Rudin–Shapiro sequence

(Redirected from Rudin-Shapiro sequence)

In mathematics, the Rudin–Shapiro sequence, also known as the Golay–Rudin–Shapiro sequence, is an infinite 2-automatic sequence named after Marcel Golay, Harold S. Shapiro, and Walter Rudin who investigated its properties.[1]

Definition edit

Each term of the Rudin–Shapiro sequence is either   or  . If the binary expansion of   is given by

 

then let

 

(So   is the number of times the block 11 appears in the binary expansion of  .)

The Rudin–Shapiro sequence   is then defined by

 

Thus   if   is even and   if   is odd.[2][3][4]

The sequence   is known as the complete Rudin–Shapiro sequence, and starting at  , its first few terms are:

0, 0, 0, 1, 0, 0, 1, 2, 0, 0, 0, 1, 1, 1, 2, 3, ... (sequence A014081 in the OEIS)

and the corresponding terms   of the Rudin–Shapiro sequence are:

+1, +1, +1, −1, +1, +1, −1, +1, +1, +1, +1, −1, −1, −1, +1, −1, ... (sequence A020985 in the OEIS)

For example,   and   because the binary representation of 6 is 110, which contains one occurrence of 11; whereas   and   because the binary representation of 7 is 111, which contains two (overlapping) occurrences of 11.

Historical motivation edit

The Rudin–Shapiro sequence was introduced independently by Golay,[5][6] Rudin,[7] and Shapiro.[8] The following is a description of Rudin's motivation. In Fourier analysis, one is often concerned with the   norm of a measurable function  . This norm is defined by

 

One can prove that for any sequence   with each   in  ,

 

Moreover, for almost every sequence   with each   is in  ,

 [9]

However, the Rudin–Shapiro sequence   satisfies a tighter bound:[10] there exists a constant   such that

 

It is conjectured that one can take  ,[11] but while it is known that  ,[12] the best published upper bound is currently  .[13] Let   be the n-th Shapiro polynomial. Then, when  , the above inequality gives a bound on  . More recently, bounds have also been given for the magnitude of the coefficients of   where  .[14]

Shapiro arrived at the sequence because the polynomials

 

where   is the Rudin–Shapiro sequence, have absolute value bounded on the complex unit circle by  . This is discussed in more detail in the article on Shapiro polynomials. Golay's motivation was similar, although he was concerned with applications to spectroscopy and published in an optics journal.

Properties edit

The Rudin–Shapiro sequence can be generated by a 4-state automaton accepting binary representations of non-negative integers as input.[15] The sequence is therefore 2-automatic, so by Cobham's little theorem there exists a 2-uniform morphism   with fixed point   and a coding   such that  , where   is the Rudin–Shapiro sequence. However, the Rudin–Shapiro sequence cannot be expressed as the fixed point of some uniform morphism alone.[16]

There is a recursive definition[3]

 

The values of the terms rn and un in the Rudin–Shapiro sequence can be found recursively as follows. If n = m·2k where m is odd then

 
 

Thus u108 = u13 + 1 = u3 + 1 = u1 + 2 = u0 + 2 = 2, which can be verified by observing that the binary representation of 108, which is 1101100, contains two sub-strings 11. And so r108 = (−1)2 = +1.

A 2-uniform morphism   that requires a coding   to generate the Rudin-Shapiro sequence is the following:

 
 

The Rudin–Shapiro word +1 +1 +1 −1 +1 +1 −1 +1 +1 +1 +1 −1 −1 −1 +1 −1 ..., which is created by concatenating the terms of the Rudin–Shapiro sequence, is a fixed point of the morphism or string substitution rules

+1 +1 +1 +1 +1 −1
+1 −1 +1 +1 −1 +1
−1 +1 −1 −1 +1 −1
−1 −1 −1 −1 −1 +1

as follows:

+1 +1 +1 +1 +1 −1 +1 +1 +1 −1 +1 +1 −1 +1 +1 +1 +1 −1 +1 +1 −1 +1 +1 +1 +1 −1 −1 −1 +1 −1 ...

It can be seen from the morphism rules that the Rudin–Shapiro string contains at most four consecutive +1s and at most four consecutive −1s.

The sequence of partial sums of the Rudin–Shapiro sequence, defined by

 

with values

1, 2, 3, 2, 3, 4, 3, 4, 5, 6, 7, 6, 5, 4, 5, 4, ... (sequence A020986 in the OEIS)

can be shown to satisfy the inequality

 [1]

Let   denote the Rudin–Shapiro sequence on  , in which case   is the number, modulo 2, of occurrences (possibly overlapping) of the block   in the base-2 expansion of  . Then the generating function

 

satisfies

 

making it algebraic as a formal power series over  .[17] The algebraicity of   over   follows from the 2-automaticity of   by Christol's theorem.

The Rudin–Shapiro sequence along squares   is normal.[18]

The complete Rudin–Shapiro sequence satisfies the following uniform distribution result. If  , then there exists   such that

 

which implies that   is uniformly distributed modulo   for all irrationals  .[19]

Relationship with one-dimensional Ising model edit

Let the binary expansion of n be given by

 

where  . Recall that the complete Rudin–Shapiro sequence is defined by

 

Let

 

Then let

 

Finally, let

 

Recall that the partition function of the one-dimensional Ising model can be defined as follows. Fix   representing the number of sites, and fix constants   and   representing the coupling constant and external field strength, respectively. Choose a sequence of weights   with each  . For any sequence of spins   with each  , define its Hamiltonian by

 

Let   be a constant representing the temperature, which is allowed to be an arbitrary non-zero complex number, and set   where   is Boltzmann's constant. The partition function is defined by

 

Then we have

 

where the weight sequence   satisfies   for all  .[20]

See also edit

Notes edit

  1. ^ a b John Brillhart and Patrick Morton, winners of a 1997 Lester R. Ford Award (1996). "A Case Study in Mathematical Research: The Golay–Rudin–Shapiro Sequence". Amer. Math. Monthly. 103 (10): 854–869. doi:10.2307/2974610. JSTOR 2974610.{{cite journal}}: CS1 maint: numeric names: authors list (link)
  2. ^ Weisstein, Eric W. "Rudin–Shapiro Sequence". MathWorld.
  3. ^ a b Pytheas Fogg (2002) p.42
  4. ^ Everest et al (2003) p.234
  5. ^ Golay, M.J.E. (1949). "Multi-slit spectrometry". Journal of the Optical Society of America. 39 (437–444): 437–444. doi:10.1364/JOSA.39.000437. PMID 18152021.
  6. ^ Golay, M.J.E. (1951). "Static multislit spectrometry and its application to the panoramic display of infrared spectra". Journal of the Optical Society of America. 41 (7): 468–472. doi:10.1364/JOSA.41.000468. PMID 14851129.
  7. ^ Rudin, W. (1959). "Some theorems on Fourier coefficients". Proceedings of the American Mathematical Society. 10 (6): 855–859. doi:10.1090/S0002-9939-1959-0116184-5.
  8. ^ Shapiro, H.S. (1952). "Extremal problems for polynomials and power series". Master's Thesis, MIT.
  9. ^ Salem, R.; Zygmund, A. (1954). "Some properties of trigonometric series whose terms have random signs". Acta Mathematica. 91: 245–301. doi:10.1007/BF02393433. S2CID 122999383.
  10. ^ Allouche and Shallit (2003) p. 78–79
  11. ^ Allouche and Shallit (2003) p. 122
  12. ^ Brillhart, J.; Morton, P. (1978). "Über Summen von Rudin–Shapiroschen Koeffizienten". Illinois Journal of Mathematics. 22: 126–148. doi:10.1215/ijm/1256048841.
  13. ^ Saffari, B. (1986). "Une fonction extrémale liée à la suite de Rudin–Shapiro". C. R. Acad. Sci. Paris. 303: 97–100.
  14. ^ Allouche, J.-P.; Choi, S.; Denise, A.; Erdélyi, T.; Saffari, B. (2019). "Bounds on Autocorrelation Coefficients of Rudin-Shapiro Polynomials". Analysis Mathematica. 45 (4): 705–726. arXiv:1901.06832. doi:10.1007/s10476-019-0003-4. S2CID 119168430.
  15. ^ Finite automata and arithmetic, Jean-Paul Allouche
  16. ^ Allouche and Shallit (2003) p. 192
  17. ^ Allouche and Shallit (2003) p. 352
  18. ^ Müllner, C. (2018). "The Rudin–Shapiro sequence and similar sequences are normal along squares". Canadian Journal of Mathematics. 70 (5): 1096–1129. arXiv:1704.06472. doi:10.4153/CJM-2017-053-1. S2CID 125493369.
  19. ^ Allouche and Shallit p. 462–464
  20. ^ Allouche and Shallit (2003) p. 457–461

References edit