This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Limits of Cubic Differentials and Buildings

John Loftin, Andrea Tamburelli and Michael Wolf
Abstract.

In the Labourie-Loftin parametrization of the Hitchin component of surface group representations into SL(3,)\mathrm{SL}(3,\mathbb{R}), we prove an asymptotic formula for holonomy along rays in terms of local invariants of the holomorphic differential defining that ray. Globally, we show that the corresponding family of equivariant harmonic maps to a symmetric space converge to a harmonic map into the asymptotic cone of that space. The geometry of the image may also be described by that differential: it is weakly convex and a (one-third) translation surface. We define a compactification of the Hitchin component in this setting for triangle groups that respects the parametrization by Hitchin differentials.

1. Introduction

In a pioneering work in the 1980’s and early 1990’s, Hitchin and others developed a beautiful non-Abelian Hodge theory for character varieties of surface groups in higher rank Lie groups. The subject has been intensely studied ever since and while new perspectives, for example more synthetic/algebro-geometric [FG06] or dynamical [Lab06, Gui08], have emerged, there remain some basic questions about how to properly geometrically interpret Hitchin’s original parametrizaton of a principal component of the character variety in terms of the holomorphic data he used. (Indeed, in [Hit92], Hitchin remarks, “Unfortunately, the analytical point of view used for the proofs gives no indication of the geometrical significance of the Teichmüller component.”) The goal of this paper is to relate the holonomy of a representation in a Hitchin component to the local synthetic geometry of the holomorphic differential that Hitchin associates to it, at least in an asymptotic sense.

More precisely, we focus on the Hitchin component Hit3\mathrm{Hit}_{3} of surface group representations into SL(3,)\mathrm{SL}(3,\mathbb{R}). Hitchin ([Hit92]) parametrizes this component in terms of pairs (q2,q3)(q_{2},q_{3}) of quadratic and cubic differentials q2H0(X0,KX02),q3H0(X0,KX03)q_{2}\in H^{0}(X_{0},K_{X_{0}}^{2}),q_{3}\in H^{0}(X_{0},K_{X_{0}}^{3}) on a fixed Riemann surface X0X_{0}. A parametrization, invariant under the action of the mapping class group, was given independently by Labourie ([Lab07]) and Loftin ([Lof01]): from their perspective, Hit3\mathrm{Hit}_{3} may be seen as the cubic differential bundle 𝒞\mathcal{C} over the Teichmüller space 𝒯(S)\mathcal{T}(S).

We study families defined by rays in this parametrization, and in particular the asymptotics. Of course, a ray is defined as multiples sq0sq_{0}, for q0q_{0} a fixed cubic differential on a fixed Riemann surface Σ=(S,J)\Sigma=(S,J) and s>0s>0, and so has holomorphic invariants that are projectively fixed; on the other hand, via the Labourie-Loftin parametrization, the ray defines a family of holonomies hol(s)=Hol(ρs)\mathrm{hol}(s)=\mathrm{Hol}(\rho_{s}) of representations ρs\rho_{s}. We relate the asymptotics of the holonomies Hols\mathrm{Hol}_{s} to the holomorphic invariants of the form q0q_{0}: we give a formula for the leading term of the holonomy Hols([γ])\mathrm{Hol}_{s}([\gamma]) of a curve class [γ][\gamma] in terms of the intersection number of [γ][\gamma] with the form q0q_{0}. The class [γ][\gamma] may be represented by a geodesic cγc_{\gamma} in the metric |q0|23|q_{0}|^{\frac{2}{3}}: this metric is flat away from the zeroes of q0q_{0}, with cone points of total angle 2π(1+13)degp(q0)2\pi(1+\frac{1}{3})\deg_{p}(q_{0}) at a zero pp of q0q_{0}. The segments, known as saddle connections, between the zeroes are denoted c1,.,clc_{1},....,c_{l}.111Saddle connections in this paper are just Euclidean geodesic segments, not restricted to be horizontal in some way. Of course, in such a zero-free region, the cubic differential q0q_{0} has three well-defined cube roots ϕ1,ϕ2,ϕ3\phi_{1},\phi_{2},\phi_{3} and we may compute intersection numbers 223eδϕj-2^{\frac{2}{3}}\mathcal{R}e\int_{\delta}\phi_{j} of curves δ\delta against these roots. Let νi\nu^{i} denote the largest of the real parts of the three periods; this is equivalent to the logarithm of the largest eigenvalue of the holonomy of a natural development of that saddle connection into affine space. Then our main results on asymptotic holonomy may be summarized in the theorem below: we comment later on what is elided in the statement as well as some of the subtleties in its statement and hence proof.

Theorem A.

For every curve class [γ][\gamma], we have

lims+logHols([γ])s13=i=1νi\lim_{s\to+\infty}\frac{\log\|\mathrm{Hol}_{s}([\gamma])\|}{s^{\frac{1}{3}}}=\sum_{i=1}\nu^{i} (1.1)

where \|\cdot\| is any submultiplicative matrix norm.

In particular, if σj(Hols(cγ))\sigma_{j}(\mathrm{Hol}_{s}(c_{\gamma})) denotes the jj-th largest singular value of the flat geodesic cγc_{\gamma} homotopic to c~γ\tilde{c}_{\gamma} with saddle connections c1,,cl{c}_{1},\dots,{c}_{l}, then

lims+log(σj(Hols(cγ)))s13=i=1llims+log(σj(Hols(ci)))s13\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}(c_{\gamma})))}{s^{\frac{1}{3}}}=\sum_{i=1}^{l}\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}} (1.2)

for j=1,2,3j=1,2,3.

As suggested previously, the main qualitative result is that the leading term of the holonomy, in this regime of a ray, is visible through the local expressions of the Hitchin holomorphic parametrization. That the right-hand-side of (1.1) is expressed as a sum of maxima is consistent with other “tropical” expressions regarding asymptotics, see for example [Foc98].

There are some nuances. First, we observe the role of the zeros of the holomorphic form q0q_{0}. They have no explicit presence in either formula (1.1) or (1.2).

Yet, there is a subtlety in that we add the dominant eigenvalue for each saddle connection, so these must get aligned as the geodesic cγc_{\gamma} transitions from a saddle connection coming into a zero. Here, two considerations collide: first we must understand the general form of the unipotents that arise as a path crosses the Stokes lines that emanate from a zero [DW15]. Second, we must show that the matrix permutations defined by these unipotents match up the dominant eigenvalues of saddle connections: this occurs because of the geometry of the unipotents that occur in the limits. The resulting agreement of directions between incoming and outgoing saddle connections is a linchpin of the current work. It is somewhat remarkable that it holds not only generically but also in the special cases where the saddle connections are in the directions of Stokes lines or walls of Weyl chambers.

We also note that, in contrast to some treatments (e.g. [MSWW16], [OSWW20]), we do not restrict to simple zeroes. In general, this present formula might be seen as an extension of the work of the first author in [Lof07]. (See also the extension of that work by Collier-Li [CL17] on more general cyclic Higgs bundles.)

Finally, we relate these considerations to the harmonic map hs:Σ~Xh_{s}\!:\tilde{\Sigma}\to X defined via the solution to Hitchin’s equations. Now, equation (1.2) gives asymptotics of the singular values, and we recall that the singular values of an element MSL(3,)M\in\mathrm{SL}(3,\mathbb{R}) define the distance in the symmetric space X=SL(3,)/SO(3)X=\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3) between the origin and [M][M], with asymptotics then defining distance in the asymptotic cone. We then study the geometry of the limiting harmonic map to the asymptotic cone, obtained by taking an ω\omega-limit of XX, rescaling by the growth of the co-diameter of the image of hsh_{s}.

Theorem B.

Upon rescaling by s13s^{-\frac{1}{3}}, the family of harmonic maps hs:Σ~Xh_{s}\!:\tilde{\Sigma}\to X converges to a Lipschitz harmonic map h:Σ~Coneω(X).h_{\infty}\!:\tilde{\Sigma}\to\mathrm{Cone}_{\omega}(X). The corresponding rescaled family of holonomy representations ρs:π1ΣSL(3,)\rho_{s}\!:\pi_{1}\Sigma\to\mathrm{SL}(3,\mathbb{R}) converges to a representation ρ\rho_{\infty} to the isometry group of Coneω(X)\mathrm{Cone}_{\omega}(X). The harmonic map hh_{\infty} is equivariant with respect to ρ\rho_{\infty}. The cubic differential q0q_{0} induces a 13\frac{1}{3}-translation surface structure on the image h(Σ~)Coneωh_{\infty}(\tilde{\Sigma})\subset\mathrm{Cone}_{\omega}, which is compatible with the local geometry of Coneω(X)\mathrm{Cone}_{\omega}(X).

More precise statements and proofs are contained in Proposition 8.6 and Theorems 8.7 and 8.9 below.

The structure on h(Σ~)h_{\infty}(\tilde{\Sigma}) involves natural local models uk:B(X,d)u_{k}:B\to(X,d), where BB is a ball and uku_{k} is a conformal harmonic map to the asymptotic cone Coneω(X)\mathrm{Cone}_{\omega}(X) defined in terms of the cubic differential ψk=zkdz3\psi_{k}=z^{k}dz^{3} and its three cube roots ϕ1,ϕ2,ϕ3\phi_{1},\phi_{2},\phi_{3}. In particular, the image of the punctured ball B{0}B\setminus\{0\} comprises 2(k+3)2(k+3) flat sectors WiW_{i} which meet only consecutively on geodesics and to which uku_{k} is defined by

uk(x)=(223e(0xϕ1),223e(0xϕ2),223e(0xϕ3))u_{k}(x)=\left(-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{0}^{x}\phi_{1}\right),-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{0}^{x}\phi_{2}\right),-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{0}^{x}\phi_{3}\right)\right)

for xx in a sector of BB. We prove in Theorems 8.7 and 8.9 that the harmonic map hh_{\infty} has this local structure, where ψk\psi_{k} is a restriction of q0q_{0} to BB.

Now, group actions on buildings have arisen in the work of several authors ([Par12], [BIPP21]) with harmonic maps to these buildings having some prominence ([BIPP21], [KNPS15], [MOT21]). Here one might compare the lower rank constructions of surface group actions on real trees in the context of SL(2,)\mathrm{SL}(2,\mathbb{R}) character varieties: see [Bes88], [Wol89], [Wol95]. In the present context, it is worth focusing on the papers of Katzarkov-Noll-Pandit-Simpson ([KNPS15], [KNPS17]), in which the authors outline an approach to compactifying character varieties of surface group representations in SL(d,)\mathrm{SL}(d,\mathbb{C}). Of course, the present paper can be seen as demonstrating a part of that program in a real setting.

Moreover, though, in a companion paper ([LTW22]), we prove a uniqueness theorem for conformal equivariant harmonic maps to buildings which applies in our situation. Thus we find that in settings in which the harmonic maps have an image in the asymptotic cone of SL(3,)/SO(3)\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3), those harmonic maps coincide with the maps described in this paper as endpoints of rays. In particular, those maps would be definable in terms of cubic differentials projectively approximated by the Hitchin differentials for the approximating representations. The results of Theorems 8.7 and 8.9, and the uniqueness results in [LTW22], then in some sense unify some of the various approaches to asymptotic holonomy of representations and the limiting buildings.

In the concluding section of this paper, we provide an example of a full compactification in a specific example, that of the SL(3,)\mathrm{SL}(3,\mathbb{R}) Hitchin component of most (p,q,r)(p,q,r)-triangle groups.

Two features of our technique limit the scope of these results. First, we rely heavily on the cyclic nature of the representations, and the resulting substantial symmetries in the Hitchin system. Indeed, in the present work, that system is but a single scalar equation. That is somewhat less of a limitation than it may seem, as some analogous results are available in the case of Sp(4,)\mathrm{Sp}(4,\mathbb{R}) ([OT20], [TW19]). Second, here we fix the conformal structure of the domain: considering asymptotics where both the domain Riemann surface and the representation degenerate seems to require an analysis finer than what we present here. Finally, we are greatly aided by the asymptotic cone being two-dimensional, and hence the same dimension as the domain Σ\Sigma. This restricts the flexibility of the limiting harmonic maps.

Organization.

In the second section, we define our notation and present some background material. Section 3 is devoted to presenting some required analysis of the Hitchin partial differential equation which governs the harmonic maps. In Section 4, we analyze the holonomy near a zero of q0q_{0}: we are interested in the holonomy along a generic saddle connection and along a portion of an arc that links the zero that crosses a Stokes line. Section 5 assembles these partial holonomies of individual saddle connections into a preliminary description of the asymptotic holonomy. Then, in Section 6. we discuss the phenomenon that the unipotents that describe the transition between incoming and outgoing saddle connections intertwine the dominant eigenvalues. In Section 7, we collect all of the ingredients from the previous sections and prove the main result on asymptotic holonomy displayed above. Section 8 pivots to describe how the endpoint of a ray is a harmonic map to a building, displaying the local structure and induced metric to q0q_{0}, and showing that the image is “weakly convex” in the sense of Parreau [Par21]. Finally, Section 9 displays a corollary of our work in a very special case: the compactification of Hit3\mathrm{Hit}_{3} in the case of a triangle group, where we can provide a somewhat complete account of the compactification using our methods.

Acknowledgements.

The authors acknowledge support from U.S. National Science Foundation grants DMS 1107452, 1107263, 1107367 RNMS: GEometric structures And Representation varieties (the GEAR Network). In addition, some of this work was supported by NSF grant DMS-1440140 administered by the Mathematical Sciences Research Institute while the authors were in residence during the period August 12-December 13, 2019 for the program Holomorphic Differentials in Mathematics and Physics. The second author acknowledges support from the U.S. National Science Foundation under grant NSF DMS-2005501. The research of the third author was supported by the National Science Foundation with grant DMS-2005551 and the Simons Foundation. The authors appreciate useful conversations with David Dumas, Michael Kapovich, Anne Parreau, Marc Burger, Alessandra Iozzi and Beatrice Pozzetti.

2. Background material

The geometry relating cubic differentials and representations in Hit3\mathrm{Hit}_{3} is that of a special sort of surface, a hyperbolic affine sphere HH, in 3\mathbb{R}^{3}. There are natural affine invariants on a hyperbolic affine sphere, a Riemannian metric called the Blaschke metric gg, and a cubic differential qq which is holomorphic with respect to the conformal structure induced by the Blaschke metric, related by the compatibility relation

κ(g)=1+2qg2,\kappa(g)=-1+2\|q\|^{2}_{g}, (2.1)

where κ(g)\kappa(g) is the Gauss curvature, and qg\|q\|_{g} is the pointwise norm of qq with respect to the metric gg. Let f:𝒟3f\!:\mathcal{D}\to\mathbb{R}^{3} be a parametrization of HH conformal with respect to gg, from a domain 𝒟\mathcal{D}\subset\mathbb{C}. Then ff is transverse to the tangent plane TfHT_{f}H and we have the following structure equations for the frame F=(ffzfz¯)F=(f\,\,f_{z}\,\,f_{\bar{z}}) in 3\mathbb{R}^{3}, for zz a conformal coordinate and g=eϕ|dz|2g=e^{\phi}|dz|^{2}

F1dF=(0012eϕ1zϕ00qeϕ0)dz+(012eϕ000q¯eϕ10z¯ϕ)dz¯.F^{-1}dF=\left(\begin{array}[]{ccc}0&0&\frac{1}{2}e^{\phi}\\ 1&\partial_{z}\phi&0\\ 0&qe^{-\phi}&0\end{array}\right)dz+\left(\begin{array}[]{ccc}0&\frac{1}{2}e^{\phi}&0\\ 0&0&\bar{q}e^{-\phi}\\ 1&0&\partial_{\bar{z}}\phi\end{array}\right)d\bar{z}. (2.2)

Equation 2.2 then gives the connection form of the pullback \nabla of the flat connection on 3\mathbb{R}^{3} to the rank-3 bundle 𝟙T𝒟\mathbbm{1}\oplus T\mathcal{D} over 𝒟\mathcal{D}, for 𝟙\mathbbm{1} the trivial line bundle. Equation 2.1 and z¯q=0\partial_{\bar{z}}q=0 ensure \nabla is flat. On a compact Riemann surface, the holonomy of \nabla gives a Hitchin representation into SL(3,)\mathrm{SL}(3,\mathbb{R}).

Alternately, given a holomorphic cubic differential qq over Σ\Sigma, we may use this data to define a stable rank-3 Higgs bundle, which then induces an equivariant conformal harmonic map from Σ~\tilde{\Sigma} to the symmetric space X=SL(3,)/SO(3)X=\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3). We need not avail ourselves of the details of this construction, as the harmonic map can be constructed directly from the hyperbolic affine sphere. To see this, identify XX with the space of all metrics (positive-definite quadratic forms) on 3\mathbb{R}^{3} of determinant 1. The Blaschke lift hh at a given point ff on HH is then the metric on 3\mathbb{R}^{3} for which

  • ff has norm 1,

  • ff is orthogonal to TfHT_{f}H,

  • the metric restricted to TfHT_{f}H is the Blaschke metric gg.

To relate the Blaschke lift hh to the frame FF, it is useful to use an orthonormal frame for hh. For z=x+iyz=x+iy, the frame

F^=(ffx|fx|gfy|fy|g)=F(1000eϕ/2ieϕ/20eϕ/2ieϕ/2)\hat{F}=\left(\begin{array}[]{ccc}f&\frac{f_{x}}{|f_{x}|_{g}}&\frac{f_{y}}{|f_{y}|_{g}}\end{array}\right)=F\left(\begin{array}[]{ccc}1&0&0\\ 0&e^{-\phi/2}&ie^{-\phi/2}\\ 0&e^{-\phi/2}&-ie^{-\phi/2}\end{array}\right)

is orthonormal. Thus

h=(F^)1F^1=(F)1(1000012eϕ012eϕ0)F1.h=(\hat{F}^{\top})^{-1}\hat{F}^{-1}=(F^{\top})^{-1}\left(\begin{array}[]{ccc}1&0&0\\ 0&0&\frac{1}{2}e^{\phi}\\ 0&\frac{1}{2}e^{\phi}&0\end{array}\right)F^{-1}.

Cheng-Yau ([CY86], [CY77]) proved that any hyperbolic affine sphere with a complete Blaschke metric is asymptotic to a properly convex cone in 3\mathbb{R}^{3}, which we take to have vertex at the origin. Likewise, for every such convex cone, there is a unique hyperbolic affine sphere asymptotic to the cone invariant under special linear automorphisms of the cone and with complete Blaschke metric. By projecting 32\mathbb{R}^{3}\to\mathbb{RP}^{2}, we may identify HH with a properly convex domain in 2\mathbb{RP}^{2}. On a compact Riemann surface Σ\Sigma of genus at least 2 equipped with a cubic differential qq, there is a unique solution to Equation 3.1 below (this is the global version of Equation 2.1 above). Thus the universal cover Σ~\tilde{\Sigma} is identified with a convex domain Ω\Omega in 2\mathbb{RP}^{2}, and Σ\Sigma itself admits a quotient of Ω\Omega by projective automorphisms. In other words, we have induced a convex 2\mathbb{RP}^{2} structure on Σ\Sigma. Choi-Goldman ([Gol90],[CG93]) show that convex 2\mathbb{RP}^{2} structures are equivalent to Hitchin representations.

Dumas and the third author [DW15] address the case of polynomial cubic differentials on \mathbb{C}, and show there is a unique complete Blaschke metric on that plane. The convex domain in this case is a convex polygon with d+3d+3 sides, for dd the degree of the polynomial. Indeed polynomial cubic differentials on \mathbb{C} are equivalent to convex polygons, up to appropriate equivalences. In this work, we are primarily interested in cubic differentials of the form zddz3z^{d}\,dz^{3}, which corresponds to the regular convex polygon of d+3d+3 sides. In general, the analysis of the boundary of the polygon follows by comparison to the simpler geometry of inscribed and circumscribed triangles around each vertex and edge.

Example 2.1.

The triangle case (for a constant cubic differential on \mathbb{C} and in which the Blaschke metric is flat) can be worked out explicitly and goes back to Ţiţeica ([Tzi08]). We see that this is a fundamental model for us, and so we describe some of its features. In terms of a natural local coordinate in which the cubic differential q=dw3q=dw^{3}, for w=reiθw=re^{i\theta}, there are “Stokes” directions for rays θ=π6+π3k\theta=\frac{\pi}{6}+\frac{\pi}{3}k for kk an integer. In the sectors bounded by θ=π3+2π3k\theta=\frac{\pi}{3}+\frac{2\pi}{3}k, the rays limit on a vertex of the triangle; the rays parallel to those angles fill out the sides of the polygon.

For a general convex polygon, in each sector in between the Stokes directions, the affine sphere is well approximated by an appropriate Ţiţeica surface. Upon crossing a Stokes ray, the approximating Ţiţeica surface changes by the action of a unipotent transformation determined by the geometry of the convex polygon: the unique unipotent transformation in 3\mathbb{R}^{3} whose projective action on 2\mathbb{RP}^{2} transforms a given inscribed (circumscribed) triangle in the polygon to the subsequent circumscribed (inscribed) triangle by moving a single vertex.

We will also encounter other special directions in \mathbb{C}, which represent the walls of Weyl chambers at θ=π3k\theta=\frac{\pi}{3}k, upon identifying \mathbb{C} with the maximum torus of the Lie algebra of SL(3,)\mathrm{SL}(3,\mathbb{R}).

3. Asymptotics of the Blaschke metric

Consider a ray of cubic differentials qs=sq0q_{s}=sq_{0} on a fixed Riemann surface Σ=(S,J)\Sigma=(S,J) with conformal hyperbolic metric σ\sigma. Let gs=eμsσg_{s}=e^{\mu_{s}}\sigma be the Blaschke metric on the associated affine sphere. The real-valued functions μs\mu_{s} are solutions of Wang equation

Δσμs=2eμs4e2μs|qs|2σ3+2κ(σ).\Delta_{\sigma}\mu_{s}=2e^{\mu_{s}}-4e^{-2\mu_{s}}\frac{|q_{s}|^{2}}{\sigma^{3}}+2\kappa(\sigma). (3.1)

Near each zero of qq, Nie has studied rescaled limits of the associated convex 2\mathbb{RP}^{2} structure, which converge to a regular polygon [Nie22]. Our approach focuses on precise estimates comparing the corresponding affine spheres.

3.1. Asymptotics far from the zeros

We compare the Blashke metrics gsg_{s} with the flat metric with cone singularities |qs|23|q_{s}|^{\frac{2}{3}}. The following results are well-known.

Lemma 3.1 ([Lof04], [DW15]).

The Blaschke metric gsg_{s} satisfies gs>213|qs|23.g_{s}>2^{\frac{1}{3}}|q_{s}|^{\frac{2}{3}}.

Lemma 3.2 ([OT21]).

The area of the Blaschke metric satisfies

213qsArea(S,gs)213qs+2π|χ(S)|2^{\frac{1}{3}}\|q_{s}\|\leq\mathrm{Area}(S,g_{s})\leq 2^{\frac{1}{3}}\|q_{s}\|+2\pi|\chi(S)|\,

where q=S|q|23\|q\|=\int_{S}|q|^{\frac{2}{3}}.

We consider now the quantity

s=μs13log(2|qs|2σ3)\mathcal{F}_{s}=\mu_{s}-\frac{1}{3}\log\left(\frac{2|q_{s}|^{2}}{\sigma^{3}}\right)

in order to compare gsg_{s} and |qs|23|q_{s}|^{\frac{2}{3}} outside the zeros of qsq_{s}. Outside the zeros of q0q_{0}, the function s\mathcal{F}_{s} satisfies the PDE

Δ|q0|23s=243s23(ese2s).\Delta_{|q_{0}|^{\frac{2}{3}}}\mathcal{F}_{s}=2^{\frac{4}{3}}s^{\frac{2}{3}}(e^{\mathcal{F}_{s}}-e^{-2\mathcal{F}_{s}}).

By Lemma 3.1, s>0\mathcal{F}_{s}>0 and Δσs>0\Delta_{\sigma}\mathcal{F}_{s}>0. Hence s\mathcal{F}_{s} is subharmonic.

Lemma 3.3 (Coarse bound on s\mathcal{F}_{s}).

Let pSp\in S and let r0r_{0} be the radius of a ball around pp for the flat metric |q0|23|q_{0}|^{\frac{2}{3}} which does not contain any zeros of q0q_{0}. Then

s(p)log(Area(S,gs)213πs23r02).\mathcal{F}_{s}(p)\leq\log\left(\frac{\mathrm{Area}(S,g_{s})}{2^{\frac{1}{3}}\pi s^{\frac{2}{3}}r_{0}^{2}}\right).
Proof.

The ball BB of radius s13r0s^{\frac{1}{3}}r_{0} for the flat metric qsq_{s} does not contain any zeros. By subharmonicity of s\mathcal{F}_{s} and Jensen’s Inequality, we have

eseBs𝑑AqsBes𝑑Aqs=213Beμs𝑑AσArea(S,gs)213πs23r02.e^{\mathcal{F}_{s}}\leq e^{\fint_{B}\mathcal{F}_{s}dA_{q_{s}}}\leq\fint_{B}e^{\mathcal{F}_{s}}dA_{q_{s}}=2^{-\frac{1}{3}}\fint_{B}e^{\mu_{s}}dA_{\sigma}\leq\frac{\mathrm{Area}(S,g_{s})}{2^{\frac{1}{3}}\pi s^{\frac{2}{3}}r^{2}_{0}}.

Lemma 3.4 (Error decay).

Let pSp\in S be a point at q0q_{0}-distance r0r_{0} from the zeros of q0q_{0}. Then for any δ>1\delta>1, there is a D>0D>0 so that

s(p)Ds16e3223ass13r0/δ,\mathcal{F}_{s}(p)\leq Ds^{\frac{1}{6}}e^{-\sqrt{3}\cdot 2^{\frac{2}{3}}a_{s}s^{\frac{1}{3}}r_{0}/\delta},

with as1a_{s}\to 1 as s+s\to+\infty.

Proof.

Consider the ball centered at pp of radius s13r0/δs^{\frac{1}{3}}r_{0}/\delta for the flat metric qsq_{s}. Since this ball does not contain any zeros of q0q_{0}, we may choose coordinates on the ball so that q0=dz3q_{0}=dz^{3} and pp is at z=0z=0. Then s\mathcal{F}_{s} on this ball satisfies

Δs=243s232es/2sinh(32s).\Delta\mathcal{F}_{s}=2^{\frac{4}{3}}s^{\frac{2}{3}}\cdot 2e^{-\mathcal{F}_{s}/2}\sinh\left(\frac{3}{2}\mathcal{F}_{s}\right).

By Lemma 3.3 and Lemma 3.2, the function es/2e^{-\mathcal{F}_{s}/2} is uniformly bounded below by a constant c>0c>0. Then

Δs3243s23cs.\Delta\mathcal{F}_{s}\geq 3\cdot 2^{\frac{4}{3}}s^{\frac{2}{3}}c\mathcal{F}_{s}.

Similarly, the function s\mathcal{F}_{s} is bounded above by a constant A=A(δ)>0A=A(\delta)>0 on the boundary of the ball. Let η\eta be the solution of the system

{Δη=3243s23cη,η|=A.\left\{\begin{array}[]{l}\Delta\eta=3\cdot 2^{\frac{4}{3}}s^{\frac{2}{3}}c\eta,\\ \eta|_{\partial}=A.\end{array}\right.

We know η(r)=AI0(3223cs13r)I0(3223cs13r0/δ)\eta(r)=A\frac{I_{0}(\sqrt{3}\cdot 2^{\frac{2}{3}}\sqrt{c}\cdot s^{\frac{1}{3}}r)}{I_{0}(\sqrt{3}\cdot 2^{\frac{2}{3}}\sqrt{c}\cdot s^{\frac{1}{3}}r_{0}/\delta)}, where I0(κr)I_{0}(\kappa r) is the only radial solution of

{ΔI0=κ2I0,I0|=1.\left\{\begin{array}[]{l}\Delta I_{0}=\kappa^{2}I_{0},\\ I_{0}|_{\partial}=1.\end{array}\right.

It is well-known ([AS64]) that I0(κx)I_{0}(\kappa x) is the Bessel function of the first kind and has asymptotic behavior I0(κx)eκxxI_{0}(\kappa x)\sim\frac{e^{\kappa x}}{\sqrt{x}} as x+x\to+\infty. By the maximum principle,

s(0)η(0)=AI0(3213cs13r0/δ)Ds16e3223cs13r0/δ.\mathcal{F}_{s}(0)\leq\eta(0)=\frac{A}{I_{0}(\sqrt{3}\cdot 2^{\frac{1}{3}}\sqrt{c}\cdot s^{\frac{1}{3}}r_{0}/\delta)}\leq Ds^{\frac{1}{6}}e^{-\sqrt{3}\cdot 2^{\frac{2}{3}}\sqrt{c}\cdot s^{\frac{1}{3}}r_{0}/\delta}.

This implies that s\mathcal{F}_{s} decays exponentially outside the zeros of q0q_{0}. Now, remember that the constant cc comes from the bound on es/2e^{-\mathcal{F}_{s}/2}. From the decay of s\mathcal{F}_{s}, we can improve this bound to es/2ase^{-\mathcal{F}_{s}/2}\geq a_{s} with as1a_{s}\to 1 as s+s\to+\infty, and the lemma is proved. ∎

Notation.

For simplicity, we denote by m(s)m(s) the exponent appearing in Lemma 3.4, i.e.

m(s)=m(s,r0,δ)=3223ass13r0/δ.m(s)=m(s,r_{0},\delta)=\sqrt{3}\cdot 2^{\frac{2}{3}}a_{s}s^{\frac{1}{3}}r_{0}/\delta.

3.2. Estimates around a zero

We now move to the study of the asymptotic behavior of μs\mu_{s} around a zero of the Pick differential. We denote by hs=eνs|dz|2h_{s}=e^{\nu_{s}}|dz|^{2} the Blaschke metric of the affine sphere with polynomial cubic differential qs=szkdz3q_{s}=sz^{k}dz^{3} on \mathbb{C}. We want to compare μs\mu_{s} and νs\nu_{s} on the ball B={|z|<ϵ}B=\{|z|<\epsilon\}. We rewrite the Blaschke metric gs=eμsσg_{s}=e^{\mu_{s}}\sigma with respect to the background flat metric |dz|2|dz|^{2} on BB: gs=eϕs|dz|2g_{s}=e^{\phi_{s}}|dz|^{2}, where ϕs\phi_{s} is a solution of the PDE (only defined on the closure of BB)

Δϕs=2eϕs4e2ϕs|qs|2.\Delta\phi_{s}=2e^{\phi_{s}}-4e^{-2\phi_{s}}|q_{s}|^{2}. (3.2)

Thus νs\nu_{s} is a solution of the same equation as ϕs\phi_{s}, but with different boundary values.

Lemma 3.5.

On B\partial B, we have ϕsνs=O(s16em(s))\phi_{s}-\nu_{s}=O(s^{\frac{1}{6}}e^{-m(s)}) as s+s\to+\infty.

Proof.

We know from Lemma 3.4 that

ϕs|B=μs+log(σ)=13log(2s2ϵ2k)+O(s16em(s))\phi_{s}|_{\partial B}=\mu_{s}+\log(\sigma)=\frac{1}{3}\,\log(2s^{2}\epsilon^{2k})+O(s^{\frac{1}{6}}e^{-m(s)})

for r0=3k+3ϵk+33r_{0}=\frac{3}{k+3}\epsilon^{\frac{k+3}{3}}, where we recall that σ\sigma is the hyperbolic metric on the fixed Riemann surface Σ\Sigma. Thus it is sufficient to show that

νs=13log(2s2ϵ2k)+o(s16em(s)).\nu_{s}=\frac{1}{3}\,\log(2s^{2}\epsilon^{2k})+o(s^{\frac{1}{6}}e^{-m(s)}).

Changing complex coordinates to w=s1k+3zw=s^{\frac{1}{k+3}}z, the ball BB can be rewritten as B={|w|<s1k+3ϵ}B=\{|w|<s^{\frac{1}{k+3}}\epsilon\}. In these coordinates, eνs|dz|2=eψs|dw|2e^{\nu_{s}}|dz|^{2}=e^{\psi_{s}}|dw|^{2} with ψs=νs2k+3log(s)\psi_{s}=\nu_{s}-\frac{2}{k+3}\log(s), and qs=wkdw3q_{s}=w^{k}dw^{3}. The estimates of ([DW15, Theorem 5.7]) then tell us that

ψs|B=13log(2s2kk+3ϵ2k)+O(em(s)s16).\psi_{s}|_{\partial B}=\frac{1}{3}\,\log(2s^{\frac{2k}{k+3}}\epsilon^{2k})+O\left(\frac{e^{-m(s)}}{s^{\frac{1}{6}}}\right).

Hence, on B\partial B,

νs\displaystyle\nu_{s} =\displaystyle= ψs+2k+3log(s)\displaystyle\psi_{s}+\frac{2}{k+3}\,\log(s)
=\displaystyle= 13log(2s2kk+3ϵ2k)+13log(s6k+3)+O(em(s)s16)\displaystyle\frac{1}{3}\,\log(2s^{\frac{2k}{k+3}}\epsilon^{2k})+\frac{1}{3}\,\log(s^{\frac{6}{k+3}})+O\left(\frac{e^{-m(s)}}{s^{\frac{1}{6}}}\right)
=\displaystyle= 13log(2s2ϵ2k)+O(em(s)s16).\displaystyle\frac{1}{3}\,\log(2s^{2}\epsilon^{2k})+O\left(\frac{e^{-m(s)}}{s^{\frac{1}{6}}}\right).

Lemma 3.6.

On the ball BB, we have ϕsνs=O(s16em(s))\phi_{s}-\nu_{s}=O(s^{\frac{1}{6}}e^{-m(s)}) as s+s\to+\infty.

Proof.

Define ηs=ϕsνs\eta_{s}=\phi_{s}-\nu_{s}. It satisfies the PDE

Δηs=ΔϕsΔνs=2eϕs4e2ϕs|qs|22eνs+4e2νs|qs|2\Delta\eta_{s}=\Delta\phi_{s}-\Delta\nu_{s}=2e^{\phi_{s}}-4e^{-2\phi_{s}}|q_{s}|^{2}-2e^{\nu_{s}}+4e^{-2\nu_{s}}|q_{s}|^{2}

on BB. Dividing by eνse^{\nu_{s}}, we find, upon setting hs=eνs|dz|2h_{s}=e^{\nu_{s}}|dz|^{2}, that

Δhsηs\displaystyle\Delta_{h_{s}}\eta_{s} =\displaystyle= eνsΔηs=2eηs4e2ϕsνs|qs|22+4e3νs|qs|2\displaystyle e^{-\nu_{s}}\Delta\eta_{s}=2e^{\eta_{s}}-4e^{-2\phi_{s}-\nu_{s}}|q_{s}|^{2}-2+4e^{-3\nu_{s}}|q_{s}|^{2}
=\displaystyle= 2(eηs1)4|qs|2e3νs(e2ηs1).\displaystyle 2(e^{\eta_{s}}-1)-4\frac{|q_{s}|^{2}}{e^{3\nu_{s}}}(e^{-2\eta_{s}}-1).

By the maximum principle,

|ηs|max(|max(ηs|B)|,|min(ηs|B)|),|\eta_{s}|\leq\max(|\max(\eta_{s}|_{\partial B})|,|\min(\eta_{s}|_{\partial B})|),

which gives the desired estimate by Lemma 3.5. ∎

4. Comparison between affine spheres

We denote by Fs(z)F_{s}(z) the frame field of the affine sphere arising from the data (σ,sq0)(\sigma,sq_{0}) on the surface SS restricted to the ball B={|z|<ϵ}B=\{|z|<\epsilon\}. We normalize the affine sphere so that Fs(0)=IdF_{s}(0)={\rm Id} for all s>0s>0. With this choice FsF_{s} will be complex-valued, belonging to a subgroup of SL(3,)\mathrm{SL}(3,\mathbb{C}) isomorphic to SL(3,)\mathrm{SL}(3,\mathbb{R}). Recall that FsF_{s} is the solution to the ODE

{Fs1dFs=Usdz+Vsdz¯,Fs(0)=Id,\left\{\begin{array}[]{l}F_{s}^{-1}dF_{s}=U_{s}dz+V_{s}d\bar{z},\\ F_{s}(0)={\rm Id},\end{array}\right.

where Us,VsU_{s},V_{s} are the matrices arising from the structure equations of the affine sphere. Precisely,

Us=(0012eϕs1zϕs00qseϕs0),Vs=(012eϕs000q¯seϕs10z¯ϕs).U_{s}=\left(\begin{array}[]{ccc}0&0&\frac{1}{2}e^{\phi_{s}}\\ 1&\partial_{z}\phi_{s}&0\\ 0&q_{s}e^{-\phi_{s}}&0\end{array}\right),\qquad V_{s}=\left(\begin{array}[]{ccc}0&\frac{1}{2}e^{\phi_{s}}&0\\ 0&0&\bar{q}_{s}e^{-\phi_{s}}\\ 1&0&\partial_{\bar{z}}\phi_{s}\end{array}\right). (4.1)

We denote by FM(w)F_{M}(w) the frame field of the (model) affine sphere over \mathbb{C} with polynomial cubic differential wkdw3w^{k}dw^{3} normalized so that FM(0)=IdF_{M}(0)={\rm Id}. Note FM(w)F_{M}(w) solves

{FM1dFM=UMdw+VMdw¯,FM(0)=Id.\left\{\begin{array}[]{l}F_{M}^{-1}dF_{M}=U_{M}dw+V_{M}d\bar{w},\\ F_{M}(0)={\rm Id}.\end{array}\right.

We compare FsF_{s} and FMF_{M} on the ball B={|z|<ϵ}B=\{|z|<\epsilon\} centered at a zero of order kk for q0q_{0}. Recall that w=s1k+3zw=s^{\frac{1}{k+3}}z, so we consider

Gs(z)=Fs(z)FM1(s1k+3z).G_{s}(z)=F_{s}(z)F_{M}^{-1}(s^{\frac{1}{k+3}}z).

The matrices GsG_{s} satisfy the differential equation

Gs1dGs\displaystyle G_{s}^{-1}dG_{s} =\displaystyle= FM(s1k+3z)Fs1(z)[dFsFM1(s1k+3z)Fs(z)FM1(s1k+3z)dFMFM1(s1k+3z)s1k+3]\displaystyle F_{M}(s^{\frac{1}{k+3}}z)F_{s}^{-1}(z)[dF_{s}\cdot F_{M}^{-1}(s^{\frac{1}{k+3}}z)-F_{s}(z)F_{M}^{-1}(s^{\frac{1}{k+3}}z)dF_{M}F_{M}^{-1}(s^{\frac{1}{k+3}}z)s^{\frac{1}{k+3}}]
=\displaystyle= FM(s1k+3z)[Usdz+Vsdz¯UMdzVMdz¯]FM1(s1k+3z)\displaystyle F_{M}(s^{\frac{1}{k+3}}z)[U_{s}dz+V_{s}d\bar{z}-U_{M}dz-V_{M}d\bar{z}]F_{M}^{-1}(s^{\frac{1}{k+3}}z)
=\displaystyle= FM(s1k+3z)[(UsUM)dz+(VsVM)dz¯]FM1(s1k+3z)\displaystyle F_{M}(s^{\frac{1}{k+3}}z)[(U_{s}-U_{M})dz+(V_{s}-V_{M})d\bar{z}]F_{M}^{-1}(s^{\frac{1}{k+3}}z)
=\displaystyle= FM(s1k+3z)ΘFM1(s1k+3z),\displaystyle F_{M}(s^{\frac{1}{k+3}}z)\Theta F_{M}^{-1}(s^{\frac{1}{k+3}}z),

when written in the zz coordinate, where

Θ(z)\displaystyle\Theta(z) =\displaystyle= (00(eϕseνs)/20z(ϕsνs)00szk(eϕseνs)0)dz+\displaystyle\left(\begin{array}[]{ccc}0&0&(e^{\phi_{s}}-e^{\nu_{s}})/2\\ 0&\partial_{z}(\phi_{s}-\nu_{s})&0\\ 0&sz^{k}(e^{-\phi_{s}}-e^{-\nu_{s}})&0\end{array}\right)dz+{}
(0(eϕseνs)/2000sz¯k(eϕseνs)00z¯(ϕsνs))dz¯.\displaystyle\left(\begin{array}[]{ccc}0&(e^{\phi_{s}}-e^{\nu_{s}})/2&0\\ 0&0&s\bar{z}^{k}(e^{-\phi_{s}}-e^{-\nu_{s}})\\ 0&0&\partial_{\bar{z}}(\phi_{s}-\nu_{s})\\ \end{array}\right)d\bar{z}.

where νs\nu_{s} was defined in section 3.2.

Lemma 4.1.

On BB, we have ΘCs53em(s)\|\Theta\|_{\infty}\leq Cs^{\frac{5}{3}}e^{-m(s)} as s+s\to+\infty.

Proof.

We handle the various entries in Θ\Theta one-by-one. First of all, the Mean Value Theorem implies for |z|<ϵ|z|<\epsilon that

eϕs(z)eνs(z)=ep(z)(ϕs(z)νs(z))e^{\phi_{s}(z)}-e^{\nu_{s}(z)}=e^{p(z)}(\phi_{s}(z)-\nu_{s}(z))

for some p(z)p(z) between ϕs(z)\phi_{s}(z) and νs(z)\nu_{s}(z). A straightforward application of the Maximum Principle applied to equation (3.2) on BB, together with a boundary estimate from Lemma 3.4, then gives

ϕs(z)13log(2s2ϵ2k)+o(1),\phi_{s}(z)\leq\frac{1}{3}\log(2s^{2}\epsilon^{2k})+o(1),

and the same is true for νs(z)\nu_{s}(z). Then Lemma 3.6 shows

eϕs(z)eνs(z)=O(s23s16em(s))=O(s56em(s)).e^{\phi_{s}(z)}-e^{\nu_{s}(z)}=O(s^{\frac{2}{3}}\cdot s^{\frac{1}{6}}e^{-m(s)})=O(s^{\frac{5}{6}}e^{-m(s)}).

Similarly,

eϕs(z)eνs(z)=ep(z)(ϕs(z)νs(z))e^{-\phi_{s}(z)}-e^{-\nu_{s}(z)}=-e^{-p(z)}(\phi_{s}(z)-\nu_{s}(z))

for some p(z)p(z) between ϕs(z)\phi_{s}(z) and νs(z)\nu_{s}(z). By considering the change of coordinate w=s1k+3zw=s^{\frac{1}{k+3}}z, we see for all zz\in\mathbb{C}

νs(z)=1k+3log(s2)+ν1(s1k+3z).\nu_{s}(z)=-\frac{1}{k+3}\,\log(s^{2})+\nu_{1}(s^{-\frac{1}{k+3}}z).

Now ν1\nu_{1} is bounded below, as in [DW15, Corollary 5.2], which implies

eνs(z)=O(s2k+3).e^{-\nu_{s}(z)}=O(s^{\frac{2}{k+3}}).

Lemma 3.6 then shows

eϕs(z)eνs(z)=O(s2k+3+16em(s)).e^{-\phi_{s}(z)}-e^{-\nu_{s}(z)}=O(s^{\frac{2}{k+3}+\frac{1}{6}}e^{-m(s)}).

The result then follows if we show a decay estimate holds for z(ϕsνs)\partial_{z}(\phi_{s}-\nu_{s}) and z¯(ϕsνs)\partial_{\bar{z}}(\phi_{s}-\nu_{s}). Set ηs=ϕsνs\eta_{s}=\phi_{s}-\nu_{s}. We know from the computations in Lemma 3.6 that

Δhsηs=2(eηs1)4|qs|2e3νs(e2ηs1).\Delta_{h_{s}}\eta_{s}=2(e^{\eta_{s}}-1)-4\frac{|q_{s}|^{2}}{e^{3\nu_{s}}}(e^{-2\eta_{s}}-1).

Then |qs|2/e3νs|q_{s}|^{2}/e^{3\nu_{s}} is uniformly bounded, because |qs|2=s2|z|2k|q_{s}|^{2}=s^{2}|z|^{2k} and, using the same notation as in Lemma 3.5 and the subsolution for ψs\psi_{s} found in [DW15, Theorem 5.1],

3νs\displaystyle 3\nu_{s} =3ψs+6k+3log(s)\displaystyle=3\psi_{s}+\frac{6}{k+3}\log(s)
log(2s2kk+3|z|2k)+log(s6k+3)\displaystyle\geq\log(2s^{\frac{2k}{k+3}}|z|^{2k})+\log(s^{\frac{6}{k+3}})
=log(2s2|z|2k).\displaystyle=\log(2s^{2}|z|^{2k})\ .

So

ΔhsηsCs16em(s)\|\Delta_{h_{s}}\eta_{s}\|_{\infty}\leq Cs^{\frac{1}{6}}e^{-m(s)}

and, using the supersolution for νs\nu_{s} in [DW15, Theorem 5.1],

zz¯ηseνsΔhsηsC(|qs|2+a)13s16em(s)Cs56em(s).\|\partial_{z}\partial_{\bar{z}}\eta_{s}\|_{\infty}\leq\|e^{\nu_{s}}\|_{\infty}\|\Delta_{h_{s}}\eta_{s}\|_{\infty}\leq C(|q_{s}|^{2}+a)^{\frac{1}{3}}s^{\frac{1}{6}}e^{-m(s)}\leq Cs^{\frac{5}{6}}e^{-m(s)}.

The bounds on z(ϕsνs)\partial_{z}(\phi_{s}-\nu_{s}) and z¯(ϕsνs)\partial_{\bar{z}}(\phi_{s}-\nu_{s}) then follow from the Schauder and LpL^{p} estimates. ∎

Proposition 4.2.

Let UU be a neighborhood of the union of the Stokes rays on BB. For every γ>0\gamma>0, there is s0>0s_{0}>0 so that for all zBUz\in B\setminus U and s>s0s>s_{0},

FM(s1k+3z)ΘFM1(s1k+3z)γ.\|F_{M}(s^{\frac{1}{k+3}}z)\Theta F_{M}^{-1}(s^{\frac{1}{k+3}}z)\|_{\infty}\leq\gamma.
Proof.

Let FTF_{T} denote the frame field of the standard Ţiţeica surface (see Example 2.1) with cubic differential dx3dx^{3} on the plane, which can be explicitly be written as

FT(x)=Sexp(223e(x)000223e(x/ω)000223e(x/ω2))S1F_{T}(x)=S\,\exp\left(\begin{array}[]{ccc}2^{\frac{2}{3}}\mathcal{R}e(x)&0&0\\ 0&2^{\frac{2}{3}}\mathcal{R}e(x/\omega)&0\\ 0&0&2^{\frac{2}{3}}\mathcal{R}e(x/\omega^{2})\end{array}\right)\,S^{-1}

for ω=e2πi/3\omega=e^{2\pi i/3} and some choice of conjugating matrix SS.
From [DW15, Lemma 6.4], we know that, outside a compact set KK\subset\mathbb{C}, we have FM(3k+3wk+33)=(A+o(1))FT(3k+3wk+33)F_{M}(\frac{3}{k+3}w^{\frac{k+3}{3}})=(A+o(1))F_{T}(\frac{3}{k+3}w^{\frac{k+3}{3}}) as ww\to\infty, where AA is a constant matrix that only depends on the sector in the complement of the Stokes rays containing ww. Here we use 3k+3wk+33\frac{3}{k+3}w^{\frac{k+3}{3}} because the result of [DW15] is stated in terms of natural coordinates for q0q_{0}, instead of the coordinate, say ww, centered at the zero of q0q_{0}. Therefore,

FM(3k+3wk+33)CFT(3k+3wk+33)for all w.\left\|F_{M}\left(\frac{3}{k+3}w^{\frac{k+3}{3}}\right)\right\|_{\infty}\leq C\left\|F_{T}\left(\frac{3}{k+3}w^{\frac{k+3}{3}}\right)\right\|_{\infty}\qquad\mbox{for all }w\in\mathbb{C}.

By [DW15, page 1768], we conclude that

FM(s1k+3z)FM1(s1k+3z)Cec(θ)s13ϵk+33,\|F_{M}(s^{\frac{1}{k+3}}z)\|_{\infty}\|F_{M}^{-1}(s^{\frac{1}{k+3}}z)\|_{\infty}\leq Ce^{c(\theta)s^{\frac{1}{3}}\epsilon^{\frac{k+3}{3}}},

where c(θ)2233c(\theta)\leq 2^{\frac{2}{3}}\sqrt{3} and achieves this maximum value when θ\theta corresponds to a Stokes direction. In particular, when zBUz\in B\setminus U, there is α>0\alpha>0 such that c(θ)2233αc(\theta)\leq 2^{\frac{2}{3}}\sqrt{3}-\alpha. In the definition of m(s)m(s), we may choose both δ\delta sufficiently close to 11 and ss sufficiently large so that

s53ec(θ)s13ϵk+33em(s)eβs13s^{\frac{5}{3}}e^{c(\theta)s^{\frac{1}{3}}\epsilon^{\frac{k+3}{3}}}e^{-m(s)}\leq e^{-\beta s^{\frac{1}{3}}}

for some β>0\beta>0. Therefore, by Lemma 4.1,

FM(s1k+3z)ΘFM1(s1k+3z)Ceβs13γ\|F_{M}(s^{\frac{1}{k+3}}z)\Theta F_{M}^{-1}(s^{\frac{1}{k+3}}z)\|_{\infty}\leq Ce^{-\beta s^{\frac{1}{3}}}\leq\gamma

for ss sufficiently large, independently of zBUz\in B\setminus U. ∎

Corollary 4.3.

There exists s0>0s_{0}>0 such that for all s>s0s>s_{0} and zBUz\in B\setminus U, we have

Gs(z)=Id+o(Id)ass+.G_{s}(z)={\rm Id}+o(\rm Id)\qquad\mbox{as}\quad s\to+\infty.
Proof.

Let z=teiθBUz=te^{i\theta}\in B\setminus U. Then by Lemma B.2 in [DW15] applied to B(t)=FM(s1k+3teiθ)Θ(t)FM1(s1k+3teiθ)B(t)=F_{M}(s^{\frac{1}{k+3}}te^{i\theta})\Theta(t)F_{M}^{-1}(s^{\frac{1}{k+3}}te^{i\theta}) with t[0,ϵ]t\in[0,\epsilon] and Proposition 4.2, we have

Gs(teiθ)IdCγ,\|G_{s}(te^{i\theta})-{\rm Id}\|_{\infty}\leq C\gamma,

which can be made arbitrarily small as s+s\to+\infty. ∎

In particular, this implies that for all zBUz\in B\setminus U,

Fs(z)=Gs(z)FM(s1k+3z)=(Id+o(Id))FM(s1k+3z)ass+.F_{s}(z)=G_{s}(z)F_{M}(s^{\frac{1}{k+3}}z)=({\rm Id}+o({\rm Id}))F_{M}(s^{\frac{1}{k+3}}z)\qquad\mbox{as}\quad s\to+\infty.

Combining this with the fact that FM(s1k+3z)=(A+o(Id))FT(s1k+3z)F_{M}(s^{\frac{1}{k+3}}z)=(A+o(\rm Id))F_{T}(s^{\frac{1}{k+3}}z), where AA only depends on the sector in the complement of the Stokes line that contains zz, we obtain

Corollary 4.4.

For every zBz\in B inside the ithi^{th} sector in the complement of the Stokes rays,

Fs(z)FT1(s1k+3z)s+Ai.F_{s}(z)\cdot F_{T}^{-1}(s^{\frac{1}{k+3}}z)\stackrel{{\scriptstyle s\to+\infty}}{{\longrightarrow}}A_{i}\ .

Here, of course, we have adapted our choice of UU to our choice of zz.

Corollary 4.5.

Let [θ0,θ1][\theta_{0},\theta_{1}] contain one Stokes direction in its interior. Then

A01A1=SUS1,A_{0}^{-1}A_{1}=SUS^{-1},

where UU is one of the unipotents introduced in [DW15].

Proof.

The matrices AiA_{i} are defined as the limit of FM(s1k+3eiθi)FT1(s1k+3eiθi)F_{M}(s^{\frac{1}{k+3}}e^{i\theta_{i}})F_{T}^{-1}(s^{\frac{1}{k+3}}e^{i\theta_{i}}) as ss\to\infty. Thus the result follows from [DW15, Lemma 6.5]. ∎

Theorem 4.6 (holonomy along arcs).

Let γ(t)=ϵeit\gamma(t)=\epsilon e^{it} with t[θ0,θ1]t\in[\theta_{0},\theta_{1}] as in Corollary 4.5. Consider Gt(s)=Fs(γ(t))FT1(s1k+3γ(t))G_{t}(s)=F_{s}(\gamma(t))F_{T}^{-1}(s^{\frac{1}{k+3}}\gamma(t)). Then

lims+Gθ01(s)Gθ1(s)=SUS1,\lim_{s\to+\infty}G_{\theta_{0}}^{-1}(s)G_{\theta_{1}}(s)=SUS^{-1},

where UU is the same unipotent as in Corollary 4.5.

Proof.

This follows immediately from Corollaries 4.4 and 4.5 because

Gθ0(s)\displaystyle G_{\theta_{0}}(s) =\displaystyle= Fs(ϵeiθ0)FT1(s1k+3ϵeiθ0)A0,\displaystyle F_{s}(\epsilon e^{i\theta_{0}})F_{T}^{-1}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{0}})\to A_{0},
Gθ1(s)\displaystyle G_{\theta_{1}}(s) =\displaystyle= Fs(ϵeiθ1)FT1(s1k+3ϵeiθ1)A1,\displaystyle F_{s}(\epsilon e^{i\theta_{1}})F_{T}^{-1}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{1}})\to A_{1},

and A01A1=SUS1A_{0}^{-1}A_{1}=SUS^{-1}. ∎

We remark that when the interval [θ0,θ1][\theta_{0},\theta_{1}] contains more than one Stokes direction we can still apply Corollary 4.5 and Theorem 4.6 after splitting the interval [θ0,θ1][\theta_{0},\theta_{1}] into subintervals containing only one Stokes direction and thus satisfying the assumptions of Corollary 4.5. Because the holonomy is multiplicative along concatenation of paths, we will have

lims+Gθ01(s)Gθ1(s)=SUS1,\lim_{s\to+\infty}G_{\theta_{0}}^{-1}(s)G_{\theta_{1}}(s)=SUS^{-1},

where UU is now a product of unipotents depending on the Stokes directions the arc crosses.

5. Asymptotic holonomy

We want to compute the asymptotic holonomy of the flat connection s\nabla^{s} on the rank-3 bundle E=𝒪TΣE=\mathcal{O}\oplus T_{\mathbb{C}}\Sigma along a |q0|23|q_{0}|^{\frac{2}{3}}-geodesic path that may cross some of the zeros of the cubic differential q0q_{0}. We say such a geodesic path γ\gamma is regular in that each segment away from the zeros of q0q_{0} are

  • not in the directions of the walls of a Weyl chamber, so that e(γϕi)e(γϕj)\mathcal{R}e(\gamma^{*}\phi_{i})\neq\mathcal{R}e(\gamma^{*}\phi_{j}) for iji\neq j. Here ϕi\phi_{i} is a root of p(λ)=λ3q0p(\lambda)=\lambda^{3}-q_{0}.

  • not in the Stokes directions.

We later remove these hypotheses.

It is convenient to work in the universal cover of SS. Equip SS with the conformal hyperbolic metric and identify S~\tilde{S} with the strip model of the hyperbolic plane

2={w=μ+iν:|μ|<π/2}\mathbb{H}^{2}=\{w=\mu+i\nu\in\mathbb{C}:|\mu|<\pi/2\}

with metric g2=dμ2+dν2cos2(μ)g_{\mathbb{H}^{2}}=\frac{d\mu^{2}+d\nu^{2}}{\cos^{2}(\mu)}. The vertical line μ=0\mu=0 with arc-length parameter ν\nu is a geodesic; so a hyperbolic deck transformation can be represented, up to conjugation, by the transformation T(w)=w+iLT(w)=w+iL, where LL is the translation length.

Remark 5.1.

We want to use this model because it gives a way of defining a frame on EE which we can use to compute parallel transport. The bundle EE lifts to a bundle E~\tilde{E} over 2\mathbb{H}^{2} which we now trivialize using the global frame ={1,w,w¯}\mathcal{F}=\{1,\partial_{w},\partial_{\bar{w}}\}. This frame is not parallel with respect to s\nabla^{s}. However, because it is globally defined, we can define the holonomy of ~s\tilde{\nabla}^{s} along an arc as a comparison between the terminal parallel transport of a frame in the fixed basis and the frame at the terminal point. Given [γ]π1(S)[\gamma]\in\pi_{1}(S), we can assume that the hyperbolic isometry corresponding to [γ][\gamma] is Tγ(w)=w+iLT_{\gamma}(w)=w+iL for some LL\in\mathbb{R}. Fix w02w_{0}\in\mathbb{H}^{2} and let w1=Tγ(w0)w_{1}=T_{\gamma}(w_{0}). Note that (w0)=Tγ(w1)\mathcal{F}(w_{0})=T_{\gamma}^{*}\mathcal{F}(w_{1}). Let cγ(t)c_{\gamma}(t) be the |q0|23|q_{0}|^{\frac{2}{3}}-geodesic connecting w0w_{0} and w1w_{1} with t[0,1]t\in[0,1]. If we have a matrix representation M(cγ(t))M(c_{\gamma}(t)) of the parallel transport along cγ(t)c_{\gamma}(t) with respect to the frame \mathcal{F}, then the matrix M(cγ(1))M(c_{\gamma}(1)) represents the holonomy of the flat connection between the final and initial points as their frames are identified in the quotient.

Remark 5.2.

We can assume that w0w_{0} and w1w_{1} are not zeros of q0q_{0}, so that the geodesic path cγc_{\gamma} starts and ends with a segment not containing any zeros.

The path cγc_{\gamma} will in general cross some of the zeros p1,,pp_{1},\dots,p_{\ell} of the cubic differential q0q_{0} with multiplicities k1,,kk_{1},\dots,k_{\ell}. In fact, we write cγc_{\gamma} as the union of c1,,cc_{1},\dots,c_{\ell}, where each cic_{i} is the straight line path in the flat coordinates for |q0|23|q_{0}|^{\frac{2}{3}} from pip_{i} to pi1p_{i-1}. Each cic_{i} does not intersect any zeros of q0q_{0} except at its endpoints. We fix ϵ>0\epsilon>0 and identify a neighborhood NiN_{i} of each zero and a conformal coordinate zz so that q0=zkidz3q_{0}=z^{k_{i}}dz^{3} on NiN_{i} and the |q0|23|q_{0}|^{\frac{2}{3}}-radius of NiN_{i} is ϵ\epsilon. Note for ϵ\epsilon small the closures Ni¯\overline{N_{i}} do not intersect. We modify cγc_{\gamma} to form a new path c~γ\tilde{c}_{\gamma} by deleting each cγNic_{\gamma}\cap N_{i} and replacing it with an arc βi\beta_{i} in Ni\partial N_{i} so that c~γ\tilde{c}_{\gamma} is continuous and homotopic to the original geodesic. Now cγiNi¯c_{\gamma}\setminus\cup_{i}\overline{N_{i}} consists of a number of line segments c~ici\tilde{c}_{i}\subset c_{i} in the flat q0q_{0}-coordinates. Divide each line segment c~i\tilde{c}_{i} between NiN_{i} and Ni+1N_{i+1} into two segments δ~i\tilde{\delta}_{i} and α~i1\tilde{\alpha}_{i-1}, so that each Ni¯\overline{N_{i}} has an incoming line segment α~i\tilde{\alpha}_{i} and an outgoing one δ~i\tilde{\delta}_{i}.

Refer to caption
Figure 1. Definition of the subpaths.

In total, c~γ\tilde{c}_{\gamma} is the concatenation of α~,β,δ~,,α~1,β1,δ~1\tilde{\alpha}_{\ell},\beta_{\ell},\tilde{\delta}_{\ell},\dots,\tilde{\alpha}_{1},\beta_{1},\tilde{\delta}_{1}. The basepoint is p=α~δ~1p=\tilde{\alpha}_{\ell}\cap\tilde{\delta}_{1}. We also denote by αi{\alpha}_{i} and δi\delta_{i} the prolongments of α~i\tilde{\alpha}_{i} and δ~i\tilde{\delta}_{i} to their forward and backward zero respectively. Since c~γ\tilde{c}_{\gamma} and cγc_{\gamma} are homotopic, the holonomies along these paths are the same. Then

Hols(c~γ)=Hols(δ~1)Hols(β1)Hols(α~1)Hols(δ~)Hols(β)Hols(α~).\mathrm{Hol}_{s}(\tilde{c}_{\gamma})=\mathrm{Hol}_{s}(\tilde{\delta}_{1})\mathrm{Hol}_{s}(\beta_{1})\mathrm{Hol}_{s}(\tilde{\alpha}_{1})\cdots\mathrm{Hol}_{s}(\tilde{\delta}_{\ell})\mathrm{Hol}_{s}(\beta_{\ell})\mathrm{Hol}_{s}(\tilde{\alpha}_{\ell}).

We want to find estimates for each factor and arrive at something of the form

Hols(c~γ)=A(s)+E(s),E(s)=o(A(s))\mathrm{Hol}_{s}(\tilde{c}_{\gamma})=A(s)+E(s),\qquad\|E(s)\|=o(\|A(s)\|)

with A(s)A(s) explicit and depending only on q0q_{0} and on the geodesic path cγc_{\gamma}.

Remark 5.3.

Let z=z(w)z=z(w) be a conformal change of coordinates. For instance, let zz be a natural coordinate for qsq_{s}. The coordinate zz induces a new frame 𝒢={1,z,z¯}\mathcal{G}=\{1,\partial_{z},\partial_{\bar{z}}\}. There is a diagonal matrix d(z)d(z) depending on the derivatives of zz so that =𝒢d(z)\mathcal{F}=\mathcal{G}\,d(z). Moreover, if z1z_{1} and z2z_{2} are two natural coordinates at a point, then z1z_{1} and z2z_{2} differ by a translation and a multiplication by a third root of unity. If we choose the natural coordinates so that they induce the same frame on the overlaps, we can multiply the matrices representing the parallel transport along consecutive arcs.

Remark 5.4.

If (U,z1)(U,z_{1}) and (U,z2)(U,z_{2}) are two natural coordinate charts that cover a path γ\gamma and overlap at a point aa, we note that z1z_{1} and z2z_{2} induce the same frame at aa if and only if the path γ\gamma makes the same angle with the positive horizontal axis, as seen in the coordinates z1z_{1} and z2z_{2}.

Let a,a2a,a^{\prime}\in\mathbb{H}^{2} and denote by Ta,aT_{a,a^{\prime}} the parallel transport from aa to aa^{\prime} for the lift ~s\tilde{\nabla}^{s} of the flat connection. Assume aa and aa^{\prime} are not zeros of q0q_{0} and are in the same natural coordinate zz . Let 𝒢(a)={1,z,z¯}\mathcal{G}(a)=\{1,\partial_{z},\partial_{\bar{z}}\} be the standard frame induced by zz. The frame 𝒢\mathcal{G} is defined at aa^{\prime} as well; so we can find a matrix Ψs(a)\Psi_{s}(a^{\prime}) such that

Ta,a(𝒢(a))Ψs(a)=𝒢(a).T_{a,a^{\prime}}(\mathcal{G}(a))\Psi_{s}(a^{\prime})=\mathcal{G}(a^{\prime}).

Let z(t)z(t) be a path connecting aa and aa^{\prime}. The parallel transport condition is equivalent to Ψs(z(t))\Psi_{s}(z(t)) being a solution of the initial value problem

{Ψs(z(0))=Id,Ψs1dΨs=Usdz+Vsdz¯,\left\{\begin{array}[]{l}\Psi_{s}(z(0))={\rm Id},\\ \Psi_{s}^{-1}d\Psi_{s}=U_{s}dz+V_{s}d\bar{z},\end{array}\right.

where UsU_{s} and VsV_{s} are defined in Equation (4.1). The matrix representing the parallel transport Ta,a:E~aE~aT_{a,a^{\prime}}\!:\tilde{E}_{a}\to\tilde{E}_{a^{\prime}} with respect to the frames 𝒢(a)\mathcal{G}(a) and 𝒢(a)\mathcal{G}(a^{\prime}) is then Ψs(a)1\Psi_{s}(a^{\prime})^{-1}.

In what follows, instead of solving the initial value problem above, we compare Ψs\Psi_{s} with the solution ΨT\Psi_{T} of the initial value problem

{ΨT(z(0))=Id,ΨT1dΨT=UTdz+VTdz¯,\left\{\begin{array}[]{l}\Psi_{T}(z(0))={\rm Id},\\ \Psi_{T}^{-1}d\Psi_{T}=U_{T}dz+V_{T}d\bar{z},\end{array}\right.

where UTU_{T} and VTV_{T} are the matrices appearing in the structure equations for the affine sphere over \mathbb{C} with constant cubic differential dz3dz^{3}. We know ([Lof07]) that

ΨT(|w|eiθ)=Sexp(|z|D(θ))S1,\Psi_{T}(|w|e^{i\theta})=S\exp(|z|D(\theta))S^{-1},

where

D(θ)=(223cosθ000223cos(θ2π/3)000223cos(θ4π/3)).D(\theta)=\left(\begin{array}[]{ccc}2^{\frac{2}{3}}\cos\theta&0&0\\ 0&2^{\frac{2}{3}}\cos(\theta-2\pi/3)&0\\ 0&0&2^{\frac{2}{3}}\cos(\theta-4\pi/3)\end{array}\right). (5.1)

and SS is the conjugating matrix that appeared e.g. in Proposition 4.2.

Remark 5.5.

Note that Ψs\Psi_{s} solves the same ODE as the frame field FsF_{s} of the associated affine sphere. They differ by the value at the initial point.

The first author in [Lof07] considered the case of geodesic paths which do not hit any zeros and determined the asymptotic behavior of the eigenvalues along such paths. We would like to use Proposition 3 of [Lof07], but unfortunately the published statement must be modified to Proposition 5.6 below, as there is a gap in the proof. The final paragraph of the proof in [Lof07] is unsupported. The main theorem of [Lof07] is still true, as follows from the results presented here. The main additional technique needed, which was available at the writing of [Lof07], is the fact that the largest eigenvalue of the holonomy along a path (and the reverse path) is enough to determine all the eigenvalues in SL(3,)\mathrm{SL}(3,\mathbb{R}). The first author regrets the error.

Thus we have the following proposition. We note that Collier-Li and Mochizuki have proved stronger estimates in a more general setting in the case in which no two eigenvalues are equal [CL17, Moc16].

Proposition 5.6 (Holonomy along rays).

The parallel transports along the segments α~i\tilde{\alpha}_{i} and δ~i\tilde{\delta}_{i} with respect to the frame 𝒢\mathcal{G} induced by a natural coordinate zz for qsq_{s} are given by the matrices

Hols(α~i)\displaystyle\mathrm{Hol}_{s}(\tilde{\alpha}_{i}) =\displaystyle= Sdiag(es13μ~1i,es13μ~2i,es13μ~3i)S1+o(es13μ~i),\displaystyle S\mathrm{diag}(e^{s^{\frac{1}{3}}\tilde{\mu}_{1}^{i}},e^{s^{\frac{1}{3}}\tilde{\mu}_{2}^{i}},e^{s^{\frac{1}{3}}\tilde{\mu}_{3}^{i}})S^{-1}+o(e^{s^{\frac{1}{3}}\tilde{\mu}^{i}}),
Hols(δ~i)\displaystyle\mathrm{Hol}_{s}(\tilde{\delta}_{i}) =\displaystyle= Sdiag(es13λ~1i,es13λ~2i,es13λ~3i)S1+o(es13λ~i),\displaystyle S\mathrm{diag}(e^{s^{\frac{1}{3}}\tilde{\lambda}_{1}^{i}},e^{s^{\frac{1}{3}}\tilde{\lambda}_{2}^{i}},e^{s^{\frac{1}{3}}\tilde{\lambda}_{3}^{i}})S^{-1}+o(e^{s^{\frac{1}{3}}\tilde{\lambda}^{i}}),

as s+s\to+\infty, where

μ~ji=223e(α~iϕj),λ~ji=223e(δ~iϕj),\tilde{\mu}^{i}_{j}=-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{\tilde{\alpha}_{i}}\phi_{j}\right),\qquad\tilde{\lambda}^{i}_{j}=-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{\tilde{\delta}_{i}}\phi_{j}\right),

ϕj\phi_{j} are the roots of λ3q0\lambda^{3}-q_{0}, and

μ~i=max{μ~1i,μ~2i,μ~3i},λ~i=max{λ~1i,λ~2i,λ~3i}.\tilde{\mu}^{i}=\max\{\tilde{\mu}_{1}^{i},\tilde{\mu}_{2}^{i},\tilde{\mu}_{3}^{i}\},\qquad\tilde{\lambda}^{i}=\max\{\tilde{\lambda}_{1}^{i},\tilde{\lambda}_{2}^{i},\tilde{\lambda}_{3}^{i}\}.
Remark 5.7.

The error bounds in the previous proposition can be improved to O(es13(μ~iC))O(e^{s^{\frac{1}{3}}(\tilde{\mu}^{i}-C)}) for CC a positive constant depending the q0q_{0}-distance of the path to the zero set of q0q_{0}, by using Lemma 3.4 instead of the coarser bounds used in [Lof07].

Note that if we parametrize the path δ~i\tilde{\delta}_{i} by δ~i(t)=teiθi\tilde{\delta}_{i}(t)=te^{i\theta_{i}} with t[ϵ,Li/2]t\in[\epsilon^{\prime},L_{i}/2] for LiL_{i} the q0q_{0}-length of the geodesic segment between successive zeros of q0q_{0} and ϵ=3k+3ϵk+33\epsilon^{\prime}=\frac{3}{k+3}\epsilon^{\frac{k+3}{3}}, then

e(δ~iϕj)=e(ϵLi/2δ~iϕj)=(Li/2ϵ)cos(θi2(j1)π/3).\mathcal{R}e\left(\int_{\tilde{\delta}_{i}}\phi_{j}\right)=\mathcal{R}e\left(\int_{\epsilon^{\prime}}^{L_{i}/2}\tilde{\delta}_{i}^{*}\phi_{j}\right)=(L_{i}/2-\epsilon^{\prime})\cos(\theta_{i}-2(j-1)\pi/3).

Hence the position of the largest eigenvalue of the diagonal matrix depends only on the angle that δ~i\tilde{\delta}_{i} makes with the positive xx-axis in the chosen natural coordinates.

Remark 5.8.

There is a relation between the diagonal matrices in Proposition 5.6 and ΨT\Psi_{T}. Precisely, if w=reiθw=re^{i\theta} and γ(t)=teiθ\gamma(t)=te^{i\theta} with t[0,r]t\in[0,r], then

ΨT(w)=Sdiag(eμ1,eμ2,eμ3)S1SD(γ)S1,\Psi_{T}(w)=S\mathrm{diag}(e^{-\mu_{1}},e^{-\mu_{2}},e^{-\mu_{3}})S^{-1}\eqqcolon SD(\gamma)S^{-1},

where

μj=223e(γϕj),\mu_{j}=-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{\gamma}\phi_{j}\right),

and we choose the same conjugating matrix SS as in Section 4.2.

We now compute the parallel transport along the circular arcs βi\beta_{i}. In the zz-coordinate centered at a zero so that qs=szkidz3q_{s}=sz^{k_{i}}dz^{3}, the path βi\beta_{i} is parametrized by βi(θ)=ϵeiθ\beta_{i}(\theta)=\epsilon e^{i\theta} with θ[θi,θi+1]\theta\in[\theta_{i},\theta_{i+1}]. Let w=3k+3ωjis13zk+33w=\frac{3}{k+3}\omega^{j_{i}}s^{\frac{1}{3}}z^{\frac{k+3}{3}} be a natural coordinate for qsq_{s}. Choose ji{1,2,3}j_{i}\in\{1,2,3\} so that the angle the incoming path αi\alpha_{i} makes with the positive xx-axis coincides with the angle we saw in the previous natural coordinate chart.

Proposition 5.9 (Holonomy along arcs).

Assume θi\theta_{i} and θi+1\theta_{i+1} do not correspond to Stokes directions. Then the holonomy along βi\beta_{i} satisfies

Hols(βi)=FT1(βi(θi+1))(SU(θi,θi+1)1S1+o(Id))FT(βi(θi))as s+\mathrm{Hol}_{s}(\beta_{i})=F_{T}^{-1}(\beta_{i}(\theta_{i+1}))(SU(\theta_{i},\theta_{i+1})^{-1}S^{-1}+o(\mathrm{Id}))F_{T}(\beta_{i}(\theta_{i}))\qquad\mbox{as }s\to+\infty

with respect to the frame induced by the natural coordinate. Here U(θi,θi+1)U(\theta_{i},\theta_{i+1}) is a product of unipotent matrices depending on which Stokes rays the path βi\beta_{i} crosses.

Proof.

Recall that Hols(βi)\mathrm{Hol}_{s}(\beta_{i}) is the inverse of the matrix Ψs\Psi_{s} which solves the initial value problem

{Ψs1dΨs=Usdz+Vsdz¯,Ψ(βi(θi))=Id.\left\{\begin{array}[]{l}\Psi_{s}^{-1}d\Psi_{s}=U_{s}dz+V_{s}d\bar{z},\\ \Psi(\beta_{i}(\theta_{i}))={\rm Id}.\end{array}\right.

The frame field Fs(z)F_{s}(z) is a solution of the same ODE with different initial conditions, so

Ψs(z)=Fs(ϵeiθi)1Fs(z)\Psi_{s}(z)=F_{s}(\epsilon e^{i\theta_{i}})^{-1}F_{s}(z)

is the solution of the above initial value problem. Now, by Corollaries 4.4 and 4.5 and the subsequent remark,

lims+FT(s1k+3ϵeiθi)Fs1(ϵeiθi)Fs(ϵeiθi+1)FT1(s1k+3ϵeiθi+1)=Ai1Ai+1=SUS1,\lim_{s\to+\infty}F_{T}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{i}})F_{s}^{-1}(\epsilon e^{i\theta_{i}})F_{s}(\epsilon e^{i\theta_{i+1}})F_{T}^{-1}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{i+1}})=A_{i}^{-1}A_{i+1}=SUS^{-1},

where U=U(θi,θi+1)U=U(\theta_{i},\theta_{i+1}) is as in the statement. Hence

Ψs(ϵeiθi+1)=Fs(ϵeiθi)1Fs(ϵeiθi+1)=FT1(s1k+3ϵeiθi)(SUS1+o(1))FT(s1k+3ϵeiθi+1).\Psi_{s}(\epsilon e^{i\theta_{i+1}})=F_{s}(\epsilon e^{i\theta_{i}})^{-1}F_{s}(\epsilon e^{i\theta_{i+1}})=F_{T}^{-1}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{i}})(SUS^{-1}+o(1))F_{T}(s^{\frac{1}{k+3}}\epsilon e^{i\theta_{i+1}}).

Because Hols=Ψs1\mathrm{Hol}_{s}=\Psi_{s}^{-1} the claim follows. ∎

Combining the holonomy of each subpath (Proposition 5.9 and Proposition 5.6), we obtain

Hols(c~γ)\displaystyle\mathrm{Hol}_{s}(\tilde{c}_{\gamma}) =\displaystyle= i=1(SD(δ~i)1S1+o(es13λ~i))\displaystyle\prod_{i=1}^{\ell}\left(SD(\tilde{\delta}_{i})^{-1}S^{-1}+o(e^{s^{\frac{1}{3}}\tilde{\lambda}^{i}})\right)\cdot{}
SD(δiδ~i)1S1(SU(θi,θi+1)1S1+o(Id))SD(αiα~i)1S1\displaystyle SD(\delta_{i}\setminus\tilde{\delta}_{i})^{-1}S^{-1}(SU(\theta_{i},\theta_{i+1})^{-1}S^{-1}+o(\mathrm{Id}))SD(\alpha_{i}\setminus\tilde{\alpha}_{i})^{-1}S^{-1}\cdot{}
(SD(α~i)1S1+o(es13μ~i))\displaystyle\left(SD(\tilde{\alpha}_{i})^{-1}S^{-1}+o(e^{s^{\frac{1}{3}}\tilde{\mu}^{i}})\right)

with respect to a frame induced by natural coordinates. Here es13λ~ie^{s^{\frac{1}{3}}\tilde{\lambda}^{i}} and es13μ~ie^{s^{\frac{1}{3}}\tilde{\mu}^{i}} are the largest eigenvalues of D(δ~i)1D(\tilde{\delta}_{i})^{-1} and D(α~i)1D(\tilde{\alpha}_{i})^{-1} respectively.

Remark 5.10.

The holonomy with respect to the global frame \mathcal{F} will only differ by multiplication on the left and on the right by the change of frame between the global coordinate on 2\mathbb{H}^{2} and the natural coordinate for qsq_{s}. These, however, only grow polynomially in ss, so they do not influence the estimates that follow.

We can then write

Hol(c~γ)=A(s)+E(s)\mathrm{Hol}(\tilde{c}_{\gamma})=A(s)+E(s)

by setting

A(s)=i=1SD(δi)1U(θi,θi+1)1D(αi)1S1.A(s)=\prod_{i=1}^{\ell}SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}. (5.3)

It is worth emphasizing two aspects of the definition above of A(s)A(s). First, the definition makes no reference to the radius ϵ\epsilon of the balls BiB_{i} cut out around the zeroes. Moreover, the formula for A(s)A(s) allows for an extension of the formula for A(s)A(s), first stated for the modified path cγ~\tilde{c_{\gamma}} to the q0q_{0}-geodesic cγc_{\gamma}. These points will be essential for computing final holonomy formulas, e.g. in Theorem A, which do not depend on ϵ\epsilon or other constructions around the zeroes.

Lemma 5.11.

Along any regular geodesic, the highest order term in (5.3) is

A(s)=i=1cji,kies13(μkii+λjii)SEji,ki(Id+o(Id))S1,A(s)=\prod_{i=1}^{\ell}c_{j_{i},k_{i}}e^{s^{\frac{1}{3}}({\mu}^{i}_{k_{i}}+{\lambda}^{i}_{j_{i}})}SE_{j_{i},k_{i}}({\rm Id}+o(\mathrm{Id}))S^{-1}, (5.4)

where Ej,kE_{j,k} denotes the elementary matrix with 1 in position (j,k)(j,k) and cji,kic_{j_{i},k_{i}} is a non-zero constant. Here μkii=μi\mu^{i}_{k_{i}}=\mu^{i} and λjii=λi\lambda^{i}_{j_{i}}=\lambda^{i} are s13s^{-\frac{1}{3}} times the logarithms of the largest eigenvalues of D(αi)1D(\alpha_{i})^{-1} and D(δi)1D(\delta_{i})^{-1} respectively.

Proof.

This is a consequence of Proposition 6.3 below, whose precise statement we defer until later, as it depends on terminology developed in Section 6. This proposition shows that the element in position (ji,ki)(j_{i},k_{i}) of U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} is non-zero. We then factor each term in the triple product in Equation (5.3). ∎

Remark 5.12.

Two consecutive terms in the above product have the property that Eji,kiEji+1,ki+1=Eji,ki+1E_{j_{i},k_{i}}\cdot E_{j_{i+1},k_{i+1}}=E_{j_{i},k_{i+1}} (since ki=ji+1k_{i}=j_{i+1}) because the position of the highest eigenvalue only depends on the angle the path makes with the xx-axis in a natural coordinate and our choices of coordinates keep this angle constant when the coordinate patches cover the same straight path. Note, indeed, that αi{\alpha}_{i} and δi+1{\delta}_{i+1} (with indices intended modulo \ell) are part of the same straight line segment ci+1c_{i+1}.

We also give an argument to address the special case in which the the angle of a geodesic segment is in a Stokes direction, extending Lemma 5.11 to the case when Proposition 5.9 does not hold. Define c~i=δ~iα~i1\tilde{c}_{i}=\tilde{\delta}_{i}\cup\tilde{\alpha}_{i-1} to be the geodesic segment in c~γ\tilde{c}_{\gamma} corresponding to this geodesic arc. Then the estimates of [DW15] fail to hold at its endpoints c~iβi\tilde{c}_{i}\cap\beta_{i} and c~iβi1\tilde{c}_{i}\cap\beta_{i-1}. We will modify c~i,βi,βi1\tilde{c}_{i},\beta_{i},\beta_{i-1} slightly by moving the endpoints. Recall ci=δiαi1c_{i}=\delta_{i}\cup\alpha_{i-1} is the corresponding geodesic segment between the zeros in γ\gamma.

We rewrite (5.3) as

A(s)=SD(α)[i=1D(ci)1U(θi,θi+1)1]D(α)1S1.A(s)=SD(\alpha_{\ell})\left[\prod_{i=1}^{\ell}D(c_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}\right]D(\alpha_{\ell})^{-1}S^{-1}. (5.5)
Proposition 5.13.

Lemma 5.11 holds for homotopy classes of free loops for which the flat geodesic’s saddle connection segments are all either regular or travel along Stokes rays: in other words, if no saddle connection is contained in a wall of a Weyl chamber.

Proof.

It suffices to address the case of a single saddle connection along a Stokes ray. Each endpoint of the c~i\tilde{c}_{i} is in the flat coordinate ϵeiθ\epsilon e^{i\theta} for θ\theta a Stokes direction. We modify the angle by ±η\pm\eta for a small positive constant η\eta to avoid these directions. So define βiη\beta_{i}^{\eta} to be the new arc formed by replacing the endpoint ϵeiθ\epsilon e^{i\theta} by ϵei(θ±η)\epsilon e^{i(\theta\pm\eta)}, and similarly define βi1η\beta_{i-1}^{\eta}. Define c~iη\tilde{c}_{i}^{\eta} to be the geodesic path between these endpoints of βiη\beta_{i}^{\eta} and βi1η\beta_{i-1}^{\eta}. By choosing η\eta small enough we can ensure the straight line homotopy between c~i\tilde{c}_{i} and c~iη\tilde{c}_{i}^{\eta} does not cross any other zeros of the cubic differential.

Refer to caption
Figure 2. Path modified by η\eta.

In certain cases we also need to specify the signs ±η\pm\eta. At each zero along the geodesic γ\gamma, the incoming and outgoing rays must make an angle of π\geq\pi with respect the flat metric, when measured in clockwise and counterclockwise directions around the zero. Proposition 6.3 below requires that each arc begins and ends away from a Stokes ray and must subtend an angle >π>\pi (and so at least 3 Stokes rays will be transversed by the arc). For each endpoint of c~i\tilde{c}_{i} choose ±η\pm\eta so that the arcs βiη\beta_{i}^{\eta} and βi1η\beta_{i-1}^{\eta} both subtend an angle >π>\pi. This is possible since the total angle around a zero of order kk is 2π+2πk/32\pi+2\pi k/3. Note that in some cases we are free to choose either +η+\eta or η-\eta; then there is a different holonomy matrix along the arc depending on the sign. Lemma 6.5 below shows that this matrix leaves the relevant entries unchanged, in terms of the leading order terms.

Define ciηc_{i}^{\eta} to be the union of c~iη\tilde{c}_{i}^{\eta} and the two radial paths from the zeros pi1,pip_{i-1},p_{i} to the endpoints of c~iη\tilde{c}_{i}^{\eta}. See Figure 2. Now by Proposition 5.6 and Remark 5.8, the contribution for ciηc_{i}^{\eta} in (5.5) is given by

Sdiag(es13ν1i,η,es13ν2i,η,es13ν3i,η)S1+o(es13νi,η)S{\rm diag}(e^{s^{\frac{1}{3}}\nu_{1}^{i,\eta}},e^{s^{\frac{1}{3}}\nu_{2}^{i,\eta}},e^{s^{\frac{1}{3}}\nu_{3}^{i,\eta}})S^{-1}+o(e^{s^{\frac{1}{3}}{\nu}^{i,\eta}})

where

νji,η=223e(ciηϕj)=223e(ciϕj)=νji=μji1+λji,\nu_{j}^{i,\eta}=-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{c_{i}^{\eta}}\phi_{j}\right)=-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{c_{i}}\phi_{j}\right)=\nu_{j}^{i}=\mu_{j}^{i-1}+\lambda_{j}^{i},

since ϕj\phi_{j} is closed and ciηc_{i}^{\eta} is homotopic to cic_{i}. This shows as above that the conclusion of Lemma 5.11 holds.

Note we call the entire modified path c~η\tilde{c}^{\eta}. It is obtained from c~\tilde{c} by replacing, for each appropriate ii, c~i\tilde{c}_{i} by c~iη\tilde{c}_{i}^{\eta} and βi\beta_{i} by βiη\beta_{i}^{\eta}, etc. ∎

6. No branching

In this section we exploit the geometry of the convex regular polygon to which a hyperbolic affine sphere with cubic differential zkdz3z^{k}dz^{3} project in order to analyze the non-zero entries of the unipotent matrices U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} of the previous section. The key result, Proposition 6.3, asserts that the unipotents in the holonomy formula (5) – that connect the holonomy of segments that come into a zero with the holonomy of segments that leave a zero – have an entry that allows the largest eigenvalues of those holonomies to multiply. This is crucial for the form of the formula (5).

Proposition 6.3 below will be used later to show that the induced map from the Riemann surface to the real building given by the asymptotic cone is locally injective near the zeros of the cubic differential, and thus can have no branching behavior. The corresponding phenomenon in the real tree case is called “folding,” which does occur in some situations (e.g. [DDW00], [Wol07]).

Choose a local coordinate ww on {0}\mathbb{C}\setminus\{0\} so that q=dw3q=\,dw^{3}. Consider ff the corresponding embedding of the Riemann surface into 3\mathbb{R}^{3} whose image is a hyperbolic affine sphere with Pick differential qq, and consider the frame F=(f,fw,fw¯)F=(f,f_{w},f_{\bar{w}}) of the affine sphere. Let fTf_{T}, FTF_{T} be the corresponding embedding and frame for a standard Ţiţeica surface (as described in Example 2.1). The osculation map G(z)=F(z)FT1(z)G(z)=F(z)F_{T}^{-1}(z) then has limits SLS1,SL0S1,SL+S1SL_{-}S^{-1},SL_{0}S^{-1},SL_{+}S^{-1} along rays γ(t)=teiθ\gamma(t)=te^{i\theta} of angle θ\theta for θ(π/2,π/6),\theta\in(-\pi/2,-\pi/6), (π/6,π/6),(π/6,π/2)(-\pi/6,\pi/6),(\pi/6,\pi/2) respectively ([DW15]). These matrices determine the construction of the convex polygon PP onto which the affine sphere f()f(\mathbb{C}) projects. Let us summarize the main step of the contruction. We label the vertices of PP as r0,r1,,rn12r_{0},r_{1},\dots,r_{n-1}\in\mathbb{RP}^{2} with indices in n\mathbb{Z}_{n}. Let rjrj+1¯\overline{r_{j}r_{j+1}} denote the line connecting rjr_{j} and rj+1r_{j+1}, and let eje_{j} denote the edge of PP from rjr_{j} to rj+1r_{j+1}. Three successive vertices rj1,rj,rj+1r_{j-1},r_{j},r_{j+1} form an inscribed triangle in PP around the vertex rjr_{j}. Also define points qj=rj1rj¯rjrj+1rj+2¯q_{j}=\overline{r_{j-1}r_{j}}\cap\overline{r_{j}r_{j+1}r_{j+2}} so that rj,qj,rj+1r_{j},q_{j},r_{j+1} are the vertices of a circumscribed triangle TjT_{j} of PP centered around the edge eje_{j}. See Figure 3.

Refer to caption
Figure 3. Inscribed and circumscribed triangles. Here each inscribed triangle includes a dotted edge and each circumscribed triangle contains one edge of the polygon and extends the immediate neighbors of that edge to meet a point qjq_{j} exterior to the polygon.
Proposition 6.1.

In the above setting the following holds:

  1. (1)

    rjriri+1¯r_{j}\in\overline{r_{i}r_{i+1}} if and only if j=ij=i or j=i+1j=i+1.

  2. (2)

    qjriri+1¯q_{j}\in\overline{r_{i}r_{i+1}} if and only if j=i1j=i-1 or j=i+1j=i+1.

Proof.

Statement (1) is obvious from the convexity of PP.

To prove statement (2), note we need only prove the "only if" part. So assume qjriri+1¯q_{j}\in\overline{r_{i}r_{i+1}}. As qi,ri,ri+1q_{i},r_{i},r_{i+1} are the vertices of a triangle, we see jij\neq i. To find a contradiction, assume j<i1j<i-1 or j>i+1j>i+1 and that qjriri+1¯q_{j}\in\overline{r_{i}r_{i+1}}. Recall qj=rj1rj¯rj+1rj+2¯q_{j}=\overline{r_{j-1}r_{j}}\cap\overline{r_{j+1}r_{j+2}}. By convexity, TjPT_{j}\supset P. Since PP is a convex polygon, it is the intersection of nn closed half-planes HkH_{k}, each bounded by rkrk+1¯ek\overline{r_{k}r_{k+1}}\supset e_{k}. Since qjriri+1¯q_{j}\in\overline{r_{i}r_{i+1}}, we see PP is a subset of the smaller triangle TjHiT_{j}\cap H_{i}, which cannot contain both ej1e_{j-1} and ej+1e_{j+1}. This is a contradiction. ∎

The columns of the matrices LL_{-}, L+L_{+} project to the vertices of a circumscribed triangle of PP, whereas the columns of L0L_{0} project to the vertices of an inscribed triangle. Moreover, the middle vertex of an inscribed triangle is always obtained as L0eiL_{0}e_{i} where SeiSe_{i} is the eigenvector corresponding to the highest eigenvalue of FT(z)F_{T}(z). The unipotent matrices arise then by taking the products L1L0L_{-}^{-1}L_{0} and similar. We determine these unipotent matrices explicitly in the case PP is regular. Choose coordinates in 2\mathbb{RP}^{2} so that

rj=[cos2πjn,sin2πjn,1]t.r_{j}=\left[\cos\frac{2\pi j}{n},\sin\frac{2\pi j}{n},1\right]^{t}.

We then compute that

q0=[12+12sec2πn,12tan2πn,1]t,q_{0}=\left[\frac{1}{2}+\frac{1}{2}\sec\frac{2\pi}{n},\frac{1}{2}\tan\frac{2\pi}{n},1\right]^{t},

while qjq_{j} can be computed as rjq0r_{j}q_{0}, where we view rj,q0r_{j},q_{0}\in\mathbb{C}. In other words, in the natural inhomogeneous coordinate chart {[x,y,1]t}\{[x,y,1]^{t}\} in 2\mathbb{RP}^{2}, qjq_{j} is found by rotating q0q_{0} by an angle of 2πj/n2\pi j/n.

The paper [DW15] provides a scheme of determining the projective transformations of triangles (across Stokes lines) to form the polygon, as well as the order of the largest eigenvalue corresponding to each vertex. See the tables on pages 1771-1772 in [DW15].

(r1r0r1)(q0r0r1)(r2r0r1)(r2q1r1)(r2r3r1)(r2r3q2)ms|sm|sm|ms|sm|ms|msπ60π6π3π22π35π6π7π64π33π25π311π6\begin{array}[]{c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c@{\hspace{-3pt}}c}\mapsto&\left(\begin{array}[]{c}r_{-1}\\ r_{0}\\ r_{1}\end{array}\right)&\mapsto&\left(\begin{array}[]{c}q_{0}\\ r_{0}\\ r_{1}\end{array}\right)&\mapsto&\left(\begin{array}[]{c}r_{2}\\ r_{0}\\ r_{1}\end{array}\right)&\mapsto&\left(\begin{array}[]{c}r_{2}\\ q_{1}\\ r_{1}\end{array}\right)&\mapsto&\left(\begin{array}[]{c}r_{2}\\ r_{3}\\ r_{1}\end{array}\right)&\mapsto&\left(\begin{array}[]{c}r_{2}\\ r_{3}\\ q_{2}\end{array}\right)&\mapsto\\[21.68121pt] \begin{array}[]{c}m\\ \ell\\ s\end{array}&\Bigg{|}&\begin{array}[]{c}s\\ \ell\\ m\end{array}&\Bigg{|}&\begin{array}[]{c}s\\ m\\ \ell\end{array}&\Bigg{|}&\begin{array}[]{c}m\\ s\\ \ell\end{array}&\Bigg{|}&\begin{array}[]{c}\ell\\ s\\ m\end{array}&\Bigg{|}&\begin{array}[]{c}\ell\\ m\\ s\end{array}&\Bigg{|}&\begin{array}[]{c}m\\ \ell\\ s\end{array}\\[21.68121pt] -\frac{\pi}{6}&0&\frac{\pi}{6}&\frac{\pi}{3}&\frac{\pi}{2}&\frac{2\pi}{3}&\frac{5\pi}{6}&\pi&\frac{7\pi}{6}&\frac{4\pi}{3}&\frac{3\pi}{2}&\frac{5\pi}{3}&\frac{11\pi}{6}\end{array} (6.1)

Here each \mapsto refers to crossing a Stokes line, which in standard flat coordinates are at angles π6,π6,π2,5π6,-\frac{\pi}{6},\frac{\pi}{6},\frac{\pi}{2},\frac{5\pi}{6},\dots, while each vertical line refers to crossing a wall of a Weyl chamber, at angles 0,π3,2π3,π,0,\frac{\pi}{3},\frac{2\pi}{3},\pi,\dots. The s,m,s,m,\ell refer to the smallest, medium and largest eigenvalue respectively of the frame FT(z)F_{T}(z).

There is of course more than one projective transformation which takes a triangle to a triangle, but the ones we are interested in are determined by the conditions that they are unipotent and that they fix the vector lifts to 3\mathbb{R}^{3} of the relevant rjr_{j}’s fixed by the projective transformation. The next proposition shows, at least in the regular case, that these vector lifts of rj,qjr_{j},q_{j} can be done globally, and that the relevant linear transformations corresponding to crossing several Stokes lines near a zero can be computed in terms of simple changes of bases (in contrast to a more complicated scheme of multiplying out several transformations).

Proposition 6.2.

Let PP be a regular convex polygon as above. There are vector lifts rj,qj3\vec{r}_{j},\vec{q}_{j}\in\mathbb{R}^{3} of the points rj,qj2r_{j},q_{j}\in\mathbb{RP}^{2} for which the linear transformations lifting the corresponding triangle transformations are all unipotent. The choice of such a set of lifts is unique up to a nonzero multiplicative constant.

Proof.

We may easily check that

rj=(cos2πjn,sin2πjn,1)tandq0=(cos2πn1,sin2πn,2cos2πn)t\vec{r}_{j}=\left(\cos\frac{2\pi j}{n},\sin\frac{2\pi j}{n},1\right)^{t}\quad\mbox{and}\quad\vec{q}_{0}=\left(-\cos\frac{2\pi}{n}-1,-\sin\frac{2\pi}{n},-2\cos\frac{2\pi}{n}\right)^{t}

satisfy the conditions, with qj\vec{q}_{j} again being defined by rotating q0\vec{q}_{0} by an angle of 2πj/n2\pi j/n. ∎

It follows that the matrix U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} represents the change of basis between i={v1,v2,v3}\mathcal{B}_{i}=\{v_{1},v_{2},v_{3}\} and i+1={w1,w2,w3}\mathcal{B}_{i+1}=\{w_{1},w_{2},w_{3}\} where vjv_{j} and wjw_{j} project to the vertices of inscribed or circumscribed triangles of the regular polygon PP. Therefore, if vkiv_{k_{i}} is the vector corresponding to the highest eigenvalue of D(αi)1D({\alpha}_{i})^{-1} and wjiw_{j_{i}} is the vector corresponding to the highest eigenvalue of D(δi)1D({\delta}_{i})^{-1}, the entry (ji,ki)(j_{i},k_{i}) of U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} is nonzero if and only if vkiv_{k_{i}} has a component along wjiw_{j_{i}} in the basis i+1\mathcal{B}_{i+1}. It is important to note, because of the orientations of the paths near a given zero pip_{i}, that the highest eigenvalue of D(αi)1D(\alpha_{i})^{-1} is the highest eigenvalue of FTF_{T}, while the highest eigenvalue of D(δi)1D(\delta_{i})^{-1} is the lowest eigenvalue of FTF_{T}.

As promised above in Lemma 5.11, we now prove

Proposition 6.3.

Consider any convex polygon PP. If the highest eigenvalue of D(δi)1D({\delta}_{i})^{-1} is in position jij_{i} and the highest eigenvalue of D(αi)1D({\alpha}_{i})^{-1} is in position kik_{i}, then the (ji,ki)(j_{i},k_{i})-entry of U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} is not zero.

Proof.

Refer to Figure 3 and Equation 6.1. Because the paths δi{\delta}_{i} and αi{\alpha}_{i} are part of a geodesic for the flat metric |q0|23|q_{0}|^{\frac{2}{3}}, the angle between δi{\delta}_{i} and αi{\alpha}_{i} is at least π\pi (measured in the singular flat metric) at either side. Since Stokes rays are π/3\pi/3 apart, the interval [θi,θi+1][\theta_{i},\theta_{i+1}] contains at least three Stokes directions. Crossing a Stokes direction corresponds to “flipping” the initial triangle (i.e., moving to the next triangle in the sequence displayed in (6.1), accomplished geometrically between triangles that share an edge). We assume that the initial basis {v1,v2,v3}\{v_{1},v_{2},v_{3}\} projects to a circumscribed triangle and thus is of the form {rj,qj,rj+1}\{\vec{r}_{j},\vec{q}_{j},\vec{r}_{j+1}\}. In this case the eigenvector corresponding to the highest eigenvalue can project to either rjr_{j} or rj+1r_{j+1}: it projects to rjr_{j} if the direction θi\theta_{i} is in the first half of the interval determined by two Stokes directions and to rj+1r_{j+1} otherwise. We will explain the argument in detail when the highest eigenvalue is in the direction of rj\vec{r}_{j}; the case of rj+1\vec{r}_{j+1} or when the initial basis projects to an inscribed triangle are analogous and left to the reader. If the direction of the highest eigenvalue is rj\vec{r}_{j}, then after at least three flips in counter-clockwise direction, by Proposition 6.1, the point rjr_{j} does not lie in any of the lines generated by the final triangle. Hence all entries in the first column of U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} are non-zero. If we move in clockwise direction instead, then after at least five flips the point rir_{i} never lies on any of the lines generated by the vertices of the final triangle and the claim follows as before. Thus, we only need to check what happens when only three or four Stokes directions are contained in the interval [θi,θi+1][\theta_{i},\theta_{i+1}].

If the final triangle is obtained after four flips, then it is circumscribed and the vertex corresponding to the highest eigenvalue of D(δi)1D({\delta_{i}})^{-1} is qj2q_{j-2}, i.e. the one vertex outside the polygon. This is because, referencing Equation 6.1 again, we see both that the highest eigenvalue of D(δi)1D({\delta_{i}})^{-1} is the lowest eigenvalue of FT(z)F_{T}(z) and that the corresponding eigenvector always projects to the vertex exterior to the polygon. Because rjrj1qj2¯r_{j}\in\overline{r_{j-1}q_{j-2}} (see for example Figure 3), the coordinates of rj\vec{r}_{j} with respect the final basis {rj1,qj2,rj2}\{\vec{r}_{j-1},\vec{q}_{j-2},\vec{r}_{j-2}\} have a nonzero component along qj2\vec{q}_{j-2}, hence the corresponding entry in U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} is nonzero. Finally, if the final basis is obtained after only three flips, it is easy to see that the highest eigenvalues of D(αi+1)1D({\alpha_{i+1}})^{-1} and D(δi)1D({\delta_{i}})^{-1} are always in the same position: this follows again by a careful reading of equation (6.1). (We add some details. Recall that we chose the initial point to be in the first half of the region between Stokes lines, i.e. in the region between the Stokes line and the Weyl chamber. For example, we are focused on a point in the region between 3π2\frac{3\pi}{2} and 5π3\frac{5\pi}{3}; after three flips, this point lands in the region between π2\frac{\pi}{2} and 2π3\frac{2\pi}{3}. By inspection of the corresponding bases in (6.1), we see that the initial and terminal basis element, r2r_{2}, corresponding to the highest eigenvalue is unchanged, so the corresponding entry in the unipotent is non-vanishing.) In general the chain in Equation 6.1 then shows that the triple (rj,qj,rj+1)(r_{j},q_{j},r_{j+1}) is sent to (rj,rj1,rj2)(r_{j},r_{j-1},r_{j-2}), so the elements on the diagonal are all non-zero.

Finally, we note that in the argument above, all the conditions checked involve only incidences of the given points and lines in 2\mathbb{RP}^{2}, and not the choices of vector lifts in 3\mathbb{R}^{3}. Thus the proposition holds not only for regular polygons, but for all convex polygons. ∎

Below in Proposition 6.8, we extend Proposition 6.3 to determine the signs of all the relevant entries. These signs will be useful in handling the special cases in which the angle of a geodesic segment in the flat metric is at a Stokes line or wall of a Weyl chamber.

We recall Cramer’s Rule from linear algebra:

Lemma 6.4.

If vv is a vector in m\mathbb{R}^{m}, and w1,,wmw_{1},\dots,w_{m} is a basis, then v=αλαwαv=\sum_{\alpha}\lambda^{\alpha}w_{\alpha}, where

λα=det(w1,,wα1,v,wα+1,,wm)det(w1,,wm)𝒟v𝒟.\lambda^{\alpha}=\frac{\det(w_{1},\dots,w_{\alpha-1},v,w_{\alpha+1},\dots,w_{m})}{\det(w_{1},\dots,w_{m})}\eqqcolon\frac{\mathcal{D}_{v}}{\mathcal{D}}.

Assume for now that the polygon is regular, so that all the rj,qjr_{j},q_{j} lift to vectors rj,qj\vec{r}_{j},\vec{q}_{j} as in Proposition 6.2. It is useful to set up some notation for the following results. We let v=rkv=\vec{r}_{k} be the eigenvector with the largest eigenvalue of D(αi~)1D(\tilde{\alpha_{i}})^{-1} for the incoming ray αi~\tilde{\alpha_{i}}. The frame for the outgoing ray is {wα}\{w_{\alpha}\}, and w{wα}w\in\{w_{\alpha}\} is the one with the largest eigenvalue for D(δi~)1D(\tilde{\delta_{i}})^{-1}. We can write {wα}{w}={rj,rj+1}\{w_{\alpha}\}\setminus\{w\}=\{\vec{r}_{j},\vec{r}_{j+1}\} for some jj.

Lemma 6.5.

𝒟v\mathcal{D}_{v} is unchanged if the incoming or outgoing angles vary by crossing a single Stokes line.

Proof.

Refer to equation (6.1). If the incoming angle moves across a single Stokes ray, the largest eigenvector vv of FTF_{T} is unchanged: recall that in equation (6.1), each Stokes ray, denoted by \mapsto, bisects a Weyl chamber region, whose boundaries are denoted by vertical lines. If the outgoing angle is changed by crossing a single Stokes line, then the smallest eigenvector ww of FTF_{T} is the only vector to change in the frame {wi}\{w_{i}\}. Upon replacing ww with vv, we see 𝒟v\mathcal{D}_{v} (and indeed the underlying matrix) is unchanged. ∎

Lemma 6.6.

𝒟v=det(rj,rj+1,rk)\mathcal{D}_{v}=\det(\vec{r}_{j},\vec{r}_{j+1},\vec{r}_{k}) for some j,k/nj,k\in\mathbb{Z}/n. 𝒟v>0\mathcal{D}_{v}>0 if kj,j+1k\neq j,j+1, and is 0 otherwise.

Proof.

From (6.1) we see vv, as the eigenvector of the largest eigenvalue for an incoming ray, is of the form v=rkv=\vec{r}_{k} for some kk. Similarly, upon replacing w{wi}w\in\{w_{i}\} with vv, we see, upon perhaps performing a cyclic permutation, that 𝒟v=det(rj,rj+1,rk)\mathcal{D}_{v}=\det(\vec{r}_{j},\vec{r}_{j+1},\vec{r}_{k}) for some ii.

If k=j,j+1k=j,j+1, it is obvious that 𝒟v=0\mathcal{D}_{v}=0. On the other hand, if kj,j+1k\neq j,j+1, then rj,rj+1,rkr_{j},r_{j+1},r_{k} form a counterclockwise-oriented triangle in the plane. By Proposition 6.2, rj=(rj,1)3\vec{r}_{j}=(r_{j},1)\in\mathbb{R}^{3}. So this orientation of the triangle implies 𝒟v=det(rj,rj+1,rk)>0\mathcal{D}_{v}=\det(\vec{r}_{j},\vec{r}_{j+1},\vec{r}_{k})>0. ∎

Lemma 6.7.

𝒟>0\mathcal{D}>0.

Proof.

In the case of an inscribed triangle, 𝒟=det(rj,rj+1,rj+2)\mathcal{D}=\det(\vec{r}_{j},\vec{r}_{j+1},\vec{r}_{j+2}) for some jj. Thus as in the previous lemma, 𝒟>0\mathcal{D}>0. To pass from an inscribed to a circumscribed triangle, the frame is changed by a unipotent transformation, which leaves the determinant 𝒟\mathcal{D} unchanged. ∎

Proposition 6.8.

Let PP be any convex polygon. Given the notation of Proposition 6.3, the (j,k)(j,k) entry of U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} is positive.

Proof.

Proposition 6.3 shows the relevant entry is nonzero. We have checked its positivity for the particular case of regular polygons. Then the general result follows by continuity and the connectedness of the moduli space of convex nn-gons. ∎

7. Main Theorem - asymptotics of singular values

In this section, we prove the asymptotics of singular values of the holonomy associated to geodesic paths. The singular values naturally give the distance in the symmetric space SL(3,)/SO(3)\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3). First, in Theorem 7.1, we focus on the case of regular geodesic paths (allowing Stokes directions). In these cases, we have shown there is always a unique largest element in each diagonal matrix D(ci)1D(c_{i})^{-1} in (5.5), while the relevant elements linking them together in U(θi,θi+1)1U(\theta_{i},\theta_{i+1})^{-1} are positive. Later, in Theorem 7.5, we remove the remaining restrictions to allow any geodesic path without any restriction on the angles of the segments. Finally in Corollary 7.6 we extend the analysis beyond just singular values to include eigenvalues.

Theorem 7.1.

Using the same notation as in Lemma 5.11, for every regular path c~γ\tilde{c}_{\gamma}, and indeed for every path none of whose segments is contained in a wall of a Weyl chamber, we have

lims+logHols(c~γ)s13=i=1νi,\lim_{s\to+\infty}\frac{\log\|\mathrm{Hol}_{s}(\tilde{c}_{\gamma})\|}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\nu^{i},

where \|\cdot\| is any submultiplicative matrix norm and νi\nu^{i} is the largest of the νji=223eciϕj\nu_{j}^{i}=-2^{\frac{2}{3}}\mathcal{R}e\int_{c_{i}}\phi_{j}.

Proof.

Recall from Section 5 that Hols(c~γ)=A(s)+E(s)\mathrm{Hol}_{s}(\tilde{c}_{\gamma})=A(s)+E(s). Assume for now that E(s)=o(A(s))\|E(s)\|=o(\|A(s)\|); we return to check this assumption after using it to conclude the theorem. Then

logHols(c~γ)\displaystyle\log\|\mathrm{Hol}_{s}(\tilde{c}_{\gamma})\| =\displaystyle= logA(s)+E(s)\displaystyle\log\|A(s)+E(s)\|
=\displaystyle= logA(s)+logA(s)A(s)+E(s)A(s)\displaystyle\log\|A(s)\|+\log\left\|\frac{A(s)}{\|A(s)\|}+\frac{E(s)}{\|A(s)\|}\right\|
=\displaystyle= logi=1es13(μkii+λjii)i=1cji,kiSEji,ki(Id+o(Id))S1\displaystyle\log\prod_{i=1}^{\ell}e^{s^{\frac{1}{3}}({\mu}^{i}_{k_{i}}+{\lambda}^{i}_{j_{i}})}\left\|\prod_{i=1}^{\ell}c_{j_{i},k_{i}}SE_{j_{i},k_{i}}({\rm Id}+o(\mathrm{Id}))S^{-1}\right\|
+logA(s)A(s)+E(s)A(s),\displaystyle{}+\log\left\|\frac{A(s)}{\|A(s)\|}+\frac{E(s)}{\|A(s)\|}\right\|,

where Eji,kiE_{j_{i},k_{i}} denotes the (ji,ki)(j_{i},k_{i}) elementary matrix. Now,

A(s)A(s)+E(s)A(s)N,with N=1, as E(s)A(s)0\frac{A(s)}{\|A(s)\|}+\frac{E(s)}{\|A(s)\|}\to N,\qquad\mbox{with }\|N\|=1,\,\mbox{ as }\,\frac{E(s)}{\|A(s)\|}\to 0

and

i=1cji,kiSEji,ki(Id+o(Id))S1=M+o(Id),M=S(i=1cji,kiEji,ki)S10.\prod_{i=1}^{\ell}c_{j_{i},k_{i}}SE_{j_{i},k_{i}}({\rm Id}+o(\mathrm{Id}))S^{-1}=M+o(\mathrm{Id}),\quad M=S\left(\prod_{i=1}^{\ell}c_{j_{i},k_{i}}E_{j_{i},k_{i}}\right)S^{-1}\neq 0.

Hence,

lims+logHols(c~γ)s13\displaystyle\lim_{s\to+\infty}\frac{\log\|\mathrm{Hol}_{s}(\tilde{c}_{\gamma})\|}{s^{\frac{1}{3}}} =\displaystyle= lims+[(i=1μkii+λjii)+logM+o(Id)s13\displaystyle\lim_{s\to+\infty}\Bigg{[}\left(\sum_{i=1}^{\ell}{\mu}^{i}_{k_{i}}+{\lambda}^{i}_{j_{i}}\right)+\frac{\log\|M+o(\mathrm{Id})\|}{s^{\frac{1}{3}}}
+1s13logA(s)A(s)+E(s)A(s)]\displaystyle{}+\frac{1}{s^{\frac{1}{3}}}\log\left\|\frac{A(s)}{\|A(s)\|}+\frac{E(s)}{\|A(s)\|}\right\|\Bigg{]}
=\displaystyle= i=1μkii+λjii=i=1μki1i1+λjii=i=1νi,\displaystyle\sum_{i=1}^{\ell}{\mu}^{i}_{k_{i}}+{\lambda}^{i}_{j_{i}}=\sum_{i=1}^{\ell}\mu^{i-1}_{k_{i-1}}+\lambda^{i}_{j_{i}}=\sum_{i=1}^{\ell}\nu^{i},

where as usual i1i-1 is considered modulo \ell.

We only need to check the condition on the error: E(s)=o(A(s))\|E(s)\|=o(\|A(s)\|). For this estimate we are going to use the LL^{\infty}-norm \|\cdot\|_{\infty}. Note, however, that because all matrix norms are equivalent the result holds for any matrix norm. Recall that E(s)E(s) is equal to

i=1(SD(δ~i)1S1+o(es13λ~i))SD(δiδ~i)1S1(SU(θi,θi+1)1S1+o(Id))SD(αiα~i)1S1(SD(α~i)1S1+o(es13μ~i))i=1SD(δi)1U(θi,θi+1)1D(αi)1S1.\begin{split}\prod_{i=1}^{\ell}\,&\left(SD(\tilde{\delta}_{i})^{-1}S^{-1}+o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{i}})\right)\cdot{}\\ &SD(\delta_{i}\setminus\tilde{\delta}_{i})^{-1}S^{-1}(SU(\theta_{i},\theta_{i+1})^{-1}S^{-1}+o(\mathrm{Id}))SD(\alpha_{i}\setminus\tilde{\alpha}_{i})^{-1}S^{-1}\cdot{}\\ &\left(SD(\tilde{\alpha}_{i})^{-1}S^{-1}+o(e^{s^{\frac{1}{3}}{\tilde{\mu}}^{i}})\right)\\ {}-\prod_{i=1}^{\ell}\,&SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}.\end{split}

Hence E(s)E(s) is a sum of terms in which at least one of the factors contains o(Id)o(\mathrm{Id}), o(es13μ~i)o(e^{s^{\frac{1}{3}}{\tilde{\mu}}^{i}}), or o(es13λ~i)o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{i}}). For instance, in the case where o(Id)o(\mathrm{Id}) arises once, we will have a term of the form

Em(s)=i=1m1SD(δi)1U(θi,θi+1)1D(αi)1S1SD(δm)1S1o(Id)SD(αm)1S1i=m+1SD(δi)1U(θi,θi+1)1D(αi)1S1\begin{split}E_{m}(s)=\prod_{i=1}^{m-1}&SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}\cdot{}\\ &SD(\delta_{m})^{-1}S^{-1}o(\mathrm{Id})SD(\alpha_{m})^{-1}S^{-1}\cdot{}\\ \prod_{i=m+1}^{\ell}&SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}\\ \end{split}

and so

Em(s)Co(1)i=1D(δi)1D(αi)1Co(1)i=1es13νi=o(A(s)).\|E_{m}(s)\|_{\infty}\leq C\,o(1)\prod_{i=1}^{\ell}\|D(\delta_{i})^{-1}\|_{\infty}\|D(\alpha_{i})^{-1}\|_{\infty}\leq C\,o(1)\prod_{i=1}^{\ell}e^{s^{\frac{1}{3}}\nu^{i}}=o(\|A(s)\|_{\infty}).

In this case, of course, we combined diagonal matrices in a convenient way. Yet even when the diagonal matrices do not simplify, we may obtain the same estimate. For instance, consider

Em(s)=i=1m1SD(δi)1U(θi,θi+1)1D(αi)1S1o(es13λ~m)SD(δmδ~m)1U(θm,θm+1)1D(αm)1S1i=m+1SD(δi)1U(θi,θi+1)1D(αi)1S1.\begin{split}E_{m}^{\prime}(s)=\prod_{i=1}^{m-1}\,&SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}\cdot{}\\ &o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{m}})SD(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}U(\theta_{m},\theta_{m+1})^{-1}D(\alpha_{m})^{-1}S^{-1}\\ \prod_{i=m+1}^{\ell}\,&SD(\delta_{i})^{-1}U(\theta_{i},\theta_{i+1})^{-1}D(\alpha_{i})^{-1}S^{-1}.\\ \end{split}

Then

Em(s)\displaystyle\|E_{m}^{\prime}(s)\|_{\infty} \displaystyle\leq Co(es13λ~m)D(δmδ~m)1D(αm)1\displaystyle C\,o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{m}})\|D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}\|_{\infty}\|D(\alpha_{m})^{-1}\|_{\infty}\cdot
i=1imD(δi)1D(αi)1\displaystyle\mathop{\prod_{i=1}^{\ell}}_{i\neq m}\|D(\delta_{i})^{-1}\|_{\infty}\|D(\alpha_{i})^{-1}\|_{\infty}
=\displaystyle= Co(es13λ~m)i=1imes13(μi+λi)D(δmδ~m)1D(αm)1\displaystyle C\,o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{m}})\mathop{\prod_{i=1}^{\ell}}_{i\neq m}e^{s^{\frac{1}{3}}(\mu^{i}+\lambda^{i})}\|D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}\|_{\infty}\|D(\alpha_{m})^{-1}\|_{\infty}

using Lemma 5.11, where, we recall that es13λme^{s^{\frac{1}{3}}\lambda^{m}} is the largest eigenvalue of D(δm)1D(\delta_{m})^{-1}. This last expression will be o(A(s))o(\|A(s)\|_{\infty}) if we can show that o(es13λ~m)D(δmδ~m)1=o(es13λm)o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{m}})\|D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}\|_{\infty}=o(e^{s^{\frac{1}{3}}{\lambda}^{m}}). Note that D(αm)1=es13μm\|D(\alpha_{m})^{-1}\|_{\infty}=e^{s^{\frac{1}{3}}{\mu}_{m}}. Now, if δ~m(t)=s13teiθm\tilde{\delta}_{m}(t)=s^{\frac{1}{3}}te^{i\theta_{m}} with t[ϵ,Lm/2]t\in[\epsilon,L_{m}/2], then δm(t)=s13teiθm\delta_{m}(t)=s^{\frac{1}{3}}te^{i\theta_{m}} with t[0,Lm/2]t\in[0,L_{m}/2] and

D(δ~m)1\displaystyle D(\tilde{\delta}_{m})^{-1} =\displaystyle= exp(s13(ϵLm/2)(cos(θm)cos(θm2π3)cos(θm4π3))),\displaystyle\exp\left(s^{\frac{1}{3}}(\epsilon-L_{m}/2)\left(\begin{array}[]{ccc}\cos(\theta_{m})&&\\ &\cos(\theta_{m}-\frac{2\pi}{3})&\\ &&\cos(\theta_{m}-\frac{4\pi}{3})\end{array}\right)\right),
D(δm)1\displaystyle D(\delta_{m})^{-1} =\displaystyle= exp(s13Lm/2(cos(θm)cos(θm2π3)cos(θm4π3))),\displaystyle\exp\left(-s^{\frac{1}{3}}L_{m}/2\left(\begin{array}[]{ccc}\cos(\theta_{m})&&\\ &\cos(\theta_{m}-\frac{2\pi}{3})&\\ &&\cos(\theta_{m}-\frac{4\pi}{3})\end{array}\right)\right),
D(δmδ~m)1\displaystyle D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1} =\displaystyle= exp(s13ϵ(cos(θm)cos(θm2π3)cos(θm4π3))).\displaystyle\exp\left(-s^{\frac{1}{3}}\epsilon\left(\begin{array}[]{ccc}\cos(\theta_{m})&&\\ &\cos(\theta_{m}-\frac{2\pi}{3})&\\ &&\cos(\theta_{m}-\frac{4\pi}{3})\end{array}\right)\right).

So if the largest eigenvalue of D(δ~m)1D(\tilde{\delta}_{m})^{-1} is in position jj, then

D(δ~m)1\displaystyle\|D(\tilde{\delta}_{m})^{-1}\|_{\infty} =\displaystyle= es13(ϵLm/2)cos(θm(j1)2π/3)=es13λ~m,\displaystyle e^{s^{\frac{1}{3}}(\epsilon-L_{m}/2)\cos(\theta_{m}-(j-1)2\pi/3)}=e^{s^{\frac{1}{3}}\tilde{\lambda}^{m}},
D(δmδ~m)1\displaystyle\|D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}\|_{\infty} =\displaystyle= es13ϵcos(θm(j1)2π/3),=es13(λmλ~m)\displaystyle e^{-s^{\frac{1}{3}}\epsilon\cos(\theta_{m}-(j-1)2\pi/3)},=e^{s^{\frac{1}{3}}(\lambda^{m}-\tilde{\lambda}^{m})}
D(δm)1\displaystyle\|D(\delta_{m})^{-1}\|_{\infty} =\displaystyle= es13(Lm/2)cos(θm(j1)2π/3)=es13λm,\displaystyle e^{-s^{\frac{1}{3}}(L_{m}/2)\cos(\theta_{m}-(j-1)2\pi/3)}=e^{s^{\frac{1}{3}}\lambda^{m}},

Thus, o(es13λ~m)D(δmδ~m)1=o(es13λm)o(e^{s^{\frac{1}{3}}{\tilde{\lambda}}^{m}})\|D(\delta_{m}\setminus\tilde{\delta}_{m})^{-1}\|_{\infty}=o(e^{s^{\frac{1}{3}}{\lambda}^{m}}), as required.

The remaining cases are analogous, involving smaller error terms. ∎

In particular, if we choose the submultiplicative matrix norm

M2:=σ1(M)\|M\|_{2}:=\sigma_{1}(M)

where σ1(M)\sigma_{1}(M) denotes the highest singular value of MM, in other words the highest eigenvalue of MtM\sqrt{M^{t}M}, we obtain

Corollary 7.2.

For every regular path cγc_{\gamma} that is the concatenation of saddle connections c,,c1{c}_{\ell},\dots,{c}_{1} we have

lims+log(σ1(Hols(cγ)))s13=i=1llims+log(σ1(Hols(ci)))s13.\lim_{s\to+\infty}\frac{\log(\sigma_{1}(\mathrm{Hol}_{s}(c_{\gamma})))}{s^{\frac{1}{3}}}=\sum_{i=1}^{l}\lim_{s\to+\infty}\frac{\log(\sigma_{1}(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}\ .
Proof.

Because the paths cγc_{\gamma} and c~γ\tilde{c}_{\gamma} have the same holonomy, by definition of μkii{\mu}^{i}_{k_{i}} and λjii{\lambda}^{i}_{j_{i}}, Proposition 5.6 and the fact that the saddle connection ci{c}_{i} is the concatenation of δi{\delta}_{i} and αi1{\alpha}_{i-1} (see Figure 1), we have

lims+log(σ1(Hols(ci)))s13=μkii+λji1i1=νi,\lim_{s\to+\infty}\frac{\log(\sigma_{1}(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}={\mu}^{i}_{k_{i}}+{\lambda}^{i-1}_{j_{i-1}}=\nu^{i},

where the indices are to be intended modulo \ell. Thus the result follows from Theorem 7.1 applied to the norm 2\|\cdot\|_{2}. ∎

We can also deduce the asymptotics of the other singular values

Corollary 7.3.

Let σj\sigma_{j} denote the jj-th largest singular value. Then

lims+log(σj(Hols(cγ)))s13=i=1lims+log(σj(Hols(ci)))s13\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}(c_{\gamma})))}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}

for j=1,2,3j=1,2,3.

Proof.

We already know that the result holds for σ1(Hols(cγ))\sigma_{1}(\mathrm{Hol}_{s}(c_{\gamma})) by Corollary 7.2. Because

log(σ3(Hols(cγ)))=log(σ1(Hols1(cγ)))=log(σ1(Hols(cγ1))),\log(\sigma_{3}(\mathrm{Hol}_{s}(c_{\gamma})))=-\log(\sigma_{1}(\mathrm{Hol}_{s}^{-1}(c_{\gamma})))=\log(\sigma_{1}(\mathrm{Hol}_{s}(c_{\gamma}^{-1})))\ ,

the statement is also true for σ3(Hols(cγ))\sigma_{3}(\mathrm{Hol}_{s}(c_{\gamma})) by applying the previous corollary to the path cγ1c_{\gamma}^{-1}. Moreover, since Hols(cγ)SL(3,)\mathrm{Hol}_{s}(c_{\gamma})\in\mathrm{SL}(3,\mathbb{R}), we have

log(σ1(Hols(cγ)))+log(σ2(Hols(cγ)))+log(σ3(Hols(cγ)))=0,\log(\sigma_{1}(\mathrm{Hol}_{s}(c_{\gamma})))+\log(\sigma_{2}(\mathrm{Hol}_{s}(c_{\gamma})))+\log(\sigma_{3}(\mathrm{Hol}_{s}(c_{\gamma})))=0\ ,

hence the result holds for σ2(Hols(cγ))\sigma_{2}(\mathrm{Hol}_{s}(c_{\gamma})) as well. ∎

We now give an argument extending the above results to the case of flat geodesics that are not regular, in that they contain segments in the wall directions of a Weyl chamber. We begin with the generic situation where each flat geodesic segment has corresponding diagonal holonomy with a unique largest eigenvalue. In that case, as above, we compute the largest asymptotic singular value and then reverse the direction of the path to find the smallest. This determines the asymptotic eigenvalue structure. In particular, the arguments above already suffice to determine the largest asymptotic singular value along any geodesic path in which each segment has a unique largest eigenvalue. We summarize the discussion in the following proposition.

Proposition 7.4.

Theorem 7.1 and Corollary 7.2 hold for flat geodesic paths all of whose segments ci{c}_{i} are such that each diagonal matrix D(ci)1D({c}_{i})^{-1} has a distinct largest eigenvalue.

Now we address the remaining case where paths may contain segments so that at least one D(ci)1D({c}_{i})^{-1} has two largest eigenvalues. In this case, Lemma 5.11 becomes

A(s)=i=1es13(μjii+λkii)S(αcjiα,kiαEjiα,kiα)(Id+o(Id))S1,A(s)=\prod_{i=1}^{\ell}e^{s^{\frac{1}{3}}(\mu^{i}_{j_{i}}+\lambda^{i}_{k_{i}})}S\Big{(}\sum_{\alpha}c_{j_{i_{\alpha}},k_{i_{\alpha}}}E_{j_{i_{\alpha}},k_{i_{\alpha}}}\Big{)}({\rm Id}+o(\mathrm{Id}))S^{-1}, (7.4)

where in each sum α\alpha ranges over one or two indices in {1,2,3}\{1,2,3\}: one if there is a unique largest eigenvalue of D(ci)1D(c_{i})^{-1}, and two if there are two largest eigenvalues. Proposition 6.8 then shows that the coefficients cjiα,kiαc_{j_{i_{\alpha}},k_{i_{\alpha}}} are all positive, and thus there are no cancellations. This is enough for the analysis above on submultiplicative matrix norms and singular values to apply.

Thus we have proved

Theorem 7.5.

For every geodesic path c~γ\tilde{c}_{\gamma}, we have

lims+logHols(c~γ)s13=i=1νi\lim_{s\to+\infty}\frac{\log\|\mathrm{Hol}_{s}(\tilde{c}_{\gamma})\|}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\nu^{i}

where \|\cdot\| is any submultiplicative matrix norm.
In particular, if σj(Hols(cγ))\sigma_{j}(\mathrm{Hol}_{s}(c_{\gamma})) denotes the jj-th largest singular value of the flat geodesic cγc_{\gamma} homotopic to c~γ\tilde{c}_{\gamma} with saddle connections c1,,cl{c}_{1},\dots,{c}_{l}, then

lims+log(σj(Hols(cγ)))s13=i=1lims+log(σj(Hols(ci)))s13\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}(c_{\gamma})))}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\lim_{s\to+\infty}\frac{\log(\sigma_{j}(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}

for j=1,2,3j=1,2,3.

Because Hols(cγ)\mathrm{Hol}_{s}(c_{\gamma}) is diagonalizable with positive eigenvalues, Theorem 7.5 also implies a similar asymptotic formula for the eigenvalues of Hols(cγ)\mathrm{Hol}_{s}(c_{\gamma}).

Corollary 7.6.

For every closed curve γπ1(S)\gamma\in\pi_{1}(S), let cγc_{\gamma} be the geodesic representative for the flat metric |q0|23|q_{0}|^{\frac{2}{3}}. Assume that cγc_{\gamma} is the concatenation of saddle connections c1,,cl{c}_{1},\dots,{c}_{l}. Then

lims+log(Λ(Hols(γ)))s13=i=1lims+log(Λ(Hols(ci)))s13=i=1νi\lim_{s\to+\infty}\frac{\log(\Lambda(\mathrm{Hol}_{s}(\gamma)))}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\lim_{s\to+\infty}\frac{\log(\Lambda(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\nu^{i}

where Λ(M)\Lambda(M) denotes the spectral radius of MM.
In particular, Corollary 7.3 holds when replacing singular values with eigenvalues.

Proof.

It is well known that the spectral radius, i.e. the absolute values of the largest (possibly complex) eigenvalue of a matrix MM, can be computed as

Λ(M)=limr+Mr1r.\Lambda(M)=\lim_{r\to+\infty}\|M^{r}\|^{\frac{1}{r}}\ .

Since Hols(c~γ)\mathrm{Hol}_{s}(\tilde{c}_{\gamma}) has all real and positive eigenvalues, its spectral radius coincides with its largest eigenvalue. Moreover, because cγc_{\gamma} and c~γ\tilde{c}_{\gamma} are in the same free homotopy class, we can do this computation for Hols(c~γ)\mathrm{Hol}_{s}(\tilde{c}_{\gamma}). Now, we know that

Hols(c~γ)=A(s)+E(s)\mathrm{Hol}_{s}(\tilde{c}_{\gamma})=A(s)+E(s)

with E(s)=o(A(s))\|E(s)\|=o(\|A(s)\|) as s+s\to+\infty and

A(s)=i=1es13(μjii+λkii)S(αcjiα,kiαEjiα,kiα)(Id+o(Id))S1A(s)=\prod_{i=1}^{\ell}e^{s^{\frac{1}{3}}(\mu^{i}_{j_{i}}+\lambda^{i}_{k_{i}})}S\Big{(}\sum_{\alpha}c_{j_{i_{\alpha}},k_{i_{\alpha}}}E_{j_{i_{\alpha}},k_{i_{\alpha}}}\Big{)}({\rm Id}+o(\mathrm{Id}))S^{-1} (7.5)

for some positive constants cjiα,kiαc_{j_{i_{\alpha}},k_{i_{\alpha}}}. Fix δ>0\delta>0 small and let s0s_{0} be such that for all ss0s\geq s_{0} we have E(s)δA(s)\|E(s)\|\leq\delta\|A(s)\|. Then, for all ss0s\geq s_{0}

Hols(c~γ)r=A(s)r+E(s)\mathrm{Hol}_{s}(\tilde{c}_{\gamma})^{r}=A(s)^{r}+E^{\prime}(s)

with E(s)2rδA(s)r\|E^{\prime}(s)\|\leq 2r\delta\|A(s)\|^{r} for every integer r>1r>1. Therefore,

Λ(Hols(c~γ))\displaystyle\Lambda(\mathrm{Hol}_{s}(\tilde{c}_{\gamma})) =\displaystyle= limr+Hols(c~γ)r1r\displaystyle\lim_{r\to+\infty}\|\mathrm{Hol}_{s}(\tilde{c}_{\gamma})^{r}\|^{\frac{1}{r}}
=\displaystyle= limr+A(s)r1rA(s)rA(s)r+E(s)A(s)r1r\displaystyle\lim_{r\to+\infty}\|A(s)^{r}\|^{\frac{1}{r}}\left\|\frac{A(s)^{r}}{\|A(s)^{r}\|}+\frac{E^{\prime}(s)}{\|A(s)^{r}\|}\right\|^{\frac{1}{r}}
=\displaystyle= Λ(A(s))C(s)\displaystyle\Lambda(A(s))C(s)

for all ss0s\geq s_{0}, where

C(s)=limr+A(s)rA(s)r+E(s)A(s)r1r.C(s)=\lim_{r\to+\infty}\left\|\frac{A(s)^{r}}{\|A(s)^{r}\|}+\frac{E^{\prime}(s)}{\|A(s)^{r}\|}\right\|^{\frac{1}{r}}\ .

First we compute the spectral radius Λ(A(s))\Lambda(A(s)) of the matrix A(s)A(s). From (7.5), we find

A(s)r\displaystyle A(s)^{r} =\displaystyle= exp(i=1rs13νi)(i=1S(αcjiα,kiαEjiα,kiα)(Id+o(Id))S1)r\displaystyle\exp\left(\sum_{i=1}^{\ell}rs^{\frac{1}{3}}\nu^{i}\right)\left(\prod_{i=1}^{\ell}S\left(\sum_{\alpha}c_{j_{i_{\alpha}},k_{i_{\alpha}}}E_{j_{i_{\alpha}},k_{i_{\alpha}}}\right)(\mathrm{Id}+o(\mathrm{Id}))S^{-1}\right)^{r}
=\displaystyle= exp(i=1rs13νi)(M+o(Id))r\displaystyle\exp\left(\sum_{i=1}^{\ell}rs^{\frac{1}{3}}\nu^{i}\right)(M+o(\mathrm{Id}))^{r}

as s+s\to+\infty, where

M=i=1S(αcjiα,kiαEjiα,kiα)S10M=\prod_{i=1}^{\ell}S\left(\sum_{\alpha}c_{j_{i_{\alpha}},k_{i_{\alpha}}}E_{j_{i_{\alpha}},k_{i_{\alpha}}}\right)S^{-1}\neq 0

Hence, for ss sufficiently large,

Λ(A(s))=limr+A(s)r1r=exp(i=1s13νi)Λ(M+o(Id)).\Lambda(A(s))=\lim_{r\to+\infty}\|A(s)^{r}\|^{\frac{1}{r}}=\exp\left(\sum_{i=1}^{\ell}s^{\frac{1}{3}}\nu^{i}\right)\Lambda(M+o(\mathrm{Id}))\ .

We then observe that the function C(s)C(s) is uniformly bounded for ss0s\geq s_{0}, because

A(s)rA(s)r+E(s)A(s)r\displaystyle\left\|\frac{A(s)^{r}}{\|A(s)^{r}\|}+\frac{E^{\prime}(s)}{\|A(s)^{r}\|}\right\| \displaystyle\leq 1+E(s)A(s)r\displaystyle 1+\frac{\|E^{\prime}(s)\|}{\|A(s)^{r}\|}
\displaystyle\leq 1+2rδA(s)rA(s)r\displaystyle 1+\frac{2r\delta\|A(s)\|^{r}}{\|A(s)^{r}\|}
\displaystyle\leq (1+2rδ)M+o(Id)r(M+o(Id))r\displaystyle\frac{(1+2r\delta)\|M+o(\mathrm{Id})\|^{r}}{\|(M+o(\mathrm{Id}))^{r}\|}

which implies that C(s)M+o(Id)Λ(M+o(Id))1C(s)\leq\|M+o(\mathrm{Id})\|\Lambda(M+o(\mathrm{Id}))^{-1}. Therefore,

lims+log(Λ(Hols(c~γ)))s13=lims+log(Λ(A(s)))s13=i=1νi.\lim_{s\to+\infty}\frac{\log(\Lambda(\mathrm{Hol}_{s}(\tilde{c}_{\gamma})))}{s^{\frac{1}{3}}}=\lim_{s\to+\infty}\frac{\log(\Lambda(A(s)))}{s^{\frac{1}{3}}}=\sum_{i=1}^{\ell}\nu^{i}\ .

Now, the saddle connection ci{c}_{i} is the concatenation of δi{\delta}_{i} and αi1{\alpha}_{i-1}, where indices are intended modulo \ell (see Figure 1), so by Theorem 4.6,

lims+log(Λ(Hols(ci)))s13=νi,\lim_{s\to+\infty}\frac{\log(\Lambda(\mathrm{Hol}_{s}({c}_{i})))}{s^{\frac{1}{3}}}=\nu^{i}\ ,

which gives the desired asymptotics of the largest eigenvalue. Repeating the same argument as in Corollary 7.3, the formula actually holds for all eigenvalues of Hols(γ)\mathrm{Hol}_{s}(\gamma). ∎

8. Harmonic map to the real building

By work of Hitchin ([Hit92]), the Hitchin representation ρs:π1(S)SL(3,)\rho_{s}:\pi_{1}(S)\rightarrow\mathrm{SL}(3,\mathbb{R}) arising from the ray of cubic differentials qs=sq0q_{s}=sq_{0} are constructed along with an associated ρs\rho_{s}-equivariant conformal harmonic map hs:Σ~SL(3,)/SO(3)h_{s}:\tilde{\Sigma}\rightarrow\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3) to the symmetric space (see Section 2). In this section we study both the asymptotic behavior of hsh_{s} around a zero and also describe the geometry of the limiting harmonic map h:Σ~h_{\infty}:\tilde{\Sigma}\rightarrow\mathcal{B} to an \mathbb{R}-building.

8.1. Generalities on Euclidean buildings

We recall here the definition and main properties of \mathbb{R}-buildings. We direct the interested reader to [KL97] for a more thorough discussion.

Let 𝔸\mathbb{A} denote a finite-dimensional affine Euclidean space. The Tits boundary of 𝔸\mathbb{A} is a sphere, denoted by Tits𝔸\partial_{Tits}\mathbb{A}. A subgroup WaffIsom(𝔸)W_{aff}\subset\mathrm{Isom}(\mathbb{A}) is an affine Weyl group if it is generated by reflections across hyperplanes of 𝔸\mathbb{A}, called walls, and its linear part WlinW_{lin} is finite. The pair (𝔸,Waff)(\mathbb{A},W_{aff}) is a Euclidean Coxeter complex. We denote by Δmod\Delta_{mod} the quotient Tits𝔸/Wlin\partial_{Tits}\mathbb{A}/W_{lin}.

An oriented geodesic ray determines a point in Tits𝔸\partial_{Tits}\mathbb{A}. Its Δmod\Delta_{mod}-direction is its projection to Δmod\Delta_{mod}. A Weyl chamber with tip at p𝔸p\in\mathbb{A} is a complete cone with vertex at pp for which its Tits boundary is a Δmod\Delta_{mod} chamber. A germ of a Weyl chamber based at p𝔸p\in\mathbb{A} is an equivalence class of Weyl chambers based at pp for the following equivalence relation: WW and WW^{\prime} are equivalent if their intersection is a neighborhood of pp in both WW and WW^{\prime}. The germ of a Weyl chamber is denoted by ΔpW\Delta_{p}W. We say that two germs ΔpW\Delta_{p}W and ΔpW\Delta_{p}W^{\prime} are opposite if one is the image of the other under the longest element in the Weyl group WlinW_{lin}. We say that two Weyl chambers based at pp are opposite if their germs are.

Definition 8.1.

A Euclidean \mathbb{R}-building modeled on a Euclidean Coxeter complex (𝔸,Waff)(\mathbb{A},W_{aff}) is a CAT(0)CAT(0) space \mathcal{B} that satisfies the following axioms:

  1. a)

    Each oriented geodesic segment xy¯\overline{xy} is assigned a Δmod\Delta_{mod}-direction θ(xy¯)Δmod\theta(\overline{xy})\in\Delta_{mod}. For any pair of oriented geodesic segments xy¯\overline{xy} and xz¯\overline{xz} emanating from the same point xx\in\mathcal{B}, the difference of their Δmod\Delta_{mod}-directions is smaller than their comparison angle;

  2. b)

    Given δ1,δ2Δmod\delta_{1},\delta_{2}\in\Delta_{mod}, denote by D(δ1,δ2)D(\delta_{1},\delta_{2}) the finite set given by all of the possible distances between points in their WaffW_{aff} orbit. The angle between any two geodesic segments xy¯\overline{xy} and xz¯\overline{xz} lies in the finite set D(θ(xy¯),θ(xz¯))D(\theta(\overline{xy}),\theta(\overline{xz})).

  3. c)

    There is a collection 𝒜\mathcal{A} of isometric embeddings ιA:𝔸\iota_{A}:\mathbb{A}\rightarrow\mathcal{B} that preserve Δmod\Delta_{mod}-directions and that is closed under precomposition by isometries in WaffW_{aff}. Each image A=ιA(𝔸)A=\iota_{A}(\mathbb{A}) is called an apartment of \mathcal{B}. Each geodesic segment, ray and complete geodesic is contained in an apartment.

  4. d)

    Coordinate charts {ιA}A𝒜\{\iota_{A}\}_{A\in\mathcal{A}} are compatible in the sense that, when defined, ιA1ιA21\iota_{A_{1}}\circ\iota_{A_{2}}^{-1} is the restriction of an isometry in the Weyl group WaffW_{aff}.

Remark 8.2.

Many different, though equivalent, sets of axioms of Euclidean \mathbb{R}-buildings appear in the literature. For a detailed discussion we refer the reader to [BS14].

It follows immediately from the axioms that any two points x,yx,y\in\mathcal{B} are contained in a common apartment AA and the distance between them coincides with the Euclidean distance computed inside AA. Moreover, we can define a (germ of a) Weyl chamber in \mathcal{B} as the image of a (germ of a) Weyl chamber in 𝔸\mathbb{A} under some chart ιA\iota_{A}. The following property will be useful.

Proposition 8.3 ([Par12]).

Two opposite Weyl chambers based at pp\in\mathcal{B} are contained in a unique apartment.

The boundary at infinity of \mathcal{B} is defined as the set of equivalence classes of geodesic rays, where two rays are equivalent if they remain at bounded distance. Given any ξ\xi\in\partial_{\infty}\mathcal{B} and pp\in\mathcal{B}, there is a unique geodesic ray ξp\xi_{p} in the equivalence class of ξ\xi starting at pp.

Given a point pp\in\mathcal{B}, and two geodesic segments c1,c2;[0,1]c_{1},c_{2};[0,1]\rightarrow\mathcal{B} such that cj(0)=pc_{j}(0)=p, the angle between them is the quantity

p(c1,c2)=lims,t0~p(c1(s),c2(t))\angle_{p}(c_{1},c_{2})=\lim_{s,t\to 0}\widetilde{\angle}_{p}(c_{1}(s),c_{2}(t))

where ~p\widetilde{\angle}_{p} denotes the angle of the Euclidean comparison triangle. This induces a distance on the set Σp\Sigma_{p}\mathcal{B} of equivalence classes of geodesic segments emanating from pp, where two segments are identified if the angle between them is zero.

8.2. Asymptotic cone of SL(3,)/SO(3)\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3)

We denote by (X,d)(X,d) the symmetric space SL(3,)/SO(3)\mathrm{SL}(3,\mathbb{R})/\mathrm{SO}(3) endowed with the distance induced by its homogeneous Riemannian metric. The construction of the asymptotic cone of (X,d)(X,d) and, more generally, of any metric space relies on the choice of a non-principal ultrafilter.

Definition 8.4.

A non-principal ultrafilter is a finitely additive probability measure ω\omega on 𝒫()\mathcal{P}(\mathbb{R}) such that

  1. (1)

    ω(S){0,1}\omega(S)\in\{0,1\} for every S𝒫()S\in\mathcal{P}(\mathbb{R}) ;

  2. (2)

    ω(S)=0\omega(S)=0 for every finite subset SS.

As we are interested in the behavior as s+s\to+\infty, we only consider non-principal ultrafilters each supported on a countable subset of \mathbb{R} whose only limit point in [,+][-\infty,+\infty] is ++\infty. Non-principal ultafilters allow us to consistently define limits of bounded sequences without passing to subsequences. Precisely, a family of points {ys}s\{y_{s}\}_{s\in\mathbb{R}} in a topological space YY is said to have a ω\omega-limit yy, denoted by y=limωysy=\lim_{\omega}y_{s} if for each neighborhood UU of yy we have ω({s|ysU})=1\omega(\{s\in\mathbb{R}\ |\ y_{s}\in U\})=1.

Definition 8.5.

Let * be a base point in (X,d)(X,d) and let λs\lambda_{s}\to\infty be a sequence of scaling factors. Fix a non-principal ultrafilter ω\omega. The asymptotic cone of (X,λs1d,)(X,\lambda_{s}^{-1}d,*) is the metric space (Coneω(X,λs,),dω)(\mathrm{Cone}_{\omega}(X,\lambda_{s},*),d_{\omega}) where

  1. (1)

    points in Coneω(X,λs,)\mathrm{Cone}_{\omega}(X,\lambda_{s},*) are equivalence classes of families xsXx_{s}\in X such that λs1d(xs,)\lambda_{s}^{-1}d(x_{s},*) is bounded. Here, two families xs,ysx_{s},y_{s} are equivalent if limωλs1d(xs,ys)=0\lim_{\omega}\lambda_{s}^{-1}d(x_{s},y_{s})=0 ;

  2. (2)

    the distance between two points [xs][x_{s}] and [ys][y_{s}] is defined as

    dω([xs],[ys]):=limωλs1d(xs,ys).d_{\omega}([x_{s}],[y_{s}]):=\lim_{\omega}\lambda_{s}^{-1}d(x_{s},y_{s})\ .

By work of [Tho02], the asymptotic cone of the symmetric space (X,d)(X,d) is actually, up to isometries, independent of the choice of the ultrafilter ω\omega (if we assume the continuum hypothesis) and of the base point *. Moreover, the asymptotic cone of the symmetric space (X,d)(X,d) can also be interpreted as the Gromov-Hausdorff limit of the pointed sequence of metric spaces Xs=(X,λs1d,)X_{s}=(X,\lambda_{s}^{-1}d,*) ([Gro81] [KL97]) and it is a non-discrete Euclidean building modelled on the affine Weyl group of 𝔰𝔩(3,)\mathfrak{sl}(3,\mathbb{R}). In particular, we are going to identify the model Euclidean plane 𝔸\mathbb{A} with

𝔸={(x1,x2,x3)3|x1+x2+x3=0}.\mathbb{A}=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}\ |\ x_{1}+x_{2}+x_{3}=0\}\ .

Then the linear part of the Weyl group consists of reflections across the walls of equation xixj=0x_{i}-x_{j}=0 for all iji\neq j. In particular, we identify Δmod\Delta_{mod} with the boundary at infinity of the Weyl chamber

𝔞+={(x1,x2,x3)3|x1>x2>x3}.\mathfrak{a}^{+}=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}\ |\ x_{1}>x_{2}>x_{3}\}\ .

Given x,yConeω(X)x,y\in\mathrm{Cone}_{\omega}(X), we can find an apartment ιA:𝔸Coneω(X)\iota_{A}:\mathbb{A}\rightarrow\mathrm{Cone}_{\omega}(X) such that ιA(0)=x\iota_{A}(0)=x and ιA1(y)𝔞+\iota_{A}^{-1}(y)\in\mathfrak{a}^{+}. Then the distance between xx and yy is

dω(x,y)=d𝔸(0,ιA1(y))=y12+y22+y32,d_{\omega}(x,y)=d_{\mathbb{A}}(0,\iota_{A}^{-1}(y))=\sqrt{y_{1}^{2}+y_{2}^{2}+y_{3}^{2}}\ ,

where yiy_{i} are the coordinates of ιA1(y)\iota_{A}^{-1}(y).

Beside the Euclidean distance on an apartment, it is also useful to consider the 𝔞+\mathfrak{a}^{+}-valued distance defined by

dω𝔞+(x,y)=d𝔸𝔞+(0,ιA1(y))=(y1,y2,y3).d_{\omega}^{\mathfrak{a}^{+}}(x,y)=d_{\mathbb{A}}^{\mathfrak{a}^{+}}(0,\iota_{A}^{-1}(y))=(y_{1},y_{2},y_{3})\ .

By a theorem of Parreau ([Par12]), the 𝔞+\mathfrak{a}^{+}-valued distance on Coneω(X)\mathrm{Cone}_{\omega}(X) is the ω\omega-limit of the analogously defined 𝔞+\mathfrak{a}^{+}-valued distance d𝔞+d^{\mathfrak{a}^{+}} on XX rescaled by λs1\lambda_{s}^{-1}. (Here this distance d𝔞+(x,y)d^{\mathfrak{a}^{+}}(x,y) on XX relies on finding a flat that contains xx and yy.) In other words, if x=[xs],y=[ys]Coneω(X)x=[x_{s}],y=[y_{s}]\in\mathrm{Cone}_{\omega}(X) with xs,ysXx_{s},y_{s}\in X, then

dω𝔞+(x,y)=limωλs1d𝔞+(xs,ys).d_{\omega}^{\mathfrak{a}^{+}}(x,y)=\lim_{\omega}\lambda_{s}^{-1}d^{\mathfrak{a}^{+}}(x_{s},y_{s})\ .

An apartment in Coneω(X)\mathrm{Cone}_{\omega}(X) can be obtained as the ω\omega-limit of a sequence of flats in XX. Precisely, if ιFs:𝔸FsX\iota_{F_{s}}:\mathbb{A}\rightarrow F_{s}\subset X are isometric parametrizations of a sequence of maximal flats FsF_{s} of XX with the property that d(Fs,)=O(λs)d(F_{s},*)=O(\lambda_{s}), then the family ιFs\iota_{F_{s}} has an ω\omega-limit ιF:𝔸Coneω(X)\iota_{F}:\mathbb{A}\rightarrow\mathrm{Cone}_{\omega}(X) which defines an apartment in Coneω(X)\mathrm{Cone}_{\omega}(X). See [KL97] for more details.

A family GsSL(3,)G_{s}\in\mathrm{SL}(3,\mathbb{R}) of isometries of XX also induces an isometric action on Coneω(X)\mathrm{Cone}_{\omega}(X) provided that d(Gs(),)=O(λs)d(G_{s}(*),*)=O(\lambda_{s}) by setting

Gs[xs]:=[Gs(xs)]G_{s}\cdot[x_{s}]:=[G_{s}(x_{s})]

for any [xs]Coneω(X)[x_{s}]\in\mathrm{Cone}_{\omega}(X).

8.3. Limiting harmonic map to the building

Given a ray of cubic differentials sq0sq_{0}, we consider the family of conformal harmonic maps hs:Σ~Xh_{s}:\tilde{\Sigma}\rightarrow X that are equivariant under the corresponding Hitchin representations ρs:π1(Σ)SL(3,)\rho_{s}:\pi_{1}(\Sigma)\rightarrow\mathrm{SL}(3,\mathbb{R}). We fix a non-principal ultrafilter ω\omega, a base point X*\in X and the sequence of scaling factors λs=s13\lambda_{s}=s^{\frac{1}{3}}. We can consider the maps hsh_{s} to take values in the re-scaled metric spaces Xs=(X,s13d)X_{s}=(X,s^{-\frac{1}{3}}d).

Proposition 8.6.

The family hs:Σ~Xsh_{s}:\tilde{\Sigma}\rightarrow X_{s} converges to a Lipschitz equivariant harmonic map h:Σ~Coneω(X)h_{\infty}:\tilde{\Sigma}\rightarrow\mathrm{Cone}_{\omega}(X). The family of holonomy maps ρs:XsXs\rho_{s}\!:X_{s}\to X_{s} ω\omega-converges to an isometry ρ\rho_{\infty} of Coneω(X)\mathrm{Cone}_{\omega}(X), and hh_{\infty} is equivariant with respect to ρ\rho_{\infty}.

Proof.

By [DM06, Theorem 1.2], it is sufficient to show that energy of the maps hs:Σ~Xh_{s}:\tilde{\Sigma}\rightarrow X grows as O(s23)O(s^{\frac{2}{3}}). Since hsh_{s} is conformal, this amounts to estimating the area of hs(D)h_{s}(D), where DΣ~D\subset\tilde{\Sigma} is a compact fundamental domain for the action of π1(S)\pi_{1}(S). Now, the induced metric g^s\hat{g}_{s} on the minimal surfaces hs(Σ~)h_{s}(\tilde{\Sigma}) can be written in terms of the Blaschke metric gsg_{s} as (see [DL19])

g^s=2(1+e3s2)gs,\hat{g}_{s}=2\left(1+\frac{e^{-3\mathcal{F}_{s}}}{2}\right)g_{s}\ , (8.1)

hence g^s\hat{g}_{s} is uniformly bi-Lipshitz to gsg_{s} and the result follows from Lemma 3.2.

The ω\omega-convergence of the holonomy maps follows from Remark 3.19 in [Par12]. The ρ\rho_{\infty}-equivariance of hh_{\infty} follows from the fact that each hsh_{s} is equivariant with respect to ρs\rho_{s}. ∎

The behavior of the limiting harmonic map h:Σ~Coneω(X)h_{\infty}:\tilde{\Sigma}\rightarrow\mathrm{Cone}_{\omega}(X) is well-known outside the zeros of the cubic differential q0q_{0}.

Theorem 8.7.

For any pΣ~p\in\tilde{\Sigma} that is not a lift of a zero of the cubic differential q0q_{0}, there is a neighborhood UpU_{p} centered at pp and an apartment ιA:𝔸Coneω(X)\iota_{A}:\mathbb{A}\rightarrow\mathrm{Cone}_{\omega}(X) with ιA(p)=0\iota_{A}(p)=0 such that

  1. i)

    the induced distance on h(Up)h_{\infty}(U_{p}) is 3216dq0\sqrt{3}\cdot 2^{\frac{1}{6}}d_{q_{0}};

  2. ii)

    the limiting harmonic map hh_{\infty} sends UpU_{p} inside A=ιA(𝔸)A=\iota_{A}(\mathbb{A});

  3. iii)

    for any qUpq\in U_{p} we have

    ιA1h(q)=(223e(pqϕ1),223e(pqϕ2),223e(pqϕ3))\iota_{A}^{-1}\circ h_{\infty}(q)=\left(-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{p}^{q}\phi_{1}\right),-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{p}^{q}\phi_{2}\right),-2^{\frac{2}{3}}\mathcal{R}e\left(\int_{p}^{q}\phi_{3}\right)\right)

    where ϕi\phi_{i} are the cube roots of q0q_{0} (which are well-defined in UpU_{p}).

Proof.

Let UpU_{p} be a q0q_{0}-disk around pp that avoids neighborhoods of zeroes of q0q_{0}. From Equation 8.1 and Lemma 3.4, we know that the induced metric g^s\hat{g}_{s} rescaled by s23s^{-\frac{2}{3}} converges to 3213|q~0|233\cdot 2^{\frac{1}{3}}|\tilde{q}_{0}|^{\frac{2}{3}} uniformly on h(Up)h_{\infty}(U_{p}). To conclude that hh_{\infty} sends UpU_{p} inside a single apartment, it is sufficient to show that h(Up)h_{\infty}(U_{p}) is totally geodesic: this follows from an extendibility feature of flat neighborhoods in buildings (see [AB08, Theorem 11.53]). To this aim we show that for every q,qUpq,q^{\prime}\in U_{p} we have that

dω(h(q),h(q))=3216dq0(q,q).d_{\omega}(h_{\infty}(q),h_{\infty}(q^{\prime}))=\sqrt{3}\cdot 2^{\frac{1}{6}}d_{q_{0}}(q,q^{\prime}). (8.2)

Since the asymptotic cone is a CAT(0)CAT(0) space, this implies that the unique geodesic connecting h(q)h_{\infty}(q) and h(q)h_{\infty}(q^{\prime}) is entirely contained in the image of hh_{\infty}, hence h(Up)h_{\infty}(U_{p}) is totally geodesic inside Coneω(X)\mathrm{Cone}_{\omega}(X). We are thus left to prove Equation 8.2. Fix a natural coordinate ww on UpU_{p} and let w0w_{0} and w0w_{0}^{\prime} be the coordinates of qq and qq^{\prime} in this chart. We then parametrize the geodesic connecting w0w_{0} and w0w_{0}^{\prime} as γ(t)=w0+teiθ\gamma(t)=w_{0}+te^{i\theta} with t[0,L]t\in[0,L] so that w0=w0+Leiθw_{0}^{\prime}=w_{0}+Le^{i\theta}. Recall that the map hh_{\infty} is the ω\omega-limit of the maps hs:Σ~Xsh_{s}:\tilde{\Sigma}\rightarrow X_{s} that can be expressed as

hs(w)=Ps(0)Fs(w)Ps(w)1h_{s}(w)=P_{s}(0)F_{s}(w)P_{s}(w)^{-1}

for some PsSL(3,)P_{s}\in\mathrm{SL}(3,\mathbb{C}). Indeed, from Section 2, we know that the equivariant harmonic map hsh_{s} is simply given by the frame field FsF_{s} of the affine sphere fs:3f_{s}:\mathbb{C}\rightarrow\mathbb{R}^{3} whose columns form at each point ww\in\mathbb{C} a real basis of 3\mathbb{R}^{3} that is orthonormal for (the lift of) the Blaschke metric. The matrices PsP_{s} represent the change of frame between a real othonormal basis and the basis {1,w,w¯}\{1,\partial_{w},\partial_{\bar{w}}\} induced by the natural coordinate ww. Therefore,

dω(h(w0),h(w0))\displaystyle d_{\omega}(h_{\infty}(w_{0}),h_{\infty}(w_{0}^{\prime})) =\displaystyle= lims+s13d(hs(w0),hs(w0))\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(h_{s}(w_{0}),h_{s}(w_{0})^{\prime})
=\displaystyle= lims+s13d(Fs(w0)Ps1(w0),Fs(w0)Ps(w0)1)\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(F_{s}(w_{0})P_{s}^{-1}(w_{0}),F_{s}(w_{0}^{\prime})P_{s}(w_{0}^{\prime})^{-1})
=\displaystyle= lims+s13d(Ps(w0)Fs(w0)1Fs(w0)Ps1(w0),Id)\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(P_{s}(w_{0}^{\prime})F_{s}(w_{0}^{\prime})^{-1}F_{s}(w_{0})P_{s}^{-1}(w_{0}),\mathrm{Id})
=\displaystyle= lims+s13d(Hols(γ),Id)\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(\mathrm{Hol}_{s}(\gamma),\mathrm{Id})

By Proposition 5.6 (see also [Lof07]) and Corollary 7.3 the singular values σj(s)\sigma_{j}(s) of Hols(γ)\mathrm{Hol}_{s}(\gamma) satisfy

lims+s13log(σj(s))=223Lλj\lim_{s\to+\infty}s^{-\frac{1}{3}}\log(\sigma_{j}(s))=-2^{\frac{2}{3}}L\lambda_{j}

where (λ1,λ2,λ3)(\lambda_{1},\lambda_{2},\lambda_{3}) is a reordering of (cos(θ),cos(θ2π/3),cos(θ4π/3))(\cos(\theta),\cos(\theta-2\pi/3),\cos(\theta-4\pi/3)) such that λ1λ2λ3\lambda_{1}\geq\lambda_{2}\geq\lambda_{3}. Therefore, using that the distance dd in the symmetric space XX from the identity is given as the Euclidean distance to the logarithms of the singular values, we see

dω(h(w0),h(w0))\displaystyle d_{\omega}(h_{\infty}(w_{0}),h_{\infty}(w_{0}^{\prime})) =lims+s13j=13log2(σj(s))\displaystyle=\lim_{s\to+\infty}s^{-\frac{1}{3}}\sqrt{\sum_{j=1}^{3}\log^{2}(\sigma_{j}(s))}
=lims+s13j=13s23243L2λj2+o(s23)\displaystyle=\lim_{s\to+\infty}s^{-\frac{1}{3}}\sqrt{\sum_{j=1}^{3}s^{\frac{2}{3}}2^{\frac{4}{3}}L^{2}\lambda_{j}^{2}+o(s^{\frac{2}{3}})}
=223Lj=13cos2(θ2π(j1)3)\displaystyle=2^{\frac{2}{3}}L\sqrt{\sum_{j=1}^{3}\cos^{2}\left(\theta-\frac{2\pi(j-1)}{3}\right)}
=3216L.\displaystyle=\sqrt{3}\cdot 2^{\frac{1}{6}}L\ .

Since L=dq0(q,q)L=d_{q_{0}}(q,q^{\prime}) the proof of Equation 8.2 is complete.
Part iii)iii) is a direct consequence of part ii)ii) and the fact that the coordinates inside an apartment are given by the rescaled limit of the singular values of hs(w)h_{s}(w). ∎

We note, in particular, that outside the zeros of q0q_{0} the map hh_{\infty} is smooth, so the set of its singular points is discrete, and is locally injective. We can also describe how these flats combine outside the zeros.

Proposition 8.8.

Let γ:[0,1]Σ\gamma:[0,1]\rightarrow\Sigma be a geodesic path which avoids all zeros of q0q_{0} and which is not in the direction of a wall of the Weyl chamber. Then there is an apartment AConeω(X)A\subset\mathrm{Cone}_{\omega}(X) such that h(γ(t))Ah_{\infty}(\gamma(t))\in A for all t[0,1]t\in[0,1].

Proof.

Let J={tI|there is apartmentAConeω(X)such thath(γ([0,t]))A}J=\{t\in I\ |\ \text{there is apartment}\ A\subset\mathrm{Cone}_{\omega}(X)\ \text{such that}\ h_{\infty}(\gamma([0,t]))\in A\}. We want to show that J=IJ=I. First, we note that JJ is not empty because there is an apartment containing h(γ(0))h_{\infty}(\gamma(0)) by axiom c)c) in the definition of buildings. Moreover, it is clear that if tJt\in J and sts\leq t then sJs\in J. Let t0=sup(J)t_{0}=\sup(J) and suppose by contradiction that t0<1t_{0}<1. Let q=γ(t0)q=\gamma(t_{0}). By Theorem 8.7, there is a neighborhood UqΣU_{q}\subset\Sigma and an apartment AqA_{q} such that h(Uq)Aqh_{\infty}(U_{q})\subset A_{q}. Up to choosing a smaller UqU_{q}, we can assume that Uqγ(I)=γ(J0)U_{q}\cap\gamma(I)=\gamma(J_{0}) for some open interval J0J_{0} containing t0t_{0}. Let t1J0Jt_{1}\in J_{0}\cap J and let p=γ(t1)p=\gamma(t_{1}). Because t1<t0t_{1}<t_{0}, there is an apartment ApConeω(X)A_{p}\subset\mathrm{Cone}_{\omega}(X) such that h(γ([0,t1]))Aph_{\infty}(\gamma([0,t_{1}]))\subset A_{p}. Let WpW_{p}^{-} denote the Weyl chamber with tip at x=h(p)x=h_{\infty}(p) containing h(γ([0,t1]))h_{\infty}(\gamma([0,t_{1}])). Since h(γ(J0))Aqh_{\infty}(\gamma(J_{0}))\subset A_{q}, we can find two opposite Weyl chambers Wq±W_{q}^{\pm} with tip at xx such that h(γ(t))Wq+h_{\infty}(\gamma(t))\in W_{q}^{+} for tt1t\geq t_{1} and h(γ(t))Wqh_{\infty}(\gamma(t))\in W_{q}^{-} for tt1t\leq t_{1}. Note that the germs of the sectors WpW_{p}^{-} and Wq+W_{q}^{+} are opposite because the Weyl chambers WpW_{p}^{-} and WqW_{q}^{-} are equivalent and Wq+W_{q}^{+} is clearly opposite to WqW_{q}^{-}. Hence, by Proposition 8.3, there is a unique apartment AA\subset\mathcal{B} that contains WpW^{-}_{p} and Wq+W^{+}_{q}. Therefore, we can find t2J0Jt_{2}\in J_{0}\cap J with t2>t0t_{2}>t_{0} such that h(γ([0,t2]))Ah_{\infty}(\gamma([0,t_{2}]))\subset A. Hence t2Jt_{2}\in J contradicting the fact that t0=sup(J)t_{0}=\sup(J). ∎

We intend to complete the description of hh_{\infty} by studying its behavior in a neighborhood of a zero of the cubic differential. Let us fix a coordinate chart around a zero of order kk of q0q_{0} such that q0=zkdz3q_{0}=z^{k}dz^{3} on the ball B={|z|<ϵ}B=\{|z|<\epsilon\}.

Theorem 8.9.

The image h(B)h_{\infty}(B) consists of the union of 2(k+3)2(k+3) (cyclically ordered) bounded, closed sectors {Wi}i=12(k+3)\{W_{i}\}_{i=1}^{2(k+3)} of angle π3\frac{\pi}{3} with tip at x=h(0)x=h_{\infty}(0) such that

  1. i)

    WiWj=W_{i}\cap W_{j}=\emptyset for all ji±1j\neq i\pm 1 (with indices intended modulo 2(k+3)2(k+3));

  2. ii)

    WiWi+1W_{i}\cap W_{i+1} is a geodesic segment.

Proof.

Let ϵ>0\epsilon>0 small and denote by ξi\xi_{i} for i=1,,2(k+3)i=1,\dots,2(k+3) the Stokes directions emanating from 0B0\in B. Let CiC_{i} be the sector in BB with tip at 0 bounded by the directions ξi+ϵ\xi_{i}+\epsilon and ξi+1ϵ\xi_{i+1}-\epsilon. The ball BB can be covered by (k+3)(k+3) standard half-planes obtained from the natural coordinates w=3k+3zk+33w=\frac{3}{k+3}z^{\frac{k+3}{3}}. Note that two such half-planes intersect in a sector of angle π3\frac{\pi}{3} and in these coordinates the Stokes directions correspond to the angles ±π6\pm\frac{\pi}{6} and ±π2\pm\frac{\pi}{2}. By Corollary 4.4, we can write

hs(w)=Ps(0)Ai(Id+o(Id))FT(s13w)Ps(w)1.h_{s}(w)=P_{s}(0)A_{i}(\mathrm{Id}+o(\mathrm{Id}))F_{T}(s^{\frac{1}{3}}w)P_{s}(w)^{-1}\ . (8.4)

Since

FT(w)=Sexp(223e(w)000223e(w/ω)000223e(w/ω2))S1,F_{T}(w)=S\,\exp\left(\begin{array}[]{ccc}2^{\frac{2}{3}}\mathcal{R}e(w)&0&0\\ 0&2^{\frac{2}{3}}\mathcal{R}e(w/\omega)&0\\ 0&0&2^{\frac{2}{3}}\mathcal{R}e(w/\omega^{2})\end{array}\right)\,S^{-1}\ ,

by the same argument as in the proof of Theorem 8.7 the ω\omega-limit hTh_{T} of the map Ps(0)(Id+o(Id))FT(s13w)Ps(w)1P_{s}(0)(\mathrm{Id}+o(\mathrm{Id}))F_{T}(s^{\frac{1}{3}}w)P_{s}(w)^{-1} sends CiC_{i} inside an apartment in Coneω(X)\mathrm{Cone}_{\omega}(X). Moreover, the image of a radial path in CiC_{i} is a geodesic in the building. Indeed, if we parameterize such a path by γ(t)=teiθ\gamma(t)=te^{i\theta} with t[0,L]t\in[0,L] in the natural coordinate ww, then for all t[0,L]t\in[0,L] we have as in (8.3)

dω𝔞+(hT(0)),hT(γ(t))\displaystyle d_{\omega}^{\mathfrak{a}^{+}}(h_{T}(0)),h_{T}(\gamma(t)) =lims+s13d𝔞+(Id,Ps(0)(Id+o(Id))FT(s13teiθ)Ps(w)1)\displaystyle=\lim_{s\to+\infty}s^{-\frac{1}{3}}d^{\mathfrak{a}^{+}}(\mathrm{Id},P_{s}(0)(\mathrm{Id}+o(\mathrm{Id}))F_{T}(s^{\frac{1}{3}}te^{i\theta})P_{s}(w)^{-1})
=t 223(λ1,λ2,λ3)\displaystyle=t\,2^{\frac{2}{3}}(\lambda_{1},\lambda_{2},\lambda_{3}) (8.5)

where (λ1,λ2,λ3)(\lambda_{1},\lambda_{2},\lambda_{3}) is a permutation of (cos(θ),cos(θ2π/3),cos(θ4π/3))(\cos(\theta),\cos(\theta-2\pi/3),\cos(\theta-4\pi/3)) such that λ1λ2λ3\lambda_{1}\geq\lambda_{2}\geq\lambda_{3}. Because this reordering only depends on θ\theta, which is constant along a radial path, and we already know that the image of hTh_{T} is entirely contained in a flat, we deduce that the image hT(γ)h_{T}(\gamma) is a straight segment of length

dω(hT(0),hT(γ(L)))\displaystyle d_{\omega}(h_{T}(0),h_{T}(\gamma(L))) =223Lj=13cos2(θ2(j1)π3)\displaystyle=2^{\frac{2}{3}}L\sqrt{\sum_{j=1}^{3}\cos^{2}\left(\theta-\frac{2(j-1)\pi}{3}\right)}
=3216L\displaystyle=\sqrt{3}\cdot 2^{\frac{1}{6}}L

Since ϵ\epsilon is arbitrary and hh_{\infty} is continuous, we can conclude that the image of each open sector between two consecutive Stokes rays must be contained in a closed sector WiW_{i} with tip at x=h(0)x=h_{\infty}(0) of angle π3\frac{\pi}{3}.
Now, recalling that Ai1Aj=SUi,jS1A_{i}^{-1}A_{j}=SU_{i,j}S^{-1} for some product of unipotents Ui,jU_{i,j}, we can write

h|Cj=Ps(0)AiSUi,jS1hTPs(0)1.{h_{\infty}}_{|_{C_{j}}}=P_{s}(0)A_{i}SU_{i,j}S^{-1}h_{T}P_{s}(0)^{-1}\ .

We first use this to show that the interiors of WiW_{i} and Wi+1W_{i+1} are disjoint. This immediately implies that WiWi+1W_{i}\cap W_{i+1} is a geodesic segment because hh_{\infty} is continuous and each WiW_{i} is circular sector. The previous computation (8.3) about the behavior of radial paths under hh_{\infty} shows that we may have h(wi)=h(wi+1)h_{\infty}(w_{i})=h_{\infty}(w_{i+1}) for some wiCiw_{i}\in C_{i} and wi+1Ci+1w_{i+1}\in C_{i+1} only if they are at the same distance from the zero. However, in this present case we compute from the expression Ai1Aj=SUi,jS1A_{i}^{-1}A_{j}=SU_{i,j}S^{-1} and Equation 8.4,

dω(h(wi),h(wi+1))\displaystyle d_{\omega}(h_{\infty}(w_{i}),h_{\infty}(w_{i+1}))
=\displaystyle= lims+s13d(FT(s13wi)Ps(wi)1,(Id+o(Id))SUi,i+1S1FT(s13wi+1)Ps(wi+1)1)\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(F_{T}(s^{\frac{1}{3}}w_{i})P_{s}(w_{i})^{-1},(\mathrm{Id}+o(\mathrm{Id}))SU_{i,i+1}S^{-1}F_{T}(s^{\frac{1}{3}}w_{i+1})P_{s}(w_{i+1})^{-1})
=\displaystyle= lims+s13d(Ps(wi+1)FT1(s13wi+1)SUi,i+11S1(Id+o(Id))FT(s13wi)Ps(wi)1,Id)\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}d(P_{s}(w_{i+1})F_{T}^{-1}(s^{\frac{1}{3}}w_{i+1})SU_{i,i+1}^{-1}S^{-1}(\mathrm{Id}+o(\mathrm{Id}))F_{T}(s^{\frac{1}{3}}w_{i})P_{s}(w_{i})^{-1},\mathrm{Id})
\displaystyle\geq lims+s13log33Ps(wi+1)FT1(s13wi+1)SUi,i+11S1(Id+o(Id))FT(s13wi)Ps(wi)1\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}\log\left\|\frac{\sqrt{3}}{3}P_{s}(w_{i+1})F_{T}^{-1}(s^{\frac{1}{3}}w_{i+1})SU_{i,i+1}^{-1}S^{-1}(\mathrm{Id}+o(\mathrm{Id}))F_{T}(s^{\frac{1}{3}}w_{i})P_{s}(w_{i})^{-1}\right\|_{\infty}
=\displaystyle= lims+s13logPs(wi+1)SD1(s13wi+1)Ui,i+11(Id+o(Id))D(s13wi)S1Ps(wi)1\displaystyle\lim_{s\to+\infty}s^{-\frac{1}{3}}\log\|P_{s}(w_{i+1})SD^{-1}(s^{\frac{1}{3}}w_{i+1})U_{i,i+1}^{-1}(\mathrm{Id}+o(\mathrm{Id}))D(s^{\frac{1}{3}}w_{i})S^{-1}P_{s}(w_{i})^{-1}\|_{\infty}

where we use again that the distance dd in the symmetric space XX from the identity is given as the Euclidean distance to the logarithms of the singular values. Now, note that in D1(s13wi+1)Ui,i+11D(s13wi)D^{-1}(s^{\frac{1}{3}}w_{i+1})U_{i,i+1}^{-1}D(s^{\frac{1}{3}}w_{i}), at least one element on the diagonal is of the form ecs13e^{cs^{\frac{1}{3}}} for some c>0c>0: here we use that we can express D(s13wi+1)D(s^{\frac{1}{3}}w_{i+1}) and D(s13wi)D(s^{\frac{1}{3}}w_{i}) in a single coordinate, observing that the matrices are distinct, as well as that the unipotent has but a single off-diagonal nonzero entry. Hence,

dω(h(wi),h(wi+1))c>0\displaystyle d_{\omega}(h_{\infty}(w_{i}),h_{\infty}(w_{i+1}))\geq c>0

and h(wi)h(wi+1)h_{\infty}(w_{i})\neq h_{\infty}(w_{i+1}).
The same argument shows that h(Ci)h_{\infty}(C_{i}) and h(Cj)h_{\infty}(C_{j}) are disjoint as long as iji\neq j and the natural coordinates wiw_{i} and wjw_{j} do not coincide. Note that in this case the bound

dω(h(wi),h(wi+1))cd_{\omega}(h_{\infty}(w_{i}),h_{\infty}(w_{i+1}))\geq c

can be made independent of ϵ\epsilon as the diagonal terms in D1(s13wj)D^{-1}(s^{\frac{1}{3}}w_{j}) and D(s13wi)D(s^{\frac{1}{3}}w_{i}) never multiply to 11 in the sectors containing CiC_{i} and CjC_{j}. Hence, in this case WiW_{i} and WjW_{j} are disjoint.
The only case that remains to be checked is when the natural coordinates on CiC_{i} and CjC_{j} are the same, when |ij||i-j| is a multiple of six. This happens when the angle between these sectors for the flat metric |q0|23|q_{0}|^{\frac{2}{3}} is at least 2π2\pi. We can then apply Proposition 6.3 to guarantee that if the highest eigenvalue es13λie^{s^{\frac{1}{3}}\lambda_{i}} of D1(s13wi)D^{-1}(s^{\frac{1}{3}}w_{i}) is in position kik_{i} and the highest eigenvalue es13λje^{s^{\frac{1}{3}}\lambda_{j}} of D(s13wj)D(s^{\frac{1}{3}}w_{j}) is in position kjk_{j}, then the (ki,kj)(k_{i},k_{j})-entry of Ui,j1U_{i,j}^{-1} is not zero. Therefore,

dω(h(wi),h(wj))λi+λj>0.d_{\omega}(h_{\infty}(w_{i}),h_{\infty}(w_{j}))\geq\lambda_{i}+\lambda_{j}>0\ .

Again the bound can be made independent of ϵ\epsilon, hence WiW_{i} and WjW_{j} are disjoint in this case as well. ∎

We are then able to describe the global behavior of the harmonic map hh_{\infty}.

Corollary 8.10.

Let q~0\tilde{q}_{0} denote the lift of the cubic differential q0q_{0} to the universal cover Σ~\tilde{\Sigma}. Let dd_{\infty} be the path distance induced by dωd_{\omega} on h(Σ~)Coneω(X)h_{\infty}(\tilde{\Sigma})\subset\mathrm{Cone}_{\omega}(X). Then h:(Σ~,3216dq~0)(h(Σ~),d)h_{\infty}:(\tilde{\Sigma},\sqrt{3}\cdot 2^{\frac{1}{6}}d_{\tilde{q}_{0}})\rightarrow(h_{\infty}(\tilde{\Sigma}),d_{\infty}) is an isometry.

Proof.

From Equation 8.1 and Lemma 3.4, we know that the induced metric g^s\hat{g}_{s} rescaled by s23s^{-\frac{2}{3}} converges pointwise to 3213|q~0|233\cdot 2^{\frac{1}{3}}|\tilde{q}_{0}|^{\frac{2}{3}} and uniformly on every compact set on the complement of the zeros of q~0\tilde{q}_{0}. Let BB be a ball centered at a zero of order kk as in the setting of Theorem 8.9. Then the induced metric at h(0)h_{\infty}(0) is singular since the total angle is 2π+2kπ32\pi+\frac{2k\pi}{3}. Moreover, in the proof of Theorem 8.9 we showed that radial paths from the origin of q0q_{0}-length LL are sent to geodesic arcs in the building of length 3216L\sqrt{3}\cdot 2^{\frac{1}{6}}L. We conclude that h(B)h_{\infty}(B) is isometric to (B,3213|q0|23)(B,3\cdot 2^{\frac{1}{3}}|q_{0}|^{\frac{2}{3}}) and the statement follows. ∎

Not only is the image of the limiting harmonic map intrinsically a singular flat surface, but h(Σ~)h_{\infty}(\tilde{\Sigma}) inherits from the building a 13\frac{1}{3}-translation surface structure as well, which allows us to reconstruct the original cubic differential q0q_{0}, up to a positive multiplicative constant.

Corollary 8.11.

The image h(Σ~)h_{\infty}(\tilde{\Sigma}) is naturally a 13\frac{1}{3}-translation surface. This structure is induced precisely by the cubic differential 3213q~03\cdot 2^{\frac{1}{3}}\tilde{q}_{0}.

Proof.

By Corollary 8.10, we know that h(Σ~)h_{\infty}(\tilde{\Sigma}) is a singular flat surface with metric 3213|q~0|233\cdot 2^{\frac{1}{3}}|\tilde{q}_{0}|^{\frac{2}{3}}. This defines on h(Σ~)h_{\infty}(\tilde{\Sigma}) a holomorphic cubic differential up to multiplication by eiθe^{i\theta}. On the other hand, a neighborhood of a regular point pp of h(Σ~)h_{\infty}(\tilde{\Sigma}) is contained in an apartment ApA_{p} by Theorem 8.7 and the directions of the walls of the Weyl-chambers based at pp define six special directions. If we identify ApA_{p} with

𝔸={(x1,x2,x3)3|x1+x2+x3=0},\mathbb{A}=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}\ |\ x_{1}+x_{2}+x_{3}=0\}\ ,

then these directions correspond to the lines xixj=0x_{i}-x_{j}=0 for iji\neq j. These can be further divided into two groups, defined in accordance with which pair of the triple of coordinates coincide: if we identify the positive Weyl-chamber with

𝔞+={(x1,x2,x3)3|x1>x2>x3}\mathfrak{a}^{+}=\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}\ |\ x_{1}>x_{2}>x_{3}\}

the two walls are given by x1x2=0x_{1}-x_{2}=0 and x2x3=0x_{2}-x_{3}=0 and the orbits of these lines under the Weyl-group WW divide the six walls into two categories, which we call type I and type II. There is only one choice of θ\theta so that along the directions of type II the differential eiθq~0e^{i\theta}\tilde{q}_{0} is real and positive. ∎

Remark 8.12.

Note that the only scaling factors λs\lambda_{s} that guarantee the existence of the harmonic map hh_{\infty} by Proposition 8.6 are λss13\lambda_{s}\cong s^{\frac{1}{3}} . Different choices of such permissible λs\lambda_{s} lead only to homothetic asymptotic cones, hence the projective class of the translation surface h(Σ~)h_{\infty}(\tilde{\Sigma}) is independent of such a choice of λs\lambda_{s}.

Finally, we relate the geometry of h(Σ~)h_{\infty}(\tilde{\Sigma}) with the notion of weak-convexity introduced by Anne Parreau ([Par21]). In brief, she considered the 𝔞+\mathfrak{a}^{+}-valued distance dω𝔞+d_{\omega}^{\mathfrak{a}^{+}} on Coneω(X)\mathrm{Cone}_{\omega}(X) and defined dω𝔞+d_{\omega}^{\mathfrak{a}^{+}}-geodesics as those paths γ:[0,1]Coneω(X)\gamma:[0,1]\rightarrow\mathrm{Cone}_{\omega}(X) such that for all t[0,1]t\in[0,1]

dω𝔞+(γ(0),γ(1))=dω𝔞+(γ(0),γ(t))+dω𝔞+(γ(t),γ(1))d_{\omega}^{\mathfrak{a}^{+}}(\gamma(0),\gamma(1))=d_{\omega}^{\mathfrak{a}^{+}}(\gamma(0),\gamma(t))+d_{\omega}^{\mathfrak{a}^{+}}(\gamma(t),\gamma(1))

Theorem 7.5 shows that the image under hh_{\infty} of a geodesic for the flat metric |q0|23|q_{0}|^{\frac{2}{3}} is a geodesic for the distance dω𝔞+d_{\omega}^{\mathfrak{a}^{+}}. We deduce the following:

Corollary 8.13.

The surface h(Σ~)Coneω(X)h_{\infty}(\tilde{\Sigma})\subset\mathrm{Cone}_{\omega}(X) is weakly convex.

The argument proves a conjecture of Katzarkov, Noll, Pandit and Simpson ([KNPS17], Conjecture 8.7) in the context of harmonic maps to buildings arising from limits of Hitchin representations in SL(3,)\mathrm{SL}(3,\mathbb{R}) along rays of holomorphic cubic differentials. (Formally, this conjecture concerns the “Finsler” distance, obtained by taking a maximum of the vector-valued distance described above, and asks that the image be geodesic in that distance. Here the geodesic nature holds for the vector-valued distance by Theorem 7.5, and such “geodesics” for the two vector-valued distance are clearly geodesics for the Finsler distance. The statement for the Finsler distance then follows.)

Remark 8.14.

Note that h(Σ~)h_{\infty}(\tilde{\Sigma}) is not totally geodesic in Coneω(X)\mathrm{Cone}_{\omega}(X). One can see this by considering a q~0\tilde{q}_{0}-geodesic arc which passes through a zero of q~0\tilde{q}_{0} and makes an angle of more than π\pi at that zero. The geodesic connecting the endpoints of that arc lies in an apartment containing the endpoints and is necessarily Euclidean, i.e. has no interior point with an angle between incoming and outgoing directions other than π\pi.

9. Epilogue: Triangle groups

We conclude with an example. Of course, for a closed surface XX of high genus, the Labourie-Loftin parametrization of the Hitchin component Hit3\mathrm{Hit}_{3} is given as the cubic differential bundle over the Teichmüller space of XX. Thus, even with the results of this paper, an analysis of the limits of this component would require an understanding of the dependence of a diverging sequence of representations on the rays that include them. On the other hand, for triangle groups, we may completely describe the compactification of the SL(3,)\mathrm{SL}(3,\mathbb{R}) Hitchin component.

9.1. The Hitchin component for triangle groups.

Consider the oriented (p,q,r)(p,q,r) triangle group Γp,q,r\Gamma_{p,q,r}. It is hyperbolic if p1+q1+r1<1p^{-1}+q^{-1}+r^{-1}<1. Denote the quotient of 2\mathbb{H}^{2} by Γ\Gamma by Σp,q,r\Sigma^{p,q,r}. Choi-Goldman have shown that the space of convex real projective structures has real dimension 2 for 3p,q<3\leq p,q<\infty and 4r<4\leq r<\infty [CG93]. We call these triangle groups projectively deformable. The general case of the dimension of the SL(n,)\mathrm{SL}(n,\mathbb{R}) Hitchin component for triangle groups was settled by Long-Thistlethwaite [LT19]. Recently, more closely to our point of view, Alessandrini-Lee-Schaffhauser [ALS22] use Higgs bundles to study Hitchin components for orbifolds. In particular, for each projectively deformable (p,q,r)(p,q,r) group, the space of cubic differentials has complex dimension one, and we can extend our techniques to give a compactification of the Hitchin component in these cases.

We briefly address cubic differentials for oriented triangle groups.

Lemma 9.1.

Let p,q,rp,q,r be integers at least 2. The complex dimension of the space of cubic differentials on Σp,q,r\Sigma^{p,q,r} is 1 if and only if p,q,r3p,q,r\geq 3 and is 0 if any of p,q,rp,q,r is 2.

Proof.

Consider a local coordinate zz with z=0z=0 mapping to an orbifold point of order pp. Then w=zpw=z^{p} is a holomorphic coordinate on the orbifold. The condition for a holomorphic cubic differential near z=0z=0 to descend locally to the orbifold is that it can be written as a holomorphic cubic differential in ww away from w=0w=0. Thus zndz3z^{n}\,dz^{3} descends if and only if n0n\geq 0 and n+3n+3 is a multiple of pp.

The case p>2p>2 and n=p3n=p-3 leads to a pole of order 2 in ww, as zndz3=p3w2dw3z^{n}\,dz^{3}=p^{-3}w^{-2}\,dw^{3}. The Riemann surface formed by treating the 3 orbifold points of Σp,q,r\Sigma^{p,q,r} as smooth points has genus zero and 3 distinguished points at the triangle vertices of order p,q,rp,q,r. Thus we seek cubic differentials on 1\mathbb{CP}^{1} with poles of order at most 2 at 3 points, thus in a family with one complex parameter. Other values of p,np,n require lower order poles (or zeros) at the relevant points. There are no nonzero cubic differentials in these cases, in particular when any of p,q,rp,q,r is 2. ∎

Proposition 9.2.

For 3p,q<3\leq p,q<\infty and 4r<4\leq r<\infty, the real projective structures on Σp,q,r\Sigma^{p,q,r} are parametrized by the complex scalings of the nonzero cubic differential q0q_{0}.

The fundamental domain of an oriented triangle orbifold consists of two adjacent triangles, each with vertices at the p,q,rp,q,r points.

Lemma 9.3.

Let 3p,q,r<3\leq p,q,r<\infty. A singular Euclidean structure on the orbifold Σp,q,r\Sigma^{p,q,r} is induced by a nonzero holomorphic cubic differential if and only if the triangles with vertices p,q,rp,q,r are equilateral.

Proof.

Consider the Euclidean structure given by forcing the given triangles to be equilateral, and let zz be a flat conformal coordinate on one such triangle. Analytically continue the cubic differential dz3dz^{3} to the other triangle of the orbifold. Then we may check the resulting cubic differential is holomorphic on the orbifold, as any monodromy around the orbifold points amounts to translations and rotations by third roots of unity. Such a scalar multiple αdz3\alpha\,dz^{3}, for α\alpha\in\mathbb{C}, is also a holomorphic cubic differential, and together these form the one-dimensional complex vector space of all cubic differentials. ∎

Remark 9.4.

On each of these triangle orbifolds Σp,q,r\Sigma^{p,q,r}, it is useful to choose a particular representative q0q_{0}. Give a consistent orientation to the equilateral triangles forming Σp,q,r\Sigma^{p,q,r}. We choose q0q_{0} so that the sides of the triangles have length 1 and at any given vertex the outgoing edges correspond to the directions on which q0q_{0} is real and positive.

Theorem 9.5.

Consider a family of cubic differentials seiθq0se^{i\theta}q_{0} on a closed surface Σ\Sigma, where s+s\to+\infty and θθ\theta\to\theta_{\infty}. Let ρs,θ\rho_{s,\theta} be the corresponding family of Hitchin representations. Then the limiting data in terms of singular values and eigenvalues converge to those determined in the limit by the ray seiθq0se^{i\theta_{\infty}}q_{0}. In particular, Theorems A and B hold in this setting.

Outline of proof.

Assume for simplicity that θ=0\theta_{\infty}=0. For a given free homotopy class of loops in Σ\Sigma, consider the cubic differential eiθq0=q0e^{i\theta_{\infty}}q_{0}=q_{0} and the piecewise geodesic path c~η\tilde{c}^{\eta} considered in Figure 2, modified from a geodesic path along saddle connections in Stokes directions. We compute the holonomy along c~η\tilde{c}^{\eta} as in Theorems 7.1 and 7.5.

The main consideration is to ensure that we have uniform estimates in θ\theta as ss\to\infty. Note that the Blaschke metric is independent of θ\theta, and the connection form in (2.2) depends on θ\theta only through the cubic differential qq. The upshot is that θ\theta varies in Equation 5.1 and thus in terms of the holonomy along linear paths in Proposition 5.6. It is straightforward to show that the error terms in Proposition 5.6 are uniform in θ\theta as ss\to\infty, as can be seen in terms of the proof in [Lof07]. In other words, the error term of the form o(es13μ~i)o(e^{-s^{\frac{1}{3}}\tilde{\mu}^{i}}) satisfies

o(es13μ~i)es13μ~i0\frac{o(e^{-s^{\frac{1}{3}}\tilde{\mu}^{i}})}{e^{-s^{\frac{1}{3}}\tilde{\mu}^{i}}}\to 0 (9.1)

uniformly in θ\theta as ss\to\infty.

We also check that the estimates of [DW15] are uniform for θ\theta near θ\theta_{\infty} as ss\to\infty. Recall, for the path c~\tilde{c}, in the case in which at least one arc βiη\beta_{i}^{\eta} has a q0q_{0}-Stokes direction as an endpoint, we modify the path to c~η\tilde{c}^{\eta} to avoid all such endpoints. Thus the same is true in a neighborhood of θ\theta_{\infty} as θθ\theta\to\theta_{\infty}. There the estimates of [DW15] are uniform in compact sets away from the Stokes directions, in terms of the frames needed in Corollary 4.4 and Theorem 4.6. The unipotent terms UU are also continuous as θθ\theta\to\theta_{\infty}, as they are determined by the geometry of the limiting (in this case, regular) polygon, which varies continuously.

With these estimates in hand, the quantity we consider is the entry-wise LL^{\infty} norm of the holonomy matrix, as given in (7.4). Our first concern is that the largest terms in each matrix match up with positive terms cj,kc_{j,k} in the adjacent unipotent matrices in the product in (5.3). This remains true by continuity of the cj,kc_{j,k}. If cηc^{\eta} contains any saddle connection cic_{i} along a wall of the Weyl chamber, then there is more than one largest eigenvalue of the holonomy along cic_{i} for q0q_{0}. This is no longer true if θθ\theta\neq\theta_{\infty}. We must consider the possibility then that this entry-wise LL^{\infty} norm may jump by a positive bounded factor, in the case that cj,kαcjα,kαc_{j,k}\to\sum_{\alpha}c_{j_{\alpha},k_{\alpha}} in (7.4). Fortunately the limit we take in Theorem 7.5, in terms of taking a logarithm and dividing by s13s^{\frac{1}{3}}, is insensitive to multiplication by a positive quantity bounded away from 0 and \infty.

Finally, we must ensure that the largest terms in the product of holonomy matrices, upon taking the limit, are not affected by the various error terms we accumulate. This is exactly the uniformity condition (9.1). ∎

Corollary 9.6.

The previous theorem holds for Σ\Sigma a closed oriented orbifold of hyperbolic type.

Proof.

Consider a smooth finite orbifold cover ΣˇΣ\check{\Sigma}\to\Sigma and lift the cubic differential to Σˇ\check{\Sigma}. Then each complex scalar multiple of the cubic differential is also invariant under the orbifold deck transformations. ∎

We next define a compactification of the Hitchin component Hit3(Δ)\mathrm{Hit}_{3}(\Delta) of representations of a deformable triangle group Δ=Γp,q,r\Delta=\Gamma_{p,q,r} into SL(3,)\mathrm{SL}(3,\mathbb{R}).

First we identify the Hitchin component Hit3(Γp,q,r)\mathrm{Hit}_{3}(\Gamma_{p,q,r}) with \mathbb{C} as in Proposition 9.2. We then consider the map

𝔞:\displaystyle\mathcal{L}_{\mathfrak{a}}:\mathbb{C} (𝔞Γp,q,r)\displaystyle\rightarrow\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}})
seiθ\displaystyle se^{i\theta} {(log(σ1(ρs,θ(γ))),log(σ2(ρs,θ(γ))),log(σ3(ρs,θ(γ))))}γΓp,q,r\displaystyle\mapsto\{(\log(\sigma_{1}(\rho_{s,\theta}(\gamma))),\log(\sigma_{2}(\rho_{s,\theta}(\gamma))),\log(\sigma_{3}(\rho_{s,\theta}(\gamma))))\}_{\gamma\in\Gamma_{p,q,r}}

where σj\sigma_{j} denotes the jj-largest singular value and ρs,θ\rho_{s,\theta} is the representation corresponding to seiθq0se^{i\theta}q_{0}. By [Kim04, Theorem B], this map is injective.

Next we also embed S1S^{1} in (𝔞Γp,q,r)\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}) as follows. For each eiθS1e^{i\theta}\in S^{1}, we consider the harmonic map h^θ:(Σ~)Coneω(X)\hat{h}_{\theta}:(\tilde{\Sigma})\to\mathrm{Cone}_{\omega}(X) that results as the ω\omega-limit of the family of harmonic maps hs:Σ~Xh_{s}:\tilde{\Sigma}\rightarrow X parameterized by a ray of cubic differentials seiθq0se^{i\theta}q_{0}. Then for each such map h^θ\hat{h}_{\theta}, we compute the Weyl-chamber lengths of a representative a curve class [γ][\gamma] by considering the image in the principal Weyl chamber 𝔞\mathfrak{a} of h^θ([γ])\hat{h}_{\theta}([\gamma]), following the prescription in Corollary 8.11. Naturally this defines a map S1(𝔞Γp,q,r)S^{1}\to\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}).

This map eiθ(𝔞Γp,q,r)e^{i\theta}\mapsto\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}) is also an embedding: to see this, consider the canonical cubic differential q0q_{0} defined in Remark 9.4, and a fixed curve class γ0\gamma_{0} whose whose |q0||q_{0}|-geodesic representative makes an angle of θ0\theta_{0} with the positive xx-axis in a fixed natural coordinate chart. (Here we take θ0[π3,π3]\theta_{0}\in[-\frac{\pi}{3},\frac{\pi}{3}] so that the largest entry in the vector is cosθ0\cos\theta_{0}.) Then for the “rotated” cubic differential eiθq0e^{i\theta}q_{0}, the largest entry changes to cos(θ0+θ3)\cos(\theta_{0}+\frac{\theta}{3}). We regard S1S^{1} as the boundary of the complex plane \mathbb{C} in the usual way, and assert that the usual compactification S1\mathbb{C}\cup S^{1} is taken homeomorphically to its image in (𝔞Γp,q,r)\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}). This is proved in the next corollary.

Corollary 9.7.

Consider a projectively deformable triangle group. Then the map above S1(𝔞Γp,q,r)\mathbb{C}\cup S^{1}\to\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}) provides a compactification of the SL(3,)\mathrm{SL}(3,\mathbb{R}) Hitchin component Hit3(Γp,q,r)\mathrm{Hit}_{3}(\Gamma_{p,q,r})\cong\mathbb{C} of representations of Γp,q,r\Gamma_{p,q,r}.

Proof.

We need to show that there is a subatlas of boundary charts for the compactification comprising images of a segment TT of the circle S1S^{1} and a sector in \mathbb{C} defined by that range TT of angles; this amounts to showing that for a family seiθse^{i\theta}, as ss\to\infty and θθ\theta\to\theta_{\infty}, the images in (𝔞Γp,q,r)\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}) of the representations associated to seiθse^{i\theta} converge to those of the harmonic map h^θ\hat{h}_{\theta_{\infty}}. This is the content of Theorem 9.5 and that the map S1(𝔞Γp,q,r)S^{1}\to\mathbb{P}(\mathfrak{a}^{\Gamma_{p,q,r}}) defined on the limiting circle S1S^{1} defined by Corollary 8.11 depended only on the limit of the representations associated to the ray seiθse^{i\theta_{\infty}}. ∎

A compactification in terms of harmonic maps is more delicate. Proposition 8.6 associates to a family of harmonic maps hs:Σ~Xh_{s}:\tilde{\Sigma}\rightarrow X parameterized by a ray of cubic differentials seiθq0se^{i\theta}q_{0} a limiting harmonic map h:Σ~Coneω(X)h_{\infty}:\tilde{\Sigma}\rightarrow\mathrm{Cone}_{\omega}(X), whose image (cf. Corollary 8.11) is a 13\frac{1}{3}-translation surface, isometrically embedded into Coneω(X)\mathrm{Cone}_{\omega}(X). However, if more general diverging families of harmonic maps are considered, corresponding to families seiθsq0se^{i\theta_{s}}q_{0} with ss\to\infty and θsθ\theta_{s}\to\theta_{\infty}, their ω\omega-limit h^\hat{h}_{\infty} (which exists by the same argument as in Proposition 8.6) are not precluded from depending on the particular family and not only on θ\theta_{\infty}. We may refer to the more classical Teichmüller theory case, where equivariant harmonic maps us:Σ~2u_{s}:\tilde{\Sigma}\rightarrow\mathbb{H}^{2} are parameterized by a family of quadratic differentials QsQ_{s}. In that case, if Qs=sQ0Q_{s}=sQ_{0} is a ray, usu_{s} always converges to an equivariant harmonic map u:Σ~Tu_{\infty}:\tilde{\Sigma}\rightarrow T to a real tree given by projection onto the leaves of the vertical foliation of Q0Q_{0} ([Wol95]). What is independent of the family is the projective class of the vertical foliation of the Hopf differential of the limiting harmonic map. In our setting, the harmonic map approach identifies the boundary points of a compactification of the Hitchin component for Γp,q,r\Gamma_{p,q,r} with projective classes of 13\frac{1}{3}-translation surfaces.

(The Euclidean (3,3,3)(3,3,3) triangle group also can studied from this point of view. There is no hyperbolic structure on this orbifold, but there is still a nowhere-vanishing cubic differential, which explicitly leads on the orbifold universal cover to the Ţiţeica example. As the cubic differential scales to infinity, the Ţiţeica scales as well, and the limiting harmonic map into the real building simply covers a single apartment.)

Remark 9.8.

We conclude with an informal remark. In the setting of these triangle groups, the limiting harmonic maps to buildings take on enough of a combinatorial nature that we may display how some of the constructions in Section 8 apply.

The simplest triangle group for this is the (3,3,4)(3,3,4) group. One can visualize the action of this group on the hyperbolic plane in terms of a triangulation: two of the vertices of each triangle are vertices with a star of 6 triangles, and the remaining, say special, vertex is a vertex with a star of eight triangles. Distinct special vertices of adjacent triangles share the same opposite edge.

For the cubic differential q0q_{0} defined in Remark 9.4, the harmonic map takes each triangle to an equilateral triangle in a building. In terms of the local geometry of the map, the six triangles in the image of the star around a non-special vertex will lie in a common apartment in the building, but the eight triangles in the image around a special vertex cannot, due for example to cone angle considerations. We return to this local geometry momentarily.

More globally, we comment a bit on some apartments which meet the image of the harmonic map. Note that a geodesic segment between (images of) special vertices in adjacent triangles – this segment will meet the opposite side orthogonally – may be extended to an infinite geodesic which subtends one of the angles, choose it always to be on the right, at each special vertex of exactly π\pi. This infinite (oriented) geodesic is the boundary of a Euclidean strip of parallel geodesics in the image of the harmonic map whose preimages limits on a pair of distinct endpoints. Each of these strips embed isometrically in an apartment which meets the image of the harmonic map in a region that contains the strips. Since there is a geodesic segment between images of special vertices that bisects each triangle in the star of the image of a special vertex, a neighborhood of the image of a special vertex is covered by eight such strips, and two (non-disjoint) strips meet in a rhombus (angles alternately π6\frac{\pi}{6} and π3\frac{\pi}{3}) whose vertices are images of special vertices.

Note that each such strip meets four of the triangles in the star around the image of a special vertex so that the apartment containing this strip meets those four triangles, i.e. those four adjacent triangles around the image of a special vertex are in an embedded flat in the building. On the other hand, five adjacent triangles around that image vertex cannot be in an embedded flat (nor apartment) since that flat would then force there to be a sixth triangle with one edge shared with the terminal triangle and one shared with the initial triangle. That sixth triangle would combine with the three remaining image triangles in the image to form an embedded cone in the building of cone angle 4π3\frac{4\pi}{3}, which cannot exist in an NPC simplicial complex (e.g. points equidistant but on opposite sides of the cone point are joined by distinct geodesics on opposite sides of the cone point).

References

  • [AB08] Peter Abramenko and Kenneth S. Brown. Buildings, volume 248 of Graduate Texts in Mathematics. Springer, New York, 2008. Theory and applications.
  • [ALS22] Daniele Alessandrini, Gye-Seon Lee, and Florent Schaffhauser. Hitchin components for orbifolds. J. European Math. Soc., 2022.
  • [AS64] Milton Abramowitz and Irene A. Stegun. Handbook of mathematical functions with formulas, graphs, and mathematical tables, volume 55 of National Bureau of Standards Applied Mathematics Series. For sale by the Superintendent of Documents, U.S. Government Printing Office, Washington, D.C., 1964.
  • [Bes88] Mladen Bestvina. Degenerations of the hyperbolic space. Duke Math. J., 56(1):143–161, 1988.
  • [BIPP21] Marc Burger, Alessandra Iozzi, Anne Parreau, and Maria Beatrice Pozzetti. The real spectrum compactification of character varieties: characterizations and applications. Comptes Rendus. Mathématique, 359(4):439–463, 2021.
  • [BS14] Curtis D. Bennett and Petra N. Schwer. On axiomatic definitions of non-discrete affine buildings. Adv. Geom., 14(3):381–412, 2014. With an appendix by Koen Struyve.
  • [CG93] Suhyoung Choi and William M. Goldman. Convex real projective structures on closed surfaces are closed. Proc. Amer. Math. Soc., 118(2):657–661, 1993.
  • [CL17] Brian Collier and Qiongling Li. Asymptotics of Higgs bundles in the Hitchin component. Adv. Math., 307:488–558, 2017.
  • [CY77] Shiu Yuen Cheng and Shing Tung Yau. On the regularity of the Monge-Ampère equation det(2u/xisxj)=F(x,u){\rm det}(\partial^{2}u/\partial x_{i}\partial sx_{j})=F(x,u). Comm. Pure Appl. Math., 30(1):41–68, 1977.
  • [CY86] Shiu Yuen Cheng and Shing-Tung Yau. Complete affine hypersurfaces. I. The completeness of affine metrics. Comm. Pure Appl. Math., 39(6):839–866, 1986.
  • [DDW00] G. Daskalopoulos, S. Dostoglou, and R. Wentworth. On the Morgan-Shalen compactification of the SL(2,𝐂){\rm SL}(2,{\bf C}) character varieties of surface groups. Duke Math. J., 101(2):189–207, 2000.
  • [DL19] Song Dai and Qiongling Li. Minimal surfaces for Hitchin representations. J. Differential Geom., 112(1):47–77, 2019.
  • [DM06] Georgios Daskalopoulos and Chikako Mese. Harmonic maps from 2-complexes. Comm. Anal. Geom., 14(3):497–549, 2006.
  • [DW15] David Dumas and Michael Wolf. Polynomial cubic differentials and convex polygons in the projective plane. Geom. Funct. Anal., 25(6):1734–1798, 2015.
  • [FG06] Vladimir Fock and Alexander Goncharov. Moduli spaces of local systems and higher Teichmüller theory. Publ. Math. Inst. Hautes Études Sci., (103):1–211, 2006.
  • [Foc98] V.V. Fock. Dual teichmüller spaces, 1998.
  • [Gol90] William M. Goldman. Convex real projective structures on compact surfaces. J. Differential Geom., 31(3):791–845, 1990.
  • [Gro81] Mikhael Gromov. Groups of polynomial growth and expanding maps. Inst. Hautes Études Sci. Publ. Math., (53):53–73, 1981.
  • [Gui08] Olivier Guichard. Composantes de Hitchin et représentations hyperconvexes de groupes de surface. J. Differential Geom., 80(3):391–431, 2008.
  • [Hit92] N. J. Hitchin. Lie groups and Teichmüller space. Topology, 31(3):449–473, 1992.
  • [Kim04] Inkang Kim. Rigidity on symmetric spaces. Topology, 43(2):393–405, 2004.
  • [KL97] Bruce Kleiner and Bernhard Leeb. Rigidity of quasi-isometries for symmetric spaces and Euclidean buildings. Inst. Hautes Études Sci. Publ. Math., (86):115–197 (1998), 1997.
  • [KNPS15] Ludmil Katzarkov, Alexander Noll, Pranav Pandit, and Carlos Simpson. Harmonic maps to buildings and singular perturbation theory. Comm. Math. Phys., 336(2):853–903, 2015.
  • [KNPS17] Ludmil Katzarkov, Alexander Noll, Pranav Pandit, and Carlos Simpson. Constructing buildings and harmonic maps. pages 203–260, 2017.
  • [Lab06] François Labourie. Anosov flows, surface groups and curves in projective space. Invent. Math., 165(1):51–114, 2006.
  • [Lab07] François Labourie. Flat projective structures on surfaces and cubic holomorphic differentials. Pure Appl. Math. Q., 3(4, Special Issue: In honor of Grigory Margulis. Part 1):1057–1099, 2007.
  • [Lof01] John C. Loftin. Affine spheres and convex n\mathbb{RP}^{n}-manifolds. Amer. J. Math., 123(2):255–274, 2001.
  • [Lof04] John C. Loftin. The compactification of the moduli space of convex 2\mathbb{R}\mathbb{P}^{2} surfaces. I. J. Differential Geom., 68(2):223–276, 2004.
  • [Lof07] John Loftin. Flat metrics, cubic differentials and limits of projective holonomies. Geometriae Dedicata, 128(1):97–106, 2007.
  • [LT19] D. D. Long and M. B. Thistlethwaite. The dimension of the Hitchin component for triangle groups. Geom. Dedicata, 200:363–370, 2019.
  • [LTW22] John Loftin, Andrea Tamburelli, and Michael Wolf. Some uniqueness and harmonic maps title. In preparation, 2022.
  • [Moc16] Takuro Mochizuki. Asymptotic behaviour of certain families of harmonic bundles on Riemann surfaces. J. Topol., 9(4):1021–1073, 2016.
  • [MOT21] Giuseppe Martone, Charles Ouyang, and Andrea Tamburelli. A closed ball compactification of a maximal component via cores of trees. arXiv:2110.06106, 2021.
  • [MSWW16] Rafe Mazzeo, Jan Swoboda, Hartmut Weiss, and Frederik Witt. Ends of the moduli space of Higgs bundles. Duke Math. J., 165(12):2227–2271, 2016.
  • [Nie22] Xin Nie. Limit polygons of convex domains in the projective plane. Int. Math. Res. Not. IMRN, (7):5398–5424, 2022.
  • [OSWW20] Andreas Ott, Jan Swoboda, Richard Wentworth, and Michael Wolf. Higgs bundles, harmonic maps, and pleated surfaces. arXiv:2004.06071, 2020.
  • [OT20] Charles Ouyang and Andrea Tamburelli. Length spectrum compactification of the SO(2,3)-Hitchin component. arXiv:2010.03499,, 2020.
  • [OT21] Charles Ouyang and Andrea Tamburelli. Limits of Blaschke metrics. Duke Math. J., 170(8):1683–1722, 2021.
  • [Par12] Anne Parreau. Compactification d’espaces de représentations de groupes de type fini. Mathematische Zeitschrift, 272(1-2):51–86, 2012.
  • [Par21] Anne Parreau. Invariant subspaces for some surface groups acting on A2-euclidean buildings. To apper in Trans. Amer. Math. Soc., 2021+.
  • [Tho02] Blake Thornton. Asymptotic cones of symmetric spaces. ProQuest LLC, Ann Arbor, MI, 2002. Thesis (Ph.D.)–The University of Utah.
  • [TW19] Andrea Tamburelli and Michael Wolf. Planar minimal surfaces with polynomial growth in the Sp(4,R)-symmetric space. 2019.
  • [Tzi08] M. Georges Tzitzéica. Sur une nouvelle classe de surfaces. Rendiconti del Circolo Matematico di Palermo (1884-1940), 25:180–187, 1908.
  • [Wol89] Michael Wolf. The Teichmüller theory of harmonic maps. J. Differential Geom., 29(2):449–479, 1989.
  • [Wol95] Michael Wolf. Harmonic maps from surfaces to \mathbb{R}-trees. Math. Z., 218(4):577–593, 1995.
  • [Wol07] Michael Wolf. Minimal graphs in H2×H^{2}\times\mathbb{R} and their projections. Pure Appl. Math. Q., 3(3, Special Issue: In honor of Leon Simon. Part 2):881–896, 2007.

JL: Department of Mathematics and Computer Science, Rutgers-Newark
E-mail address: [email protected]


AT: Department of Mathematics, University of Pisa
E-mail address: [email protected]


MW: School of Mathematics, Georgia Institute of Technology
E-mail address: [email protected]