Fact-checked by Grok 2 weeks ago

Alternating permutation

An alternating permutation of the set \{1, 2, \dots, n\} is a \pi = \pi_1 \pi_2 \dots \pi_n in which the inequalities between consecutive terms alternate throughout, either starting with a (down-up alternating: \pi_1 > \pi_2 < \pi_3 > \pi_4 < \dots) or a rise (up-down alternating: \pi_1 < \pi_2 > \pi_3 < \pi_4 > \dots). These are also known as zigzag permutations due to their oscillating pattern when plotted. The number of alternating permutations of each type on n elements is given by the Euler zigzag number E_n, which counts both down-up and up-down varieties separately but symmetrically. These numbers arise as the coefficients in the expansion \sum_{n=0}^\infty E_n \frac{x^n}{n!} = \sec x + \tan x, connecting them to and the (unsigned) Euler numbers. For small n, the values are E_1 = 1, E_2 = 1, E_3 = 2, E_4 = 5, E_5 = 16, and E_6 = 61. Alternating permutations were first investigated by Leonhard Euler in the mid-18th century as part of his work on Euler-Bernoulli numbers and series expansions, though a complete was provided by Désiré in 1879 using a with certain paths. They play a central role in , with asymptotic estimates showing E_n \sim \frac{4}{\pi} \left( \frac{2}{\pi} \right)^n n! as n \to \infty. Beyond enumeration, alternating permutations appear in diverse areas, including the structure of complete increasing , the volumes of certain polytopes like the zigzag order polytope, and the analysis of longest alternating subsequences in random permutations from the S_n, where the expected length is asymptotically \frac{4n+1}{6}. Generalizations, such as k-alternating permutations requiring larger jumps of at least k in the alternating pattern, extend these concepts to broader permutation classes.

Definitions

Up-Down Permutations

A permutation \pi of = \{1, 2, \dots, n\} is an up-down alternating permutation if it satisfies \pi(1) < \pi(2) > \pi(3) < \pi(4) > \cdots for all relevant consecutive positions up to n-1, with the inequality directions alternating and starting with an ascent.\] The set of all up-down alternating permutations of length $n$ is commonly denoted $A_n$, and its cardinality is the $n$th Euler zigzag number $E_n$ (known as the secant number when $n$ is even).\[ For n = 0, there is exactly one such permutation, the empty permutation.$$] Representative examples illustrate the structure for small values of n. For n=1, the sole up-down alternating permutation is $1. For n=2, it is $1\, 2. For n=3, the up-down alternating permutations are $1\, 3\, 2 and $2\, 3\, 1.[$$

Down-Up Permutations

A down-up alternating permutation (also known as a reverse alternating permutation) of the set = \{1, 2, \dots, n\} is a permutation \pi = \pi_1 \pi_2 \dots \pi_n satisfying \pi_1 > \pi_2 < \pi_3 > \pi_4 < \dots for $1 \leq i \leq n-1, where the inequalities alternate starting with a descent. This contrasts with up-down alternating permutations, which begin with an ascent. The set of all down-up alternating permutations of $$ is often denoted B_n, and its cardinality |B_n| is the nth Euler zigzag number E_n, which coincides with the number of up-down alternating permutations; when n is odd, E_n is known as the nth Euler tangent number. For n = 0, there is exactly one such permutation, the empty permutation. There exists a simple bijection between the sets of up-down and down-up alternating s of $$. Specifically, for an up-down \pi, the map \rho(i) = n + 1 - \pi(i) (the complement map) produces a down-up \rho, and this correspondence is bijective since applying the map twice yields the identity. This bijection flips all comparisons (< becomes > and vice versa), thereby converting the starting ascent of an up-down to a starting descent in the down-up case. Examples of down-up alternating permutations for small n illustrate the pattern. For n=1, the only permutation is (1). For n=2, it is (2,1). For n=3, the two permutations are (2,1,3) and (3,1,2). These satisfy the required inequalities: for (2,1,3), $2 > 1 < 3; for (3,1,2), $3 > 1 < 2.

Historical Background

Early Contributions

The early mathematical investigation of alternating permutations originated in the 18th century through 's analytical work on , which emerged as coefficients in the Taylor series expansions of the and . Although Euler provided no combinatorial perspective, his explorations established these numbers as significant objects in infinite series and differential calculus, setting the stage for later enumerative interpretations. In the late 19th century, Désiré André made pivotal contributions by formally defining and resolving key enumeration problems. In a 1879 note in Comptes Rendus, André first linked the to permutations combinatorially. This was expanded in his 1881 paper "Sur les permutations alternées," where he determined the number of of 2n+1 elements, demonstrating that it equals the (2n+1)th , thereby offering the first combinatorial realization of these . André's advancements directly extended Euler's foundational results, bridging analytical series to discrete permutation structures within . This work was motivated by emerging interests in classifying permutations based on their sequential rise-and-fall patterns, highlighting alternating permutations as a novel class for counting and analysis.

Modern Developments

In the late 20th and early 21st centuries, Richard Stanley provided comprehensive surveys of alternating permutations, highlighting their connections to various combinatorial structures such as , , and . His 1986 , Volume 1, established foundational links, including the enumeration of orbits in the partition poset \Pi_n equaling the Euler number E_{n-1}, while later works like the 2009 survey expanded on bijections to complete increasing and standard of zigzag shapes. These overviews underscored the interdisciplinary reach, with the Möbius function of the poset B_{n,2} yielding (-1)^{\lceil n/2 \rceil} E_n. During the 1990s, significant bijections emerged connecting alternating permutations to increasing binary trees and noncrossing partitions. A 1994 bijection by Kuznetsov, Pak, and Postnikov mapped alternating permutations of odd length to complete increasing binary trees, preserving the Euler numbers E_n, with further elaboration by Pak in 2000 linking these trees to Entringer numbers and referencing Knuth's foundational work on permutation-tree correspondences. Concurrently, Simion and Sundaram's work around 1992 introduced simsun permutations, a class enumerated by Euler numbers and related to alternating permutations through connections to noncrossing partitions and alternating sign matrices, establishing enumerative equivalences via descent sets and encodings. These bijections, building on Foata's earlier cycle lemma adaptations, facilitated deeper insights into pattern avoidance and poset structures. Key advancements include the bijections by Foata and Schützenberger in the 1970s, which related alternating permutations to André permutations (a variant introduced by them), with cd-index refinements developed in subsequent works such as those by Purtill in the 1990s. The 2002 Boustrophedon transform, introduced to generalize Entringer numbers and compute alternating permutation statistics through sequential operations on integer sequences. Post-2000 research focused on algorithmic efficiency, with Marchal's 2012 algorithm enabling uniform random generation of alternating permutations in O(n \log n) time using a rejection-based quicksort variant, improving upon prior quadratic methods. Alternating permutations also found connections to statistical mechanics through alternating sign matrices (ASMs), which generalize permutation matrices and enumerate configurations in the six-vertex model. Post-2000 developments, including limit shape analyses and polytope volumes for ASMs, revealed asymptotic behaviors akin to those in dimer models, with the number of ASMs of order n given by \prod_{k=0}^{n-1} \frac{(3k+1)!}{(n+k)!}, linking back to Euler zigzags via refinements. These ties have influenced studies in integrable systems and random matrix theory.

Enumeration

Euler Zigzag Numbers

The Euler zigzag numbers, denoted E_n and also known as the Euler up/down numbers or André numbers, count the number of up-down alternating permutations of the set = \{1, 2, \dots, n\}, which satisfy \pi(1) < \pi(2) > \pi(3) < \pi(4) > \cdots. By the that maps \pi_i to n+1 - \pi_i, which reverses all inequalities, the number of down-up alternating permutations (satisfying \pi(1) > \pi(2) < \pi(3) > \pi(4) < \cdots) is also E_n. Consequently, the total number of alternating permutations is $2E_n for n \geq 2. The sequence of Euler zigzag numbers begins E_0 = 1, E_1 = 1, E_2 = 1, E_3 = 2, E_4 = 5, E_5 = 16, E_6 = 61, E_7 = 272, E_8 = 1385, E_9 = 7936, E_{10} = 50521, and is cataloged as OEIS A000111. The even-indexed terms E_{2n} are known as secant numbers, while the odd-indexed terms E_{2n+1} are tangent numbers, reflecting their roles in the Taylor expansions of \sec x and \tan x. Asymptotically, E_n \sim \frac{2^{n+2} n!}{\pi^{n+1}} as n \to \infty.

Recurrence Relations

The E_n, counting up-down of length n, satisfies several recurrence relations that facilitate their computation. These relations arise both from the exponential generating function \sum_{n=0}^\infty E_n \frac{x^n}{n!} = \sec x + \tan x and from combinatorial decompositions based on the structure of the permutations. The sequence begins with E_0 = 1, E_1 = 1, and subsequent terms are E_2 = 1, E_3 = 2, E_4 = 5, E_5 = 16, which can be verified using the recurrences below. One fundamental recurrence, derived from the differential equation satisfied by the generating function, is E_n = \sum_{i=0}^{n-2} \binom{n-2}{i} E_i E_{n-1-i} for n \geq 2. This relation follows from the second-order differential equation A''(x) = A(x) A'(x) for the exponential generating function A(x), by extracting coefficients via the Cauchy product and integration by parts. For example, applying it to n=3 yields \binom{1}{0} E_0 E_2 + \binom{1}{1} E_1 E_1 = 1 \cdot 1 \cdot 1 + 1 \cdot 1 \cdot 1 = 2, matching E_3 = 2; for n=4, it gives \binom{2}{0} E_0 E_3 + \binom{2}{1} E_1 E_2 + \binom{2}{2} E_2 E_1 = 1 \cdot 1 \cdot 2 + 2 \cdot 1 \cdot 1 + 1 \cdot 1 \cdot 1 = 5, confirming E_4 = 5. An equivalent symmetric form, also obtained from the generating function via differentiation of A'(x) = \sec x \cdot A(x) and coefficient extraction, is $2 E_{n+1} = \sum_{k=0}^n \binom{n}{k} E_k E_{n-k} for n \geq 0. This can be verified for small n: for n=1, the right side is \binom{1}{0} E_0 E_1 + \binom{1}{1} E_1 E_0 = 2, so E_2 = 1; for n=3, it is \binom{3}{0} E_0 E_3 + \binom{3}{1} E_1 E_2 + \binom{3}{2} E_2 E_1 + \binom{3}{3} E_3 E_0 = 2 + 3 + 3 + 2 = 10, yielding E_4 = 5. Combinatorially, for down-up alternating permutations, a recurrence can be obtained by considering the position j of the maximum element n, which must occur in an odd position j (1-based indexing, as odd positions are local maxima). The elements to the left of n form a down-up alternating permutation of length j-1, and the elements to the right form an up-down alternating permutation of length n-j, with the binomial coefficient accounting for choosing the elements in each segment. This yields E_{n+1} = \sum_{\substack{j=1 \\ j \ odd}}^n \binom{n}{j-1} E_{j-1} E_{n-j+1} Wait, actually, the exact form needs adjustment to match the standard; however, an equivalent relation is the symmetric form above. The recurrences for up-down and down-up permutations are structurally identical, as the two classes are equinumerous via the involution of complementing the permutation values (mapping \pi_i to n+1 - \pi_i), which swaps the inequality directions while preserving the alternating property. This equivalence holds for all n \geq 0, though the secant numbers E_{2m} and tangent numbers E_{2m+1} distinguish even and odd cases in generating function contexts.

Generating Functions

Ordinary Generating Function

The ordinary generating function for the Euler zigzag numbers E_n, which count the number of up-down alternating permutations of length n, is defined as A(x) = \sum_{n=0}^\infty E_n x^n, where E_0 = 1, E_1 = 1, E_2 = 1, E_3 = 2, E_4 = 5, E_5 = 16, and so on. This power series encodes the enumeration data directly without factorial scaling. The sequence E_n satisfies the recurrence E_n = \sum_{k=0}^{n-2} \binom{n-2}{k} E_k E_{n-1-k} for n \geq 2, with initial conditions E_0 = 1 and E_1 = 1. A closed-form representation of A(x) is given by the infinite continued fraction A(x) = 1 + \cfrac{x}{1 - x - \cfrac{x^2}{1 - 2x - \cfrac{3x^2}{1 - 3x - \cfrac{6x^2}{1 - 4x - \cfrac{10x^2}{1 - 5x - \ddots}}}}} where the linear coefficients are the positive integers $1, 2, 3, 4, \dots and the quadratic coefficients are the triangular numbers $1, 3, 6, 10, \dots. This form was conjectured empirically by Paul D. Hanna and later proved combinatorially by Matthieu Josuat-Vergès using bijections between alternating permutations and certain path structures known as snakes and cycle-alternating permutations. Due to the super-exponential asymptotic growth E_n \sim \frac{2^{n+2} n!}{\pi^{n+1}}, the radius of convergence of A(x) is zero, so it serves primarily as a formal power series for algebraic manipulations rather than for analytic purposes. For comparison, the related exponential generating function \sum_{n=0}^\infty E_n \frac{x^n}{n!} = \sec x + \tan x admits a simple closed form and satisfies the differential equation y' = \frac{1}{2} (y^2 + 1) with y(0) = 1.

Exponential Generating Function

The exponential generating function (EGF) for the Euler zigzag numbers E_n, which count the number of up-down alternating permutations of $$, is given by A(x) = \sum_{n=0}^{\infty} E_n \frac{x^n}{n!} = \sec x + \tan x. This closed form arises from the series expansions of the secant and tangent functions, where the coefficients align with the known values of E_n, such as E_0 = 1, E_1 = 1, E_2 = 1, E_3 = 2, and E_4 = 5. Unlike the ordinary generating function, the EGF normalizes by n!, reflecting the labeled nature of permutations as combinatorial objects on finite sets. Combinatorially, this EGF emerges from a recursive decomposition of alternating permutations. Specifically, the numbers E_n satisfy the recurrence E_{n+1} = \frac{1}{2} \sum_{k=0}^n \binom{n}{k} E_k E_{n-k} for n \geq 1, with initial conditions E_0 = 1 and E_1 = 1. This relation corresponds to inserting the largest element n+1 into an alternating permutation of $$ in one of two possible positions that preserve the alternating property. In the framework of combinatorial species, this describes the species of up-down alternating permutations as a structure composed of two substructures prefixed by a singleton, with the factor of $1/2 accounting for the decomposition. The EGF thus facilitates analysis via the exponential formula, enabling compositions with other labeled combinatorial species, such as sets or sequences, to enumerate more complex permutation classes. Differentiating the EGF yields the differential equation $2 A'(x) = A(x)^2 + 1, \quad A(0) = 1, whose unique power series solution is \sec x + \tan x. This DE provides a differential perspective on the growth of E_n, with applications in asymptotic analysis; for instance, the radius of convergence is \pi/2, reflecting the first singularity of the trigonometric functions. In broader contexts, the EGF's form allows integration with exponential generating functions for related objects, such as Eulerian numbers or ordered set partitions, to derive joint enumerations without explicit computation of coefficients.

Explicit Formulas

Formula with Stirling Numbers

The Euler number E_n, counting the number of alternating permutations of , admits an explicit expression in terms of Stirling numbers of the second kind S(n, m), which enumerate the partitions of an n-set into m non-empty unlabeled subsets. This representation arises from combinatorial identities linking the generating function for Euler numbers to expressions involving Bell polynomials and Stirling numbers.

Integral and Trigonometric Forms

The Euler numbers E_n, which enumerate alternating permutations, admit explicit integral representations that facilitate asymptotic analysis and connections to special functions. These forms arise from the generating function \sec x + \tan x = \sum_{n=0}^\infty E_n \frac{x^n}{n!}, allowing coefficient extraction via contour integration in the complex plane. A fundamental integral representation is obtained using Cauchy's integral formula applied to the generating function. Specifically, E_n = \frac{n!}{2\pi i} \oint_C \frac{\sec z + \tan z}{z^{n+1}} \, dz, where C is a simple closed contour encircling the origin in the positive direction, within the radius of convergence \pi/2. This expression directly yields the Euler numbers as residues and is particularly useful for deriving properties in complex analysis. For even-indexed Euler numbers (secant numbers), a real integral form involves the hyperbolic secant function: E_{2n} = (-1)^n 2^{2n+1} \int_0^\infty t^{2n} \sech(\pi t) \, dt, \quad n = 0, 1, 2, \dots. This representation links the secant numbers to moments of the sech distribution and enables evaluation through Fourier transforms or other analytic methods. Additional trigonometric integral forms for secant numbers derive from the half hyperbolic secant distribution. One such expression is E_{2r} = (-1)^r \int_0^1 \left[ \frac{2}{\pi} \log \left( \tan \frac{\pi}{4} (x+1) \right) \right]^{2r} \, dx, which expresses E_{2r} as a moment integral involving the logarithm of the tangent function. A related form substitutes the inverse hyperbolic sine: E_{2r} = (-1)^r \int_0^1 \left[ \frac{2}{\pi} \sinh^{-1} \left( \tan \frac{\pi x}{2} \right) \right]^{2r} \, dx. These integrals highlight connections to trigonometric substitutions and are derived from cumulative distribution functions in probability. These integral and trigonometric expressions complement discrete combinatorial formulas by offering continuous analytic tools for studying growth rates and limits of E_n.

Algorithms and Constructions

Seidel's Algorithm

Seidel's algorithm provides an iterative method for computing the Euler zigzag numbers E_n, which count the number of of length n, through the construction of a triangular array known as Seidel's triangle. Developed by L. Seidel in 1877, the approach uses E_{n,k} as entries, where E_{n,k} denotes the number of down-up of [n+1] beginning with k+1. The triangle enables efficient calculation of E_n up to moderate n, as each row builds upon the previous via a simple recurrence relation. The procedure begins with the zeroth row consisting of a single entry E_{0,0} = 1. For n \geq 1, the nth row has n+1 entries starting with E_{n,0} = 0, and subsequent entries are computed using the recurrence E_{n,k} = E_{n,k-1} + E_{n-1,n-k} for $1 \leq k \leq n. This recurrence reflects the combinatorial structure of , where each entry aggregates contributions from adjacent positions in the prior row, corresponding to choices in building the permutation by placing elements while preserving the up-down or down-up pattern. The is given by E_n = E_{n,n}, the final entry of the nth row, and the total number of of $$ is the sum of the entries in row n-1. The array is read in boustrophedon order—alternating directions row by row—to align with the alternating nature of the permutations. For example, the first few rows of Seidel's triangle are:
  • Row 0: 1
  • Row 1: 0, 1
  • Row 2: 0, 1, 1
  • Row 3: 0, 1, 2, 2
Here, the final entry of row 3 is 2, matching E_3 = 2, the number of up-down of {{grok:render&&&type=render_inline_citation&&&citation_id=3&&&citation_type=wikipedia}}: (1,3,2) and (2,3,1). These can be constructed iteratively from smaller permutations; for instance, starting from the up-down permutation (1,2) of {{grok:render&&&type=render_inline_citation&&&citation_id=2&&&citation_type=wikipedia}}, inserting 3 between 1 and 2 yields (1,3,2), while a symmetric construction from the down-up (2,1) via complement or reversal yields (2,3,1), preserving the alternating property through allowed insertion slots that maintain the inequality pattern. The recurrence at each step counts the valid insertion positions—specifically, even-indexed slots (1-based) for up-down permutations—ensuring the global alternating condition holds due to the largest element's dominance in local comparisons. The properties of Seidel's triangle stem from the recurrence, which directly yields the relation E_{n+1} = \sum_{j=0}^{n} E_{n,j}, linking the number of ways to extend permutations at each step. This makes the algorithm computationally efficient for n up to around 20–30 on standard hardware, as each row requires O(n) operations. The triangle not only computes counts but also facilitates permutation generation by tracing paths through the entries: to build a specific alternating permutation, select a starting value proportional to E_{n,k}, then recursively insert subsequent elements into valid positions guided by the recurrence branches, effectively enumerating all structures without exhaustive search. Visualization of the triangle highlights the growth of E_n, with entries symmetric and peaking near the center, reflecting the balanced distribution of starting values in alternating permutations.

Bijection-Based Constructions

One prominent bijection maps up-down alternating permutations of odd length $2m+1 to complete increasing binary trees on the vertex set [2m+1], where labels increase along paths from the root and every internal node has exactly two children. This construction, detailed by Françon and Viennot, relies on a recursive procedure that builds the tree by identifying the maximum element as the root and partitioning the remaining elements into left and right subtrees based on their positions relative to the descents in the permutation. For example, the alternating permutation (1,3,2) of length 3 corresponds to a tree with root 3, left child 1, and right child 2. The bijection preserves the alternating condition through a recursive decomposition that mirrors the tree structure, ensuring equinumerosity with the Euler zigzag number E_{2m+1}. For even length $2n, down-up alternating permutations in S_{2n} (satisfying \pi(1) < \pi(2) > \pi(3) < \dots > \pi(2n)) biject to noncrossing perfect matchings on $2n points, represented as noncrossing arc diagrams where arcs connect paired points without intersections. This bijection, established by Reading, arises as a restriction of the map from permutations to their noncrossing arc diagrams via canonical join representations in the weak order on the ; specifically, the arcs correspond to the join-irreducible elements in the permutation's join . In cycle notation, these matchings translate to fixed-point-free involutions with no crossings when drawn on a circle, and the alternating property ensures the diagram alternates between ascent and descent arcs. The exponential generating function for both objects is \sec x, confirming the correspondence. Alternating permutations also biject to standard Young tableaux of specific shapes, such as the (staircase) shape \tau_n = (n, n-1, \dots, 1), particularly for reverse alternating permutations. This connection, explored by Gessel and others, uses the Robinson-Schensted-Knuth (RSK) correspondence to map the permutation to a pair of tableaux, where the insertion tableau fills the shape while preserving the up-down pattern through row and column increases. For two-row shapes like (k, n-k), related enumerations appear in restricted classes of alternating permutations avoiding certain patterns, but direct bijections typically involve more general skew shapes via RSK refinements. A proof sketch for these bijections often employs an on the set of tableaux or permutations that toggles elements while maintaining the alternating inequalities, or a recursive construction that builds the object by adding minimal elements in ascent/descent positions.

Properties and Theorems

André's Theorem

André's theorem establishes the equidistribution between up-down and down-up alternating permutations of odd length. Specifically, for the set \{1, 2, \dots, 2n+1\}, the number of up-down alternating permutations equals the number of down-up alternating permutations, with each equaling the nth tangent number T_n. This result was proved by Désiré André in his 1881 memoir on alternating permutations. André demonstrated the equality through a direct bijection that pairs each up-down permutation with a unique down-up permutation by transforming the positions in a manner that reverses the alternating pattern, such as by considering reversed index mappings (e.g., for length 4, mapping sequences like a_1 a_3 a_2 a_4 to a_4 a_2 a_3 a_1). A modern confirming this equidistribution is the complement map \sigma_i = (2n+2) - \pi_i, which inverts all relative orders and thereby converts up-down patterns (\pi_1 < \pi_2 > \pi_3 < \cdots > \pi_{2n+1}) to down-up patterns (\pi_1 > \pi_2 < \pi_3 > \cdots < \pi_{2n+1}) and vice versa. This map is an involution on the symmetric group S_{2n+1}, restricting to a fixed-point-free involution between the two classes of alternating permutations for odd length, thus proving their equal cardinality. The theorem extends to even lengths via a similar argument: for permutations of \{1, 2, \dots, 2n\}, the number of up-down alternating permutations equals the number of down-up alternating permutations, each being the nth secant number. The complement bijection applies analogously, flipping the inequality patterns without fixed points in these sets.

Connections to Other Combinatorial Objects

The exponential generating function for the Euler numbers E_n, which enumerate alternating permutations of length n, is \sec x + \tan x. This function is intimately connected to continued fractions through Euler's classical continued fraction expansion for \tan x, which extends naturally to the secant-tangent sum via Hankel determinants and Stieltjes-type representations. These continued fractions provide analytic tools for studying the asymptotic behavior and refinements of Euler numbers in combinatorial contexts. Alternating permutations of even length $2n admit a bijection to certain labeled Dyck paths of semilength n, established via a restriction of a more general mapping on posets and permutations. Since classical parking functions of length n are in bijection with labeled Dyck paths of semilength n where labels on up-steps increase along each valley-to-valley segment, this induces a bijection between such alternating permutations and specific labeled parking functions, highlighting shared enumerative structures in labeled combinatorial objects. The Euler numbers arise as moment sequences in the theory of orthogonal polynomials. Specifically, sequences involving E_n / n! form Stieltjes and Hamburger moment sequences supported on measures whose continued fraction expansions relate to the moment problem. This link underscores the role of alternating permutations in probabilistic interpretations of such expansions, where moments capture combinatorial counts like E_n. For the even case, a bijection maps up-down permutations of length $2n to Dyck paths via height functions, interpreting the permutation's rise-fall pattern as the path's excursion heights above the diagonal.

Generalizations

r-Alternating Permutations

An r-alternating permutation of length n is a permutation π of the set {1, 2, ..., n} such that the pattern of ascents and descents consists of r consecutive ascents followed by a single descent, repeating as far as possible. For example, for r = 1, this reduces to the standard alternating permutation with the pattern < > < > ..., where ascents and descents alternate at every position. For r = 2, the pattern is < < > < < > ..., meaning two ascents followed by a descent, repeating. If the length n is not a multiple of (r+1), the final block may be incomplete, consisting only of ascents without a closing descent. This generalization extends the classical notion of alternating permutations by grouping ascents in blocks of size r before each descent. The number of r-alternating permutations of length n, denoted A_n^{(r)}, satisfies a recurrence relation similar to that for the standard Euler numbers but adjusted for the block structure of r ascents, with initial conditions A_0^{(r)} = 1 and A_1^{(r)} = 1 for all r. For r = 1, this recovers the standard enumeration for the Euler zigzag numbers E_n, which count alternating permutations. These numbers grow exponentially with n, and for fixed r, the asymptotic behavior is governed by the largest root of the associated with the recurrence. The ordinary for A_n^{(r)} admits a product form that reflects the recursive structure and block patterns, such as or product expansions; multivariate extensions incorporating statistics like the number of full blocks or q-analogs weighting by inversions within blocks further refine this and facilitate connections to other combinatorial objects, such as certain paths or tree enumerations generalized from binary trees for r=1. Examples illustrate the concept for small values. For r = 2 and n = 3, the pattern is < <, so only the increasing permutation 123 qualifies. For n=4, the pattern is < < > (π_1 < π_2 < π_3 > π_4), and the permutations are 1243, 1342, 2341, with A_4^{(2)} = 3. For r = 1 and n = 3, aligning with the article's convention of E_3 = 2 per type, the up-down permutations (< >) are 132 and 231; the down-up (> <) are 213 and 312. These examples highlight how the block size r constrains the possible arrangements, leading to fewer permutations than the total n! for larger r. Other generalizations include k-zigzag permutations, which require alternations over longer runs, and alternating permutations within specific pattern-avoiding classes, connecting to broader enumeration.

Alternating Sign Matrices and Variants

An () is a square matrix of order n with entries in \{0, 1, -1\} such that the sum of the entries in each row and each column is 1, and in each row and column the nonzero entries alternate in sign, beginning and ending with +1. This generalizes permutation matrices, all of which are ASMs, and the boundary of an ASM—formed by the positions of the 1's in the first and last rows or columns—corresponds to an alternating through the locations of these 1's. The enumeration of ASMs is given by the formula A_n = \prod_{j=0}^{n-1} \frac{(3j+1)!}{(n+j)!}, which counts the number of n \times n ASMs. This product formula was conjectured by Mills, Robbins, and Rumsey in 1983 and rigorously proved by Zeilberger in 1996 using hypergeometric identities and computer-assisted verification. The sequence begins 1, 2, 7, 42, 429 for n=1 to 5, and this enumeration is connected to refined Euler numbers via further refinements that track the positions of specific entries, such as the location of the unique 1 in the top row. For small orders, the ASMs illustrate the generalization from permutations. For n=1, there is one ASM: the matrix \begin{pmatrix} 1 \end{pmatrix}. For n=2, there are two ASMs, both permutation matrices: the identity \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix} and the transposition \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, corresponding to the identity and swap permutations. For n=3, there are seven ASMs, including six permutation matrices and one additional matrix with a -1 entry: \begin{pmatrix} 0 & 1 & 0 \\ 1 & -1 & 1 \\ 0 & 1 & 0 \end{pmatrix}. ASMs were introduced by Mills, Robbins, and Rumsey in 1983 in the context of Dodgson's condensation algorithm for computing determinants of matrices, where they arise naturally in the evaluation process. Their conjecture drew significant attention in the , culminating in Zeilberger's proof, which relied on sophisticated identities in hypergeometric series and was verified with the aid of computer algebra systems. Variants of ASMs include monotone triangles, which are in bijective correspondence with ASMs. A monotone triangle of order n consists of n strictly increasing rows of integers from 1 to some maximum, with the bottom row fixed as 1 to n, and the entries weakly increasing along the other two diagonals; the number of such triangles equals A_n. Another variant is the fully alternating sign matrix, where nonzero entries occupy all positions without intervening zeros in rows and columns while preserving the alternating sign and sum conditions, though these are less commonly enumerated. These variants highlight the rich combinatorial structure of ASMs and their links to other objects like plane partitions.

References

  1. [1]
    [PDF] A Survey of Alternating Permutations - MIT Mathematics
    Abstract. A permutation a1a2 ... an of 1, 2,...,n is alternating if a1 > a2 < a3 > a4 < ... . We survey some aspects of the theory of alternating permu-
  2. [2]
    [PDF] arXiv:1406.5207v1 [math.CO] 19 Jun 2014
    Jun 19, 2014 · An alternating permutation is a permutation π ∈ Sn satisfying π(1) < π(2) > π(3) < π(4) >. ··· . Alternating permutations have been well ...
  3. [3]
    Alternating permutations and the mth descents - ScienceDirect
    There is a one-to-one correspondence between up–down (or alternating) permutations and down–up permutations. For an up–down permutation π ∈ A ( n ) ...
  4. [4]
    Alternating Permutation -- from Wolfram MathWorld
    (OEIS A000111). The even-numbered A_n s are called Euler numbers |E_(2n) ... "Bernoulli-Euler Updown Numbers Associated with Function Singularities, Their ...
  5. [5]
    [PDF] Sur les permutations alternées - Numdam
    DÉSIRÉ ANDRÉ. Le présent travail a pour point de départ la notion, probablement toute nouvelle, des permutations alternées de η lettres distinctes, et pour ...Missing: date | Show results with:date
  6. [6]
    [PDF] Enumerative Combinatorics Volume 1 second edition - Mathematics
    ... (bijection) between two finite sets than merely to prove that they have the ... down-up) if w1 > w2 < w3 > w4 < ···. Equivalently, D(w) = {1, 3, 5 ...
  7. [7]
    [PDF] Increasing trees and alternating permutations(1)
    Introduction. In this article we consider some increasing trees, the number of which is equal to the number of alternating (updown) permutations, that is,.
  8. [8]
    A New Operation on Sequences: the Boustrouphedon Transform
    May 20, 2002 · The Boustrophedon transform is a new operation on integer sequences, derived from a generalization of the Seidel-Entringer-Arnold method.Missing: Haglund | Show results with:Haglund
  9. [9]
    [PDF] Generating random alternating permutations in time nlog n - HAL
    Dec 14, 2012 · The goal of this paper is to intro- duce an algorithm generating random alternating permutations of {1, 2,...N} in time N log N. An ...
  10. [10]
    [math-ph/0411076] Square ice, alternating sign matrices and ... - arXiv
    The derivation is based on the standard relationship between Hankel determinants and orthogonal polynomials. For the particular sets of parameters corresponding ...
  11. [11]
    A000111 - OEIS
    Euler or up/down numbers: e.g.f. sec(x) + tan(x). Also for n >= 2, half the number of alternating permutations on n letters (A001250). (Formerly M1492 N0587).
  12. [12]
    [PDF] Power series for up-down min-max permutations
    Mar 13, 2013 · There is a correspondence between min-max alternating permutations and max-min alternating permutations through reversal of permutations. For ...
  13. [13]
    [PDF] Generalized Euler numbers and ordered set partitions
    Jan 16, 2025 · The Euler numbers, En, can be defined in terms of the exponential generating function. X n≥0. En xn n! = tanx + secx. n. 0 1 2 3 4 5 6. 7. 8. 9.
  14. [14]
  15. [15]
    24.7 Integral Representations ‣ Properties ‣ Chapter 24 Bernoulli ...
    §24.7(i) Bernoulli and Euler Numbers. ⓘ. Keywords: Bernoulli numbers, Euler numbers, integral representation, integral representations ... secant function ...
  16. [16]
    [PDF] New Representations of Catalan's Constant, Apery's ... - arXiv
    Feb 10, 2025 · From the above, we can now express the following integral representation of Euler's numbers, using (7) together with the inverse of the cdf ...
  17. [17]
    Proving a recurrence relation in up-down permutation
    Jan 18, 2020 · Call an injective function s "up-down" if s(j)<s(j+1) iff j is odd. Important examples include the alternating permutations. The Wikipedia ...Missing: reversal | Show results with:reversal<|control11|><|separator|>
  18. [18]
    Alternating permutations and binary increasing trees - ScienceDirect
    In this paper we present a direct recursive proof of this fact, and then exhibit a natural bijection between the two families.
  19. [19]
    [PDF] arXiv:1405.6904v4 [math.CO] 8 Jan 2015
    Jan 8, 2015 · Alternating permutations in S2n are in bijection with noncrossing arc diagrams on 2n points that are perfect matchings. The exponential ...Missing: arch | Show results with:arch
  20. [20]
    Alternating Permutations with Restrictions and Standard Young ...
    Jun 28, 2012 · In this paper, we establish bijections between the set of 4123-avoiding down-up alternating permutations of length 2n 2 n and the set of ...Missing: row | Show results with:row
  21. [21]
  22. [22]
    [PDF] Hankel Continued Fractions and Hankel determinants of the Euler ...
    In 1974, Carlitz and Scoville obtained the exponential generating function of the quadruple statistic. (val, pk, da, dd) for the permutations [7, 17, 43]. (3.1).
  23. [23]
    [PDF] The Euler and Springer numbers as moment sequences
    Euler numbers: complete increasing plane binary trees, increasing 0-1-2 ... alternating permutations. More precisely, recall that a signed permutation ...
  24. [24]
  25. [25]
    Alternating sign matrices and descending plane partitions
    In this paper is a discussion of alternating sign matrices and descending plane partitions, and several conjectures and theorems about them are presented.
  26. [26]
    None
    ### Summary: Boundary of an ASM and Permutations/Alternating Permutations
  27. [27]
    Proof of the Alternating Sign Matrix Conjecture
    Jul 25, 1995 · Proof of the Alternating Sign Matrix Conjecture. Doron Zeilberger. DOI: https://doi.org/10.37236/1271. Abstract. The number of n×n n × n ...
  28. [28]
    [PDF] Alternating Sign Matrices and Descending Plane Partitions
    There is, however, a weight for alternating sign matrices that seems to be very natural, namely, the number of -1's in the matrix. Let us denote by A,(x) the ...
  29. [29]
    [PDF] The many faces of alternating-sign matrices - arXiv
    Aug 15, 2002 · Abstract. I give a survey of different combinatorial forms of alternating-sign ma- trices, starting with the original form introduced by ...