Integration on complex Grassmannians, deformed monotone Hurwitz numbers, and interlacing phenomena
Department of Mathematics, The University of Auckland, Auckland 1142 New Zealand
School of Mathematics, Monash University, VIC 3800 Australia
School of Mathematics, Monash University, VIC 3800 Australia
Email: [email protected], [email protected], [email protected]
Abstract. We introduce a family of polynomials, which arise in three distinct ways: in the large expansion of a matrix integral, as a weighted enumeration of factorisations of permutations, and via the topological recursion. More explicitly, we interpret the complex Grassmannian as the space of idempotent Hermitian matrices of rank and develop a Weingarten calculus to integrate products of matrix elements over it. In the regime of large and fixed ratio , such integrals have expansions whose coefficients count factorisations of permutations into monotone sequences of transpositions, with each sequence weighted by a monomial in . This gives rise to the desired polynomials, which specialise to the monotone Hurwitz numbers when .
These so-called deformed monotone Hurwitz numbers satisfy a cut-and-join recursion, a one-point recursion, and the topological recursion. Furthermore, we conjecture on the basis of overwhelming empirical evidence that the deformed monotone Hurwitz numbers are real-rooted polynomials whose roots satisfy remarkable interlacing phenomena.
An outcome of our work is the viewpoint that the topological recursion can be used to “topologise” sequences of polynomials, and we claim that the resulting families of polynomials may possess interesting properties. As a further case study, we consider a weighted enumeration of dessins d’enfant and conjecture that the resulting polynomials are also real-rooted and satisfy analogous interlacing properties.
Acknowledgements. X.C. was supported by a University of Auckland Doctoral Scholarship. N.D. was supported by the Australian Research Council grant DP180103891. E.M. was supported by an Australian Government Research Training Program (RTP) Scholarship and a Monash University Postgraduate Publication Award. N.D. would like to thank Maksim Karev for planting the seed of an idea that led to this work by asking the question: “What does the topological recursion for Schröder numbers calculate?”
2020 Mathematics Subject Classification. 05A15, 05E10, 15B52, 60B20
1 Introduction
In this paper, we introduce a family of polynomials, which arise in three distinct ways: in the large expansion of a matrix integral, as a weighted enumeration of factorisations of permutations, and via the topological recursion. Our construction simultaneously generalises the Narayana polynomials and the monotone Hurwitz numbers, both of which have garnered considerable attention in the literature [33, 46]. We prove or conjecture that the family of polynomials we introduce satisfies a number of remarkable properties concerning their coefficients and roots, such as symmetry, unimodality, real-rootedness and interlacing.
This work is inspired by the known relations between Weingarten calculus on unitary groups, Jucys–Murphy elements in the symmetric group algebra, and monotone Hurwitz numbers [44, 33]. Our starting point is the space of idempotent Hermitian matrices of rank , where . This space admits the following three descriptions, where denotes the identity matrix and denotes the matrix whose first diagonal entries are 1 and whose remaining entries are 0.
The unitary group acts transitively on by conjugation, thus endowing it with the structure of a homogeneous space. Since the stabiliser of is , we may identify with the complex Grassmannian , which parametrises -dimensional subspaces of an -dimensional complex vector space.
As a compact homogeneous space, inherits a normalised -invariant Haar measure, which we denote succinctly by . For , define the function corresponding to a matrix element. Taking our cue from the general theory of Weingarten calculus, we consider integrals of the form
where . Our primary goal is to study such matrix integrals in the regime of large and fixed ratio .
In recent decades, Weingarten calculus has developed into a rich theory concerned with integration on compact groups and related objects, with respect to the Haar measure [14]. Following the usual paradigm, we define a Weingarten function that takes as input a permutation and outputs the following elementary integral, where our notation suppresses the dependence on and .
Modern approaches to Weingarten calculus typically rely on abstract algebraic tools such as Schur–Weyl duality [12, 15]. These are not as amenable to the current setting as the ideas rooted in the pioneering work of Weingarten [49] and revisited more recently by Collins and Matsumoto [13], which we use as inspiration to obtain the following.111This introduction includes only a cursory discussion of our main results and conjectures. The reader is encouraged to consult the main text for the relevant definitions, precise statements, and further details.
Theorem A (Weingarten calculus).
-
•
Convolution formula (Theorem 2.3)
Arbitrary integrals of monomials in the matrix elements of reduce to elementary integrals via the equation -
•
Orthogonality relations (Theorem 2.4)
For each permutation , the Weingarten function satisfies the relationHere, for any permutation that fixes , we write to denote the permutation that agrees with on the set .
The orthogonality relations allow for at least two distinct approaches to calculate values of the Weingarten function. By solving the non-degenerate linear system provided by the orthogonality relations, one can explicitly obtain values of the Weingarten function, which are rational functions of and . Alternatively, by iteratively and infinitely applying the orthogonality relations, one obtains the following large expansion for the Weingarten function, for fixed ratio .
Theorem B (Large expansion, Theorem 2.12).
For a permutation , let denote the weighted enumeration of tuples of transpositions in such that
-
•
;
-
•
the sequence is monotone — that is, if we write with , then ; and
-
•
the weight of is .
The Weingarten function has the following large expansion, for fixed .
Monotone Hurwitz numbers count monotone sequences of transpositions with prescribed length and with product of a prescribed cycle type, usually with an additional transitivity assumption. They are known to arise in the Weingarten calculus for unitary groups and in the large expansion of the HCIZ matrix integral [33]. The theorem above motivates a “deformed” version of the monotone Hurwitz numbers, in which each sequence of transpositions is weighted by the monomial . The resulting polynomials in are referred to as deformed monotone Hurwitz numbers and denoted by . Their precise definition appears in Definition 3.1 and they are the central objects introduced and studied in the present work.
The deformed monotone Hurwitz numbers obey various recursions, from which known results for the usual monotone Hurwitz numbers are recovered when [32, 18, 9].
Theorem C (Recursions).
-
•
Cut-and-join recursion (Theorem 3.7)
The deformed monotone Hurwitz numbers can be computed from the base case and the recursion -
•
One-point recursion (Theorem 3.12)
The one-point deformed monotone Hurwitz numbers — in other words, those with — satisfy -
•
Topological recursion (Theorem 4.1)
The deformed monotone Hurwitz numbers are governed by the topological recursion on the genus zero spectral curve .
These recursions are all effective and can be used to produce explicit data, some of which is contained in Appendix A. At the level of coefficients, each deformed monotone Hurwitz number is symmetric (the sequence of coefficients is palindromic) and unimodal (the sequence of coefficients increases to a point and then decreases). We prove this in Proposition 3.9 using the cut-and-join recursion and note that these properties are not immediate from the combinatorial deformation of the deformed monotone Hurwitz numbers.
At the level of roots, it appears that the deformed monotone Hurwitz numbers are real-rooted and that they exhibit interlacing phenomena. Two real-rooted polynomials are said to interlace if their degrees differ by one and their roots weakly alternate on the real number line. We have gathered overwhelming numerical evidence to support the following conjectures.
Conjecture D (Roots).
-
•
Real-rootedness (Conjecture 3.13)
The deformed monotone Hurwitz number is a real-rooted polynomial in . -
•
Interlacing (Conjecture 3.14)
The polynomial interlaces each of the polynomials
In the case , the deformed monotone Hurwitz numbers recover the sequence of Narayana polynomials via the equation
(1) |
Thus, we consider the deformed monotone Hurwitz numbers to be a “topological generalisation” of the Narayana polynomials. We propose the topological recursion as a mechanism to “topologise” sequences of polynomials more generally. In particular, we claim that doing so can preserve interesting behaviour in the polynomials, such as symmetry, unimodality, real-rootedness, and interlacing properties. This is not only the case for the deformed monotone Hurwitz numbers, but we also observe these phenomena in the weighted enumeration of dessins d’enfant — in other words, bicoloured maps — in which each black vertex in a dessin d’enfant is assigned a multiplicative weight .
Our results concerning integration on complex Grassmannians and deformed monotone Hurwitz numbers suggest various avenues for further research. It would be natural to consider integration on real Grassmannians , also in the regime of large with fixed ratio , and the first two authors are currently pursuing this line of investigation. Matsumoto considered Weingarten calculus on compact symmetric spaces [39] and the particular case of the symmetric space AIII bears a strong resemblance to the present work. It would be interesting to further develop the parallels between these two settings. The real-rootedness and interlacing conjectures for deformed monotone Hurwitz numbers (Conjectures 3.13 and 3.14) and the weighted dessin d’enfant enumeration (Conjecture 4.11) not only require proof, but also invite a deeper exploration of how common these phenomena might be.
The structure of the paper is as follows.
-
•
In Section 2, we develop the Weingarten calculus for integration over . This includes a convolution formula (Theorem 2.3) and orthogonality relations (Theorem 2.4). We use the latter to express the Weingarten function of a permutation as a weighted enumeration of monotone sequences of transpositions (Theorem 2.12). As a consequence, the Weingarten function can be written succinctly in terms of Jucys–Murphy elements in the symmetric group algebra (Proposition 2.14).
-
•
In Section 3, we define the notion of deformed monotone Hurwitz numbers, which are polynomials in the deformation parameter , motivated by the results of the previous section. We prove “deformed” analogues of existing results concerning the usual monotone Hurwitz numbers, such as a character formula (Proposition 3.3), a cut-and-join recursion (Theorem 3.7), and a one-point recursion (Theorem 3.12). It follows from the cut-and-join recursion that the coefficients of deformed monotone Hurwitz numbers are symmetric and unimodal (Proposition 3.9). On the basis of extensive numerical evidence, we conjecture that these polynomials are real-rooted (Conjecture 3.13) and that their roots satisfy remarkable interlacing phenomena (Conjecture 3.14).
-
•
In Section 4, we briefly introduce the topological recursion of Chekhov, Eynard and Orantin [11, 25] and then use a powerful result of Bychkov, Dunin-Barkowski, Kazarian and Shadrin [8] to prove that topological recursion on the spectral curve governs the deformed monotone Hurwitz numbers (Theorem 4.1). We propose the topological recursion as a mechanism to produce topological generalisations of sequences of polynomials with interesting properties. As a case study, we consider the weighted enumeration of dessins d’enfant — in other words, bicoloured maps — in which each black vertex receives a multiplicative weight . This produces a family of polynomials that also satisfies a cut-and-join recursion (Proposition 4.8) and the topological recursion (Theorem 4.10). In analogy with the case of deformed monotone Hurwitz numbers studied in the previous section, we conjecture that these polynomials exhibit real-rootedness and interlacing phenomena (Conjecture 4.11).
2 Weingarten calculus
2.1 Convolution formula and orthogonality relations
As mentioned in the introduction, the present work is concerned with integration on the complex Grasmannian for . We interpret this Grassmannian as the space of idempotent Hermitian matrices of rank , which admits three equivalent descriptions as per the following definition.
Definition 2.1.
Let denote the identity matrix and denote the matrix whose first diagonal entries are 1 and whose remaining entries are 0. For , define the space
The unitary group acts transitively on by conjugation, thus endowing it with the structure of a compact homogeneous space. Thus, the Haar measure on induces a -invariant normalised Haar measure on , which we denote succinctly by . (See [16] for an introduction to Haar measures.) The fact that the stabiliser of is allows us to identify with the complex Grassmannian .
For , define the function corresponding to the matrix element. Our primary goal is to calculate integrals of the form
where . We impose the technical assumption that for future convenience. However, this assumption has little bearing on our work, since we will study these matrix integrals in the regime of large and fixed ratio . The following elementary integrals will be of particular importance.
Definition 2.2 (Weingarten function).
For each permutation , define the integral
We refer to the function as the Weingarten function for . Here, we include the symmetric group , whose unique element is denoted and represents the empty permutation. In the notation for the Weingarten function, we suppress the dependence on and to avoid clutter.
The setup described above sits firmly in the realm of Weingarten calculus, which is broadly concerned with the calculation of integrals on compact groups and related objects with respect to the Haar measure [14]. Modern accounts of Weingarten calculus often rely on elegant algebraic approaches via Schur–Weyl duality [12, 15]. A direct use of such an argument is not immediately available for the case of integration over . We instead follow the approach via orthogonality relations utilised by Collins and Matsumoto [13], which is in turn inspired by the ideas contained in the seminal paper of Weingarten [49].
In the remainder of this section, we develop the Weingarten calculus for integration on in three parts. First, we prove a convolution formula that reduces general integrals of monomials in the matrix elements to the elementary ones defined in Definition 2.2. Second, we prove so-called orthogonality relations that completely determine all values of the Weingarten function. Third, we solve the linear system provided by these orthogonality relations, which connects naturally to the representation theory of the symmetric groups, particularly to the Jucys–Murphy elements in the symmetric group algebra.
Theorem 2.3 (Convolution formula).
Arbitrary integrals of monomials in the matrix elements of reduce to elementary integrals via the equation
Proof.
Write in the form for . Then the two sides of the desired equation can be equivalently expressed as integrals over .
LHS | |||
RHS |
Thus, the result would follow from the equation
However, this is a direct consequence of applying the convolution formula for [15, Corollary 2.4] to both sides. ∎
The following orthogonality relations for are inspired by the orthogonality relations obtained by Collins and Matsumoto in other settings for Weingarten calculus [13]. The proof crucially relies on the convolution formula of Theorem 2.3. For future reference, observe that we use the notational convention of composing permutations from right to left.
Theorem 2.4 (Orthogonality relations).
For each permutation , the Weingarten function satisfies the relation
Here, for any permutation that fixes , we write to denote the permutation that agrees with on the set .
Proof.
Let and consider the cases and separately.
Case 1. Suppose .
Consider the integral
() |
On the one hand, we can use and the definition of to express the integral as . On the other hand, we can apply the convolution formula directly to each summand. For the th summand, where , the convolution formula yields
For the th summand, where , the convolution formula yields
Adding these contributions over and equating with the expression we previously obtained for the integral () leads to
(2) |
Case 2. Suppose .
Let and consider the integral
() |
On the one hand, we have for all , so it follows that . Combining this observation with the convolution formula allows us to express the integral as
On the other hand, we can apply the convolution formula directly to each summand, resulting in a calculation analogous to that of Case 1. For the th summand, where , the convolution formula yields
For the th summand, where , the convolution formula yields
Adding these contributions over and equating with the expression we previously obtained for the integral () leads to
(3) |
Finally, the desired result is obtained by writing the two expressions for obtained in Sections 2.1 and 2.1 from the two separate cases in one formula, making use of the Kronecker delta notation. ∎
The orthogonality relations provide a non-degenerate linear system of equations that uniquely determines the Weingarten function. The example below shows that values of the Weingarten function can be computed explicitly and are rational functions of and .
Example 2.5.
By the orthogonality relations of Theorem 2.4 and the conjugacy invariance of , we obtain the following equations.
Using the values and , one can solve this linear system of equations to obtain the following unique solution.
In this way, one can begin with the base case and inductively obtain for in terms of for . Further values can be found in Appendix A.
2.2 Large expansion
A priori, computing the values of the Weingarten function via the integral definition is far from straightforward. However, as evidenced by the calculations of Example 2.5, the orthogonality relations of Theorem 2.4 uniquely determine the Weingarten function and imply that its values are rational functions of and . We will shortly see that their structure also leads directly to a large expansion for , with coefficients that enumerate factorisations of into transpositions that satisfy a certain monotonicity condition. This combinatorial structure can be understood in terms of paths in the Weingarten graph, which encode the result of recursively applying the orthogonality relations ad infinitum. The notion of a Weingarten graph was originally introduced by Collins and Matsumoto [13]. In the setting of integration over unitary groups, the Weingarten graph has two types of edges, which reflects the fact that the orthogonality relations in that case express the Weingarten function of a permutation in terms of two types of terms. In the setting of integration over , we have three terms appearing on the right side of the orthogonality relations, motivating the following definition.
Definition 2.6.
Define the Weingarten graph to be the infinite directed graph with vertex set and edge set , where:
-
•
the set comprises the “type ” edges, which are of the form
for and ;
-
•
the set comprises the “type ” edges, which are of the form
for with ; and
-
•
the set comprises the “type ” edges, which are of the form
for such that for some .
Remark 2.7.
The Weingarten graph in the setting of integration over appears in [13] and is the subgraph of obtained by removing all type edges.
Repeated application of the orthogonality relations leads to paths in the Weingarten graph . In this way, one is motivated to enumerate paths in the Weingarten graph in which edges of types , , receive multiplicative weights , , , respectively.
Definition 2.8.
A path in is a sequence of permutations
where is a directed edge of for each . We call the integer the length of the path and denote by the set of all paths from to . Define the weight of a path by
where denotes the number of edges of type on the path .
By construction, the Weingarten graph is a combinatorial encoding of the orthogonality relations and their repeated application. Starting with , iterating the orthogonality relations times produces the formula
By sending the number of iterations to infinity, one obtains the following large expansion of the Weingarten function.
Proposition 2.9.
For any permutation , we have the large expansion
At this point, let us consider the Weingarten function in the regime of large with fixed ratio .222One could of course consider other regimes, such as fixed , although this line of investigation did not appear to be as fruitful. For and , we have , which leads to
To develop a clearer combinatorial description of the coefficients appearing in the expansion above, we consider the correspondence between paths in the Weingarten graph and monotone factorisations. This connection already appears in the Weingarten calculus for unitary groups [13].
Definition 2.10.
A monotone factorisation of is a sequence of transpositions in such that
-
•
; and
-
•
if we write with , then .
Moreover, we call a monotone factorisation transitive if generate a transitive subgroup of . Denote the set of monotone factorisations of by and the length of by . We refer to the number of distinct elements of the multiset as the hive number of and denote this quantity by .
Given a path in the Weingarten graph from to , one can record the sequence of transpositions arising from the type edges and the type edges. The composition of these transpositions in reverse order recovers the permutation . Thus, a path gives rise to a monotone factorisation satisfying . Represent this construction via the map
In the analogous construction for the unitary case, one obtains a one-to-one correspondence between paths and monotone factorisations, but that is not the case here. Given a monotone factorisation of , there exists a unique path in containing only edges of types and . The number of pairs of consecutive – edges in the path is equal to the hive number . Any such pair can be replaced by a type edge to obtain an element of and every element of can be obtained in this way. Thus, we have . Moreover, keeping track of the effect on edge weights, we have the equality
These observations allow us to express the coefficients of the large expansion of in terms of monotone factorisations via the equation
The discussion above motivates the following definition and theorem, whose proof is immediate after setting in the previous equation.
Definition 2.11.
For and a permutation, let denote the weighted count of monotone factorisations of , where the weight of a monotone factorisation is . Let denote the analogous count restricted to transitive monotone factorisations.
Theorem 2.12 (Large expansion).
For a permutation , we have the following large expansion for fixed .
(4) |
2.3 Representation-theoretic interpretation
The unitary invariance of the Haar measure implies that the Weingarten function is a function of permutations that is constant on conjugacy classes. It is natural to ask whether it admits a natural description in the class algebra of the symmetric group. Such a description of the unitary Weingarten function is well understood and given in terms of the Jucys–Murphy elements in the symmetric group algebra [40, 44]. We will show that the Weingarten function can be expressed similarly.
The Jucys–Murphy elements are elements of the symmetric group algebra defined by
where we interpret the formula for as . They were introduced independently by Jucys [36] and Murphy [41] and their seemingly simple definition belies their remarkable properties. For example, they commute with each other and indeed, generate a maximal commutative subalgebra of . Any symmetric function of the Jucys–Murphy elements lies in the class algebra and the class expansions of such expressions are of significant interest, appearing in various contexts [30]. Furthermore, the Jucys–Murphy elements are essential elements of the Okounkov–Vershik approach to the representation theory of symmetric groups [45]. We will subsequently require the following results that date back to the seminal work of Jucys.
Proposition 2.13 (Jucys [36]).
-
(a)
The Jucys–Murphy elements satisfy
-
(b)
Let be a partition of and let the corresponding irreducible character of the symmetric group. We adopt the usual abuse of notation and consider as an element of the class algebra . If is a symmetric polynomial in variables, then
where denotes the multiset of contents in the Young diagram for . (The content of a box in row and column of a Young diagram is the number .)
It was shown by Novak [44] that the unitary Weingarten function can be naturally expressed in terms of the Jucys–Murphy elements via the formula
By considering the right side as a series in , one observes that the coefficients are homogeneous symmetric functions of the Jucys–Murphy elements. This leads directly to the notion of monotone Hurwitz numbers, which count monotone factorisations of a given permutation with a prescribed length. The analogue of the equation above for the Weingarten function is the following, which leads to a deformation of the monotone Hurwitz numbers, which are now weighted counts of monotone factorisations with weight equal to to the power of the hive number. This idea was already briefly introduced in Definition 2.11, but will be studied in further detail in Section 3.
Proposition 2.14.
For each positive integer , we have the following equality in , the centre of the symmetric group algebra.
Proof.
This is essentially an algebraic reformulation of the orthogonality relations of Theorem 2.4 and we will prove it by induction on . The base case corresponds to the fact that , which is an immediate consequence of the orthogonality relation for .
It remains to show that, for ,
where for any permutation , we write to denote the permutation that agrees with on the set and fixes . Consider multiplying both sides of this equation by and use the definition of the Jucys–Murphy element to show that
(5) | ||||
(6) |
The second equality is more subtle than the first and relies on the observation that for ,
Finally, the desired equality between equations 5 and 6 follows directly from the orthogonality relations of Theorem 2.4, which completes the proof. ∎
Proposition 2.14 implies the following character formula for the Weingarten function.
Corollary 2.15 (Character formula).
Consider a permutation and let be the identity. Then
where denotes the content of the box in a Young diagram of a partition. Here and throughout, we use the notation to denote that is a partition of .
Proof.
By the orthogonality of characters, the identity in can be expressed as . Use part (b) of Proposition 2.13 to write
By Proposition 2.14, is the coefficient of in the above expression, and the result follows. ∎
In summary, integrals on the space of matrices admit a convolution formula involving the Weingarten function . The Weingarten function is in turn determined by orthogonality relations, from which it follows that its values have a representation-theoretic interpretation in terms of Jucys–Murphy elements. An alternative perspective shows that the large expansion of the Weingarten function for fixed has coefficients that are weighted counts of monotone factorisations. In the next section, we investigate the combinatorics of these enumerations in further detail.
3 Deformed monotone Hurwitz numbers
3.1 Deformed monotone Hurwitz numbers
Monotone Hurwitz numbers enumerate monotone factorisations of permutations and arise as coefficients in the large expansion of the unitary Weingarten function [33]. Weingarten calculus on the space naturally motivates a deformation of the monotone Hurwitz numbers, obtained by counting with a weight equal to to the power of the hive number. This construction produces a family of polynomials with remarkable properties.
Recall from Definition 2.11 that denotes the weighted count of monotone factorisations of , while restricts the count to transitive monotone factorisations. To fit the usual nomenclature concerning Hurwitz-type problems, we introduce a new notation to reindex by “topology”, consider the enumeration as a function on cycle type, and normalise by the product of cycle lengths.
Definition 3.1 (Deformed monotone Hurwitz numbers).
For , let be any permutation with cycles of lengths and define
Here and throughout, we use the notation as a shorthand for the sum . We refer to as a disconnected deformed monotone Hurwitz number and as a connected deformed monotone Hurwitz number.
Remark 3.2.
The pair here can be interpreted as encoding the topology of a surface, with denoting the genus and the number of punctures, or marked points. A monotone factorisation of can be interpreted as the monodromy of a branched cover of Riemann surfaces, with representing the monodromy over . So and enumerate certain branched covers of by a genus surface with preimages of , the latter restricting the enumeration to connected covers. Consistent with this interpretation, we have that unless is a non-negative integer, while can be non-zero when is a negative integer. Note that the genus of a disconnected surface may be negative, since we consider the Euler characteristic to be additive over disjoint union, rather than the genus itself.
The character formula of Corollary 2.15 translates into one for deformed monotone Hurwitz numbers as follows.
Proposition 3.3 (Character formula).
Let denote the coefficient of in the series for at and let denote the value of the symmetric group character on a permutation of cycle type . Then
Proof.
Equate the expressions for in terms of monotone factorisations (Theorem 2.12) and characters of the symmetric group (Corollary 2.15), using the notation and .
The desired result is obtained by multiplying through by , extracting the coefficient of from both sides, and using Definition 3.1 for . ∎
The family of deformed monotone Hurwitz numbers is stored in the large expansion of the following matrix integral, which we interpret as the partition function for the enumeration. As usual, the partition function naturally stores the disconnected enumeration, while its logarithm stores the connected enumeration.
Proposition 3.4 (Matrix integral).
The large expansion of the formal matrix integral below is a partition function for the deformed monotone Hurwitz numbers. Here, we use the notation and represents the th power sum symmetric function, evaluated on the eigenvalues of the complex matrix . (It is common to see a power of in the partition function — the extra here can be removed by rescaling.)
Remark 3.5.
Recall that the parameter is encoded in the parameters of the space via the relation . The usual monotone Hurwitz numbers are recovered from their deformed counterparts defined above by setting . Perhaps surprisingly, this particular value of does not correspond to valid parameters and and in the matrix integral interpretation. It would be interesting to have a more direct connection between integration over and .
Remark 3.6.
It would also be natural to consider “double” deformed monotone Hurwitz numbers that take two partitions as arguments, rather than one. These enumerate monotone factorisations whose product with a permutation of cycle type given by the first partition produces a permutation of cycle type given by the second. Although this “double” enumeration deserves further attention, we refrain from developing this theory in the present work, since they do not give rise to polynomials with the same remarkable properties as the “single” enumeration.
3.2 Cut-and-join recursion
The notion of cut-and-join analysis was originally developed for single Hurwitz numbers, before being adapted to work in the monotone case [34, 32]. Deformed monotone Hurwitz numbers are amenable to a similar analysis, resulting in the following recursion.
Theorem 3.7 (Cut-and-join recursion).
Apart from the initial condition , the deformed monotone Hurwitz numbers satisfy
(7) |
where we use the notation and for .
Proof.
The proof is based on the case for the usual monotone Hurwitz numbers [32], with some additional care required to keep track of the deformation parameter . Consider multiplying Theorem 3.7 by and consider the right side to be composed of four terms in the obvious way.
Fix a permutation of cycle type and consider the quantity for some . Note that a transitive monotone factorisation of must have for some . So a transitive monotone factorisation of with length is equivalent to a monotone factorisation of with length for some , where is transitive. This can be expressed via the equation
where we introduce the boldface notation to denote the set , excluding multiplicities, for .
The sum on the right side can be broken down into further sums, depending on various cases. For a fixed value of , either is in the same cycle of as , or is in a different cycle. The transposition is referred to as a cut or a join of depending on these two cases, respectively. Moreover, if is a join of , then any monotone factorisation of , with transitive, must be transitive itself. Based on these considerations, each term in the equation above falls into one of three cases: for each and with transitive, either
-
•
is a join of and is transitive;
-
•
is a cut of and is transitive; or
-
•
is a cut of and is not transitive.
The first two cases can be expressed as one term to give the equation
(8) |
In the usual notation for deformed monotone Hurwitz numbers, the first term here precisely gives rise to the first and third terms of Theorem 3.7, while the second term here almost gives rise to the fourth term of Theorem 3.7. The remainder of the proof explains the meaning of “almost” here and how the second term of Theorem 3.7 accounts correctly for this anomaly.
If falls into the third case for some value of , then the transitive orbit of is split into two orbits of — one containing and one containing . In fact, these orbits are given by a partition of the disjoint cycles of into two classes, thus giving rise to the quadratic terms in Theorem 3.7. The only instance that and differ is when is in an orbit of size under , which then requires an additional weight of to be contributed. The factor in the second term of Theorem 3.7 precisely handles this adjustment, since we wish to remove one of the contributions appearing in the fourth term and weight it by an additional factor of .
Thus, in the correct final accounting of all terms, we see that the first term of equation 8 corresponds to the first and third terms of Theorem 3.7, while the second term of equation 8 corresponds to the second and fourth terms of Theorem 3.7. ∎
The cut-and-join recursion provides an effective algorithm for calculating deformed monotone Hurwitz numbers, which is more efficient than naive approaches. With this improved computational power comes the ability to make large-scale empirical observations on the structure of the deformed monotone Hurwitz numbers and a recursive means to prove them. The most evident of these observations are the symmetric and unimodal nature of the coefficients of .
Definition 3.8.
A polynomial with real coefficients is said to be:
-
•
symmetric if the sequence is palindromic;
-
•
unimodal if there exists an integer such that ; and
-
•
a -polynomial if it is a non-zero polynomial with non-negative coefficients that is both palindromic and unimodal.
If is a -polynomial, then there exists a unique integer such that . The quantity is called the darga of .
Proposition 3.9.
For , and , the deformed monotone Hurwitz number is a -polynomial of degree and darga .
Proof.
We use strong induction on the value of . The only deformed monotone Hurwitz number corresponding to is , which is indeed a -polynomial of degree and darga . Now consider a deformed monotone Hurwitz number corresponding to and rewrite the cut-and-join recursion of Theorem 3.7 as
Here, we have simply removed the two terms including from the final summation and incorporated them into the second term on the right side.
By the inductive hypothesis, all deformed monotone Hurwitz numbers appearing on the right side of the cut-and-join recursion are -polynomials. To prove that is also one, we rely on the following closure properties of -polynomials, proofs of which can be found in [47].
-
•
If and are -polynomials, then the product is a -polynomial with .
-
•
If and are -polynomials of the same darga , then is a -polynomial of darga .
It follows that each of the four terms on the right side of the equation above are -polynomials of darga , or possibly equal to the zero polynomial. So the entire right side, and hence , is a -polynomial of darga .
For any permutation for , one can construct a transitive monotone factorisation of with hive number . So the polynomial has a non-zero linear term, but no constant term. Combined with the fact that it is a -polynomial of darga , it follows that is a polynomial of degree . ∎
Remark 3.10.
We have thus far provided two interpretations for the polynomials that we refer to as deformed monotone Hurwitz numbers — one as weighted counts of monotone factorisations (Definition 3.1) and the other as coefficients in the large expansion of a matrix integral (Proposition 3.4). The symmetry of these polynomials does not appear to be immediately evident from either of these manifestations. In the matrix integral interpretation, the transformation encodes the transformation . This is natural from the perspective of integration on the Grassmannian, since , although further investigation of this symmetry is warranted. In the monotone factorisation interpretation, the symmetry of the polynomials implies that the number of transitive monotone factorisations of a permutation with hive number is equal to the number of transitive monotone factorisations of with hive number . This does not appear to be immediately obvious, so it would be interesting to have a direct combinatorial proof of this fact. It is worth remarking that the disconnected enumeration does not lead to symmetric polynomials, so the transitivity condition is important here.
Remark 3.11.
There are various other properties of polynomial coefficients that are often mentioned alongside symmetry and unimodality. The deformed monotone Hurwitz numbers constitute a family of polynomials that appears to satisfy many of these. As examples, we pose the following questions. Do the coefficients of the deformed monotone Hurwitz numbers exhibit log-concavity? Do they exhibit asymptotic normality?
As we will see in Section 4, deformed monotone Hurwitz numbers belong to a large class of enumerative problems governed by the topological recursion. In many such instances, the one-point invariants — in other words, those with — are governed by a recursion that is linear, rather than quadratic like the cut-and-join recursion. In previous work of Chaudhuri and the second author [9], it was shown that such one-point recursions exist for “weighted Hurwitz numbers”, to use the terminology of Alexandrov, Chapuy, Eynard and Harnad [1]. Furthermore, they gave an explicit algorithmic approach to obtain these recursions that is effective in many examples. In the case of deformed monotone Hurwitz numbers, one obtains the following result.
Theorem 3.12 (One-point recursion).
The one-point deformed monotone Hurwitz numbers — in other words, those with — satisfy
Observe that setting above recovers the known one-point recursion for monotone Hurwitz numbers [9]. Furthermore, setting and combining with equation 1 recovers the known linear recursion for Narayana polynomials [31].
3.3 Interlacing phenomena
Root conjectures
The cut-and-join recursion of Theorem 3.7 can be used to effectively compute a large number of deformed monotone Hurwitz numbers, some of which are shown in Appendix A. We previously stated in Proposition 3.9 that the coefficients of these polynomials are symmetric and unimodal. We now turn our attention to their roots, which exhibit rather striking behaviour that remains largely conjectural at present. The most apparent property concerning roots of deformed monotone Hurwitz numbers is the following.
Conjecture 3.13 (Real-rootedness).
For all , and , the deformed monotone Hurwitz number is a real-rooted polynomial in .
The roots of the deformed monotone Hurwitz numbers are not only real, but also possess interesting structure relative to each other. We say that a polynomial interlaces a polynomial if
-
•
the degree of is and the degree of is for some positive integer ;
-
•
has real roots and has real roots , allowing for multiplicity; and
-
•
.
If the inequalities above are strict, then we say that the polynomial strictly interlaces the polynomial . By convention, we will also say that a polynomial (strictly) interlaces a polynomial if is constant and is affine.
It is known that the sequence of Narayana polynomials interlace in the sense that interlaces for every positive integer [31]. Given equation 1, one may wonder whether an analogous property persists more generally across the whole family of deformed monotone Hurwitz numbers. Overwhelming empirical evidence suggests that this is indeed the case and we have the following.
Conjecture 3.14 (Interlacing).
The polynomial interlaces each of the polynomials
Conjecture 3.14 states that for fixed , one obtains a lattice of interlacing polynomials. It has been checked computationally for many cases, including the following, amounting to 1430 independent checks:
-
•
, , ;
-
•
, , ;
-
•
, , .
One can observe rich structure in the roots of the deformed monotone Hurwitz numbers and write down further conjectures. As an example, consider the following, which asserts that the roots of behave well for fixed and increasing . A great deal of data can be generated to support this conjecture, some examples of which appear in the tables of Figures 2 and 3.
Conjecture 3.15.
Fix positive integers whose sum is . As , the roots of the polynomial , in order from smallest to largest, respectively approach the values
Moreover, the convergence to a number less than is increasing from below, while the convergence to a non-zero number greater than is decreasing from above.
-5.0012679418 | -2.0003283655 | -0.49991792208 | -0.19994929518 | |||
-5.0012679418 | -2.0003283655 | -0.49991792208 | -0.19994929518 | |||
-5.0010564352 | -2.0002736067 | -0.49993160767 | -0.19995775151 | |||
-5.0012326847 | -2.0003192382 | -0.49992020316 | -0.19995070476 | |||
-5.0010564383 | -2.0002736143 | -0.49993160576 | -0.19995775139 | |||
-5.0010564362 | -2.0002736092 | -0.49993160703 | -0.19995775147 | |||
-5.0011739286 | -2.0003040277 | -0.49992400462 | -0.19995305387 | |||
-5.0010564383 | -2.0002736143 | -0.49993160576 | -0.19995775139 | |||
-5.0008802353 | -2.0002279852 | -0.49994301017 | -0.19996479678 | |||
-5.0010270659 | -2.0002660064 | -0.49993350723 | -0.19995892579 | |||
-5.0011004921 | -2.0002850163 | -0.49992875605 | -0.19995599000 | |||
-5.0008802353 | -2.0002279852 | -0.49994301017 | -0.19996479678 | |||
-5.0009781178 | -2.0002533320 | -0.49993667500 | -0.19996088293 | |||
-5.0010189060 | -2.0002638930 | -0.49993403543 | -0.19995925206 | |||
-5.0010189060 | -2.000263893081 | -0.49993403543 | -0.19995925206 |
An interlacing result
There is a rich theory concerning interlacing polynomials, as evidenced by the tome of Fisk [31]. The following lemma is a slight adaptation of [31, Lemma 1.82], and will be used to prove the and cases of Conjectures 3.13 and 3.14.
Lemma 3.16.
Let be a sequence of polynomials with non-negative coefficients, such that . Suppose that strictly interlaces , and that the sequence of polynomials satisfies the relation
(9) |
where is an affine function and is a quadratic function that is positive for . Then all polynomials in the sequence are real-rooted and strictly interlaces for every positive integer .
Proof.
Suppose that strictly interlaces . Since has non-negative integer coefficients, its roots can be written as . Then equation 9 evaluated at one of these roots yields
which implies that and have different signs. Since strictly interlaces , its values at must alternate in sign, leading to the following table. The entries indicate the signs of and , evaluated at and in the limit .
The intermediate value theorem implies that has real roots and that strictly interlaces . The result then follows by induction on . ∎
Proposition 3.17.
Conjectures 3.13 and 3.14 hold for and .
Proof.
Setting in the one-point recursion of Theorem 3.12 yields
(10) |
Divide both sides by to obtain a recursion governing the polynomials of the form of equation 9. It follows from Lemma 3.16 that this sequence is real-rooted and strictly interlacing. Therefore, the sequence is a sequence of real-rooted polynomials in which each polynomial interlaces the next.
For the genus 1 case, we argue in exactly the same way, using the as yet unproven claim that
(11) |
It remains to prove this relation, which we do using a generating function approach. Multiply both sides by and sum over all positive integers . Setting , the claim is equivalent to
(12) |
We borrow Theorem 4.1 from the next section, which allows us to explicitly compute via that
(13) |
where the coordinates and are related by . One can now check that the expression for of equation 13 satisfies the differential equation of equation 12 using a computer algebra system, thereby proving equation 11. ∎
Remark 3.18.
It is natural to seek higher genus analogues of equations 10 and 11, in order to extend the result and proof of Proposition 3.17. However, one can prove that there exists no recursion of the form
where are affine functions of , is an affine function of , and is a quadratic function of . So the exact argument used in the proof of Proposition 3.17 cannot extend to genus 2 and higher without some alteration.
4 Topological recursion
4.1 Topological recursion for deformed monotone Hurwitz numbers
The topological recursion arose in the mathematical physics literature as a formalism to capture loop equations in matrix models [11, 25]. There is now an extensive theory around topological recursion and it is known to govern a wide range of enumerative-geometric problems beyond the original matrix model context. These include: Hurwitz numbers and various generalisations [6, 4, 19, 27, 2, 23], monotone Hurwitz numbers [18], lattice points in moduli spaces of curves [42, 10], intersection theory on moduli spaces of curves [25, 24], Gromov–Witten theory of [43, 22], Gromov–Witten theory of toric Calabi–Yau threefolds [5, 26, 29], and cohomological field theories [22]. The topological recursion is also conjectured to govern certain quantum knot invariants, such as the large asymptotics of coloured Jones polynomials [17, 3] and coloured HOMFLY-PT polynomials [35].
In its original formulation due to Chekhov, Eynard and Orantin [11, 25], the topological recursion takes as input a spectral curve comprising a compact Riemann surface equipped with two meromorphic functions and a bidifferential with a double pole along the diagonal. These are required to satisfy mild technical assumptions, which we do not mention here. Indeed, rather than fully define the topological recursion, we refer the reader to the many existing expositions in the literature [18, 28]. For the present discussion, it suffices to note that the topological recursion uses the initial data to define the base cases and , and then produces so-called correlation differentials via the following recursive formula.333The reader should note that the various appearances of the topological recursion formula in the literature differ subtly, particularly in terms of the expression for , the definition of the recursion kernel , and the choice of sign for . Ultimately, the first two differences are inconsequential and a change of sign for simply sends each correlation differential to .
In various settings, the correlation differential acts as a generating function for enumerative-geometric quantities, where represents the genus of a surface and its number of punctures, or boundaries.
It was previously shown that the monotone Hurwitz numbers are governed by the topological recursion in the following sense [18]. Topological recursion on the spectral curve444The topological recursion is not sensitive to reparametrisation of the underlying plane curve, whose global description in this case is . We have expressed the spectral curve here using a different rational parametrisation to that appearing in [18], in order to align more closely with the statement and proof of Theorem 4.1.
produces correlation differentials satisfying
Here, represents the exterior derivative in the th coordinate and denotes a monotone Hurwitz number.
It is thus natural to consider whether the deformed monotone Hurwitz numbers satisfy the topological recursion on a deformation of the spectral curve above. Perhaps unsurprisingly, this is indeed the case and follows from the representation-theoretic interpretation for deformed monotone Hurwitz numbers of Corollary 2.15 and a powerful result of Bychkov, Dunin-Barkowski, Kazarian and Shadrin [8], which builds on previous work of Alexandrov, Chapuy, Eynard and Harnad [1].
Theorem 4.1.
Topological recursion on the spectral curve
produces correlation differentials satisfying
Proof.
The main theorem of Bychkov, Dunin-Barkowski, Kazarian and Shadrin in [8] states that the topological recursion governs the coefficients of certain KP tau functions of hypergeometric type. We present here a simplified version of their result that is fit for purpose.
Let be the series expansion of a rational function satisfying and let be a rational function satisfying . From this data, define the generating function
where is the set of partitions, represents the Schur function expressed in terms of power sum symmetric functions, and denotes the content of the box in the Young diagram of a partition. This is a KP tau function of hypergeometric type and there exist “weighted Hurwitz numbers” for , and such that
(14) |
Bychkov, Dunin-Barkowski, Kazarian and Shadrin [8] prove that topological recursion on the spectral curve
produces correlation differentials satisfying
We simply specialise this result to and . It remains to check that the monotone deformed Hurwitz numbers agree with the weighted Hurwitz numbers with this particular choice of data.
The exponential of equation 14 transforms the connected generating function to the disconnected generating function, so it suffices to show that the disconnected deformed monotone Hurwitz numbers satisfy
Here, denotes the subgroup of permutations in that fix . This automorphism factor arises because the summation in equation 14 is over all tuples of positive integers, rather than partitions.
Now use in the expression
to obtain
That this is equal to is precisely the content of Proposition 3.3, and this concludes the proof. ∎
Remark 4.2.
The meromorphic functions of Theorem 4.1 satisfy
which is the global form of the spectral curve. The general theory of topological recursion asserts that, for or , the correlation differential is a rational multidifferential on this curve with poles only at the zeroes of and further conditions on the pole structure [25]. This in turn leads to a polynomiality structure theorem for the deformed monotone Hurwitz numbers. Furthermore, the relation between topological recursion and cohomological field theories should lead to an interpretation of deformed monotone Hurwitz numbers as intersection numbers on moduli spaces of curves [24, 22]. However, we do not pursue this investigation further in the present work.
4.2 Topologising sequences of polynomials
From Catalan curves to Narayana curves
As previously mentioned in equation 1, the case of the deformed monotone Hurwitz numbers recovers the Narayana polynomials in the following way.
Hence, we consider the deformed monotone Hurwitz numbers to be a “topological generalisation” of the Narayana polynomials. This viewpoint appears throughout the literature, such as in Dumitrescu and Mulase’s exposition on generalised Catalan numbers [21].
We propose that the topological recursion can be used as a mechanism to “topologise” sequences of polynomials. One would apply topological recursion to a spectral curve with a deformation parameter to obtain a family of polynomials for , and such that recovers the original sequence of polynomials in . Furthermore, we claim that interesting properties of the original sequence of polynomials can persist into the topological generalisation. Examples of such properties are the symmetry and unimodality of coefficients as well as the interlacing of roots. These both hold for Narayana polynomials and are respectively proven (Proposition 3.9) and conjectured (Conjecture 3.14) to hold for deformed monotone Hurwitz numbers more generally.
There is more than one way to topologise a given sequence of polynomials. In order to obtain a topological generalisation of the Narayana numbers, it is natural to consider spectral curves that store Catalan numbers in the correlation differential and to consider a suitable -deformation. The literature suggests three natural candidates for such “Catalan spectral curves” and these curves govern monotone Hurwitz numbers, map enumeration, and dessin d’enfant enumeration. This leads to the following three “Narayana spectral curves” with , , and the meromorphic functions satisfying the following.
-
•
Monotone Hurwitz numbers. We have
which is the global form of the spectral curve for monotone Hurwitz numbers [18]. Promoting the Catalan numbers to the Narayana polynomials, one obtains
This is the global form of the spectral curve for deformed monotone Hurwitz numbers, which appeared above in Theorem 4.1.
-
•
Map enumeration. We have
which is the global form of the spectral curve for the enumeration of maps [11, 25]. Promoting the Catalan numbers to the Narayana polynomials, one obtains
To the best of our knowledge, this is the global form of a spectral curve that has not yet been studied in the literature. We do not investigate it further in the present work, largely due to the fact that it has genus 1 for generic values of . The topological recursion on spectral curves of positive genus is much less straightforward than on those of genus 0. Nevertheless, it would be interesting to understand this spectral curve and whether it governs a natural enumerative problem.
-
•
Dessin d’enfant enumeration. We have
which is the global form of the spectral curve for the enumeration of dessins d’enfant [20, 37]. Promoting the Catalan numbers to the Narayana polynomials, one obtains
This is the global form of a spectral curve with genus 0. It leads to a topological generalisation of the Narayana polynomials that differs from the deformed monotone Hurwitz numbers. We will consider its properties in the next section.
Remark 4.3.
We certainly do not claim that the spectral curves above provide the only natural topological generalisations of the Catalan numbers or the Narayana polynomials. Indeed, one could study spectral curves arising from or for judicious choices of and , and it is not unlikely that these would lead to interesting enumerative problems.
Deforming the dessin d’enfant enumeration
The previous discussion motivates the study of the spectral curve whose global form is . In parametrised form, this corresponds to the spectral curve
By construction, this spectral curve should govern a -deformation of the usual dessin d’enfant enumeration. In the following, we prove that this enumeration arises by attaching a multiplicative weight to each black vertex of a dessin d’enfant. Moreover, we observe that the polynomials arising from this construction empirically satisfy real-rootedness and interlacing properties analogous to those observed for deformed monotone Hurwitz numbers in Conjectures 3.13 and 3.14. Although the weighted enumeration of dessins d’enfant has been studied previously [37], these real-rootedness and interlacing properties have not been observed in the literature, to the best of our knowledge. Our discussion of the weighted enumeration of dessins d’enfant should be considered as a further case study promoting the viewpoint that topological recursion produces interesting topological generalisations of sequences of polynomials.
First, let us define dessins d’enfant — in other words, bicoloured maps — and their enumeration.
Definition 4.4.
A map is a finite graph — possibly with loops or multiple edges — embedded in a compact oriented surface such that the complement of the graph is a disjoint union of topological disks. We refer to these disks as faces and require that they are labelled , where denotes the number of faces.
A dessin d’enfant is a map whose vertices have been coloured black or white such that each edge is incident to one black vertex and one white vertex. The degree of a face is defined to be half the number of edges incident to it.
An equivalence between two maps (or dessins d’enfant) is a bijection between their respective vertices, edges and faces that is realised by an orientation-preserving homeomorphism of their underlying surfaces and preserves all adjacencies, incidences, labels (and colours). An automorphism of a map (or dessin d’enfant) is an equivalence from the map (or dessin d’enfant) to itself.
Definition 4.5.
For , and , let denote the weighted enumeration of connected genus dessins d’enfant with labelled faces of degrees . The weight attached to a dessin d’enfant is , where denotes the number of black vertices of and denotes the automorphism group of . Let denote the analogous count, including possibly disconnected dessins d’enfant.
Example 4.6.
Consider the dessin d’enfant enumeration . The only dessins d’enfant that contribute are the five pictured below. We have represented these as plane graphs, with the face labelled 1 corresponding to the bounded region and the face labelled 2 corresponding to the unbounded region.
The central dessin d’enfant has two automorphisms, while the remaining dessins d’enfant have one, so we have .
Proposition 4.7.
If , then . If , then is a polynomial of degree whose coefficients are symmetric.
Proof.
Consider a dessin d’enfant with genus and faces with degrees . By an Euler characteristic calculation, the number of vertices is . Since there must be at least one black vertex and at least one white vertex, we must have have .
We will show that, under the assumption , there exists a dessin d’enfant with exactly one white vertex. Take a polygon with sides, whose vertices are alternately coloured black and white. By identifying opposite edges, one obtains a dessin d’enfant on a genus surface, with one black vertex, one white vertex, and one face with degree . By adding appropriate edges from the black vertex to the white vertex, one can obtain a genus dessin d’enfant with one black vertex, one white vertex, and faces with any degrees satisfying . We can relax this condition to by adding appropriate black vertices with degree 1 that are adjacent to the unique white vertex. Such a dessin d’enfant contributes to degree in , and there can be no contributions in higher degree.
Finally, observe that switching the colours of the vertices of a dessin d’enfant changes the number of black vertices from to , leading to the desired symmetry in the coefficients of . ∎
There is an analogue of the cut-and-join recursion for deformed monotone Hurwitz numbers of Theorem 3.7 in the context of the dessin d’enfant enumeration. We omit the proof, since it can be found in the literature [37].
Proposition 4.8 (Cut-and-join recursion).
Apart from the initial condition , the dessin d’enfant enumeration satisfies
where we use the notation and for .
It is well-known that dessins d’enfant correspond to pairs of permutations acting on the edges — one permutation rotates edges around black vertices and the other rotates edges around white vertices. (Equivalent descriptions in the literature often use triples of permutations that compose to give the identity.) This leads to an expression for the disconnected enumeration of dessins d’enfant in terms of characters of the symmetric group, via Frobenius’s formula. For details, we recommend the book of Lando and Zonkin [38]. To incorporate the parameter , we observe that its exponent should equal the number of cycles in the permutation that rotates edges around black vertices. As a consequence of Jucys’s results described in part (a) of Proposition 2.13, we obtain the following result, for which we omit the proof.
Proposition 4.9.
The disconnected enumeration of dessins d’enfant is given by the formula
We are now in a position to prove that topological recursion governs the weighted enumeration of dessins d’enfant.555Topological recursion for a more general weighted enumeration of dessins d’enfant was previously studied by Kazarian and Zograf [37], although they obtain a different form for the spectral curve.
Theorem 4.10.
Topological recursion on the spectral curve
produces correlation differentials satisfying
Proof.
We again invoke the result of Bychkov, Dunin-Barkowski, Kazarian and Shadrin [8], noting that the more restricted result of Alexandrov, Chapuy, Eynard and Harnad [1] would suffice in this particular context. See the proof of Theorem 4.1 above for a statement of the result.
The representation-theoretic interpretation of the dessin d’enfant enumeration given in Proposition 4.9 implies that they are weighted Hurwitz numbers with the choice of data and . Hence, topological recursion on the spectral curve
produces correlation differentials satisfying
This does not yet match our spectral curve, since the enumeration is stored as an expansion at rather than at . We obtain the desired spectral curve by performing the following three transformations in order:
The first transformation changes the expansion at to an expansion at ; the second transformation preserves the correlation differentials [28, Section 4.2]; and the third transformation preserves the coefficients of the expansion, due to the symmetry of the coefficients of and the homogeneity property of topological recursion [28, Section 4.1]. Thus, we arrive at the desired result. ∎
Although the enumeration of dessins d’enfant has been studied previously [37], the following conjectural properties have not yet been observed in the literature, to the best of our knowlege.
Conjecture 4.11 (Real-rootedness and interlacing).
For all , and , the dessin d’enfant enumeration is a real-rooted polynomial in . Furthermore, the polynomial interlaces each of the polynomials
As with the deformed monotone Hurwitz numbers, we can effectively calculate using the cut-and-join recursion of Proposition 4.8 and explicitly check for real-rootedness and interlacing. Again, one finds a wealth of data to support Conjecture 4.11. We consider this as evidence that the real-rootedness and interlacing observed for deformed monotone Hurwitz numbers is not simply a sporadic artefact, but may be a more widespread phenomenon with a deep reason underlying it.
Remark 4.12.
As previously mentioned, dessins d’enfant correspond to pairs of permutations acting on the edges, or equivalently, triples of permutations that compose to give the identity. It follows from Proposition 2.13 that each permutation has a unique strictly monotone factorisation, defined analogously to a monotone factorisation in Definition 2.10, but with the stronger monotonicity requirement that . Using this result on the permutation that rotates edges around black vertices leads to the fact that the enumeration of dessins d’enfant can be interpreted as strictly monotone Hurwitz numbers.
Appendix A Data
Deformed monotone Hurwitz numbers
The deformed monotone Hurwitz numbers can be computed using the cut-and-join recursion or the topological recursion. These were both implemented in SageMath to produce the following table [48].
(1) | ||
---|---|---|
(2) | ||
(3) | ||
(4) | ||
(5) | ||
(1, 1) | ||
(2, 1) | ||
(3, 1) | ||
(2, 2) | ||
(4, 1) | ||
(3, 2) | ||
(1, 1, 1) | ||
(2, 1, 1) | ||
(3, 1, 1) | ||
(2, 2, 1) |
Dessin d’enfant enumeration
The weighted dessin d’enfant enumeration can be computed using the cut-and-join recursion or the topological recursion. These were both implemented in SageMath to produce the following table [48].
Weingarten functions
The value of can be calculated by inverting the orthogonality relations or by the character formula. The value of is the leading coefficient of , considered as a polynomial in .
References
- [1] A. Alexandrov, G. Chapuy, B. Eynard, and J. Harnad. Weighted Hurwitz numbers and topological recursion. Comm. Math. Phys., 375(1):237–305, 2020.
- [2] Gaëtan Borot, Norman Do, Maksim Karev, Danilo Lewański, and Ellena Moskovsky. Double Hurwitz numbers: polynomiality, topological recursion and intersection theory. Math. Ann., 2022.
- [3] Gaëtan Borot and Bertrand Eynard. All order asymptotics of hyperbolic knot invariants from non-perturbative topological recursion of A-polynomials. Quantum Topol., 6(1):39–138, 2015.
- [4] Vincent Bouchard, Daniel Hernández Serrano, Xiaojun Liu, and Motohico Mulase. Mirror symmetry for orbifold Hurwitz numbers. J. Differential Geom., 98(3):375–423, 2014.
- [5] Vincent Bouchard, Albrecht Klemm, Marcos Mariño, and Sara Pasquetti. Remodeling the B-model. Comm. Math. Phys., 287(1):117–178, 2009.
- [6] Vincent Bouchard and Marcos Mariño. Hurwitz numbers, matrix models and enumerative geometry. In From Hodge theory to integrability and TQFT tt*-geometry, volume 78 of Proc. Sympos. Pure Math., pages 263–283. Amer. Math. Soc., Providence, RI, 2008.
- [7] Francesco Brenti. Unimodal polynomials arising from symmetric functions. Proc. Amer. Math. Soc., 108(4):1133–1141, 1990.
- [8] Boris Bychkov, Petr Dunin-Barkowski, Maxim Kazarian, and Sergey Shadrin. Topological recursion for Kadomtsev–Petviashvili tau functions of hypergeometric type. arXiv:2012.14723 [math-ph], 2021.
- [9] Anupam Chaudhuri and Norman Do. Generalisations of the Harer-Zagier recursion for 1-point functions. J. Algebraic Combin., 53(2):469–503, 2021.
- [10] Anupam Chaudhuri, Norman Do, and Ellena Moskovsky. Local topological recursion governs the enumeration of lattice points in . arXiv:1906.06964 [math.GT], 2019.
- [11] Leonid Chekhov and Bertrand Eynard. Hermitian matrix model free energy: Feynman graph technique for all genera. J. High Energy Phys., (3):014, 18, 2006.
- [12] Benoît Collins. Moments and cumulants of polynomial random variables on unitary groups, the Itzykson-Zuber integral, and free probability. Int. Math. Res. Not., (17):953–982, 2003.
- [13] Benoît Collins and Sho Matsumoto. Weingarten calculus via orthogonality relations: new applications. ALEA Lat. Am. J. Probab. Math. Stat., 14(1):631–656, 2017.
- [14] Benoît Collins, Sho Matsumoto, and Jonathan Novak. The Weingarten calculus. Notices Amer. Math. Soc., 69(5):734–745, 2022.
- [15] Benoît Collins and Piotr Śniady. Integration with respect to the Haar measure on unitary, orthogonal and symplectic group. Comm. Math. Phys., 264(3):773–795, 2006.
- [16] Joe Diestel and Angela Spalsbury. The joys of Haar measure, volume 150 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2014.
- [17] Robbert Dijkgraaf, Hiroyuki Fuji, and Masahide Manabe. The volume conjecture, perturbative knot invariants, and recursion relations for topological strings. Nuclear Phys. B, 849(1):166–211, 2011.
- [18] Norman Do, Alastair Dyer, and Daniel V. Mathews. Topological recursion and a quantum curve for monotone Hurwitz numbers. J. Geom. Phys., 120:19–36, 2017.
- [19] Norman Do, Oliver Leigh, and Paul Norbury. Orbifold Hurwitz numbers and Eynard-Orantin invariants. Math. Res. Lett., 23(5):1281–1327, 2016.
- [20] Norman Do and Paul Norbury. Topological recursion for irregular spectral curves. J. Lond. Math. Soc. (2), 97(3):398–426, 2018.
- [21] Olivia Dumitrescu and Motohico Mulase. Lectures on the topological recursion for Higgs bundles and quantum curves. In The geometry, topology and physics of moduli spaces of Higgs bundles, volume 36 of Lect. Notes Ser. Inst. Math. Sci. Natl. Univ. Singap., pages 103–198. World Sci. Publ., Hackensack, NJ, 2018.
- [22] P. Dunin-Barkowski, N. Orantin, S. Shadrin, and L. Spitz. Identification of the Givental formula with the spectral curve topological recursion procedure. Comm. Math. Phys., 328(2):669–700, 2014.
- [23] Petr Dunin-Barkowski, Reinier Kramer, Alexandr Popolitov, and Sergey Shadrin. Loop equations and a proof of Zvonkine’s -ELSV formula. arXiv:1905.04524 [math.AG], 2020.
- [24] B. Eynard. Invariants of spectral curves and intersection theory of moduli spaces of complex curves. Commun. Number Theory Phys., 8(3):541–588, 2014.
- [25] B. Eynard and N. Orantin. Invariants of algebraic curves and topological expansion. Commun. Number Theory Phys., 1(2):347–452, 2007.
- [26] B. Eynard and N. Orantin. Computation of open Gromov-Witten invariants for toric Calabi-Yau 3-folds by topological recursion, a proof of the BKMP conjecture. Comm. Math. Phys., 337(2):483–567, 2015.
- [27] Bertrand Eynard, Motohico Mulase, and Bradley Safnuk. The Laplace transform of the cut-and-join equation and the Bouchard-Mariño conjecture on Hurwitz numbers. Publ. Res. Inst. Math. Sci., 47(2):629–670, 2011.
- [28] Bertrand Eynard and Nicolas Orantin. Topological recursion in enumerative geometry and random matrices. J. Phys. A, 42(29):293001, 117, 2009.
- [29] Bohan Fang, Chiu-Chu Melissa Liu, and Zhengyu Zong. On the remodeling conjecture for toric Calabi-Yau 3-orbifolds. J. Amer. Math. Soc., 33(1):135–222, 2020.
- [30] Valentin Féray. On complete functions in Jucys-Murphy elements. Ann. Comb., 16(4):677–707, 2012.
- [31] Steve Fisk. Polynomials, roots, and interlacing. arXiv:math/0612833 [math.CA], 2008.
- [32] I. P. Goulden, Mathieu Guay-Paquet, and Jonathan Novak. Monotone Hurwitz numbers in genus zero. Canad. J. Math., 65(5):1020–1042, 2013.
- [33] I. P. Goulden, Mathieu Guay-Paquet, and Jonathan Novak. Monotone Hurwitz numbers and the HCIZ integral. Ann. Math. Blaise Pascal, 21(1):71–89, 2014.
- [34] I. P. Goulden, D. M. Jackson, and A. Vainshtein. The number of ramified coverings of the sphere by the torus and surfaces of higher genera. Ann. Comb., 4(1):27–46, 2000.
- [35] Jie Gu, Hans Jockers, Albrecht Klemm, and Masoud Soroush. Knot invariants from topological recursion on augmentation varieties. Comm. Math. Phys., 336(2):987–1051, 2015.
- [36] A.-A. A. Jucys. Symmetric polynomials and the center of the symmetric group ring. Rep. Mathematical Phys., 5(1):107–112, 1974.
- [37] Maxim Kazarian and Peter Zograf. Virasoro constraints and topological recursion for Grothendieck’s dessin counting. Lett. Math. Phys., 105(8):1057–1084, 2015.
- [38] Sergei K. Lando and Alexander K. Zvonkin. Graphs on surfaces and their applications, volume 141 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2004. With an appendix by Don B. Zagier, Low-Dimensional Topology, II.
- [39] Sho Matsumoto. Weingarten calculus for matrix ensembles associated with compact symmetric spaces. Random Matrices Theory Appl., 2(2):1350001, 26, 2013.
- [40] Sho Matsumoto and Jonathan Novak. Jucys-Murphy elements and unitary matrix integrals. Int. Math. Res. Not. IMRN, (2):362–397, 2013.
- [41] G. E. Murphy. A new construction of Young’s seminormal representation of the symmetric groups. J. Algebra, 69(2):287–297, 1981.
- [42] Paul Norbury. String and dilaton equations for counting lattice points in the moduli space of curves. Trans. Amer. Math. Soc., 365(4):1687–1709, 2013.
- [43] Paul Norbury and Nick Scott. Gromov-Witten invariants of and Eynard-Orantin invariants. Geom. Topol., 18(4):1865–1910, 2014.
- [44] Jonathan I. Novak. Jucys-Murphy elements and the unitary Weingarten function. In Noncommutative harmonic analysis with applications to probability II, volume 89 of Banach Center Publ., pages 231–235. Polish Acad. Sci. Inst. Math., Warsaw, 2010.
- [45] Andrei Okounkov and Anatoly Vershik. A new approach to representation theory of symmetric groups. Selecta Math. (N.S.), 2(4):581–605, 1996.
- [46] Robert A. Sulanke. Counting lattice paths by Narayana polynomials. Electron. J. Combin., 7:Research Paper 40, 9, 2000.
- [47] Hua Sun, Yi Wang, and Hai Xia Zhang. Polynomials with palindromic and unimodal coefficients. Acta Math. Sin. (Engl. Ser.), 31(4):565–575, 2015.
- [48] The Sage Developers. SageMath, the Sage Mathematics Software System (Version 9.8), 2023. https://www.sagemath.org.
- [49] Don Weingarten. Asymptotic behavior of group integrals in the limit of infinite rank. J. Mathematical Phys., 19(5):999–1001, 1978.
- [50] Doron Zeilberger. A one-line high school algebra proof of the unimodality of the Gaussian polynomials for . In -series and partitions (Minneapolis, MN, 1988), volume 18 of IMA Vol. Math. Appl., pages 67–72. Springer, New York, 1989.