The Weil–Petersson volumes of moduli spaces of hyperbolic surfaces with geodesic boundaries are known to be given by polynomials in the boundary lengths. These polynomials satisfy Mirzakhani’s recursion formula, which fits into the general framework of topological recursion. We generalize the recursion to hyperbolic surfaces with any number of special geodesic boundaries that are required to be tight. A special boundary is tight if it has minimal length among all curves that separate it from the other special boundaries. The Weil–Petersson volume of this restricted family of hyperbolic surfaces is shown again to be polynomial in the boundary lengths. This remains true when we allow conical defects in the surface with cone angles in (0, π) in addition to geodesic boundaries. Moreover, the generating function of Weil–Petersson volumes with fixed genus and a fixed number of special boundaries is polynomial as well, and satisfies a topological recursion that generalizes Mirzakhani’s formula. This work is largely inspired by recent works by Bouttier, Guitter, and Miermont [Ann. Henri Lebesgue 5, 1035–1110 (2022)] on the enumeration of planar maps with tight boundaries. Our proof relies on the equivalence of Mirzakhani’s recursion formula to a sequence of partial differential equations (known as the Virasoro constraints) on the generating function of intersection numbers. Finally, we discuss a connection with Jackiw–Teitelboim (JT) gravity. We show that the multi-boundary correlators of JT gravity with defects are expressible in the tight Weil–Petersson volume generating functions, using a tight generalization of the JT trumpet partition function.

In the celebrated work1 Mirzakhani established a recursion formula for the Weil–Petersson volume Vg,n(L) of the moduli space of genus-g hyperbolic surfaces with n labeled boundaries of lengths L=(L1,,Ln)R>0n. Denoting [n] = {1, 2, …, n} and using the notation LI=(Li)iI, I ⊂ [n], for a subsequence of L and LÎ=L[n]\I, the recursion can be expressed for (g, n) ∉ {(0, 3), (1, 1)} as
(1)
where
(2)
Together with V0,3(L) = 1 and V1,1(L)=L1248+π212 this completely determines Vg,n as a symmetric polynomial in L12,,Ln2 of degree 3g − 3 + n.

This recursion formula remains valid1,2 when we replace one or more of the boundaries by cone points with cone angle αi ∈ (0, π) if we assign to it an imaginary boundary length Li = i. Cone points with angles in (0, π) are called sharp, as opposed to blunt cone points that have angle in (π, 2π). The Weil–Petersson volume of the moduli space of genus-g surfaces with n geodesic boundaries or sharp cone points is thus correctly computed by the polynomial Vg,n(L).

It was recognized by Eynard and Orantin3 that Mirzakhani’s recursion (in the case of geodesic boundaries) fits the general framework of topological recursion. To state this result explicitly one introduces for any g, n ≥ 0 satisfying 3g − 3 + n ≥ 0 the Laplace transformed4 Weil–Petersson volumes
(3)
which are even polynomials in z11,,zn1 of degree 6g − 4 + 2n, while setting
(4)
Then Mirzakhani’s recursion (1) translates into the recursion (Ref. 3, Theorem 2.1)
(5)
valid when g, n ≥ 1 and 3g − 3 + n ≥ 0, which one may recognize as the recursion for the invariants ωg,n(0)(z) of the complex curve
(6)
The main purpose of the current work is to generalize these recursion formulas to hyperbolic surfaces with so-called tight boundaries, which we introduce now.
Let Sg,n be a fixed topological surface of genus g with n boundaries and Tg,n(L) the Teichmüller space of hyperbolic structures on Sg,n with geodesic boundaries of lengths L=(L1,,Ln)R>0n. Denote the boundary cycles5 by 1, …, n and the free homotopy class of a cycle γ in Sg,n by [γ]Sg,n. For a hyperbolic surface XTg,n(L) and a cycle γ, we denote by γ(X) the length of γ, in particular i(X)=Li. The mapping class group of Sg,n is denoted Modg,n and the quotient of Tg,n(L) by its action leads to the moduli space
Let us denote by Sg,n,pSg,n+p the topological surface obtained from Sg,n+p by capping off the last p boundaries with disks (Fig. 1). Note that the free homotopy classes []Sg,n+p of Sg,n+p are naturally partitioned into the free homotopy classes []Sg,n,p of Sg,n,p. In particular, [j]Sg,n+p for j = n + 1, …, n + p are all contained in the null-homotopy class of Sg,n,p. For i = 1, …, n the boundary i of XMg,n+p(L) is said to be tight in Sg,n,p if i is the only simple cycle γ in [i]Sg,n,p of length γ(X) ≤ Li. Remark that both [i]Sg,n+p and [i]Sg,n,p for i = 1, …, n are Modg,n+p-invariant, so these classes are well-defined at the level of the moduli space. This allows us to introduce the moduli space of tight hyperbolic surfaces
(7)
Note that Mg,n,0tight(L)=Mg,0,ntight(L)=Mg,n(L), while M0,1,ptight(L)= because 1 is null-homotopic and M0,2,ptight(L)= because [1]S0,2,p=[2]S0,2,p and therefore 1 and 2 can never both be the unique shortest cycle in their class. In general, it is an open subset of Mg,n+p(L) and therefore it inherits the Weil–Petersson symplectic structure and Weil–Petersson measure dμWP from Mg,n+p(L). The corresponding tight Weil–Petersson volumes are denoted
(8)
such that Tg,n,0(L) = Tg,0,n(L) = Vg,n(L) and T0,1,p(L) = T0,2,p(L) = 0.
FIG. 1.

A surface Sg,n+p (left) and the result Sg,n,p (right) of capping off the last p boundaries, with in this case g = 1, n = 4 and p = 6. If the original surface was equipped with a hyperbolic structure, it is natural to think of the resulting surface as the Riemannian manifold obtained by attaching constantly curved hemispheres of appropriate diameter to the p geodesic boundaries.

FIG. 1.

A surface Sg,n+p (left) and the result Sg,n,p (right) of capping off the last p boundaries, with in this case g = 1, n = 4 and p = 6. If the original surface was equipped with a hyperbolic structure, it is natural to think of the resulting surface as the Riemannian manifold obtained by attaching constantly curved hemispheres of appropriate diameter to the p geodesic boundaries.

Close modal

We can extend this definition to the case in which one or more of the boundaries n+1, …, n+p is replaced by a sharp cone point with cone angle αi ∈ (0, π). In this case we make the usual identification Li = i, and still denote the corresponding Weil–Petersson volume by Tg,n,p(L). Our first result is the following.

Proposition 1.

For g, n, p ≥ 0 such that 2g − 3 + n ≥ 0, the tight Weil–Petersson volume Tg,n,p(L) of genus g surfaces with n tight boundaries and p geodesic boundaries or sharp cone points is a polynomial in L12,,Ln+p2 of degree 3g − 3 + n + p that is symmetric in L1, …, Ln and symmetric in Ln+1, …, Ln+p.

For most of the upcoming results we maintain the intuitive picture that the tight boundaries are the “real” boundaries of the surface, whose number and lengths we specify, while we allow for an arbitrary number of other boundaries or cone points that we treat as defects in the surface. To this end we would like to encode the volume polynomials in generating functions that sum over the number of defects with appropriate weights. A priori it is not entirely clear what is the best way to organize such generating functions, so to motivate our definition we take a detour to a natural application of Weil–Petersson volumes in random hyperbolic surfaces.

If we fix g, n and L[0,i(0,π))n, then upon normalization by 1/Vg,n(L) the Weil–Petersson measure dμWP provides a well-studied probability measure on Mg,n(L) defining the Weil–Petersson random hyperbolic surface, see e.g. Refs. 6–10. A natural way to extend the randomness to the boundary lengths or cone angles is by choosing a (Borel) measure μ on [0, ) ∪ i(0, π) and first sampling L[0,i(0,π))n from the probability measure
(9)
and then sampling a Weil–Petersson random hyperbolic surface on Mg,n(L). If the genus-g partition function11 
(10)
converges, we can furthermore make the size n ≥ 0 random by sampling it with the probability μn(Vg,n)/(n!Fg[μ]). The resulting random surface (of random size) is called the genus-g Boltzmann hyperbolic surface with weight μ. See the upcoming work12 for some of its statistical properties.
A natural extension is to consider the genus-g Boltzmann hyperbolic surface with n tight boundaries of length L = (L1, …, Ln), where the number p of defects and their boundary lengths/cone angles K = (K1, …, Kp) are random. The corresponding partition function is
(11)
If it is finite, we can sample p with probability μp(Tg,n,p)(L)/(p!Tg,n(Lμ]) and then K from the probability measure
(12)
and then finally a random tight hyperbolic surface from the probability measure dμWP/Tg,n,p(L, K) on Mg,n,ptight(L,K). Note that for μ = 0 the genus-g Boltzmann hyperbolic surface with n tight boundaries reduces to the Weil–Petersson random hyperbolic surface we started with.

The important observation for the current work is that the partition functions Fg[μ] and Tg,n(Lμ] of these random surfaces can be thought of as (multivariate, exponential) generating functions of the volumes Vg,n(L) and Tg,n,p(L, K) if we treat μ as a formal generating variable. Since we will not be concerned with the details of the measures (9) and (12) and Fg[μ] and Tg,n(Lμ] only depend on the even moments ∫L2kdμ(L), we can instead take these moments as the generating variables.

To be precise, we let a weight μ be a real linear function on the ring of even, real polynomials (i.e., μR[K2]*). For an even real polynomial f we use the suggestive notation
(13)
making it clear that the notion of weight generalizes the Borel measure described in the intermezzo above. For L ∈ [0, ) ∪ i(0, π), the Borel measure given by the delta measure δL at L gives a simple example of a weight μ = δL satisfying δL(f) = f(L). The choice of weight μ is clearly equivalent to the choice of a sequence of times (t0,t1,)RZ0 recording the evaluations of μ on the even monomials, up to a conventional normalization,
(14)
Naturally we can interpret μp(R[K2]*)p as an element of (R[K2]p)*R[K12,,Kp2]* by setting
for even polynomials f1, …, fp and extending by linearity. More generally, we can view μp as a linear map R[L12,,Ln2,K12,,Kp2]R[L12,,Ln2][K12,,Kp2]R[L12,,Ln]. We use the notation
(15)
One can then naturally introduce the generating function F[μ] of a collection of symmetric, even polynomials f1(L1), f2(L1, L2), … via F[μ]=p0μp(fp)p!. Then the generating function of tight Weil–Petersson volumes is defined to be
(16)
which we interpret in the sense of a formal power series, so we do not have to worry about convergence. We could make this more precise by fixing a weight μ and considering Tg,n(L;xμ]R[[x]] as a univariate formal power series in x. Or we could view Tg,n(Lμ] as a multivariate formal power series in the times (t0, t1, …) defined in (14).
What is important is that we can make sense of the functional derivative δδμ(L) on these types of series defined by
(17)
In particular, if f(L, K), with L = (L1, …, Ln) and K = (K1, …, Kp), is an even polynomial that is symmetric in K1, …, Kp then
(18)
At the level of the generating function we thus have
(19)
In terms of formal power series in the times we may instead identify the functional derivative in terms of the formal partial derivatives as
(20)
To state our main results about Tg,n(Lμ], we need to introduce the generating function R[μ] as the unique formal power series solution satisfying R[μ] = dμ(L) + O(μ2) to
(21)
where I0 and J1 are (modified) Bessel functions. Let also the moments Mk[μ] be the defined recursively via
(22)
where the reciprocal in the first identity makes sense because δRδμ(0)=1+O(μ). Alternatively, for k ≥ 0 we may express Mk as
(23)
where Z(k+1)(rμ] denotes the (k + 1)th derivative of Z(rμ] with respect to r.
We further consider the series
(24)
which we both interpret as formal power series in u with coefficients that are formal power series in μ. The reciprocal in the second definition is well-defined because η(0) = M0[μ] = 1 + O(μ). We can now state our main result that generalizes Mirzakhani’s recursion formula.

Theorem 2.
The tight Weil–Petersson volume generating functions Tg,n(Lμ] satisfy
(25)
which is the same recursion formula (1) as for the Weil–Petersson volumes Vg,n(L) except that the kernel K0(x, t) is replaced by the “convolution”
(26)
where X(z) = X(zμ] is a measure on R determined by its two-sided Laplace transform
(27)
Furthermore, we have
(28)
(29)

We will not specify precisely what it means to have a measure X(zμ] that itself is a formal power series in μ. Importantly its moments dX(z)zp=(1)pp![up]X̂(u;μ] are formal power series in μ, so for any x,tR
(30)
is a formal power series in μ as well.
In the case μ = 0, it is easily verified that
(31)
(32)
so X̂(u;0]=1 and X(z; 0] = δ0(z) and therefore one retrieves Mirzakhani’s kernel K(x, t) = K0(x, t). Given that the form of Mirzakhani’s recursion is unchanged except for the kernel, this strongly suggests that the Laplace transforms
(33)
of the tight Weil–Petersson volumes can be obtained as invariants in the framework of topological recursion as well. When μ = 0 this reduces to the Laplace-transformed Weil–Petersson volumes ωg,n(z;0]=ωg,n(0)(z) defined in (3). The following theorem shows that this is the case in general.

Theorem 3.
Setting ω0,2(z)=(z1z2)2 and ω0,0(z) = ω0,1(z) = 0, the Laplace transforms (33) satisfy for every g ≥ 0, n ≥ 1 such that 3g − 3 + n ≥ 0 the recursion
(34)
These correspond precisely to the invariants of the curve
(35)

Another consequence of Theorem 2 is that for all g ≥ 0, n ≥ 1, such that 3g − 3 + n ≥ 0, the tight Weil–Petersson volume Tg,n(Lμ] is expressible as a rational polynomial in L12,,Ln2 and M01,M1,M2,.

Besides satisfying a recursion in the genus g and the number of tight boundaries n, these also satisfy a recurrence relation in n only.

Theorem 4.
For all g, n ≥ 0, such that n ≥ 3 for g = 0 and n ≥ 1 for g = 1, we have that
(36)
where Pg,n(L,m) is a rational polynomial in L12,,Ln2,m1,,m3g3+n. This polynomial is symmetric and of degree 3g − 3 + n in L12,,Ln2, while Pg,n(σL,σm1,σ2m2,σ3m3,) is homogeneous of degree 3g − 3 + n in σ. For all g ≥ 0, n ≥ 1 such that 2g − 3 + n > 0 the polynomial Pg,n(L,m) can be obtained from Pg,n1(L,m) via the recursion relation
(37)
where we use the shorthand notation dLLf(L,)=0Ldxxf(x,). Furthermore, we have
(38)
(39)
and Pg,0 for g ≥ 2 is given by
(40)
where τ2d2τ3d3g are the ψ-class intersection numbers on Deligne–Mumford compactification of the moduli space of curves M̄g,n with n = kdk ≤ 3g − 3 marked points.

For instance, the first few applications of the recursion yield
Note that this provides a relatively efficient way of calculating the Weil–Petersson volumes Vg,n(L) from the polynomial Pg,0, since
A simple corollary of Theorem 4 is that the volumes satisfy string and dilaton equations generalizing those for the Weil–Petersson volumes derived by Do and Norbury in Ref. 13, Theorem 2.

Corollary 5.
For all g ≥ 0 and n ≥ 1, such that 2g − 3 + n > 0, we have the identities
(41)
(42)
where the notation [L12p]Tg,n(L;μ] refers to the coefficient of L12p in the polynomial Tg,n(Lμ].

As explained in Ref. 13, the string and dilaton equations for symmetric polynomials, in particular for the Weil–Petersson volumes, give rise to a recursion in n for genus 0 and 1. Using Theorem 4, we also get such a recursion for higher genera in the case of tight Weil–Petersson volumes.

This work is largely inspired by the recent work14 of Bouttier, Guitter, and Miermont. There the authors consider the enumeration of planar maps with three boundaries, i.e., graphs embedded in the triply punctured sphere, see the left side of Fig. 2. Explicit expressions for the generating functions of such maps, also known as pairs of pants, with controlled face degrees were long known, but they show that these generating functions become even simpler when restricting to tight pairs of pants, in which the three boundaries are required to have minimal length (in the sense of graph distances) in their homotopy classes. They obtain their enumerative results on tight pairs of pants in a bijective manner by considering a canonical decomposition of a tight pair of pants into certain triangles and diangles, see Fig. 2.

FIG. 2.

An example of a tight pair of pants, i.e., a planar map with three boundaries (in blue) of minimal length in their homotopy class in the triply punctured sphere. On the right the canonical decomposition described in Ref. 14. Figure adapted from Ref. 14, Fig. 1.1.

FIG. 2.

An example of a tight pair of pants, i.e., a planar map with three boundaries (in blue) of minimal length in their homotopy class in the triply punctured sphere. On the right the canonical decomposition described in Ref. 14. Figure adapted from Ref. 14, Fig. 1.1.

Close modal

Our result (28) for the genus-0 tight Weil–Petersson volumes with three distinguished boundaries can be seen as the analogue of Ref. 14, Theorem 1.1, although less powerful because our proof is not bijective. Instead, we derive generating functions of tight Weil–Petersson volumes from known expressions in the case of ordinary Weil–Petersson volumes. The general idea is that a genus-0 hyperbolic surface with two distinguished (but not necessarily tight) boundaries can unambiguously be cut along a shortest geodesic separating those two boundaries, resulting in a pair of certain half-tight cylinders (Fig. 4). Also a genus-g surface with n distinguished (not necessarily tight) boundaries can be shown to decompose into a tight hyperbolic surface and n half-tight cylinders. The first decomposition uniquely determines the Weil–Petersson volumes of the moduli spaces of half-tight cylinders, while the second determines the tight Weil–Petersson volumes. This relation is at the basis of Proposition 1.

To arrive at the recursion formula of Theorem 2 we follow the line or reasoning of Mirzakhani’s proof15 of Witten’s conjecture16 (proved first by Kontsevich17). She observes that the recursion equation (1) implies that the generating function of certain intersection numbers satisfies an infinite family of partial differential equations, the Virasoro constraints. Mulase and Safnuk18 have observed that the reverse implication is true as well. We will demonstrate that the generating functions of tight Weil–Petersson volumes and ordinary Weil–Petersson volumes are related in a simple fashion when expressed in terms of the times (14) and that the former obey a modified family of Virasoro constraints. These constraints in turn are equivalent to the generalized recursion of Theorem 2.

Mirzakhani’s recursion formula has a bijective interpretation.1 Upon multiplication by 2L1 the left-hand side 2L1Vg,n(L) accounts for the volume of surfaces with a marked point on the first boundary. Tracing a geodesic ray from this point, perpendicularly to the boundary, until it self-intersects or hits another boundary allows one to canonically decompose the surface into a hyperbolic pair of pants (three-holed sphere) and one or two smaller hyperbolic surfaces. The terms on the right-hand side of Ref. 1 precisely take into account the Weil–Petersson volumes associated to these parts and the way they are glued.

It is natural to expect that Theorem 2 admits a similar bijective interpretation, in which the surface decomposes into a tight pair of pants (a sphere with 3 + p boundaries, three of which are tight) and one or two smaller tight hyperbolic surfaces. However, Mirzakhani’s ray shooting procedure does not generalize in an obvious way. Nevertheless, working under the assumption that a bijective decomposition exists, one is led to suspect that the generalized kernel K(x, tμ] of Theorem 2 contains important information about the geometry of tight pairs of pants. Moreover, one would hope that this geometry can be further understood via a decomposition of the tight pairs of pants themselves analogous to the planar map case of Ref. 14 described above.

Since a genus-0 surface XM0,3+p(0,0,0,L) with three distinguished cusps is always a tight pair of pants (since the zero length boundaries are obviously minimal), a consequence of a bijective interpretation of Theorem 2 is a conjectural interpretation of the series X̂(u;μ]=sin(2πu)/(2πuη) in (27) in terms of the hyperbolic distances between the three cusps. To be precise, let c1, c2, c3 be unit-length horocycles around the three cusps and Δ(X)=dhyp(c1,c2)dhyp(c1,c3) the difference in hyperbolic distance between two pairs, then it is plausible that
(43)
Or in the probabilistic terms of Sec. I C, the measure M0[μ]X(zμ] on R, which integrates to 1 due to (28), is the probability distribution of the random variable 2Δ(X) in a genus-0 Boltzmann hyperbolic surface X with weight μ. In upcoming work we shall address this conjecture using very different methods.

Another natural question to ask is whether the generalization of the spectral curve (6) of Weil–Petersson volumes to the one of tight Weil–Petersson volumes in Theorem 3 can be understood in the general framework of deformations of spectral curves in topological recursion.19 

The structure of the paper is as follows:

In Sec. II we introduce the half-tight cylinder, which allows us to do tight decomposition of surfaces, which relates the regular hyperbolic surfaces to the tight surfaces. Using the decomposition we prove Proposition 1.

In Sec. III we consider the generating functions of (tight) Weil–Petersson volumes and their relations. Furthermore, we use the Virasoro constraints to prove Theorem 2, Theorem 4, and Corollary 5.

In Sec. IV we take the Laplace transform of the tight Weil–Petersson volumes and prove Theorem 3. We also look at the relation between the disk function of the regular hyperbolic surfaces and the generating series of moment η.

Finally, in Sec. V we briefly discuss how our results may be of use in the study of Jackiw–Teitelboim (JT) gravity.

Recall that a boundary i of XMg,n+p(L) is said to be tight in Sg,n,p if i is the only simple cycle γ in [i]Sg,n,p of length γ(X) ≤ Li and we defined the moduli space of tight hyperbolic surfaces as
(44)
We noted before that when g = 0 and n = 2 we have M0,2,ptight(L)= because 1 and 2 belong to the same free homotopy class of S0,2,p and can therefore never both be the unique shortest cycle. Instead, it is useful for any p ≥ 1 to consider the moduli space of half-tight cylinders
(45)
which is non-empty whenever L1 > L2 > 0. We will also consider
(46)
and denote its Weil–Petersson volume by Hp(L).

Lemma 6.

When L1L2 > 0, Hp(L) is an open subset of M0,2+p(L), and when it is non-empty (L1 > L2) its closure is Hp̄(L). In particular, both have the same finite Weil–Petersson volume Hp(L) ≤ V0,2+p(L) when L1 > L2, but Hp(L) has 0 volume and Hp̄(L) non-zero volume Hp(L) when L1 = L2.

Proof.

For L1 > L2, Hp(L) is the intersection of the open sets {γ(X) > L2} indexed by the countable set of free homotopy classes γ in [2]S0,2,p. It is not hard to see that in a neighborhood of any XHp(L) only finitely many of these are important, so the intersection is open. Its closure is given by the countable intersection of closed sets {γ(X) ≥ L2}, which is precisely Hp̄(L).

In the case L1 = L2, Hp(L) is empty, because 1 is a curve in [2]S0,2,p of the same length as 2. On the other hand, Hp̄(L) is non-empty and of full dimension. To see this, observe, for instance, that there exists an ϵ > 0, depending only on L1, such that Hp̄(L) includes all surfaces in M0,2+p(L) that contain a pair of pants bounded by 1, 2 and a simple closed geodesic of length less than ϵ. In such a surface, each simple closed geodesics other than 1 and 2 must exit the pair of pants and can thus be ensured to be of length at least L1. Hence, Hp̄(L) has positive volume Hp(L) > 0.□

We are now ready to state the main result of this section.

Proposition 7.
The Weil–Petersson volumes Tg,n,p(L) and Hp(L) satisfy
where in the first equation it is understood that Ki = Li whenever Ii = ∅.

The remainder of this section will be devoted to proving this result. But let us first see how it implies Proposition 1.

Proof of Proposition 1.
Clearly H1(L) = V0,3(L) = 1 for L1L2 and Tg,n,0(L) = Vg,n(L). Rewriting the equations as
it is clear that they are uniquely determined recursively in terms of Vg,n(L). Moreover, by induction we easily verify that Hp(L) in the region L1L2 is a polynomial in L12,,L2+p2 of degree p − 1 that is symmetric in L3, …, L2+p, and Tg,n,p is a polynomial in L12,,Ln+p2 of degree 3g − 3 + n + p that is symmetric in L1, …, Ln and symmetric in Ln+1, …, Ln+p.□

1. Shortest cycles

The following parallels the construction of shortest cycles in maps described in Ref. 14, Sec. 6.1.

Lemma 8.

Given a hyperbolic surface XMg,n+p for g ≥ 1 or n ≥ 2, then for each i = 1, …, n there exists a unique innermost shortest cycle σSg,n,pi(X) on X, meaning that it has minimal length in [i]Sg,n,p and such that all other cycles of minimal length (if they exist) are contained in the region of X delimited by i and σSg,n,pi(X). Moreover, if g ≥ 1 or n ≥ 3, the curves σSg,n,p1(X),,σSg,n,pn(X) are disjoint.

Proof.

First note that if a shortest cycle exists, it is a simple closed geodesic. As a consequence of Ref. 20, Theorem 1.6.11, there are only finitely many closed geodesics with length Li in [i]Sg,n,p. Since i[i]Sg,n,p has length Li, this proves the existence of at least one cycle in [i]Sg,n,p with minimal length.

Regarding the existence and uniqueness of a well-defined innermost shortest cycle, suppose α,β[i]Sg,n,p are two distinct simple closed geodesics with minimal length (see left side of Fig. 3). Since α[i]Sg,n,p, cutting along α separates the surface in two disjoint parts. Therefore, α and β can only have an even number of intersections. If the number of intersections is greater than zero, we can choose two distinct intersections and combine α and β to get two distinct cycles γ1 and γ2 by switching between α and β at the chosen intersections, such that γ1 and γ2 are still in [i]Sg,n,p. Since the total length is still 2, at least one of the new cycles has length . This cycle is not geodesic, so there will be a closed cycle in [i]Sg,n,p with length <, which contradicts that α and β have minimal length. We conclude that α and β are disjoint. Since all cycles in [i]Sg,n,p with minimal length are disjoint and separating, the notion of being innermost is well-defined.

Consider αi=σSg,n,pi(X) and αj=σSg,n,pj(X) for ij (see right side of Fig. 3). Just as before, since αi is separating and αi and αj are simple, the number of intersections is even. If αi and αj are not disjoint, we can choose two distinct intersections and construct two distinct cycles γi and γj by switching between αi and αj at the chosen intersections, such that γi and γj are in [i]Sg,n,p and [j]Sg,n,p respectively. Since the total length of the cycles stays the same, there is at least one a ∈ {i, j} such that γa has length less or equal than αa. Since γa is not geodesic, there is a closed cycle in [a]Sg,n,p with length strictly smaller than αa, which is a contradiction, so the innermost shortest cycles are disjoint.□

In particular the proof implies the following criterions are equivalent:

  • A simple closed geodesic α[i]Sg,n,p is the innermost shortest cycle σSg,n,pi(X);

  • For a simple closed geodesic α[i]Sg,n,p we have (α) ≤ Li and each simple closed geodesic βσSg,n,pi(X) that is disjoint from α has length (β) ≥ (α) with equality only being allowed if β is contained in the region between α and i.

FIG. 3.

Illustrations of why two intersecting geodesics α and β cannot both have minimal length in their free-homotopy classes on Sg,n,p.

FIG. 3.

Illustrations of why two intersecting geodesics α and β cannot both have minimal length in their free-homotopy classes on Sg,n,p.

Close modal

2. Integration on Moduli space

Let us recap Mirzakhani’s decomposition of moduli space integrals in the presence of distinguished cycles (Ref. 1, Sec. 8). A multicurve Γ = (γ1, …, γk) is a collection of disjoint simple closed curves Γ = (γ1, …, γk) in Sg,n which are pairwise non-freely-homotopic. Given a multicurve, in which each curve γi may or may not be freely homotopic to a boundary j of Sg,n, one can consider the stabilizer subgroup
Note that if γi[j]Sg,n is freely homotopic to one of the boundaries j then h · γi = γi for any h ∈ Modg,n. The moduli space of hyperbolic surfaces with distinguished (free homotopy classes of) curves is the quotient
For a closed curve γ in Sg,n and XMg,n, let γ(X) be the length of the geodesic representative in the free homotopy class of γ. For K=(K1,,Kk)R>0k we can restrict the lengths of the geodesic representatives of curves in Γ by setting
If γi[j]Sg,n then this set is empty unless Ki = Lj. Denote by πΓ:Mg,n(L)ΓMg,n(L) the projection. If there are exactly p cycles among Γ that are not freely homotopic to a boundary, then this space admits a natural action of the p-dimensional torus (S1)p obtained by twisting along each of these p cycles proportional to their length. The quotient space is denoted
and is naturally equipped with a symplectic structure inherited from the Weil–Petersson symplectic structure on Mg,n(L)Γ. If we denote by Sg,n(Γ) the possibly disconnected surface obtained from Sg,n by cutting along all γi that are not freely homotopic to a boundary and by M(Sg,n(Γ),L,K) its moduli space, then according to Ref. 1, Lemma 8.3, the canonical mapping
(47)
is a symplectomorphism. Given an integrable function F:Mg,n(L)ΓR that is invariant under the action of (S1)p, there exists a naturally associated function F̃:M(Sg,n(Γ),L,K)R such that (essentially Ref. 1, Lemma 8.4)
(48)

3. Shortest multicurves

Suppose g ≥ 1 or n ≥ 3, meaning that we momentarily exclude the cylinder case (g = 0, n = 2). We consider now a special family of multicurves Γ = (γ1, …, γn) on Sg,n+p for n ≥ 1, p ≥ 0. Namely, we require that γi[i]Sg,n,p is freely homotopic to the boundary i in the capped-off surface Sg,n,p for i = 1, …, n. Then there exists a partition I0 ⊔⋯ ⊔ In = {n + 1, …, n + p} such that Sg,n+p(Γ) has n + 1 connected components s0, …, sn, where s0 is of genus g and is adjacent to all curves Γ as well as the boundaries (j)jI0 while for each i = 1, …, n, si is of genus 0 and contains the ith boundary i as well as (j)jIi and is adjacent to γi. Note that Ii = ∅ if and only if γi[i]Sg,n+p. Finally, we observe that mapping class group orbits {Modg,n+p[Γ]Sg,n+p} of these multicurves Γ are in bijection with the set of partitions {I0 ⊔ ⋯ ⊔ In = {n + 1, …, n + p}}.

With the help of Lemma 8 we may introduce the restricted moduli space in which we require γi to be (freely homotopic to) the innermost shortest cycle in [i]Sg,n,p,

Lemma 9.
The natural projection
where the disjoint union is over (representatives of) the mapping class group orbits of multicurves Γ, is a bijection.

Proof.

If X,XTg,n+p(L) are representatives of hyperbolic surfaces in M̂g,n,p(L)Γ and M̂g,n,p(L)Γ respectively, then by definition [γi]=[σSg,n,pi(X)] and [γi]=[σSg,n,pi(X)]. If X and X′ represent the same surface in Mg,n+p(L), they are related by an element h of the mapping class group, X′ = h · X, and therefore also σSg,n,pi(X)=hσSg,n,pi(X) and [γi]=h[γi]. So Γ and Γ′ belong to the same mapping class group orbit and, if Γ and Γ′ are freely homotopic, we must have hStab(Γ). Hence, X and X′ represent the same element in the set on the left-hand side, and we conclude that the projection is injective. It is also surjective since any XTg,n+p(L) is a representative of M̂g,n,p(L)Γ if we take Γ=(σSg,n,p1(X),,σSg,n,pn(X)), which is a valid multicurve due to Lemma 8.□

We can introduce the length-restricted version M̂g,n,p(L)Γ(K)Mg,n+p(L)Γ(K) as before.

Lemma 10.
The subset M̂g,n,p(L)Γ(K)Mg,n+p(L)Γ(K) is invariant under twisting [the torus-action on Mg,n+p(L)Γ(K) described above]. The image of the quotient M̂g,n,p(L)Γ*(K) under the symplectomorphism (47) is precisely
(49)

Proof.

Let XMg,n+p(L)Γ(K) be a hyperbolic surface with distinguished multicurve Γ. The lengths of the geodesics associated to Γ as well as the lengths of the geodesics that are disjoint from those geodesics are invariant under twisting X along Γ. The criterion explained just below Lemma 8 for γi to be the innermost shortest cycle σSg,n,pi(X) is thus also preserved under twisting, showing that the subset M̂g,n,p(L)Γ(K) is invariant.

Let X0Mg,n,|I0|(K,LI0) and XiM0,2+|Ii|(Li,Ki,LIi) for those i = 1, …, n for which Ii ≠ ∅ be the hyperbolic structures on the connected components s0, …, sn of Sg,n+p(Γ) obtained by cutting X along the geodesics associated to Γ. For each i = 1, …, n the criterion for γi to be the innermost shortest cycle σSg,n,pi(X) is equivalent to the following two conditions holding:

  • the ith boundary of X0 is tight in the capped-off surface associated to s0;

  • Ii = ∅ (meaning γi = i) or XiH|Ii|̄(Li,Ki,LIi) [recall the definition in (46)].

Hence, we have XM̂g,n,p(L)Γ(K) precisely when X0Mg,n,|I0|tight(K,LI0) and XiH|Ii|̄(Li,Ki,LIi) when Ii ≠ ∅. This proves the second statement of the lemma.□

It follows that the Weil–Petersson volume of M̂g,n,p(L)Γ*(K) is equal to the product of the volumes of the spaces appearing in (49). Combining with Lemma 9 and the integration formula (48) this shows that
where it is understood that Ki = Li whenever Ii = ∅. This proves the first relation of Proposition 7.

The decomposition we have just described does not work well in the case g = 0 and n = 2, because 1 and 2 are in the same free homotopy class of the capped surface S0,2,p. Instead, we should consider a multicurve Γ = (γ1) consisting of a single curve γ1 on S0,2+p in the free homotopy class [1]S0,2,p=[2]S0,2,p, see Fig. 4. In this case there exists a partition I1I2 = {3, …, p + 2} such that S0,2+p(Γ) has two connected components s1 and s2, with si a genus-0 surface with 2 + |Ii| boundaries corresponding to i, γ1 and (j)jIi. We consider the restricted moduli space
which thus treats the two boundaries 1 and 2 asymmetrically, by requiring that γ1 is the shortest curve farthest from 1. Lemma 9 goes through unchanged: the projection
where the disjoint union is over the mapping class group orbits of Γ = (γ1), is a bijection. Assuming L1L2, we cannot have γ1[1]S0,2+p so I1 ≠ ∅. There are two cases to consider:
  • γ1[2]S0,2+p and therefore I2 = ∅: this means that 2 has minimal length in [2]S0,2,p, so M̂0,2,p(L)Γ=Hp̄(L).

  • I2 ≠ ∅: by reasoning analogous to that of Lemma 10 we have that M̂0,2,p(L)Γ*(K) is symplectomorphic to
    Hence, when L1L2 we have
    This proves the second relation of Proposition 7.
FIG. 4.

A surface with two distinguished boundaries 1 and 2 naturally decomposes into a pair of half-tight cylinders.

FIG. 4.

A surface with two distinguished boundaries 1 and 2 naturally decomposes into a pair of half-tight cylinders.

Close modal
Let us define the following generating functions of the Weil–Petersson volumes, half-tight cylinder volumes and tight Weil–Petersson volumes:
(50)
(51)
(52)
Furthermore, for g ≥ 2 we recall the polynomial
(53)
where τ2d2τ3d3g are the ψ-class intersection numbers on Deligne–Mumford compactification of the moduli space of curves M̄g,n with n = kdk ≤ 3g − 3 marked points. Then according to Ref. 21, Theorem 322 
(54)
(55)
(56)
where the moments Mk[μ] are defined in Eq. (22).
In the genus-0 case we can take successive derivatives to find useful formulas for one, two or three distinguished boundaries of prescribed lengths,
(57)
(58)
(59)
The equations of Proposition 7 turn into the equations
(60)
(61)
Let us focus on the last equation, which should determine H(L1, L2μ] uniquely. The left-hand side depends on μ only through the quantity R[μ], and the dependence on R is analytic,
Hence, the same is true for H(L1, L2μ] and one may easily calculate order by order in R that
This suggests that H(L1, L2μ] depends on L1 and L2 only through the combination L12L22. Let’s prove this.

Lemma 11.
The half-tight cylinder generating function satisfies
(62)
and is therefore given by
(63)

Proof.
By construction H(L1,0;μ]=δ2F0[μ]δμ(L1)δμ(0) and the integral (58) with L2 = 0 evaluates to
(64)
The identity
which can be easily checked by calculating the derivatives, implies that
Hence, by (61) we find that
Since the leading coefficient in R of H(L1, L2μ] satisfies (62), it follows that the same is true for the higher-order coefficients in R.

As a consequence of (62), H(L1,L2;μ]=HL12L22,0;μ and the claimed expression (63) follows from (64).□

Since the work of Mirzakhani15 it is known that the Weil–Petersson volumes Vg,n(L) are expressible in terms of intersection numbers as follows. The compactified moduli space M̄g,n of genus-g curves with n marked points comes naturally equipped with the Chern classes ψ1, …, ψn associated with its n tautological line bundles, as well as the cohomology class κ1 of the Weil–Petersson symplectic structure (up to a factor 2π2). The corresponding intersection numbers are given by the integrals
where d1, …, dn ≥ 0 and n = d1 + ⋯ dn + m + 3 − 3g. For g ≥ 0 we denote the generating function of these intersection numbers by
(65)
We may sum over all genera to arrive at the generating function
(66)
In order to lighten the notation we do not write the dependence on λ explicitly here, which only serves as a formal generating variable. Note that λ is actually redundant for organizing the series, since any monomial appears in at most one of the Gg as can be seen from (65). Then the generating function of Weil–Petersson volumes can be expressed as
(67)
where the times tk[μ] are defined by
(68)
See Ref. 21, Lemma 11 based on Ref. 15, where one should be careful that some conventions differ by some factors of two compared to the current work.

We will show that the (bivariate) generating function F̃g[ν,μ] of tight Weil–Petersson volumes, defined in (52), is also related to the intersection numbers, but with different times.

Proposition 12.
The generating function of the volumes Tg,n is related to the generating function of intersection numbers via
(69)
where the shifted times τk[νμ] are defined by
(70)

We remark that this result highlights the integrable properties of the tight Weil–Petersson volumes. It is well-known that exp G(0; t0, t1, …) is a τ-function of the KdV hierarchy.17 Since F̃g[ν,μ] and Gg are related by a ν-independent shift of the times, expg=0λ2g2233gF̃g[ν,μ] viewed as a formal power series in the times tk = tk[ν] is therefore a τ-function of the KdV hierarchy as well. The associated differential equations will play an important role in the Proof of Theorem 2 in Sec. III E.

This proposition will be proved in the remainder of this subsection, relying on an appropriate substitution of the weight ν. To this end, we informally introduce a linear mapping Hμ on measures on the half line [0, ) as follows. If ρ is a measure on [0, ) we let Hμρ be the measure given by
(71)
The effect of Hμ on the times can be computed using the series expansion (63),
(72)
We observe that Hμ acts as an infinite upper-triangular matrix on the times. This matrix is easily inverted to give
(73)
This means that knowledge of the generating function F̃g[Hμρ,μ] with substituted weight Hμρ is sufficient to recover the original generating function F̃g[ν,μ]. Luckily the former is within close reach.

Lemma 13.
The generating functions for tight Weil–Petersson volumes and regular Weil–Petersson volumes are related by
(74)
where the correction term
(75)
is necessary to subtract the constant, linear and quadratic dependence on ρ in the genus-0 case.

Proof.
If g, n ≥ 0 (such that n ≥ 3 if g = 0) and L1, …, Ln ∈ [0, ) ∪ i(0, π), then Proposition 7 allows us to compute
(76)
where we use the notation L(j)=(Ln+p0++pj1+1,,Ln+p0++pj). In terms of the tight Weil–Petersson volume generating function (16) and the half-tight cylinder generating function (51) this evaluates to
(77)
where it is understood that in the argument of Tg,n(Kμ] we take Ki = Li for iJ.
Expanding Fg[ρ + μ] from its definition (10) we find
(78)
Plugging in (77) and using that T0,n(Kμ] = 0 for n < 3, yields
as claimed.□
Lemma 13 and (67) together lead to the relation
(79)
The right-hand side can be specialized, making use of a variety of identities between intersection numbers. Firstly, a relation between intersection numbers involving κ1 and pure ψ-class intersection numbers16,23 leads to the identity24 
(80)
where the shifts are
(81)
For us this gives
(82)
where we use the notation G(0; x) = G(0; x0, x1, x2, …).
This can be further refined using Witten’s observation,16 proved by Kontsevich,17 that G(0; x) satisfies the string equation
(83)
Following a computation of Itzykson and Zuber,25 it implies the following identity.

Lemma 14.
The solution to the string equation (83) satisfies a formal power series identity in the parameter r,
(84)

Proof.
For x0, x1, … fixed, let us consider the sequence of functions
(85)
such that yp(s)=δp,0yp+1(s). The string equation (83) then implies
(86)
Integrating from s = 0 to s = r gives the claimed identity.□

Before we can use this lemma, we establish a relation between τk[Hμρμ] and tk[ρ + μ].

Lemma 15.
We can rewrite
(87)
where the shifted times τk[νμ] are defined in (70).

Proof.
We first relate the moments Mi[μ] defined in (22) to the times ti[μ]. Note that Z(uμ] defined in (21) can be expressed in the times as
(88)
By taking p derivatives with respect to u, we get
(89)
Just like before in obtaining (73), this can be inverted to
(90)
The right-hand side of (87) can thus be expressed as
(91)
From (73) the last term is just tq[ρ], so we have reproduced the left-hand side of (87), since tq[ρ + μ] = tq[ρ] + tq[μ].□
The last two lemmas allow us to express (82) as
(92)
To finish the Proof of Proposition 12 we thus only need to check that the last two terms cancel.

Lemma 16.
We have
(93)

Proof.
Let us denote the right-hand side by Gcorr. By the definition (70),
(94)
Changing integration variables to r = R[μ] − s/2 gives
(95)
where in the second equality we use the series expansion of Z(r) = Z(rμ] around r = R[μ] [recall from (23) that Mk[μ]=Z(k+1)R[μ];μ], and in the third equality we expanded (2r − 2R[μ])k as a polynomial in r and made use of (73).
In terms of the weight ρ this can be written as
(96)
In the constant, linear and quadratic term in ρ we then recognize exactly the expressions (54), (57), and (58),
(97)

Recall that the new kernel is given by
(98)
where X(z) = X(zμ] is determined by its two-sided Laplace transform X̂(u;μ],
(99)
and
(100)
To prove Theorem 2, we need to relate K(x, tμ] to the moments Mk[μ], since they appear in the shifted times. We define the reverse moments βm[μ] as the coefficients of the reciprocal series
(101)
Multiplying both series shows that the moments and reverse moments obey
(102)
for each p ≥ 0. Note in particular that
(103)

Proposition 17.
For i, j ≥ 1, the new kernel satisfies
(104)
and
(105)

We need two lemmas to prove this proposition. First we examine the one-sided Laplace transforms
(106)
(107)

Lemma 18.
(108)

Proof.
To compute the integral (106), we only need positive values for x, so we assume x > 0. Since K0(x, −t) = K0(x, t) we can also assume t ≥ 0. For x < t we may expand
(109)
while for x > t we may use
(110)
This gives
(111)
When subtracting K̂0(u,t) it should be clear that the sum cancels and we easily obtain the claimed formula.□

Lemma 19.
(112)

Proof.
From the definition (107) we obtain
(113)
The first integral evaluates to X̂(u;μ](K̂0(u,t)K̂0(u,t)). By changing variables (x, z) → (−x, −z) and using the symmetry of dX(z), we observe that the second integral is unchanged when K0(x, t) is replaced by K0(−x, t). Since also K0(x, t) + K0(−x, t) = 2, the second integral can be calculated to give
(114)
Subtracting both integrals gives the desired result.□

Proof of Proposition 17.
We start by noting
(115)
Using Lemmas 19 and 18, we get for i ≥ 1
(116)
The second identity follows easily from the first by performing the integration at constant x + y, since

We will prove the tight topological recursion by retracing Mirzakhani’s proof15 of Witten’s conjecture, which relies on the observation that her recursion formula (1), expressed as an identity on the coefficients of the volume polynomials Vg,n(L), is equivalent to certain differential equations for the generating function G(sx0, x1, …) of intersection numbers (see also Ref. 18). These differential equations can be expressed as the Virasoro constraints16,18,26
(117)
Here the Virasoro operators Ṽ1,Ṽ0,Ṽ1,Ṽ2, are the differential operators acting on the ring of formal power series in x0, x1, x2, … via
(118)
They satisfy the Virasoro relations
(119)
Proposition 12 suggests introducing the shift xkxk+γ̃k in G with γ̃k=δk,121kMk1[μ], which satisfies
(120)
We use the reverse moments βm[μ] of (101) to introduce linear combinations
(121)
of these operators for all p ≥ −1, which therefore obey
(122)
Using (102) the operators Vp can be expressed as
(123)
In particular, after some rearranging (and shifting pp − 1) we observe the identity
(124)
where G is understood to be evaluated at G=G(0;x0,x1+γ̃1,x2+γ2̃,).
Substituting xk = tk[ν] such that xk+γ̃k=τk[ν,μ], Proposition 12 links G to the generating function
(125)
of tight Weil–Petersson volumes. The differential equations (124) can then be reformulated as the functional differential equation (here all partial derivates of G are evaluated at 0, τ0[νμ], τ1[νμ], …)
Inserting the integral identities of Proposition 17 this can also be expressed in terms of the kernel K(x, tμ] as
where in the last line we used (103). This equation at the level of the generating function (125) is precisely equivalent to the recursion equation on its polynomial coefficients
(126)
for (g, n) ∉ {(0, 3), (1, 1)} combined with the initial data
(127)
(128)
This completes the Proof of Theorem 2.
We follow a strategy along the lines of the Proof of Theorem 2. Recall the relation (125) between the intersection number generating function G and the tight Weil–Petersson volumes Tg,n. Let us denote by Gg,n(x0,x1,;M0,M1,) the homogeneous part of degree n in x0, x1, … of Gg(0;x0,x1+1M0,x212M1,x314M2,). In other words, they are homogeneous polynomials of degree n in x0, x1, … with coefficients that are formal power series in M0,M1,, such that
(129)
We will prove that there exist polynomials P̄g,n(x0,x1,;m1,m2) such that
(130)
and deduce a recurrence in n.
For g ≥ 2 and n = 0, the existence of a polynomial P̄g,0(m1,m2,) follows from Ref. 21, Lemma 12, since
(131)
and Gg(0; 0, 0, x2, x3, …) is polynomial by construction. Also
(132)
Let us now assume 2g − 2 + n ≥ 2 and aim to express Gg,n in tems of Gg,n−1. By construction the series Gg,n obeys for k ≥ 1,
(133)
The string equation, i.e., (120) at p = −1, written in terms of Gg,n reads
(134)
which after rearranging gives the relation
(135)
Together with (133) this is sufficient to identify the recursion relation
(136)
By induction, we now verify that Gg,n is of the form (130). If (130) is granted for Gg,n−1, then
(137)
is indeed of the form (130) provided
(138)
According to (125) the series Gg,n and the tight Weil–Petersson volume Tg,n are related via
(139)
This naturally leads to the existence of polynomials Pg,n(L,m1,m2,) such that
(140)
to get
(141)
The claims about the degree of the polynomials Pg,n are easily checked to be valid for the initial conditions and to be preserved by the recursion formula (37). This proves theorem 4.

Proof of Corollary 5.
We note that
(142)
Setting m0 = 1 this gives
(143)
and
(144)
Using Eq. (140), we get the desired result.□

Let us consider the partial derivative operator
(145)
sometimes called the loop insertion operator in the literature, on the ring of formal power series in x0, x1, … and 1/z. For later purposes we record several identities for the power series coefficients around z = , valid for a ≥ 0,
(146)
(147)
(148)
where the reverse moments β[μ] were introduced in (101).
From the definition (33) and the relation (125) we deduce that for g ≥ 1 or n ≥ 3
(149)
where γ̃k=δk,121kMk1[μ] as before. Recall the differential equation (124) satisfied by this (shifted) intersection number generating function G(0;x0,x1+γ̃1,x2+γ̃2,),
With the help of the identities (147) it can be recast in terms of the operator Δ(z) as
Extracting the genus-g contribution, which appears as the coefficient of λ2g−2, the relation (149) allows us to turn this into a recursion for ωg,n,
Finally, if we set ω0,2(z)=(z1z2)2 and ω0,0(z) = ω0,1(z) = 0, this reduces to
(150)

Due to Proposition 7, there is a relation between the regular and tight Weil–Petersson volumes. In this subsection we will look at this relation in the Laplace transformed setting.

In particular, we are interested in the Laplace transformed generating functions of (regular) Weil–Petersson volumes
where we recall that

Lemma 20.
We define xi=xi(zi;μ]=zi22R[μ]. For g ≥ 1 or n ≥ 3 we have
(151)
while for g = 0 and n = 1, 2,
(152)
(153)

Proof.
For the first identity we wish to combine (60) and (33). It requires an expression for the Laplace transform of the half-tight cylinder. Using
(154)
allows us to compute
Therefore,
which by (33) gives the first stated identity.
For the last two identities we use that the Laplace transform of the modified Bessel function I0 is given by
Then (152) follows directly from (57), while for the cylinder case (58) implies

We finish this section by giving alternative expressions for the disk function and the series η(uμ].

Lemma 21.
The disk function W0,1(z) is related to η via
valid when 4|R| < |z1|2.

Note that μ = 0 gives W0,1(z)=0 as expected.

Proof.
The starting point is the standard generating function [Ref. 27, Eq. (10.1.39)]
(155)
for the spherical Bessel functions yk(u) valid when 2|t| < |u|. Restricting to 2|R| < |x|2 and using the series expansion of the ordinary and spherical Bessel functions we find
(156)
Setting λ = 1 now gives
(157)
On the other hand we can show that
Integrating and setting λ = (iL)/(2π) gives
(158)
valid for 2|R| < |x|2.
Starting from (152) and restricting to 2|R| < |x1|2 we find the series expansion
(159)
We can use Ref. 28, 6.567.1 and 6.567.18,
(160)
(161)
This yields
(162)
On the other hand, (23) and (24) imply
(163)
Together with Z(R) = 0 we may now conclude that
(164)
valid for 2|R| < |x1|2. Substituting x1=z122R gives the desired expression.□
For convenience, we record an explicit expression for η(uμ] that follows from this proof. For a formal power series F(r, u) in r with coefficients that are Laurent polynomials in u, we denote by [u≥0]F(r, u) the formal power series obtained by dropping the negative powers of u in the coefficients of F(r, u). Then we can write η(uμ] as
(165)

The Weil–Petersson volumes play an important role in Jackiw–Teitelboim (JT) gravity,29–31 a two-dimensional toy model of quantum gravity. JT gravity has received significant attention in recent years because of the holographic perspective on the double-scaled matrix model it is dual to Ref. 31. In this section we point to some opportunities to use our results in the context of JT gravity and its extensions in which hyperbolic surfaces with defects play a role.32–34 But we start with a brief introduction to the JT gravity partition function in Euclidean signature.

JT gravity is governed by the (Euclidean) action
(166)
where ϕ is the scalar dilaton field, gμν is a two-dimensional Riemannian metric and R the corresponding Ricci scalar curvature, all living on a surface M.
Since we want this action to make sense when the manifold has boundaries of lengths β = (β1, …, βn), the boundary term
(167)
is included, where hμν is the induced metric on the boundary and K is the extrinsic curvature at the boundary. Including the topological Einstein–Hilbert term, proportional to parameter S0, gives the full (Euclidean) JT action
(168)
where χ is the Euler characteristic of the manifold.
The JT gravity partition function can formally be written as
(169)
In the partition function, the dilation field ϕ acts as a Lagrange multiplier on (R + 2), therefore enforcing a constant negative curvature R = −2 in the bulk. This is why the relevant manifolds will be hyperbolic surfaces.
Due to the Einstein–Hilbert term, we can do a topological expansion by a formal power series expansion in eS0,
(170)
It has been shown by Saad, Shenker, and Stanford31 that the JT partition functions Zg,n for 2g + n − 2 > 0 can be further decomposed by splitting the surfaces into n trumpets and a hyperbolic surface of genus g and n geodesic boundaries with lengths b = (b1, …, bn), and that the partition function measure is closely related to the Weil–Petersson measure (Fig. 5). To be precise, it satisfies the identity
(171)
where Vg,n(b) are the Weil–Petersson volumes and the trumpet contributions are given by
(172)
This formula is the link between JT gravity and Weil–Petersson volumes.
FIG. 5.

Left: Visualization of decomposing the manifold into n JT trumpets and a hyperbolic surface with n geodesic boundaries. In this case g = 1 and n = 4. Right: A conical defect on a hyperbolic surface.

FIG. 5.

Left: Visualization of decomposing the manifold into n JT trumpets and a hyperbolic surface with n geodesic boundaries. In this case g = 1 and n = 4. Right: A conical defect on a hyperbolic surface.

Close modal
There are several natural extensions of the JT action. If we only allow up to two derivatives, the most general action can be transformed to35 
(173)
In Subsection V A we will discuss a natural choice of the dilaton potential U(ϕ), which gives rise to defects in the hyperbolic surfaces.
One of the most natural dilaton potentials is
(174)
which adds a gas of conical defects of cone angle 2πα carrying weight μ each. It naturally arises32 from Kaluza–Klein instantons when performing dimensional reduction on three-dimensional black holes.
More generally, one can allow multiple types of defects by considering a measure μ on i[0, 2π) and setting
(175)
For instance, the choice μ=j=1kμjδiαj gives k types of defects with cone angles α1, …, αk ∈ [0, 2π],
(176)
The choice to consider the measure on the imaginary interval will be convenient later.
It can be shown32 that these potentials indeed lead to conical defects. For example, one can look at the term linear in μ in the integrand of the partition function for a single type of gas:
(177)
It follows that the surface has curvature R = −2 everywhere, except at point x1, where we have a conical defect with cone angle 2πα. If one includes all orders of μ, any number of defects may appear and each defect carries a weight μ.32 
As already mentioned in the introduction, the Weil–Petersson volumes for surfaces with sharp cone points (cone angle 2πα < π) are obtained1,2,13,36 from the usual Weil–Petersson volume polynomials by treating the defect angle 2πα as a geodesic boundary with imaginary boundary length 2πiα. This implies that the partition function Zg,n(β) is closely related to the generating function Fg[μ] of Weil–Petersson volumes considered in this paper. To be precise, using (10) and (60),
(178)
which we can compute using the recursions described in this paper. In particular, its topological recursion can in principle be derived from that of Tg,n in Theorem 2.
We can simplify (178) by considering the tight trumpet, which is a genus-0 hyperbolic surface with an asymptotic boundary of length β, a tight boundary of length K and an arbitrary number of extra geodesic boundaries, with the constraint that the tight boundary cannot be separated from the asymptotic one by a curve of length β. See Fig. 6. Since it can be obtained by gluing a trumpet to a half-tight cylinder, with the help of Lemma 11 we find that the partition function associated to a tight trumpet is given by
(179)
Remarkably it differs from the JT trumpet only in a factor exponential in the boundary length β. We conclude that for g ≥ 1 or n ≥ 3,
(180)
which we understand as a gluing of tight trumpets to tight hyperbolic surfaces. In the case g = 0 and n = 2, we only need to glue two tight trumpets together to find the universal two-boundary correlator
(181)
In particular, a statistical interpretation of the two-boundary partition function leads to the observation that the shortest cycle separating the two boundaries has random length with density
(182)
independent of the weight μ. This is the Rayleigh distribution with mean πβ1β2/(β1+β2).
FIG. 6.

Visualization of the tight trumpet.

FIG. 6.

Visualization of the tight trumpet.

Close modal

We note that these expressions do not apply to the case of blunt cone points (cone angle 2πα ∈ [π, 2π]). The problem is that in the presence of such defects it is no longer true that every free homotopy class of closed curves necessarily contains a geodesic, because, informally, when shortening a closed curve it can be pulled across a blunt cone point, while that never happens for a short one. However, this is not an issue when considering tight cycles, because in that setting one is considering larger homotopy classes, namely of the manifold with its defects closed off. Such homotopy classes will always contain a shortest geodesic, which generically is unique. Whereas the JT trumpet cannot always be removed from a surface with blunt defects in a well-defined manner, the removal of a tight trumpet should pose no problem. It is natural to ask whether such reasoning can be used to connect to the recent works37,38 in JT gravity dealing with blunt cone points.

Another well-studied extension of JT gravity, is the introduction of Fateev-Zamolodchikov-Zamolodchikov-Teschner (FZZT) branes. With this extension the hyperbolic surfaces can end on a FZZT brane. In the random matrix model description of JT gravity, this corresponds to fixing some eigenvalues of the random matrix.33 

In the partition function, this leads to the addition of an arbitrary number of geodesic boundaries as defects with a certain weight M(L)=ezL, where L is the length of the boundary,
(183)
Such weights have been interpreted34,39 as the action of a fermion with mass z.
Using our setup, we can rewrite this to:
(184)
with
(185)
or again using the tight trumpet
(186)
with
(187)
The behavior of R[μFZZT] depends on z and S0 and its critical points should give insight into critical phenomena of the partition function, see Ref. 40.

We thank an anonymous referee for useful suggestions. This work is supported by the START-UP 2018 program with Project No. 740.018.017 and the VIDI program with Project No. VI.Vidi.193.048, which are financed by the Dutch Research Council (NWO).

The authors have no conflicts to disclose.

Timothy Budd: Conceptualization (equal); Investigation (equal); Writing – original draft (equal). Bart Zonneveld: Conceptualization (equal); Investigation (equal); Writing – original draft (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
M.
Mirzakhani
, “
Simple geodesics and Weil-Petersson volumes of moduli spaces of bordered Riemann surfaces
,”
Invent. Math.
167
,
179
222
(
2006
).
2.
S. P.
Tan
,
Y. L.
Wong
, and
Y.
Zhang
, “
Generalizations of McShane’s identity to hyperbolic cone-surfaces
,”
J. Differ. Geom.
72
,
73
112
(
2006
).
3.
B.
Eynard
and
N.
Orantin
, “
Weil-petersson volume of moduli spaces, Mirzakhani’s recursion and matrix models
,” arXiv:0705.3600 (
2007
).
4.

Note that, due to the extra factors Li in the integrand, ωg,n(0)(z) is (−1)n times the partial derivative in each of the variables z1, …, zn of the Laplace transforms of Vg,n(L), but we will refer to ωg,n(0)(z) as the Laplace-transformed Weil–Petersson volumes nonetheless.

5.

Our constructions will not rely on an orientation of the boundary cycles, but for definiteness we may take them clockwise (keeping the surface on the left-hand side when following the boundary).

6.
M.
Mirzakhani
, “
Growth of Weil-Petersson volumes and random hyperbolic surface of large genus
,”
J. Differ. Geom.
94
,
267
300
(
2013
).
7.
L.
Guth
,
H.
Parlier
, and
R.
Young
, “
Pants decompositions of random surfaces
,”
Geom. Funct. Anal.
21
,
1069
1090
(
2011
).
8.
M.
Mirzakhani
and
B.
Petri
, “
Lengths of closed geodesics on random surfaces of large genus
,”
Comment. Math. Helvetici
94
,
869
889
(
2019
).
9.
C.
Gilmore
,
E.
Le Masson
,
T.
Sahlsten
, and
J.
Thomas
, “
Short geodesic loops and Lp norms of eigenfunctions on large genus random surfaces
,”
Geom. Funct. Anal.
31
,
62
110
(
2021
).
10.
L.
Monk
, “
Benjamini–Schramm convergence and spectra ofrandom hyperbolic surfaces of high genus
,”
Anal. PDE
15
,
727
752
(
2022
).
11.

We use the physicists’ convention of writing the argument μ in square brackets to signal it is a functional dependence (in the sense of calculus of variations).

12.
T.
Budd
and
P.
Koster
, “
Statistics of critical Boltzmann hyperbolic surfaces
” (unpublished).
13.
N.
Do
and
P.
Norbury
, “
Weil-Petersson volumes and cone surfaces
,”
Geom. Dedicata
141
,
93
107
(
2009
).
14.
J.
Bouttier
,
E.
Guitter
, and
G.
Miermont
, “
Bijective enumeration of planar bipartite maps with three tight boundaries, or how to slice pairs of pants
,”
Ann. Henri Lebesgue
5
,
1035
1110
(
2022
).
15.
M.
Mirzakhani
, “
Weil-Petersson volumes and intersection theory on the moduli space of curves
,”
J. Am. Math. Soc.
20
,
1
23
(
2006
).
16.
E.
Witten
, “
Two-dimensional gravity and intersection theory on moduli space
,” in
Surveys in Differential Geometry (Cambridge, MA, 1990)
(
Lehigh University
,
Bethlehem, PA
,
1991
), pp.
243
310
.
17.
M.
Kontsevich
, “
Intersection theory on the moduli space of curves and the matrix Airy function
,”
Commun. Math. Phys.
147
,
1
23
(
1992
).
18.
M.
Mulase
and
B.
Safnuk
, “
Mirzakhani’s recursion relations, Virasoro constraints and the KdV hierarchy
,”
Indian J. Math.
50
,
189
218
(
2008
).
19.
B.
Eynard
and
N.
Orantin
, “
Invariants of algebraic curves and topological expansion
,”
Commun. Number Theory Phys.
1
,
347
(
2007
); arXiv:math-ph/0702045.
20.
P.
Buser
,
Geometry and Spectra of Compact Riemann Surfaces
(
Birkhäuser
,
Boston
,
1992
).
21.
T.
Budd
, “
Irreducible metric maps and Weil-Petersson volumes
,”
Commun. Math. Phys.
394
,
887
917
(
2022
).
22.

Note that there has been a shift in conventions, e.g., regarding factors of 2.

23.
C.
Faber
, “
A conjectural description of the tautological ring of the moduli space of curves
,” in
Moduli of Curves and Abelian Varieties
,
Aspects of Mathematics Vol. E33
(
Friedrich Vieweg
,
Braunschweig
,
1999
), pp.
109
129
.
24.
R.
Kaufmann
,
Y.
Manin
, and
D.
Zagier
, “
Higher Weil-Petersson volumes of moduli spaces of stable n-pointed curves
,”
Commun. Math. Phys.
181
,
763
787
(
1996
).
25.
C.
Itzykson
and
J.-B.
Zuber
, “
Combinatorics of the modular group. II. The Kontsevich integrals
,”
Int. J. Mod. Phys. A
07
,
5661
5705
(
1992
).
26.
R.
Dijkgraaf
,
H.
Verlinde
, and
E.
Verlinde
, “
Loop equations and Virasoro constraints in non-perturbative two-dimensional quantum gravity
,”
Nucl. Phys. B
348
,
435
456
(
1991
).
27.
M.
Abramowitz
and
I. A.
Stegun
,
Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables
(
Courier Corporation
,
1964
).
28.
I. S.
Gradshteyn
and
I. M.
Ryzhik
, in
Table of Integrals, Series, and Products
, 8th ed. (
Elsevier; Academic Press
,
Amsterdam
,
2015
), p.
xlvi+1133
.
29.
R.
Jackiw
, “
Lower dimensional gravity
,”
Nucl. Phys. B
252
,
343
356
(
1985
).
30.
C.
Teitelboim
, “
Gravitation and Hamiltonian structure in two spacetime dimensions
,”
Phys. Lett. B
126
,
41
45
(
1983
).
31.
P.
Saad
,
S. H.
Shenker
, and
D.
Stanford
, “
JT gravity as a matrix integral
,” arXiv:hep-th/1903.11115 (
2019
).
32.
H.
Maxfield
and
G. J.
Turiaci
, “
The path integral of 3D gravity near extremality; or, JT gravity with defects as a matrix integral
,”
J. High Energy Phys.
2021
,
118
.
33.
A.
Blommaert
,
T. G.
Mertens
, and
H.
Verschelde
, “
Eigenbranes in Jackiw-Teitelboim gravity
,”
J. High Energy Phys.
2021
,
168
.
34.
K.
Okuyama
and
K.
Sakai
, “
FZZT branes in JT gravity and topological gravity
,”
J. High Energy Phys.
2021
,
191
.
35.
E.
Witten
, “
Matrix models and deformations of JT gravity
,”
Proc. R. Soc. A
476
,
20200582
(
2020
).
36.
N.
Do
, “
Moduli spaces of hyperbolic surfaces and their weil-petersson volumes
,” arXiv:1103.4674 (
2011
).
37.
G. J.
Turiaci
,
M.
Usatyuk
, and
W. W.
Weng
, “
2D dilaton-gravity, deformations of the minimal string, and matrix models
,”
Classical Quantum Gravity
38
,
204001
(
2021
).
38.
L.
Eberhardt
and
G. J.
Turiaci
, “
2D dilaton gravity and the Weil–Petersson volumes with conical defects
,”
Commun. Math. Phys.
405
,
103
(
2024
).
39.

Please note that z in our work corresponds to z/(2π) in Ref. 34.

40.
A.
Castro
, “
Critical JT gravity
,”
J. High Energy Phys.
2023
,
36
.