Chapter IV: Maximal surfaces and AdS 3-manifolds
4.3 Maximal surfaces: existence, uniqueness, deformations
With Proposition 4.2.1 in hand, we prove the main theorems. For convenience we assume the twist parameter is zeroâthe proof has no dependence on it. The existence result is immediate from the next proposition.
Proposition 4.3.1. The functionalE =Eð
1, ð
2 : T (Î) â Ris proper if and only if ð1almost strictly dominatesð
2.
Indeed, by Proposition 4.2.1, properness implies the existence of a maximal space- like immersion. Thus, the proof of the existence result reduces to showing that almost strict domination implies properness.
Mapping class groups
Preparing for the proof of Theorem 4A, we review some TeichmÃŒller theory. Set Diff+(Σ) to be the group of ð¶â orientation preserving diffeomorphisms of Σ, equipped with the ð¶â topology. Denote by Diff+0(Σ) the normal subgroup con- sisting of maps that are isotopic to the identity. The mapping class group of Σ is defined
MCG(Σ) :=ð
0(Diff+(ð)) =Diff+(ð)/Diff+0(ð). Given a collection of boundary lengthsâ = (â
1, . . . , âð) â Rð, we can consider the TeichmÃŒller space Tâ(Î) of surfaces with punctures and geodesic boundary with lengths determined by (â
1, . . . , âð). This is not a relative representation space but a union of them. The mapping class group acts on TeichmÃŒller space by pulling back classes of hyperbolic metrics:
[ð] · [ð] âŠâ [ðâð]. (4.13) The action of the mapping class group is properly discontinuous and the quotient is the moduli space of hyperbolic surfaces.
There is a wealth of metrics on TeichmÃŒller space, and the mapping class group acts by isometries on a number of them. In the work below, we set ðT to be any metric distance function on TeichmÃŒller space on which the mapping class group acts by isometries. One example is the TeichmÃŒller distance.
Selecting a basepointð§ âΣ, there is an action of the mapping class group onð
1(Σ, ð§), which identifies with Î. Given ð âDiff+(Σ) such thatð(ð§) = ð§, the induced map ðâ :ð
1(Σ, ð§) â ð
1(Σ, ð§)is an automorphism. While a generalðmay not fixð§, one can find a different ð
1 â Diff+(Σ) that does fixð§and is isotopic to ð. The isotopy gives rise to a pathðŸ
1fromð(ð§)toð§. Ifð
2âDiff+(Σ)also fixesð§and is isotopic to ð, then we get another pathðŸ
2fromð(ð§)toð§. The automorphisms(ð
1)â and(ð
2)â of ð
1(Σ, ð§) differ by the inner automorphism corresponding to the conjugation by the class ofðŸ
1·ðŸ
2. This association furnishes an injective homomorphism from MCG(Σ) âOut(ð
1(Σ)) =Aut(ð
1(Σ))/Inn(ð
1(Σ)).
Through this injective mapping, the mapping class group acts on representation spaces: we can precompose a representative of a representation with a representative of an element in Out(ð
1(Σ)), and then take the corresponding equivalence class.
On a TeichmÃŒller space, this agrees with the action (4.13). For this action, we use the similar notation
[ð] · [ð] âŠâ [ðâð] =[ðâŠðâ].
Note that it does not in general preserve relative representation spaces: a mapping class may permute the punctures. Finally, we record that MCG(Σ)acts equivariantly with respect to
⢠length spectrum: for a representation ð and [ð] â MCG(Σ), â(ðâð(ðŸ)) = â(ð(ðâ(ðŸ))). If ð is a hyperbolic metric we set âð(ðŸ) to be the ð-length of the geodesic representative of ðŸ in (Σ, ð). A restatement of the above is âðâð(ðŸ) =âð(ð(ðŸ)).
⢠Energy of harmonic maps: if ð : (ΣË,ðË) â (ð , ð) is a ð-equivariant map, then
ð(ð, ðð) =ð(ðâð, ððâŠð). (4.14) Furthermore, if ð is harmonic, then ð âŠð : (ΣË,ðËâðË) â (ð , ð) is a ðâð- equivariant harmonic map.
Existence of maximal surfaces
Suppose a simple closed curve ð is either a geodesic boundary component or a horocycle in a hyperbolic surface. If ð is a boundary, letð¶ð(ð) denote the collar aroundð consisting of points with distance toð at mostð. If it is a horocycle, put ð¶ð(ð) to be the union of the ð-collar and the enclosed cusp (both are defined for suitably smallð).
We now begin the proof of properness. Suppose ð
1almost strictly dominates ð
2. We assume there is a single peripheral curveâthere are no substantial changes in the case of many cuspsâand we setðto be either the geodesic boundary component of the convex core ofH/ð
1(Î), or a deep horocycle ofH/ð
1(Î), depending on whether the monodromy is hyperbolic or parabolic. Letðbe an optimal(ð
1, ð
2)-equivariant map. If the image of the peripheral curve under ð
1andð
2is hyperbolic, thenðhas constant speed on the boundary components. Note thatðâŠâðisð
2-equivariant, and in the hyperbolic case, converges atâ to a projection onto the geodesic at infinity in the sense of Definition 4.2.13. Thus, from Lemma 4.2.14 we have
E (ð) =
â«
Σ
ð(ð, âð) âð(ð, ðð)ððŽð â¥
â«
Σ
ð(ð, âð) âð(ð, ðâŠâð)ððŽð. One of the main advantages of replacing ððwithðâŠâð is that the integrand is now non-negative. Moving toward the proof of properness, let ( [ðð])â
ð=1 â T (Î) be such that
E (ðð) â€ðŸ
for all ð, for some large ðŸ. It suffices to show [ðð] converges in T (Î) along a subsequence. Structurally, our proof is similar to that of the classical existence result for minimal surfaces in hyperbolic manifolds (see [SY79] and also the more modern paper [GW07]): we first prove 1) compactness in moduli space, and then 2) extend to compactness in TeichmÃŒller space.
Toward 1), we show there is að¿ > 0 such that for all non-peripheral simple closed curves inÎ, we haveâð
ð(ðŸ) â¥ð¿. We argue by contradiction: suppose this is not the case, so that there is a sequence of non-peripheral simple closed curves(ðŸð)â
ð=1 â Î such that
âð :=âð
ð(ðŸð) â0
as ð â â. By the regular collar lemma, we know that in (Σ, ðð), the geodesic representingðŸðis enclosed by a collarð¶ð of widthð€ð = arcsinh( (sinhâð)â1). Up to a conjugation of the holonomy,ð¶ðis conformally equivalent to{ð¥+ð ðŠ âC: 0†ð¥ †ð€ð,0†ðŠ â€ 1}with the lines{ðŠ=0}and{ðŠ=1} identified.
Let ðð= ðð
ð,âð= âð
ð. Henceforward, conformally modify the metric ððto be flat inð¶ð. Our control over the energy functional depends on our knowledge of the local Lipschitz constant of the map ð. This in turns relies on the image of the harmonic mapâð, in particular how far it takesð¶ðintoð¶ð(ð). For allð¥ ,0< ð¡ < ðin question, set ðŽð¡
ð¥ ={ð â {ð¥} Ãð1: âð(ð¥ , ðŠ) âð¶ð¡(ð)}. We also write Lip(ð, ð¡) = max
ð¥âð¶(H/ð
1(Î))\ð¶ð¡(ð)Lipð¥(ð). We have the inequalities
E (ðð) (ðð) â¥
â«
ð¶ð
ð(ðð, âð) âð(ðð, ðâŠâð)ððŽð
ð
=
â« ð€ð
0
â« 1
0
ð(ðð, âð) âð(ðð, ðâŠâð)ððŠ ðð
⥠ð€ðmin
ð¥
â«
ðŽð¡ð¥
ð(ðð, âð) âð(ðð, ðâŠâð)ððŠ
⥠ð€ðmin
ð¥
(1âLip(ð, ð¡)2)
â«
ðŽð¡ð¥
ð(ðð, âð)ððŠ
⥠ð€ð 2 min
ð¥
(1âLip(ð, ð¡)2)
â«
ðŽð¡ð¥
ð âð(ð¥ , ðŠ)
ð ðŠ
2
ððŠ
⥠ð€ð 2 min
ð¥
(1âLip(ð, ð¡)2) (â(Að¡ð¥))â1 â«
ðŽð¡ð¥
ð âð(ð¥ , ðŠ)
ð ðŠ ððŠ
2
⥠ð€ð 2 min
ð¥
(1âLip(ð, ð¡)2)â«
ðŽð¡ð¥
ð âð(ð¥ , ðŠ)
ð ðŠ ððŠ
2 .
Here â(ðŽð¡
ð¥) is the length of the (possibly broken) segment ðŽð¥ ,ð¡, which is clearly bounded above by 1. Fuchsian representations have discrete length spectrum, and hence there is að > 0 such thatâ(ð
1(ðŸ)) ⥠ð for allðŸ â Î. Lemma 4.3.2. Setð¡ =ð/2. Then for anyð¥, the inequality
â«
ðŽð/ð¥2
|ð âð(ð¥ , ðŠ)
ð ðŠ
ððŠ â¥ min{ð/2, ð } =: ð holds.
Proof. Let ðŒ be any core curve of the form {ð¥} à ð1. If âð(ðŒ) always remains outsideð¶ð/
2(ð), then
â«
ðŽð¡ð¥
|ð âð(ð¥ , ðŠ)
ð ðŠ ððŠ=
â« 1
0
ð âð(ð¥ , ðŠ)
ð ðŠ
ððŠâ¥ â(ð
1(ðŸð)) â¥ð . Thus, we assumeâð(ðŒ) intersectsð¶ð/
2(ð). We parametrize âð(ð¥ ,·)by arc length, so that the integral in question measures the hyperbolic length on H/ð
1(Î) of the segment of âð(ðŒ) that does not enterð¶ð/
2(ð). If this length is ever less thanð/2, then the curveâð(ðŒ)is contained inð¶ð(ð). Phrased differently, it is a simple closed curve contained in an embedded cylinder. There is only one such homotopy class of curves, namely the homotopy class of the boundary geodesic. This situation is impossible, sinceâð(ðŒ)is homotopic to a non-peripheral simple closed curve, and
hence the length is at leastð/2. â¡
Returning to the inequalities above, we now have E (ðð) ⥠ð€ð
2
(1âLip(ð, ð/2)2)ð2.
We thus findE (ðð) â âasðâ â, which violates the uniform upper bound.
The lower bound on the length spectrum shows the metrics (ðð)â
ð=1 satisfy Mum- fordâs compactness criteria [Mum71] and hence project under the action of the mapping class group to a compact subset of the moduli space.
Now we promote to compactness in TeichmÃŒller space. By compactness in moduli space, after passing to a subsequence of theððâs, there exists a sequence of mapping class group representatives (ðð)â
ð=1 such that ðâ
ððð converges as ð â â to a hyperbolic metricðâ.
Lemma 4.3.3. Upon passing to a further subsequence, ( [ðð])â
ð=1 â MCG(Σ) is eventually constant.
The main thrust of the proof is to show that for each non-peripheral ðŸ â Î,â(ð
1⊠ðð(ðŸ))is uniformly bounded above inð. Our proof of this claim is by contradiction and similar in nature to the argument above. Suppose there is a class of curvesðŸ âÎ such thatâ(ð
1âŠðð(ðŸ)) â âasð â â. Modify the notation: setâð =âð
ðâŠððand ðð
ð âŠðð. These are harmonic forðâ
ðð
1andðâ
ðð
2respectively, and have the correct boundary behaviour according to [Sag19, Theorem 1.1]. From (4.14), we also have the equality
Eð
1, ð
2(ðð) =Eððâð
1,ðâðð
2(ðâ
ððð). On each (Σ, ðâ
ððð), there is a collarð¶ð of finite widthð€ð around ðŸ, and theð¶ðâs converge to some collarð¶âin (Σ, ðâ)of widthð€â < â. As before, we perturb to a flat metric and parametrizeð¶ðby [0, ð€ð] Ãð1. We note that ðis (ðâ
ðð
1, ðâ
ðð
2)- equivariant. Therefore, we can apply our previous reasoning to see that forðlarge enough, any small enoughð¥, andð¡ â [0, ð€ð],
Eð
1, ð
2(ðð) =Eðâðð
1,ðâðð
2(ðâ
ððð) ⥠ð€â 4 min
ð¥
(1âLip(ð, ð¡)2)â«
ðŽð¡ð¥
ð âð(ð¥ , ðŠ)
ð ðŠ ððŠ
2 , (4.15) where ðŽð¡
ð¥is defined as above. This leads us to the analogue of Lemma 4.3.2.
Lemma 4.3.4. Setð¡ =ð/2. Then, independent of the choice ofð¥,
â«
ðŽð/ð¥2
ð âð(ð¥ , ðŠ)
ð ðŠ
ððŠ â â asðâ â.
Proof. Choose any simple closed curveðŒof the form{ð¥} Ãð1. Parametrizeâð(ð¥ ,·) by arc length so that the integral above returns the length of the segment ofâð(ðŒ) that does not enterð¶ð/
2(ð). If âð(ðŒ) does not enterð¶ð/
2(ð), then we can argue as we did in the proof of Lemma 4.3.2, bounding the length below by theâ(ðâ
ðð
1(ðŸ)), which blows up. Thus, we restrict our discussion to ðŒ such that âð(ðŒ) intersects ð¶ð/
2(ð). Let us first assume there is a positive integer ðŸ such that every âð(ðŒ) entersð¶ð/
2(ð) at mostðŸ times. Of course, âð(ðŒ) cannot live entirely in ð¶ð/
2(ð), for then it lies in the wrong homotopy class. We construct a new curve as follows:
if we choose a basepoint not withinð¶ð/
2(ð), then every timeâð(ðŒ)entersð¶ð/
2(ð), there is a corresponding point at which it exits. We plan to erase the segment of âð(ðŒ) that connects these two points and replace it with the one of the two possible paths onðð¶ð/
2(ð), say,ðœ
1andðœ
2. Ifâð(ðŒ)spends some time going in a path along the boundary circle when it is entering or when it is exiting (we count just touching
it once as an exit), we connect the endpoint of the entrance path to the starting point of the exiting path and concatenate with the path going along the boundary.
The question is: which path to take? We argue there is a choice so that the homotopy class of the new curve is the same as that ofâð(ðŒ). To this end, orient the boundary circle so that there is a âleftâ and a ârightâ path. Arbitrarily choose one of the two paths connecting the entrance and the exit point, say, ðœ
1, and consider the loop in the cylinder obtained by concatenating this path with the piece ofâð(ðŒ) â©ð¶ð/
2(ð) that connects the endpoints. This is a simple closed curve in a cylinder, and hence there are three possible homotopy classes: trivial, non-trivial and left oriented, and non-trivial and right oriented. If the class is trivial, then using this path ðœ
1 will work. If it is non-trivial and left oriented, then ðœ
1must be the right oriented path.
Thus, if we use the opposite pathðœ
2, it will cancel the homotopy class, and therefore the new path will indeed be homotopic toâð(ðŒ). The right oriented case is similar.
Each replacement adds at most â(ðð¶ð/
2(ð)) to the length of the path. Thus, the length of the new path is (quite crudely) bounded above by
âð/
2(âð(ðŒ)) +ðŸ â(ðð¶ð/
2(ð)). Therefore,
âð/
2(âð(ðŒ)) â¥â(ðâ
ðð
1(ðŸ)) âðŸ â(ðð¶ð/
2(ð))
which tends toâasðâ â, and moreover establishes the result of the lemma.
We are left to consider the case of an unbounded number of crossings intoð¶ð/
2(ð) as we takeð â â. The idea is that each time âð(ðŒ) crosses ðð¶ð/
2(ð), there is a corresponding âdown-crossing,â a curve that connects the point at which it exits to the new point of entry. If each down crossing is denoted byðð
ð, then
â«
ðŽð/ð¥2
ð âð(ð¥ , ðŠ
ð ðŠ
ððŠ â¥ âïž
ð
â(ðð
ð).
Clearly, if the limit of these sums is infinite, then we are done. Hence, we assume the total length due to down-crossings is finite. This implies that there is an even integerðŸ > 0 such that, as we takeð â â, there are at mostðŸ down-crossing that exitð¶ð. Again using our chosen basepoint, letð
1be the first crossing intoð¶ð/
2(ð), and then let ð
2 be the first crossing out ofð¶ð/
2(ð) that exitsð¶ð. Set ð
3 to be the next entry point intoð¶ð/
2(ð), andð
4 the next exit point from ð¶ð(ð). We end up withðŸð †ðŸpointsð
1, . . . ððŸ
ð. Replace the segments ofâð(ðŒ)betweenððandðð+
1
with the correct arc on ðð¶ð/
2(ð) that does not change the homotopy class, where ð=1,3, . . . , ðŸðâ1 †ðŸâ1. As above, we end up with a curve of length at most
âð/
2(âð(ðŒ)) +ðŸ â(ðð¶ð/
2(ð)) and homotopic to â(ðâ
ðð
1(ðŸ)). Using the minimizing property of geodesics once again, we have
âð/
2(âð(ðŒ)) â¥â(ðâ
ðð
1(ðŸ)) âðŸ â(ðð¶ð/
2(ð)) â â,
and the resolution of this final case completes the proof. â¡ Returning to (4.15), this lemma showsE (ðð) â â, which is a contradiction. We can now conclude that each sequenceâ(ðâ
ðð
1(ðŸ))remains bounded above.
It is well-understood that the boundedness of the length spectrum implies that (ðâ
ðð
1) converges along a subsequence to some new Fuchsian representation ðâ : Î â PSL2(R). We now restrict the [ðð],[ðð] to this chosen subsequence. Since the mapping class group acts properly discontinuously on the TeichmÃŒller space, it follows that [ðâ
ðð
1] = [ðâ] for allðlarge enough. In particular, ððis isotopic to ðð for allð, ðlarge enough. This completes the proof of Lemma 4.3.3.
To finish the proof of Proposition 4.3.1, we know there is anðsuch that[ðð] = [ðð] for allð ⥠ð. As the mapping class group acts by isometries with respect to the metricðT,
ðT( [ðð],[(ðâ1
ð )âðâ]) =ðT( [ðâ
ððð],[ðâ]) â 0, and hence [ðð]converges to [(ðâ1
ð )âðâ] asð â â. Proposition 4.3.1 is proved.
Uniqueness of critical points
The uniqueness statement in Theorem 4A amounts to showing that critical points ofEare unique. The main step in the uniqueness proof for closed surfaces [Tho17, Theorem 1] is a local computation ([Tho17, Lemma 2.4]) that goes through in our setting. For this reason, we omit the proof of uniqueness and invite the reader to see [Tho17]. The only real difference is that we must address convergence of various integrals, and use our Lemma 4.2.14 instead of the usual energy minimizing property. Weâve shown this type of calculation throughout the chapter, so we feel comfortable leaving it to the interested reader.
The mapsΚð
We now begin the proof of Theorem 4B. Generalizing the map Κ from [Tho17, subsection 2.3]), we define the map
Κð :T (Î) ÃRepc(Î, ðº) âASDc(Î, ðº) â Tc(Î) ÃRepc(Î, ðº)
as follows. When chas no hyperbolic classes, we implicitly assume there is no ð. Begin with a hyperbolic surface (Σ, ð) and a reductive representation ð
2 : Î â PSL(2,R). Associated to this representation, we take a ð
2-equivariant harmonic map with twist parameterð, ðð : (ΣË,ðË) â (ð , ð). We proved in [Sag19, Theorem 1.4] that there exists a unique Fuchsian representation ð
1 : Î â PSL(2,R) and a unique ð
1-equivariant harmonic diffeomorphismâð : (ΣË, ð) â (H, ð) that has Hopf differential Ί(âð) = Ί(ðð). From [Sag19, Proposition 3.13], the mapping ð¹ =(âð, ðð)is a spacelike maximal immersion into the pseudo-Riemannian product, and hence ð
1almost strictly dominatesð
2. Κð is defined by Κð( [ð],[ð
2])= ( [ð
1],[ð
2]).
It is clear that the map is fiberwise in the sense of the statement of Theorem 4B.
Theorem 4A showsΚð is bijection, and the content of Theorem 4B is that Κð is a homeomorphism.
Remark 4.3.5. We do not know ifΚð = Κðâ² forð â ðâ²! Proof of Theorem 4B: continuity
From now on we refine our notation: when the representation ð is not implicit, the ð-equivariant harmonic map for a metric ð with zero twist parameter will be denoted ð
ð
ð, unless specified otherwise.
The proof of continuity is almost identical for each twist parameter, so we work only withΚ = Κ0. Let us also assume for the remainder of this section that there is a single peripheral ð. Lifting various equivalence relations, Κis described by a composition
(ð, ð
2) âŠâ (ð
ð2
ð , ð
2) âŠâ (Ί(ð
ð2
ð ), ð
2) âŠâ (ð
1, ð
2), where ð
1 is the holonomy of the hyperbolic metric ðâ² on Σ such that the associ- ated harmonic map from (Σ, ð) â (Σ, ðâ²) has Hopf differential Ί(ð
ð2
ð ). [Sag19, Theorem 1.4] implies the association fromΊ(ð
ð2
ð ) âŠâ ð
1is continuous. Here, the topology on the space of holomorphic quadratic differentials is the Fréchet topology
coming from taking ð¿1-norms over a compact exhaustion. To show continuity of ðâŠâ ð
ð2
ð âŠâΊ(ð
ð2
ð )with respect to this topology, note that, by the constructions of Section 4.2, we already have continuity in the TeichmÃŒller coordinate. Continuity will thus follow from the results below.
Lemma 4.3.6. Suppose representations (ðð)â
ð=1 converge to an irreducible ð in Hom(Î, ðº), and all such representations project to Repc(Î, ðº). Then ð
ðð
ð converges to ð
ð
ð in theð¶âsense on compacta asðâ â.
Below, we work in fixed local sections of the principal bundles ðc(Î, ðº) â Repc(Î, ðº) ðc(Î,PSL(2,R)) â Repc(Î,PSL(2,R)) (having assumed in the in- troduction that such things exist). We choose a basepoint for theð
1so that we can view these local systems through their holonomies, which are genuine representa- tions.
Proof. We may assume each ðð is irreducible and that there is a single cusp. The parabolic and elliptic cases are much simpler than the hyperbolic case, so we treat only the latter. Set ðð = ð
ðð
ð , ð = ð
ð
ð and let ðð, ð be the approximation maps for ð
ðð
ð , ð
ð
ð from 3.1. These maps project a cusp neighbourhood onto a geodesic, but now the geodesic is varying with ðð, and converging in the Gromov-Hausdorff sense to the geodesic forð. If theððâs can be chosen to vary smoothly, then we have a uniform bound on the total energies (recall they depend on a choice of a constant speed map onto their geodesic). As discussed in Section 4.2 (in particular the proof of Lemma 4.2.11), this yields a uniform total energy bound for ððon compacta. Via the argument described early in Section 4.2, the maps ððð¶â-converge on compacta along a subsequence to some limiting map ðâ. The mapping is necessarily ð- equivariant and harmonic. To see that ðâ = ð, recall that ð is characterized by the fact that its Hopf differential has a pole of order 2 at the cusp and the residue has a specific complex argument. To check this property, from the uniqueness in [Sag19, Theorem 1.1] it suffices to check ð(ðâ, ð) < â, and the standard contradiction argument will also show there is no need to pass to a subsequence. As remarked previously, the proof of [Sag19, Lemma 5.2] showsð(ðð, ðð)is maximized onðΣ2. Ifðð is the subsequence, then taking ð â â, via compactness we do win
ð(ðâ, ðâ) â€lim sup
ðââ max
ðΣ2
ð(ðð
ð, ðð
ð) â€max
ðΣ2
ð(ðâ, ðâ) +1< â.
We are left to argue that one can choose theððso thatðð âðin theð¶âsense. We realizeð, ððas monodromies of flat connectionsâ,âðrespectively on anð-bundle
ðž over Σ with structure group ðº. This is possible when all ðð lie in the same component of the relative representation space ([Lab13, Corollay 6.1.2] generalizes to relative representation spaces), which we are free to assume. The pullback bundle with respect to the universal covering ËΣ â Σidentifies with the trivial bundle and also pullsâ,âðback to flat connections Ëâ,âËð. That is, up to an isomorphism, we have a family of commutative diagrams
(ð ÃΣË,âËð) (ðž ,âð)
Î£Ë Î£.
The bundleðž and connectionsâðare constructed by choosing a good covering of Σ and depend on the local section of the principal bundle ðc(Î, ðº) âRepc(Î, ðº) (see the proof of [Lab13, Lemma 6.1.1]). We can build a good covering by glueing a good covering of a relatively compact tubular neighbourhood of(Σ2, ð)with a good covering of a tubular neighbourhood of (Σ\Σ2, ð). This ensures a lower bound on theð-radius insideΣ2(note that we cannot do this on all ofΣ). With this constraint, the local systems can be prescribed so that âð â âin the ð¶â sense on Σ2 in the affine space of connections.
The map ð induces a section ð of the bundle (ðž ,â), which can also be seen as a section ð ð of (ðž ,âð). With respect to the diagram above, ð ð pulls back to a ðð-equivariant map Ëðð :H â ð, and by our comments above, Ëðð â ðin theð¶â sense. The maps Ëðð are most likely not harmonic and may not project onto the geodesic axis of ðð(ð). However, because they are converging to ð, the geodesic curvature of Ëðð(ðΣË2) is tending to 0. Moreover,ð|ðΣË
2 is arbitrarily close to some parametrization of a geodesic. So, we setððto be the harmonic map with boundary data equal to this nearby geodesic projection. From energy control on Ëðð, we get the same for ðð, and hence convergence along a subsequence to a harmonic map ðâ. From the boundary values, we must have ðâ = ð, and as usual we can show
there is no need to pass to a subsequence. â¡
For a reducible representation, the harmonic maps are not unique but differ by translations along a geodesic. However, the Hopf differential is unique. We can choose a sequenceðð â ðas in the previous propositionâboundary values are fixed so we do have uniqueness for these harmonic maps. Then we take the harmonic maps ðð built from using the approximating mapsðð, and the proof above goes through,
with the minor modification that we must rescale the approximating maps as in [Sag19, Proposition 5.1]. ð¶â convergence implies the Hopf differentials converge.
Remark 4.3.7. For different twist parameters, we simply precompose everyððwith a fractional Dehn twist. Using (4.12), the energies are uniformly controlled.
Asymmetric metrics on TeichmÃŒller spaces
In [GK17, Section 8], Guéritaud-Kassel define asymmetric pseudo-metrics on de- formation spaces of geometrically finite hyperbolic manifolds. These metrics are natural generalizations of Thurstonâs asymmetric metric on TeichmÃŒller space. We will use such a metric in the proof of bi-continuity ofΚð.
For[ð
1],[ð
2] â Tð (ðð,ð), we set ð¶(ð
1, ð
2) =inf{Lip(ð) : ð :HâHis(ð
1, ð
2) âequivariant}. The metricðð â :Tð (ðð,ð) Ã Tð (ðð,ð) â Ris defined by
ðð â( [ð
1],[ð
2]) =log
ð¶(ð
1, ð
2)ð¿(ð
1) ð¿(ð
2)
. Hereð¿(ð
1)is the critical exponent of ð
1: ð¿(ð
1) = lim
ð ââ
1
ð log #(ð(Î) ·ðâ©ðµð(ð )),
where ðis any point inHandðµð(ð ) is the ball of radiusð centered at ð. Alterna- tively, ð¿(ð
1) is the Hausdorff dimension of the limit set of ð
1in ðâH = ð1. If all ðð =0, so that all critical exponents are 1 andTð (ðð,ð)is the ordinary TeichmÃŒller space of a surface of finite volume, then this agrees with Thurstonâs asymmetric metric introduced in [Thu98]. Here asymmetric means that in general,
ðð â( [ð
1],[ð
2]) â ðð â( [ð
2],[ð
1]).
Proposition 4.3.8(Guéritaud-Kassel, Lemma 8.1 in [GK17]). The function ðð â : Tc(Î) à Tc(Î) â Ris a continuous asymmetric metric.
Remark 4.3.9. The correction factorð¿(ð
1)/ð¿(ð
2)is needed for non-negativity. For a general hyperbolicð-manifold, continuity may fail and the generalization may be just an asymmetric pseudo-metric (ð(ð¥ , ðŠ) need not implyð¥ =ðŠ).
This metric allows us to control translation lengths nicely. Consider a left open ball aroundð
2:
ðµð â(ð
2, ð¶) ={[ð
1] â Tc(Î) : ðð â(ð
1, ð
2) < ð¶}.