elements at each position. The notion of a sequence can be generalized to an indexed family, defined as a function from an index set that may not be numbers to another set of elements.

In mathematics, a transcendental number is a number that is not algebraic—that is, not the root of a non-zero polynomial of finite degree with rational coefficients. The best known transcendental numbers are π and e.

In mathematics, specifically in real analysis, the Bolzano–Weierstrass theorem, named after Bernard Bolzano and Karl Weierstrass, is a fundamental result about convergence in a finite-dimensional Euclidean space Rn. The theorem states that each bounded sequence in Rn has a convergent subsequence. An equivalent formulation is that a subset of Rn is sequentially compact if and only if it is closed and bounded. The theorem is sometimes called the sequential compactness theorem.

Extreme value theorem Continuous real function on a closed interval has a maximum and a minimum

In calculus, the extreme value theorem states that if a real-valued function is continuous on the closed interval , then must attain a maximum and a minimum, each at least once. That is, there exist numbers and in such that:

In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou.

In mathematics, specifically functional analysis, Mercer's theorem is a representation of a symmetric positive-definite function on a square as a sum of a convergent sequence of product functions. This theorem, presented in, is one of the most notable results of the work of James Mercer (1883–1932). It is an important theoretical tool in the theory of integral equations; it is used in the Hilbert space theory of stochastic processes, for example the Karhunen–Loève theorem; and it is also used to characterize a symmetric positive semi-definite kernel.

In additive number theory, the Schnirelmann density of a sequence of numbers is a way to measure how "dense" the sequence is. It is named after Russian mathematician Lev Schnirelmann, who was the first to study it.

The Arzelà–Ascoli theorem is a fundamental result of mathematical analysis giving necessary and sufficient conditions to decide whether every sequence of a given family of real-valued continuous functions defined on a closed and bounded interval has a uniformly convergent subsequence. The main condition is the equicontinuity of the family of functions. The theorem is the basis of many proofs in mathematics, including that of the Peano existence theorem in the theory of ordinary differential equations, Montel's theorem in complex analysis, and the Peter–Weyl theorem in harmonic analysis and various results concerning compactness of integral operators.

In mathematics, specifically order theory, a well-quasi-ordering or wqo is a quasi-ordering such that any infinite sequence of elements from contains an increasing pair with .

In mathematics, Hurwitz's automorphisms theorem bounds the order of the group of automorphisms, via orientation-preserving conformal mappings, of a compact Riemann surface of genus g > 1, stating that the number of such automorphisms cannot exceed 84(g − 1). A group for which the maximum is achieved is called a Hurwitz group, and the corresponding Riemann surface a Hurwitz surface. Because compact Riemann surfaces are synonymous with non-singular complex projective algebraic curves, a Hurwitz surface can also be called a Hurwitz curve. The theorem is named after Adolf Hurwitz, who proved it in.

In mathematics and theoretical computer science, an automatic sequence is an infinite sequence of terms characterized by a finite automaton. The n-th term of an automatic sequence a(n) is a mapping of the final state reached in a finite automaton accepting the digits of the number n in some fixed base k.

In mathematics the Baum–Sweet sequence is an infinite automatic sequence of 0s and 1s defined by the rule:

Anatoly Karatsuba Russian mathematician

Anatoly Alexeyevich Karatsuba was a Russian mathematician working in the field of analytic number theory, p-adic numbers and Dirichlet series.

In mathematics, the Rudin–Shapiro sequence, also known as the Golay–Rudin–Shapiro sequence, is an infinite 2-automatic sequence named after Marcel Golay, Walter Rudin, and Harold S. Shapiro, who independently investigated its properties.

In mathematics, dimension theory is the study in terms of commutative algebra of the notion dimension of an algebraic variety. The need of a theory for such an apparently simple notion results from the existence of many definitions of the dimension that are equivalent only in the most regular cases. A large part of dimension theory consists in studying the conditions under which several dimensions are equal, and many important classes of commutative rings may be defined as the rings such that two dimensions are equal; for example, a regular ring is a commutative ring such that the homological dimension is equal to the Krull dimension.

In computer science, the complexity function of a word or string is the function that counts the number of distinct factors of that string. More generally, the complexity function of a formal language counts the number of distinct words of given length.

In mathematics and theoretical computer science, a pattern is an unavoidable pattern if it is unavoidable on any finite alphabet.

In mathematics, a summability kernel is a family or sequence of periodic integrable functions satisfying a certain set of properties, listed below. Certain kernels, such as the Fejér kernel, are particularly useful in Fourier analysis. Summability kernels are related to approximation of the identity; definitions of an approximation of identity vary, but sometimes the definition of an approximation of the identity is taken to be the same as for a summability kernel.

Cobham's theorem is a theorem in combinatorics on words that has important connections with number theory, notably transcendental numbers, and automata theory. Informally, the theorem gives the condition for the members of a set S of natural numbers written in bases b1 and base b2 to be recognised by finite automata. Specifically, consider bases b1 and b2 such that they are not powers of the same integer. Cobham's theorem states that S written in bases b1 and b2 is recognised by finite automata if and only if S is a finite union of arithmetic progressions. The theorem was proved by Alan Cobham in 1969 and has since given rise to many extensions and generalisations.

References