Hook length formula

Last updated

In combinatorial mathematics, the hook length formula is a formula for the number of standard Young tableaux whose shape is a given Young diagram. It has applications in diverse areas such as representation theory, probability, and algorithm analysis; for example, the problem of longest increasing subsequences. A related formula gives the number of semi-standard Young tableaux, which is a specialization of a Schur polynomial.

Contents

Definitions and statement

Let be a partition of . It is customary to interpret graphically as a Young diagram, namely a left-justified array of square cells with rows of lengths . A (standard) Young tableau of shape is a filling of the cells of the Young diagram with all the integers , with no repetition, such that each row and each column form increasing sequences. For the cell in position , in the th row and th column, the hook is the set of cells such that and or and . The hook length is the number of cells in .

The hook length formula expresses the number of standard Young tableaux of shape , denoted by or , as

where the product is over all cells of the Young diagram.

Examples

A tableau listing the hook length of each cell in the Young diagram

(
4
,
3
,
1
,
1
)
{\displaystyle (4,3,1,1)} Hook-length tableau December 8 2013 3.jpg
A tableau listing the hook length of each cell in the Young diagram

The figure on the right shows hook lengths for the cells in the Young diagram , corresponding to the partition 9 = 4 + 3 + 1 + 1. The hook length formula gives the number of standard Young tableaux as:

A Catalan number counts Dyck paths with steps going up (U) interspersed with steps going down (D), such that at each step there are never more preceding D's than U's. These are in bijection with the Young tableaux of shape : a Dyck path corresponds to the tableau whose first row lists the positions of the U-steps, while the second row lists the positions of the D-steps. For example, UUDDUD correspond to the tableaux with rows 125 and 346.

This shows that , so the hook formula specializes to the well-known product formula

History

There are other formulas for , but the hook length formula is particularly simple and elegant. A less convenient formula expressing in terms of a determinant was deduced independently by Frobenius and Young in 1900 and 1902 respectively using algebraic methods. [1] [2] MacMahon found an alternate proof for the Young–Frobenius formula in 1916 using difference methods. [3]

The hook length formula itself was discovered in 1953 by Frame, Robinson, and Thrall as an improvement to the Young–Frobenius formula. Sagan [4] describes the discovery as follows.

One Thursday in May of 1953, Robinson was visiting Frame at Michigan State University. Discussing the work of Staal (a student of Robinson), Frame was led to conjecture the hook formula. At first Robinson could not believe that such a simple formula existed, but after trying some examples he became convinced, and together they proved the identity. On Saturday they went to the University of Michigan, where Frame presented their new result after a lecture by Robinson. This surprised Thrall, who was in the audience, because he had just proved the same result on the same day!

Despite the simplicity of the hook length formula, the Frame–Robinson–Thrall proof is not very insightful and does not provide any intuition for the role of the hooks. The search for a short, intuitive explanation befitting such a simple result gave rise to many alternate proofs. [5] Hillman and Grassl gave the first proof that illuminates the role of hooks in 1976 by proving a special case of the Stanley hook-content formula, which is known to imply the hook length formula. [6] Greene, Nijenhuis, and Wilf found a probabilistic proof using the hook walk in which the hook lengths appear naturally in 1979. [7] Remmel adapted the original Frame–Robinson–Thrall proof into the first bijective proof for the hook length formula in 1982. [8] A direct bijective proof was first discovered by Franzblau and Zeilberger in 1982. [9] Zeilberger also converted the Greene–Nijenhuis–Wilf hook walk proof into a bijective proof in 1984. [10] A simpler direct bijection was announced by Pak and Stoyanovskii in 1992, and its complete proof was presented by the pair and Novelli in 1997. [11] [12] [4]

Meanwhile, the hook length formula has been generalized in several ways. R. M. Thrall found the analogue to the hook length formula for shifted Young Tableaux in 1952. [13] Sagan gave a shifted hook walk proof for the hook length formula for shifted Young tableaux in 1980. [14] Sagan and Yeh proved the hook length formula for binary trees using the hook walk in 1989. [15] Proctor gave a poset generalization (see below).

Probabilistic proof

Knuth's heuristic argument

The hook length formula can be understood intuitively using the following heuristic, but incorrect, argument suggested by D. E. Knuth. [16] Given that each element of a tableau is the smallest in its hook and filling the tableau shape at random, the probability that cell will contain the minimum element of the corresponding hook is the reciprocal of the hook length. Multiplying these probabilities over all and gives the formula. This argument is fallacious since the events are not independent.

Knuth's argument is however correct for the enumeration of labellings on trees satisfying monotonicity properties analogous to those of a Young tableau. In this case, the 'hook' events in question are in fact independent events.

Probabilistic proof using the hook walk

This is a probabilistic proof found by C. Greene, A. Nijenhuis, and H. S. Wilf in 1979. [7] Define

We wish to show that . First,

Corners of the Young diagram (5,3,2,1,1) Corners of young.png
Corners of the Young diagram (5,3,2,1,1)

where the sum runs over all Young diagrams obtained from by deleting one corner cell. (The maximal entry of the Young tableau of shape occurs at one of its corner cells, so deleting it gives a Young tableaux of shape .)

We define and , so it is enough to show the same recurrence

which would imply by induction. The above sum can be viewed as a sum of probabilities by writing it as

We therefore need to show that the numbers define a probability measure on the set of Young diagrams with . This is done in a constructive way by defining a random walk, called the hook walk, on the cells of the Young diagram , which eventually selects one of the corner cells of (which are in bijection with diagrams for which ). The hook walk is defined by the following rules.

  1. Pick a cell uniformly at random from cells. Start the random walk from there.
  2. Successor of current cell is chosen uniformly at random from the hook .
  3. Continue until you reach a corner cell .

Proposition: For a given corner cell of , we have

where .

Given this, summing over all corner cells gives as claimed.

Connection to representations of the symmetric group

The hook length formula is of great importance in the representation theory of the symmetric group , where the number is known to be equal to the dimension of the complex irreducible representation associated to .

Detailed discussion

The complex irreducible representations of the symmetric group are indexed by partitions of (see Specht module) . Their characters are related to the theory of symmetric functions via the Hall inner product:

where is the Schur function associated to and is the power-sum symmetric function of the partition associated to the cycle decomposition of . For example, if then .

Since the identity permutation has the form in cycle notation, , the formula says

The expansion of Schur functions in terms of monomial symmetric functions uses the Kostka numbers:

Then the inner product with is , because . Note that is equal to , so that from considering the regular representation of , or combinatorially from the Robinson–Schensted–Knuth correspondence.

The computation also shows that:

This is the expansion of in terms of Schur functions using the coefficients given by the inner product, since . The above equality can be proven also checking the coefficients of each monomial at both sides and using the Robinson–Schensted–Knuth correspondence or, more conceptually, looking at the decomposition of by irreducible modules, and taking characters. See Schur–Weyl duality.

Proof of hook formula using Frobenius formula

Source: [17]

By the above considerations

so that

where is the Vandermonde determinant.

For the partition , define for . For the following we need at least as many variables as rows in the partition, so from now on we work with variables .

Each term is equal to

(See Schur function.) Since the vector is different for each partition, this means that the coefficient of in , denoted , is equal to . This is known as the Frobenius Character Formula, which gives one of the earliest proofs. [17]

It remains only to simplify this coefficient. Multiplying

and

we conclude that our coefficient as

which can be written as

The latter sum is equal to the following determinant

which column-reduces to a Vandermonde determinant, and we obtain the formula

Note that is the hook length of the first box in each row of the Young diagram, and this expression is easily transformed into the desired form : one shows , where the latter product runs over the th row of the Young diagram.

Connection to longest increasing subsequences

The hook length formula also has important applications to the analysis of longest increasing subsequences in random permutations. If denotes a uniformly random permutation of order , denotes the maximal length of an increasing subsequence of , and denotes the expected (average) value of , Anatoly Vershik and Sergei Kerov [18] and independently Benjamin F. Logan and Lawrence A. Shepp [19] showed that when is large, is approximately equal to . This answers a question originally posed by Stanislaw Ulam. The proof is based on translating the question via the Robinson–Schensted correspondence to a problem about the limiting shape of a random Young tableau chosen according to Plancherel measure. Since the definition of Plancherel measure involves the quantity , the hook length formula can then be used to perform an asymptotic analysis of the limit shape and thereby also answer the original question.

The ideas of Vershik–Kerov and Logan–Shepp were later refined by Jinho Baik, Percy Deift and Kurt Johansson, who were able to achieve a much more precise analysis of the limiting behavior of the maximal increasing subsequence length, proving an important result now known as the Baik–Deift–Johansson theorem. Their analysis again makes crucial use of the fact that has a number of good formulas, although instead of the hook length formula it made use of one of the determinantal expressions.

The formula for the number of Young tableaux of shape was originally derived from the Frobenius determinant formula in connection to representation theory: [20]

Hook lengths can also be used to give a product representation to the generating function for the number of reverse plane partitions of a given shape. [21] If λ is a partition of some integer p, a reverse plane partition of n with shape λ is obtained by filling in the boxes in the Young diagram with non-negative integers such that the entries add to n and are non-decreasing along each row and down each column. The hook lengths can be defined as with Young tableaux. If πn denotes the number of reverse plane partitions of n with shape λ, then the generating function can be written as

Stanley discovered another formula for the same generating function. [22] In general, if is any poset with elements, the generating function for reverse -partitions is

where is a polynomial such that is the number of linear extensions of .

In the case of a partition , we are considering the poset in its cells given by the relation

.

So a linear extension is simply a standard Young tableau, i.e.

Proof of hook formula using Stanley's formula

Combining the two formulas for the generating functions we have

Both sides converge inside the disk of radius one and the following expression makes sense for

It would be violent to plug in 1, but the right hand side is a continuous function inside the unit disk and a polynomial is continuous everywhere so at least we can say

Applying L'Hôpital's rule times yields the hook length formula

Semi-standard tableaux hook length formula

The Schur polynomial is the generating function of semistandard Young tableaux with shape and entries in . Specializing this to gives the number of semi-standard tableaux, which can be written in terms of hook lengths:

The Young diagram corresponds to an irreducible representation of the special linear group , and the Schur polynomial is also the character of the diagonal matrix acting on this representation. The above specialization is thus the dimension of the irreducible representation, and the formula is an alternative to the more general Weyl dimension formula.

We may refine this by taking the principal specialization of the Schur function in the variables  :

where for the conjugate partition .

Skew shape formula

There is a generalization of this formula for skew shapes, [23]

where the sum is taken over excited diagrams of shape and boxes distributed according to .

A variation on the same theme is given by Okounkov and Olshanski [24] of the form

where is the so-called shifted Schur function .

Generalization to d-complete posets

Young diagrams can be considered as finite order ideals in the poset , and standard Young tableaux are their linear extensions. Robert Proctor has given a generalization of the hook length formula to count linear extensions of a larger class of posets generalizing both trees and skew diagrams. [25] [26]

Related Research Articles

In number theory, an arithmetic, arithmetical, or number-theoretic function is generally any function f(n) whose domain is the positive integers and whose range is a subset of the complex numbers. Hardy & Wright include in their definition the requirement that an arithmetical function "expresses some arithmetical property of n". There is a larger class of number-theoretic functions that do not fit this definition, for example, the prime-counting functions. This article provides links to functions of both classes.

The Liouville lambda function, denoted by λ(n) and named after Joseph Liouville, is an important arithmetic function. Its value is +1 if n is the product of an even number of prime numbers, and −1 if it is the product of an odd number of primes.

In mathematics, the Dirichlet convolution is a binary operation defined for arithmetic functions; it is important in number theory. It was developed by Peter Gustav Lejeune Dirichlet.

The Gram–Charlier A series, and the Edgeworth series are series that approximate a probability distribution in terms of its cumulants. The series are the same; but, the arrangement of terms differ. The key idea of these expansions is to write the characteristic function of the distribution whose probability density function f is to be approximated in terms of the characteristic function of a distribution with known and suitable properties, and to recover f through the inverse Fourier transform.

In mathematics, a Young tableau is a combinatorial object useful in representation theory and Schubert calculus. It provides a convenient way to describe the group representations of the symmetric and general linear groups and to study their properties. Young tableaux were introduced by Alfred Young, a mathematician at Cambridge University, in 1900. They were then applied to the study of the symmetric group by Georg Frobenius in 1903. Their theory was further developed by many mathematicians, including Percy MacMahon, W. V. D. Hodge, G. de B. Robinson, Gian-Carlo Rota, Alain Lascoux, Marcel-Paul Schützenberger and Richard P. Stanley.

Variational Bayesian methods are a family of techniques for approximating intractable integrals arising in Bayesian inference and machine learning. They are typically used in complex statistical models consisting of observed variables as well as unknown parameters and latent variables, with various sorts of relationships among the three types of random variables, as might be described by a graphical model. As typical in Bayesian inference, the parameters and latent variables are grouped together as "unobserved variables". Variational Bayesian methods are primarily used for two purposes:

  1. To provide an analytical approximation to the posterior probability of the unobserved variables, in order to do statistical inference over these variables.
  2. To derive a lower bound for the marginal likelihood of the observed data. This is typically used for performing model selection, the general idea being that a higher marginal likelihood for a given model indicates a better fit of the data by that model and hence a greater probability that the model in question was the one that generated the data.
<span class="mw-page-title-main">Noncentral chi-squared distribution</span> Noncentral generalization of the chi-squared distribution

In probability theory and statistics, the noncentral chi-squared distribution is a noncentral generalization of the chi-squared distribution. It often arises in the power analysis of statistical tests in which the null distribution is a chi-squared distribution; important examples of such tests are the likelihood-ratio tests.

In mathematics, Schur polynomials, named after Issai Schur, are certain symmetric polynomials in n variables, indexed by partitions, that generalize the elementary symmetric polynomials and the complete homogeneous symmetric polynomials. In representation theory they are the characters of polynomial irreducible representations of the general linear groups. The Schur polynomials form a linear basis for the space of all symmetric polynomials. Any product of Schur polynomials can be written as a linear combination of Schur polynomials with non-negative integral coefficients; the values of these coefficients is given combinatorially by the Littlewood–Richardson rule. More generally, skew Schur polynomials are associated with pairs of partitions and have similar properties to Schur polynomials.

In queueing theory, a discipline within the mathematical theory of probability, a Jackson network is a class of queueing network where the equilibrium distribution is particularly simple to compute as the network has a product-form solution. It was the first significant development in the theory of networks of queues, and generalising and applying the ideas of the theorem to search for similar product-form solutions in other networks has been the subject of much research, including ideas used in the development of the Internet. The networks were first identified by James R. Jackson and his paper was re-printed in the journal Management Science’s ‘Ten Most Influential Titles of Management Sciences First Fifty Years.’

In mathematics, the Weyl character formula in representation theory describes the characters of irreducible representations of compact Lie groups in terms of their highest weights. It was proved by Hermann Weyl. There is a closely related formula for the character of an irreducible representation of a semisimple Lie algebra. In Weyl's approach to the representation theory of connected compact Lie groups, the proof of the character formula is a key step in proving that every dominant integral element actually arises as the highest weight of some irreducible representation. Important consequences of the character formula are the Weyl dimension formula and the Kostant multiplicity formula.

<span class="mw-page-title-main">Inverse Gaussian distribution</span> Family of continuous probability distributions

In probability theory, the inverse Gaussian distribution is a two-parameter family of continuous probability distributions with support on (0,∞).

In 3-dimensional topology, a part of the mathematical field of geometric topology, the Casson invariant is an integer-valued invariant of oriented integral homology 3-spheres, introduced by Andrew Casson.

In mathematics, the Jack function is a generalization of the Jack polynomial, introduced by Henry Jack. The Jack polynomial is a homogeneous, symmetric polynomial which generalizes the Schur and zonal polynomials, and is in turn generalized by the Heckman–Opdam polynomials and Macdonald polynomials.

In mathematics, Macdonald polynomialsPλ(x; t,q) are a family of orthogonal symmetric polynomials in several variables, introduced by Macdonald in 1987. He later introduced a non-symmetric generalization in 1995. Macdonald originally associated his polynomials with weights λ of finite root systems and used just one variable t, but later realized that it is more natural to associate them with affine root systems rather than finite root systems, in which case the variable t can be replaced by several different variables t=(t1,...,tk), one for each of the k orbits of roots in the affine root system. The Macdonald polynomials are polynomials in n variables x=(x1,...,xn), where n is the rank of the affine root system. They generalize many other families of orthogonal polynomials, such as Jack polynomials and Hall–Littlewood polynomials and Askey–Wilson polynomials, which in turn include most of the named 1-variable orthogonal polynomials as special cases. Koornwinder polynomials are Macdonald polynomials of certain non-reduced root systems. They have deep relationships with affine Hecke algebras and Hilbert schemes, which were used to prove several conjectures made by Macdonald about them.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

In mathematics, the Hall–Littlewood polynomials are symmetric functions depending on a parameter t and a partition λ. They are Schur functions when t is 0 and monomial symmetric functions when t is 1 and are special cases of Macdonald polynomials. They were first defined indirectly by Philip Hall using the Hall algebra, and later defined directly by Dudley E. Littlewood (1961).

In probability theory and statistics, the generalized multivariate log-gamma (G-MVLG) distribution is a multivariate distribution introduced by Demirhan and Hamurkaroglu in 2011. The G-MVLG is a flexible distribution. Skewness and kurtosis are well controlled by the parameters of the distribution. This enables one to control dispersion of the distribution. Because of this property, the distribution is effectively used as a joint prior distribution in Bayesian analysis, especially when the likelihood is not from the location-scale family of distributions such as normal distribution.

In combinatorial mathematics, cyclic sieving is a phenomenon by which evaluating a generating function for a finite set at roots of unity counts symmetry classes of objects acted on by a cyclic group.

In mathematics, specifically in representation theory, the Frobenius formula, introduced by G. Frobenius, computes the characters of irreducible representations of the symmetric group Sn. Among the other applications, the formula can be used to derive the hook length formula.

<span class="mw-page-title-main">Representations of classical Lie groups</span>

In mathematics, the finite-dimensional representations of the complex classical Lie groups , , , , , can be constructed using the general representation theory of semisimple Lie algebras. The groups , , are indeed simple Lie groups, and their finite-dimensional representations coincide with those of their maximal compact subgroups, respectively , , . In the classification of simple Lie algebras, the corresponding algebras are

References

  1. G. Frobenius. Uber die charaktere der symmetrischer gruppe, Preuss. &ad. Wk. sitz. (1900), 516–534.
  2. A. Young. Quantitative substitutional analysis II, Proc. London Math. Sot., Ser. 1, 35 (1902), 361–397.
  3. P. A. MacMahon. “Combinatory Analysis,” Cambridge Univ. Press, London/New York, 1916; reprinted by Chelsea, New York, 1960.
  4. 1 2 Sagan, Bruce (2001). The Symmetric Group. Representations, Combinatorial Algorithms, and Symmetric Functions, 2nd edition. Springer-Verlag. ISBN   0-387-95067-2.
  5. Knuth, Donald (1973). The Art of Computer Programming, Volume 3: Sorting and Searching, 3rd Edition, Addison–Wesley, p. 63
  6. Hillman, A. P.; Grassl, R. M. (1976). "Reverse plane partitions and tableau hook numbers". Journal of Combinatorial Theory . Series A. 21 (2): 216–221. doi: 10.1016/0097-3165(76)90065-0 .
  7. 1 2 Greene, Curtis; Nijenhuis, Albert; Wilf, Herbert S. (1979). "A probabilistic proof of a formula for the number of Young tableaux of a given shape". Advances in Mathematics . 31 (1): 104–109. doi: 10.1016/0001-8708(79)90023-9 .
  8. J. B. Remmel. Bijective proofs of formulae for the number of standard Young tableaux, Linear and Multilinear Algebra 11 (1982), 45–100.
  9. Franzblau, D. S. and Zeilberger, D. (1982). A bijective proof of the hook-length formula. J. Algorithms 3, 317–343.
  10. Zeilberger, Doron (1984). "A short hook-lengths bijection inspired by the Greene–Nijenhuis–Wilf proof". Discrete Mathematics . 51 (1): 101–108. doi: 10.1016/0012-365X(84)90027-X .
  11. Pak, I. M. and Stoyanovskii, A. V. (1992). A bijective proof of the hook-length formula. Funct. Anal. Appl. 24.
  12. Novelli, J.-C., Pak, I. M. and Stoyanovskii, A. V. (1997). A direct bijective proof of the hook-length formula. Discrete Mathematics and Theoretical Computer Science 1, 1997, 53–67.
  13. R. M. Thrall. A combinatorial problem, Michigan Math. J. 1 (1952), 81–88.
  14. Sagan, B. On selecting a random shifted Young tableau. J. Algorithms 1, 3 (1980), 213–234.
  15. Sagan, B. E., and Yeh, Y. N. Probabilistic algorithms for trees. Fibonacci Quart. 27, 3 (1989), 201–208.
  16. Knuth, Donald (1973), The Art of Computer Programming, Volume 3: Sorting and Searching, 3rd Edition, Addison–Wesley, p. 63, ISBN   0-201-03803-X .
  17. 1 2 Fulton, William, 1939- (1991). Representation theory : a first course. Harris, Joe, 1951-. New York: Springer-Verlag. pp. 49–50. ISBN   0387974954. OCLC   22861245.{{cite book}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  18. Vershik, A. M.; Kerov, C. V. (1977), "Asymptotics of the Plancheral measure of the symmetric group and a limiting form for Young tableaux", Dokl. Akad. Nauk SSSR 233: 1024–1027
  19. Logan, B. F.; Shepp, L. A. (1977). "A variational problem for random Young tableaux". Advances in Mathematics . 26 (2): 206–222. doi: 10.1016/0001-8708(77)90030-5 .
  20. Knuth, Donald (1973), The Art of Computer Programming , vol. 3 (1 ed.), Addison–Wesley, pp. 61–62
  21. Stanley, Richard P. (1971), "Theory and applications of plane partitions, 2", Studies in Applied Mathematics , 50: 259–279, doi:10.1002/sapm1971503259
  22. R.P. Stanley, "Ordered Structures and Partitions" PhD Thesis, Harvard University, 1971
  23. Morales, Alejandro H.; Pak, Igor; Panova, Greta (2018). "Hook formulas for skew shapes I. q-analogues and bijections". Journal of Combinatorial Theory . Series A. 154: 350–405. arXiv: 1512.08348 . doi: 10.1016/j.jcta.2017.09.002 .
  24. Okounkov, A. and Olshanski, G. Shifted Schur functions, Algebra I Analiz, 9.2 (1997), 73-146.
  25. Proctor, Robert (1999). "Dynkin diagram classification of λ-minuscule Bruhat lattices and of d-complete posets". Journal of Algebraic Combinatorics. 9: 61–94. doi: 10.1023/A:1018615115006 .
  26. Kim, Jang Soo; Yoo, Meesue (2019). "Hook length property of d-complete posets via q-integrals". Journal of Combinatorial Theory . Series A. 162: 167–221. arXiv: 1708.09109 . doi: 10.1016/j.jcta.2018.11.003 . S2CID   57759574.