Eulerian number

From Wikipedia, the free encyclopedia

In combinatorics, the Eulerian number A(n, m) is the number of permutations of the numbers 1 to n in which exactly m elements are greater than the previous element (permutations with m "ascents"). They are the coefficients of the Eulerian polynomials:

The Eulerian polynomials are defined by the exponential generating function

The Eulerian polynomials can be computed by the recurrence

An equivalent way to write this definition is to set the Eulerian polynomials inductively by

Other notations for A(n, m) are E(n, m) and .

History[]

Shows the polynomials presently known as Eulerian polynomials in Euler's work from 1755, Institutiones calculi differentialis, part 2, p. 485/6. The coefficients of these polynomials are known as Eulerian numbers.

In 1755, in his book Institutiones calculi differentialis, Leonhard Euler investigated polynomials α1(x) = 1, α2(x) = x + 1, α3(x) = x2 + 4x + 1, etc. (see the facsimile). These polynomials are a shifted form of what are now called the Eulerian polynomials An(x).

Basic properties[]

For a given value of n > 0, the index m in A(n, m) can take values from 0 to n − 1. For fixed n there is a single permutation which has 0 ascents: (n, n − 1, n − 2, ..., 1). There is also a single permutation which has n − 1 ascents; this is the rising permutation (1, 2, 3, ..., n). Therefore A(n, 0) and A(n, n − 1) are 1 for all values of n.

Reversing a permutation with m ascents creates another permutation in which there are n − m − 1 ascents. Therefore A(n, m) = A(n, n − m − 1).

Values of A(n, m) can be calculated "by hand" for small values of n and m. For example

n m Permutations A(n, m)
1 0 (1) A(1,0) = 1
2 0 (2, 1) A(2,0) = 1
1 (1, 2) A(2,1) = 1
3 0 (3, 2, 1) A(3,0) = 1
1 (1, 3, 2) (2, 1, 3) (2, 3, 1) (3, 1, 2) A(3,1) = 4
2 (1, 2, 3) A(3,2) = 1

For larger values of n, A(n, m) can be calculated using the recursive formula

For example

Values of A(n, m) (sequence A008292 in the OEIS) for 0 ≤ n ≤ 9 are:

 m
n 
0 1 2 3 4 5 6 7 8
1 1
2 1 1
3 1 4 1
4 1 11 11 1
5 1 26 66 26 1
6 1 57 302 302 57 1
7 1 120 1191 2416 1191 120 1
8 1 247 4293 15619 15619 4293 247 1
9 1 502 14608 88234 156190 88234 14608 502 1

The above triangular array is called the Euler triangle or Euler's triangle, and it shares some common characteristics with Pascal's triangle. The sum of row n is the factorial n!.

Explicit formula[]

An explicit formula for A(n, m) is[1]

One can see from this formula, as well as from the combinatorial interpretation, that for , so that is a polynomial of degree for .

Summation properties[]

It is clear from the combinatorial definition that the sum of the Eulerian numbers for a fixed value of n is the total number of permutations of the numbers 1 to n, so

The alternating sum of the Eulerian numbers for a fixed value of n is related to the Bernoulli number Bn+1

Other summation properties of the Eulerian numbers are:

where Bn is the nth Bernoulli number.

Identities[]

The Eulerian numbers are involved in the generating function for the sequence of nth powers:

for . This assumes that 00 = 0 and A(0,0) = 1 (since there is one permutation of no elements, and it has no ascents).

Worpitzky's identity[2] expresses xn as the linear combination of Eulerian numbers with binomial coefficients:

It follows from Worpitzky's identity that

Another interesting identity is

The numerator on the right-hand side is the Eulerian polynomial.

For a fixed function which is integrable on we have the integral formula [3]

Eulerian numbers of the second order[]

The permutations of the multiset {1, 1, 2, 2, ···, n, n} which have the property that for each k, all the numbers appearing between the two occurrences of k in the permutation are greater than k are counted by the double factorial number (2n−1)!!. The Eulerian number of the second order, denoted , counts the number of all such permutations that have exactly m ascents. For instance, for n = 3 there are 15 such permutations, 1 with no ascents, 8 with a single ascent, and 6 with two ascents:

332211,
221133, 221331, 223311, 233211, 113322, 133221, 331122, 331221,
112233, 122133, 112332, 123321, 133122, 122331.

The Eulerian numbers of the second order satisfy the recurrence relation, that follows directly from the above definition:

with initial condition for n = 0, expressed in Iverson bracket notation:

Correspondingly, the Eulerian polynomial of second order, here denoted Pn (no standard notation exists for them) are

and the above recurrence relations are translated into a recurrence relation for the sequence Pn(x):

with initial condition

The latter recurrence may be written in a somehow more compact form by means of an integrating factor:

so that the rational function

satisfies a simple autonomous recurrence:

whence one obtains the Eulerian polynomials of second order as Pn(x) = (1−x)2n un(x), and the Eulerian numbers of second order as their coefficients.

Here are some values of the second order Eulerian numbers:

 m
n 
0 1 2 3 4 5 6 7 8
1 1
2 1 2
3 1 8 6
4 1 22 58 24
5 1 52 328 444 120
6 1 114 1452 4400 3708 720
7 1 240 5610 32120 58140 33984 5040
8 1 494 19950 195800 644020 785304 341136 40320
9 1 1004 67260 1062500 5765500 12440064 11026296 3733920 362880

The sum of the n-th row, which is also the value Pn(1), is (2n − 1)!!.

Indexing the second-order Eulerian numbers comes in three flavors: (sequence A008517 in the OEIS) following Riordan and Comtet, (sequence A201637 in the OEIS) following Graham, Knuth, and Patashnik and (sequence A340556 in the OEIS), extending the definition of Gessel and Stanley.

References[]

  • Eulerus, Leonardus [Leonhard Euler] (1755). Institutiones calculi differentialis cum eius usu in analysi finitorum ac doctrina serierum [Foundations of differential calculus, with applications to finite analysis and series]. Academia imperialis scientiarum Petropolitana; Berolini: Officina Michaelis.
  • Carlitz, L. (1959). "Eulerian Numbers and polynomials". Math. Mag. 32 (5): 247–260. doi:10.2307/3029225. JSTOR 3029225.
  • Gould, H. W. (1978). "Evaluation of sums of convolved powers using Stirling and Eulerian Numbers". Fib. Quart. 16 (6): 488–497.
  • Desarmenien, Jacques; Foata, Dominique (1992). "The signed Eulerian numbers". Discrete Math. 99 (1–3): 49–58. doi:10.1016/0012-365X(92)90364-L.
  • Lesieur, Leonce; Nicolas, Jean-Louis (1992). "On the Eulerian Numbers M=max (A(n,k))". Europ. J. Combinat. 13 (5): 379–399. doi:10.1016/S0195-6698(05)80018-6.
  • Butzer, P. L.; Hauss, M. (1993). "Eulerian numbers with fractional order parameters". Aequationes Mathematicae. 46 (1–2): 119–142. doi:10.1007/bf01834003. S2CID 121868847.
  • Koutras, M.V. (1994). "Eulerian numbers associated with sequences of polynomials". Fib. Quart. 32 (1): 44.
  • Graham; Knuth; Patashnik (1994). Concrete Mathematics: A Foundation for Computer Science (2nd ed.). Addison-Wesley. pp. 267–272.
  • Hsu, Leetsch C.; Jau-Shyong Shiue, Peter (1999). "On certain summation problems and generalizations of Eulerian polynomials and numbers". Discrete Math. 204 (1–3): 237–247. doi:10.1016/S0012-365X(98)00379-3.
  • Boyadzhiev, Khristo N. (2007). "Apostol-Bernoulli functions, derivative polynomials and Eulerian polynomials". arXiv:0710.1124 [math.CA].
  • Petersen, T. Kyle (2015). Eulerian Numbers. Birkhäuser Advanced Texts Basler Lehrbücher. Birkhäuser. pp. 3–18. doi:10.1007/978-1-4939-3091-3_1. ISBN 978-1-4939-3090-6.

Citations[]

  1. ^ (L. Comtet 1974, p. 243)
  2. ^ Worpitzky, J. (1883). "Studien über die Bernoullischen und Eulerschen Zahlen". Journal für die reine und angewandte Mathematik. 94: 203–232.
  3. ^ Exercise 6.65 in Concrete Mathematics by Graham, Knuth and Patashnik.

External links[]

Retrieved from ""