In this previous post here (from 2018), I described some “back of the envelope calculations” (based on private conversations with Scott Wolpert) indicating that some sectional curvatures of the Weil–Petersson (WP) metric could be at least exponentially small in terms of the distance to the boundary divisor of Deligne–Mumford compactification.
Very roughly speaking, this heuristic computation went as follows: the WP sectional curvature of any -plane can be written as the sum of three terms; for the
-planes considered in the previous post, the main term among those three seemed to be a kind of
-norm of Beltrami differentials with essentially disjoint supports; finally, this
-type norm was shown to be really small once a certain Green propagator is ignored.
Last April 2019, I met Scott during an event at Simons Center for Geometry and Physics, and I took the opportunity to tell him that one could perhaps show that the measure of the set of -planes leading to tiny WP curvatures is very small using the real-analyticity of the WP metric.
More concretely, my idea was very simple: since the Grassmannian of
-planes tangent to a point
is a compact space, the WP sectional curvature defines a real-analytic function
, and we dispose of good upper bounds for
and all of its derivatives in terms of the distance of
to the boundary (see this article here), we can hope to get reasonable estimates for the measure of the sets
using the techniques of these articles here and here (which are close in spirit to the classical fact [explained in Lemma 3.2 of Kleinbock–Margulis paper, for instance] that the measure of the sets
are small whenever
is a polynomial function on
whose degree and
-norm are bounded).
As it turns out, Scott thought that this strategy made some sense and, in particular, he promised to use my suggestion as a motivation to review his arguments concerning WP sectional curvatures.
After several email exchanges with Howard Masur and I, Scott announced that there were some mistakes in the construction of tiny WP sectional curvature: in a nutshell, one should not restrict the analysis to a single “main term” in the formula for WP sectional curvatures as a sum of three expressions, and one can not ignore the effect of the Green propagator. More importantly, Scott made a detailed study of these mistakes which ultimately led him to establish polynomial upper bounds for WP sectional curvatures at the heart of his newest preprint available here.
In this post, we will follow closely Scott’s preprint in order to give an outline of the proof of a polynomial upper bound for WP sectional curvatures:
Theorem 1 (Wolpert) Given two integers
and
with
, there exists a constant
with the following property.If
denotes the product of the lengths of the short geodesics of a hyperbolic surface
of genus
with
cusps whose systole is sufficiently small, then the sectional curvatures of the Weil-Petersson metric at
are at most
Remark 1 As it was pointed out by Scott in his preprint, it is likely that this estimate is not optimal: indeed, one expects that the best exponent should be
rather than
.
In what follows, we’ll assume some familiarity with some basic aspects of the geometry of the Weil–Petersson metric (such as those described in these posts here and here).
1. Weil–Petersson sectional curvatures
Let be a hyperbolic surface of genus
with
. If we write
, where
is the usual hyperbolic plane and
is a group of isometries of
describing the fundamental group of
, then the holomorphic tangent space at
to the moduli space
of Riemann surfaces of genus
with
punctures is naturally identified with the space
of harmonic Beltrami differentials on
(and the cotangent space is related to quadratic differentials).
In this setting, the Weil–Petersson metric is the Riemannian metric induced by the Hermitian inner product
where and
is the hyperbolic area form on
.
Remark 2 Note that
is well-defined: if
and
are Beltrami differentials, then
is a function on
.
The Riemann tensor of the Weil–Petersson metric was computed by Wolpert in 1986:
where and
is an operator related to the Laplace–Beltrami operator
on
.
Remark 3 Our choice of notation here differs from Wolpert’s preprint! Indeed, he denotes the Laplace–Beltrami operator by
and he writes
.
The Riemann tensor gives access to nice formulas for the sectional curvatures thanks to the work of Bochner. More concretely, given and
span a
-plane
in the real tangent space to
at
, let us take Beltrami differentials
and
such that
,
, and
is orthonormal. Then, Bochner showed that the sectional curvature of
is
Hence, by Wolpert’s formula for the Riemann tensor of the WP metric, we see that
2. Spectral theory of
Wolpert’s formula for the Riemann tensor of the WP metric hints that the spectral theory of plays an important role in the study of the WP sectional curvatures.
For this reason, let us review some key properties of (and we refer to Section 3 of Wolpert’s preprint for more details and references). First,
is a positive operator on
whose norm is
: these facts follow by integration by parts. Secondly,
is essentially self-adjoint on
, so that
is self-adjoint on
. Moreover, the maximum principle permits to show that
is also a positive operator on
with unit norm. Finally,
has a positive symmetric integral kernel: indeed,
where the Green propagator is the Poincaré series
associated to an appropriate Legendre function . (Here,
stands for the hyperbolic distance on
.) For later reference, we recall that
has a logarithmic singularity at
and
whenever
is large.
3. Negativity of the WP sectional curvatures
Interestingly enough, as it was first noticed by Wolpert in 1986, the spectral features of described above are sufficient to derive the negativity of WP sectional curvatures from Cauchy-Schwarz inequality. More precisely, since
is self-adjoint, i.e.,
and its integral kernel is a real function, a straightforward computation reveals that the equation (1) for the sectional curvature
of a
-plane
can be rewritten as
If we decompose the function into its real and imaginary parts, say
, then we see that
Since is a positive operator, we conclude that
and, a fortiori,
The non-positivity of the right-hand side of (2) can be established in three steps. First, the positivity of also implies that
Secondly, the fact that has a positive integral kernel
allows to apply the Cauchy–Schwarz inequality to get that
. Therefore,
Finally, the Cauchy–Schwarz inequality also says that
In particular, , so that it follows from (2) that all sectional curvatures
of the WP metric are non-positive, i.e.,
.
Actually, it is not hard to derive that at this stage: indeed,
would force a case of equality in Cauchy-Schwarz inequality and this is not possible in our context because
is orthonormal.
Remark 4 Philosophically speaking, the “analog” to this argument in the realm of Teichmüller dynamics is Forni’s proof of the spectral gap property
for the Lyapunov exponents of the Teichmüller geodesic flow. In fact, after some computations with variational formulas for the so-called Hodge norm, Forni establishes that
by ruling out an equality case in a certain Cauchy-Schwarz estimate.
4. Reduction of Theorem 1 to bounds on ‘s kernel
The discussion in the previous section says that small WP sectional curvatures correspond to almost equalities in certain Cauchy-Schwarz inequalities.
Hence, a natural strategy towards the proof of Theorem 1 consists into showing that an almost equality in (3) is impossible. In this direction, Wolpert establishes the following result:
Theorem 2 (Wolpert) There are two constants
and
with the following property. If we have an almost equality
between the terms
and
in (3), then
and
can not be almost equal:
Of course, Theorem 1 is an immediate consequence of Theorem 2 (in view of (2) and the estimate [implied by (3)]).
Thus, it remains only to prove Theorem 2. For this sake, we need further spectral information on , namely, some lower bounds on its the kernel
. In order to illustrate this point, let us now show Theorem 2 assuming the following statement.
Proposition 3 There exists a constant
such that
whenever
and
do not belong to the cusp region
of
.
Remark 5 We recall that the cusp region
of
is a finite union of portions of
which are isometric to a punctured disk
(equipped with the hyperbolic metric
).
For the sake of exposition, let us first establish Theorem 2 when is compact, i.e.,
, before explaining the extra ingredient needed to treat the general case.
4.1. Proof of Theorem 2 modulo Proposition 3 when
Suppose that for a constant
to be chosen later. In this regime, our goal is to show that
is “big” and
is “small”, so that
is necessarily “big”.
We start by quickly showing that is “big”. Since
and
are unitary tangent vectors, it follows from Proposition 3 that
Let us now focus on proving that is “small”. Since
(cf. (3)), if we write
(where
and
are the positive and negative parts of the real part
of
, then we obtain that
Since is positive, we derive that
. Thus, if
is compact, i.e.,
, then Proposition 3 says that
for all
. It follows that
By orthogonality of , we have that
, i.e.,
. By plugging this information into the previous inequality, we obtain the estimate
Next, we observe that (cf. (3)) in order to obtain that
On the other hand, Proposition 3 ensures that for
. Since
and
are unitary tangent vectors, one has
for
. By inserting this inequality into the previous estimate, we derive that
whenever has a sufficiently small systole.
This bound on
can be converted into a
bound thanks to Cauchy integral formula. More concrentely, as it is explained in Section 2 of Wolpert’s preprint, after observing that
and replacing Beltrami differentials
and
by the dual objects
and
(namely, quadratic differentials), we are led to study quartic differentials
. By Cauchy integral formula on
, one has
On the other hand, if has systole
and the cusp region
is empty, then the injectivity radius at any
is
. Thus, there exists an universal constant
such that
for all . By plugging this inequality into (7), we conclude that
for all .
Since is a positive operator on
with unit norm (cf. Section 2 above) and
, we have that the previous inequality implies the following
bound on
:
for all . By combining this estimate with (7), we conclude that
In summary, (4) and (8) imply that
for the choice of constant . This proves Theorem 2 in the absence of cusp regions.
4.2. Proof of Theorem 2 modulo Proposition 3 when
The arguments above for the case also work in the case
because the cusp regions carry only a tiny fraction of the mass of the relevant functions, Beltrami differentials, etc.
More precisely, as it is explained in Section 2 of Wolpert’s preprint, if the constant is chosen correctly, then the Cauchy integral formula and the Schwarz lemma can be used to prove that
for all holomorphic quartic differentials .
In particular, we do not lose too much information after truncating ,
, etc. to
and this allows us to repeat the arguments of the case
to the corresponding truncated objects
,
, etc. without any extra difficulty: see Section 5 of Wolpert’s preprint for more details.
5. Proof of Proposition 3
Closing this post, let us give an idea of the proof of Proposition 3 (and we refer the reader to Section 4 of Wolpert’s preprint for more details).
Since and
(cf. Section 2 above), our task is reduced to give lower bounds on the Poincaré series
For this sake, let us first recall that a hyperbolic surface has thick-thin decomposition: the thick portion is the region where the injectivity radius is bounded away from zero by a uniform constant and the thin portion is the complement of the thick region. Geometrically, the thin region is the disjoint union of the cusp region
and a finite number of collars around simple closed short geodesics: roughly speaking, a collar consisting of the points at distance
of a short simple closed geodesic
of length
.
We can provide lower bounds on in terms of the behaviours of simple geodesic arcs connecting
and
on
.
More concretely, let be the shortest geodesic connecting
and
. Since
is simple, we have that, for certain adequate choices of the constants defining the collars, one has that
can not “back track” after entering a collar, i.e., it must connect the boundaries (rather than going out via the same boundary component). Furthermore,
can not go very high into a cusp. Thus, if we decompose
according to its visits to the thick region, the collars and the cusps, then the fact that
permits to check that it suffices to study the passages of
through collars in order to get a lower bound on
.
Next, if is a subarc of
crossing a collar around a short closed geodesic
, then we can apply Dehn twists to
to get a family of simple arcs indexed by
giving a “contribution” to
of
for some constant depending only on the topology of
. In this way, the desired result follows by putting all “contributions” together.
Leave a Reply