Scalar curvature rigidity of domains in a 3-dimensional warped product
Abstract.
A warped product with a spherical factor and a logarithmically concave warping function satisfies a scalar curvature rigidity of the Llarull type. We develop a scalar curvature rigidity of the Llarull type for a general class of domains in a three dimensional spherical warped product. In the presence of rotational symmetry, we identify this class of domains as those satisfying a boundary condition analogous to the logarithmic concavity of the warping function.
Key words and phrases:
Scalar curvature, mean curvature, stable capillary surface, prescribed mean curvature, rigidity, foliation.1991 Mathematics Subject Classification:
53C24, 49Q201. Introduction
Llarull proved a scalar curvature rigidity theorem for the standard -spheres.
Theorem 1.1 ([Lla98]).
Let be a smooth metric on the n-sphere with the metric comparison and the scalar curvature comparison . Then .
An interesting feature of this scalar curvature rigidity for spheres comparing to that of torus [SY79a], the Euclidean space [SY79b] and the hyperbolic space [MO89] is the requirement of a metric comparison . A counterexample without the requirement was given in [BM11] (for half -sphere fixing the geometry of the boundary), but can be weakened, see [Lis10]. Llarull also showed that the condition can be formulated more generally as the existence of a distance non-increasing map of non-zero degree.
Recently, there were efforts in extending Llarull’s theorem to a spherical warped product
(1.1) |
by spinors [CZ24], [BBHW24], [WX23b], by -bubbles [Gro21], [HLS23] and by spacetime harmonic functions [HKKZ].
We are interested in the Llarull type theorems of domains in the spherical warped product (1.2). Although the form (1.2) can also be considered as a domain in a larger spherical warped product, our focus will be on such domains with boundaries that are not necessarily given by -level sets. Previously, this direction has been explored by Lott [Lot21], Wang-Xie [WX23a] and Chai-Wan [CW24], which all involved spinors. Gromov first suggested the use of stable capillary minimal surface in studying the scalar curvature rigidity of Euclidean balls (see Section 5.8.1 of [Gro21]; Spin-Extremality of Doubly Punctured Balls) and Li [Li20] in three-dimensional Euclidean dihedral rigidity. In this article, we make use of the stable capillary surfaces with prescribed (varying) contact angle and prescribed mean curvature, or in the terminology of Gromov [Gro21], (part of) the boundary of a stable capillary -bubble. This is also a further development of the previous work [CW23] and the recent work of Ko-Yao [KY24] in the Euclidean case.
We consider three dimensions, fix a metric of positive Gauss curvature on the -sphere , and the following
(1.2) |
We reserve for the standard round metric on the 2-sphere . We use to indicate the coordinate and to denote the coordinate of a point . Let be the Gauss curvature at .
Let , we always assume that lies between such that is non-empty. If contains only a point, we set this point to be . Let and be the unit outward normal of in with respect to . We fix
(1.3) |
and to be the dihedral angles formed by and that is given by
(1.4) |
We call the -suplevel set and -sublevel set . We also need to fix some more conventions for the direction of the unit normal, the sign of the mean curvatures and the dihedral angles. Let be a surface with boundary on and separates and , we always fix the direction of the unit normal of to be the direction which points inside of the region bounded by and . The mean curvature is then the trace of the second fundamental form . We fix to be the contact angle formed by and , that is, . For the mean curvature of , it is always computed with respect to the outward unit normal. The geometric quantity on comes with a bar unless otherwise specified (see Figure 2.1).
As is well known, the warped product metric (1.2) is conformal to a direct product metric. Indeed, let , then and
(1.5) |
where is implicitly given by .
Now we state our scalar curvature rigidity result. (For a quick feel, we refer to Theorem 1.4.)
Theorem 1.2.
Assume that is a Riemannian metric on as in (1.2). Assume that is a domain in such that is convex with respect to the conformally related metric where is given in (1.5) and is a smooth metric on . If satisfies the metric comparison
(1.6) |
and the scalar curvature comparison
(1.7) |
and the mean curvature comparison
(1.8) |
along , and further that and satisfy either of the following conditions:
-
(1)
, only contains a point where is smooth under both metrics and ;
-
(2)
, only contains a point where where the tangent cone of at is Euclidean, and the tangent cone of at is a round (solid) circular cone and that converges to its axial section as ;
-
(3)
as where is a constant such that the tangent cone of at is of non-negative Ricci curvature and admits a metric tangent cone at ;
-
(4)
on , is a disk, the mean curvatures and the dihedral angles formed by and ,
then .
Remark 1.3.
The mean curvature comparisons can be reformulated as on if all mean curvatures are computed with respect to the outward unit normal. For convergence of sequences of Riemannian manifolds and notions of tangent cones at a point of a Riemannian manifold, we refer to [BBI01, Chapter 8]. We believe that the extra condition on the tangent cone of at can be dropped in item (2).
We use a special case of to illustrate the convexity of with respect to the metric given in (1.5). Assume that is just the standard round metric, that is,
(1.9) |
using the polar coordinates and is given by
for some positive function on . In this case, the prescribed contact angle only depends on . It is easy to check that the convexity of with respect to is equivalent to . Also, we find that
(1.10) |
using (1.5) or that the angles are conformally invariant. The mean curvature of a -level set is given by
The logarithmic concavity of is equivalent to the more geometric statement that the mean curvature of the -level set is monotonically decreasing as increases. The condition can be viewed as a boundary analog of the logarithmic concavity. Geometrically, (1.10) says that the dihedral angles (1.4) formed by and monotonically decreases along the direction with respect to the metric . This condition answers a question raised by Gromov at the end of [Gro21, Section 5.8.1]. See also Chai-Wan [CW24, Theorem 1.1].
We can have various scalar curvature rigidity assuming different combinations of geometric structures at . The following is a simple corollary of Theorem 1.2 where both take the structure as specified in item (1) of Theorem 1.2.
Theorem 1.4.
Let be a smooth domain in with the metric where , on . Assume that is convex with respect to the conformally related metric where is given in (1.5) and is another metric on which satisfies the comparisons of:
-
(1)
the scalar curvatures ,
-
(2)
the mean curvatures of the boundary in ,
-
(3)
and the metrics ,
then .
Our approach toward items (1)-(3) is by construction of surfaces of prescribed mean curvature and prescribed contact angles near which serves as barriers, a concept which now introduce. The existence of barriers enables us to find a stable capillary -bubble, if fact, are natural barriers if we are in item (4) of Theorem 1.2.
Definition 1.5.
We say that a surface () whose boundary separates and is an upper (lower) barrier if () and () along (). We call and are a set of barriers if and are respectively an upper barrier and a lower barrier and is closer to than .
Our construction of barriers near are purely local, and in this way, our proof of items (1)-(3) can be reduced to the last item of Theorem 1.2, that is, the case with a set of barriers. In particular, if both take the structure as specified in item (3) of Theorem 1.2, we obtain a genuine generalization of Theorem 1.1, since in the case of round metric, , is allowed to take zero values at and . Hu-Liu-Shi [HLS23] (see also Gromov [Gro21]) used a -bubble approach for Theorem 1.1. As indicated earlier, we use the capillary version of the -bubble approach. However, our method differs from theirs in a technical manner when handling
near . (Similarly, near .) They constructed a family of perturbations on the function while our strategy is to perform a careful tangent cone analysis near or . As a result of this new strategy, we are able to generalize the Llarull Theorem 1.1 to the case where the background metric are equipped with antipodal conical points.
Theorem 1.6.
Let and be a three dimensional warped product given in (1.1) such that
If is another smooth metric on with possible cone singularity at only which satisfies the comparisons of metrics and scalar curvatures , then .
Theorem 1.6 directly follows from the proof of item (3) of Theorem 1.2 with only slight changes and we omit its proof. See Remark 3.13. Note that the condition ensures that the Ricci curvature of the tangent cone with respect to at is non-negative. The scalar curvature rigidity of the Llarull type for is an interesting question. One could also compare Theorem 1.6 with [CLZ24] where conical singularities with respect to the metric are allowed at multiple points on .
Some of the essential difficulties of items (1) and (2) are already present in [CW23, Theorem 1.2 (2) and (3)]. In light of this, we only give a sketch for their proof in Section 4 , and refer relevant details to [CW23].
It is possible that the inequalities in and the convexity of can be weakened in some cases. For instance, we can consider a convex disk in with the direct product metric . In this case, vanishes. The Llarull type rigidity Theorem 1.4 is still valid for this .
Now one could naturally ask what are other shapes of point singularities, in particular, asymptotics of near , such that Theorem 1.2 remain valid. However, it is a quite intricate matter to which we do not have an answer at the moment. It is also desirable to find a proof for higher dimensional analogs of our results using the stable capillary -bubbles. This seems to be a promising direction to investigate being aware of the recent works [CWXZ24, WWZ24].
The article is organized as follows:
In Section 2, we introduce basics of stable capillary -bubble and we use it to show item (4) of Theorem 1.2.
In Section 3, we use the tangent cone analysis at to construct barriers and reduce item (3) of Theorem 1.2 to item (4).
In Section 4, we revisit our constructions in [CW23] and use the techniques developed there to show items (2) and (1) of Theorem 1.2.
Acknowledgments X. Chai was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. RS-2024-00337418) and an NRF grant No. 2022R1C1C1013511. G. Wang was supported by the China Postdoctoral Science Foundation (No. 2024M751604).
2. Stable capillary -bubble
In this section, we introduce the functional (2.1) whose minimiser is a stable capillary -bubble. We introduce the barrier condition which combining with a maximum principle ensures the existence of a regular minimiser to (2.1). By a rigidity analysis on the second variation of (2.1), we conclude the proof of item (4) of Theorem 1.2.
2.1. Notations
We set up some notations. Let be be a set such that is a regular surface with boundary which we name it . We set
-
•
, unit normal vector of pointing inside ;
-
•
, unit normal vector of in pointing outside of ;
-
•
, unit normal vector of in pointing outside of ;
-
•
: unit normal vectors of in pointing outside of ;
-
•
: the contact angle formed by and and the magnitude of the angle is given by ,
-
•
, the inner product of vectors and with respect to the metric ;
-
•
, the inner product of vectors and with respect to the metric .
See Figure 2.1. We use a bar on every quantity to denote that the quantity is computed with respect to the metric given in (1.2).
2.2. Functional and first variation
We fix and to be given by (1.4). We define the functional
(2.1) |
where denotes the reduced boundary of and the variational problem
(2.2) |
where is the collection of contractible open subsets such that . Let be a surface with boundary such that separates . Then separates into two components and the component closer to is just . We reformulate the functional (2.1) in terms of . We define
(2.3) |
Let be a family of immersions such that and . Let and be the corresponding component separated by . Let be the vector field . Define and , then by the first variation
(2.4) |
We know that if is regular, then it is of mean curvature and meets at a prescribed angle . And is called a capillary -bubble. The second variation at such is
(2.5) |
where and
(2.6) |
We define two operators
(2.7) |
and
The surface is called stable if
(2.8) |
for all . The second variation (2.5) is closely related to the variation of and . Indeed, let , we have that the first variation of is
(2.9) | ||||
(2.10) |
And the first variation of the angle difference is
(2.11) | ||||
(2.12) |
For , Schoen-Yau [SY79b] rewrote the term as
(2.13) |
where is the Gauss curvature of . Along the boundary , we have the rewrite (see [RS97, Lemma 3.1] or [Li20, (4.13)])
(2.14) |
where is the geodesic curvature of in .
2.3. Analysis of stability
Starting from now on, we assume that is a regular stable capillary -bubble in which satisfies the assumptions of item (4) of Theorem 1.2.
Lemma 2.1.
Let be a regular stable capillary -bubble, then is a -level set.
Proof.
First, we note that the second variation as in (2.5). First, using Schoen-Yau’s rewrite (2.13) we see that
(2.15) | ||||
(2.16) | ||||
(2.17) | ||||
(2.18) |
where is the traceless part of the second fundamental form. Similarly using (2.14), we see
We obtain by letting in the (2.8) (also using (2.5) and (2.6)),
(2.19) | ||||
(2.20) | ||||
(2.21) | ||||
(2.22) |
where in the last line we have incorporated the comparisons in and on .
Now we estimate . We have that
since , so
and we get
So
For any point , the right hide side is just . Recall that is the coordinate of . This is by a direct calculation of the scalar curvature of the warped product metric (1.2). So
(2.23) |
Let and it is conformally related to via (1.5). Let be the unit outward normal of in and be the mean curvature of in with respect to . Since is conformal to , by a well known formula of conformal change of mean curvature,
(2.24) |
Similarly, the mean curvature of in is
(2.25) |
Hence, by (2.24), (2.25), (1.5) and that ,
Inserting (2.23) and the above in (2.22) yields
where we have written the area element and line length element explicitly in the metric . The rest of the proof is deferred to the next Lemma 2.2. ∎
Lemma 2.2.
Assume that . If is a surface in whose boundary is a simple smooth curve that separates and , then
(2.26) |
where equality occurs if and only if is a -level set.
Proof.
It suffices to prove (2.26) for since . In this case, , we just use . We also suppress the subscript for clarity in this proof. In addition, we assume that every -coordinate of is strictly less than , since we can increase a little. Let be the unit tangent vector of with respect to and be the unit outward normal of in with respect to . Let (resp. ) be the unit tangent vector of with respect to (resp. ). We recall an ingenious inequality of [KY24, Lemma 3.2],
(2.27) |
where denotes the inner product with respect to .
Let be an arc-length parameter of with respect to , then the length element of with respect to is given by and . Therefore,
(2.28) | ||||
(2.29) | ||||
(2.30) |
Let be the unit outward normal of in with respect to . Recall the notation , and we know that the components of satisfy
(2.31) |
in the frame . So
Since the orthonormal frames and have the same orientation, we can set and so for some and which satisfy . It then follows that
(2.32) |
by considering also that . Hence
Suppose that (we set to avoid double subscripts) and enclose a region in . By the divergence theorem,
(2.33) | ||||
(2.34) | ||||
(2.35) | ||||
(2.36) |
where is the induced connection on with respect to the metric . For , we use Gauss-Codazzi equation,
where the fact that was also used in determining the Ricci curvature. For , we use (2.32), we see
where is the geodesic curvature of in at . It seems tricky to calculate in directly at a first sight, but we can convert to terms that are easier. By the identity,
at , we see
(2.37) | ||||
(2.38) | ||||
(2.39) | ||||
(2.40) |
at . It is now a tedious task to check from the definition of , and (2.31) that the above vanishes. Therefore, and to sum up, we have shown that
(2.41) |
Now we set the region enclosed by , and to be . Let . Using the divergence free property of (twice-contracted Gauss-Codazzi equation), , and the divergence theorem,
where now also denotes the unit outward normal of with respect to and denotes the -th component of the vector field . Note that and so
and it follows from that that
(2.42) |
Finally, it follows from (2.41) and (2.42) that
An application of the Gauss-Bonnet theorem on the right hand finishes the proof of (2.26). The equality case is easy to trace. ∎
2.4. Infinitesimally rigid surface
The surface is a stable capillary -bubble has more consequences than the mere Lemma 2.1. We can conclude that is a so-called infinitesimally rigid surface. See Definition 2.4.
All inequalities are in fact equalities by Lemma 2.2 and tracing the equalities in (2.22), we arrive that
(2.43) |
and
(2.44) |
It then follows from the equality case of Lemma 2.2 that
(2.45) |
for some constant . Because is stable (equivalently ), so the eigenvalue problem
(2.46) |
has a non-negative first eigenvalue .
The analysis now is similar to [FCS80]. Letting in (2.8), using (2.43), (2.44) and (2.45), we get
(2.47) | ||||
(2.48) |
And so the first eigenvalue is zero, and the constant 1 is its corresponding eigenfunction.
2.5. Capillary foliation of constant
See for instance the works [Ye91], [BBN10] and [Amb15] on constructing CMC foliations. First, we construct a foliation with prescribed angles whose leaf is of constant . Let be a local flow of a vector field which is tangent to and transverse to and that .
In the following theorem, we only require that the scalar curvature of and the mean curvature of are bounded below.
Theorem 2.5.
Suppose is a three manifold with boundary, if is an infinitesimally rigid surface, then there exists and a function on such that for each , the surface
(2.50) |
is a surface of constant intersecting with prescribed angle . Moreover, for every and every ,
(2.51) |
Proof.
Given a function in the Hölder space (), we consider
which is a properly embedded surface if the norm of is small enough. We use the subscript to denote the quantities associated with .
Consider the space
and
Given small and , we define the map
given by
(2.52) | ||||
(2.53) |
Here, is a ball of radius centered at the zero function in . For each , the map
gives a variation with
Since is infinitesimally rigid and using also (2.10) and (2.12), we obtain that
It follows from the elliptic theory for the Laplace operator with Neumann type boundary conditions that is an isomorphism when restricted to .
Now we apply the implicit function theorem: For some smaller , there exists a function , such that and for every . In other words, the surfaces
are of constant with prescribed angles .
Let where . By definition, for every and for every . Observe that the map gives a variation of with variational vector field is given by
Since for every we have that
(2.54) | ||||
(2.55) |
by taking the derivative at we conclude that
satisfies the homogeneous Neumann problem. Therefore, it is constant on . Since
for every , by taking derivatives at again, we conclude that
Hence, . Taking small, we see that parameterize a foliation near . ∎
Theorem 2.6.
There exists a continuous function such that
Proof.
Let parameterize the foliation, , . Then
(2.56) |
and
(2.57) |
By shrinking the interval if needed, we assume that for . By multiplying of (2.56) and integrate on , we deduce by integration by parts and applying the Schoen-Yau rewrite (2.13) that
(2.58) | ||||
(2.59) | ||||
(2.60) |
Let , we have that
(2.61) | ||||
(2.62) | ||||
(2.63) | ||||
(2.64) |
where is the traceless part of . Also,
So
(2.65) | ||||
(2.66) | ||||
(2.67) | ||||
(2.68) | ||||
(2.69) |
where in the last line we have also used the bound (2.23). Now we use (2.57) and also the rewrite (2.14), we see that
(2.70) | ||||
(2.71) | ||||
(2.72) | ||||
(2.73) | ||||
(2.74) | ||||
(2.75) | ||||
(2.76) |
It follows from Lemma 2.2 and the proof of Lemma 2.1 that the second term in the big bracket is bounded below by . Using also the Gauss-Bonnet theorem on the first term in the bracket, we see that
(2.77) |
Let
(2.78) |
then note that we have assume that near , so satisfies the ordinary differential inequality
(2.79) |
We see then
So the function is non-increasing. ∎
2.6. From local foliation to rigidity
Let be the constant mean curvature surfaces with prescribed contact angles with .
Proposition 2.7.
Every constructed in Theorem 2.5 is infinitesimally rigid.
Proof.
Proof of item (4) of Theorem 1.2.
We note easily by the assumptions of item (4) of Theorem 1.2 that are a set of barriers (see Definition 1.5), by the maximum principle, there exists a minimiser to (2.2) such that is either empty or or lies entirely away from . Without loss of generality, we assume that non-empty. By [DPM15], is a regular stable surface of prescribed mean curvature and prescribed contact angle . Moreover, the second variation in (2.5) for any smooth family such that .
Let where and are as Theorem 2.5, we show first that is conformal. It suffices to show that is conformal.
Since every is infinitesimally rigid by Proposition 2.7, from (2.46) and (2.48), we know that is a constant. Let , be vector fields induced by local coordinates on , also extends to a neighborhood of via the diffeomorphism . We have . Note that are umbilical with constant mean curvature , so
and
(2.80) | ||||
(2.81) | ||||
(2.82) |
On the other hand,
(2.83) | ||||
(2.84) | ||||
(2.85) | ||||
(2.86) | ||||
(2.87) |
Combining the two equations above, we conclude that which implies that foliates a warped product under the diffeomorphism (parameterized by ). Considering that the induced metric on agrees with the induced metric from , we conclude that . ∎
3. Construction of barriers (I)
In this section, we prove item (3) of Theorem 1.2. Our strategy is to construct a surface () which serves as a lower (upper) barrier, and to use item (4) of Theorem 1.2 to finish the proof. This section is occupied by such a construction of .
3.1. Setting up coordinates and notations
For convenience, we set . As before, for any , we set to be the -level set of and to be the -sublevel set, that is, all points of which lie below . Since both and has cone structures near where where each cross-section of the cone is a topological disk and it collapses to a point which we denote by .
In the following subsections, we construct graphical perturbations of . Let be the surface which consists of points where is the unit normal of with respect to the metric at . The boundary might not lie in , we can compensate this by expanding or shrinking a little, and we still denote the resulting surface .
We use a subscript on every geometric quantity on and a subscript on every geometric quantity on . We will explicitly indicate when there was confusion or change.
3.2. Capillary foliation with constant
We assume that and have isometric tangent cones at and we construct a foliation of constant with prescribed angles near . In fact, later in Subsection 3.3, it is shown that this is the only case.
By the first variation formula of the mean curvatures
(3.1) |
where is the Laplacian with respect to the induced rescaled metric . Note that by the fact the tangent cone is . By the Taylor expansion of the function , we see that
(3.2) |
So
(3.3) |
Note that both and are finite and considering that and have isometric tangent cones at .
Remark 3.1.
We elaborate a bit more on (3.1) and its remainder term. Since the metric is close to when , we calculate the expansions with respect to the rescaled metric when computing for small . This is similar to [Ye91]. Then we rescale back and we obtain (3.1). The term involves products of which is of order with terms of order at most . That is why the remainder is only of order instead of .
Also, the variation of angles give
(3.4) | ||||
(3.5) |
where is the outward unit normal of in with respect to the rescaled induced metric (note that is of unit length with respect to ). Other geometric quantities are not rescaled. By the variation of the prescribed angle ,
(3.6) |
So
(3.7) | ||||
(3.8) | ||||
(3.9) |
Remark 3.2.
The term , however, we observe that , and since . Or we can calculate with respect to the rescaling metric as in Remark 3.1.
Since and has isometric tangent cone at , we see that the limit of the surface as is where is a scaling copy of a geodesic disk of radius in the standard 2-sphere. Consider the spaces
and
(3.10) |
Given small and , we define the map
(3.11) |
given by where , are given by
(3.12) | ||||
(3.13) |
for . Here is an open ball with radius in the norm and the integration on is with respect to the metric . We extend to by taking limits, that is,
(3.14) |
We have the following proposition.
Proposition 3.3.
For each with small enough, we can find such that and
In particular, each of the surfaces have constant and prescribed angles . Moreover, for all small .
Before proving this proposition, we give a variational lemma.
Lemma 3.4.
Suppose that is a compact manifold with piecewise smooth boundary and is a relatively open, smooth subset of . Let be a smooth family of metrics indexed by such that as , let . We now omit the subscript on . Let be the unit outward normal of in , and be the mean curvatures and the second fundamental form of in computed with respect to the unit normal pointing outward of , and be the dihedral angles formed by and with respect to the metric . We put a hat at appropriate places for the geometric quantities with respect to .
Then
(3.15) | ||||
(3.16) |
Here, we have used to denote a remainder term comparable to .
Proof.
Lemma 3.5 implies the following by taking the difference of two families of metrics.
Corollary 3.5.
Assume is the manifold from Lemma 3.5, for two family of metrics close to indexed both by a small parameter , we have
(3.19) | ||||
(3.20) |
Now we are ready to prove Proposition 3.3.
Proof of Proposition 3.3.
The proof is similar to [CW23]. We bring up only the main differences.
Because the right hand of both (3.3) and (3.9) converge to and (up to a constant) respectively, so we can first follow [CW23, Proposition 4.2] to construct a foliation near with constant and along , and then [CW23, Lemma 4.3] to obtain that
(3.21) |
Now we show that .
We consider the rescaled set with two rescaled metrics and . Since and , it is easy to see that converges to a truncated metric cone with the metric where and is some convex disk in a 2-sphere . We set . Since and has isometric tangent cone at , converges to as well. Therefore, we can view and (indexed by ) as two metrics on getting closer to as . We rescale (3.21) by a factor of , we obtain
which is equivalent to
In the above the integration done is with respect to the metric and are the mean curvature of in computed with respect to the normal pointing inside of .
All the comparisons in item (3) of Theorem 1.2 carry over to the rescaled metrics and on , and that has non-negative Ricci curvature by the assumptions of item (3) of Theorem 1.2. We use Corollary 3.5 and arrive that , that is,
Since satisfies the differential inequality (2.79) and considering the asymptotics , and in (2.78), we see that for all . ∎
Remark 3.6.
The Ricci curvature in Corollary 3.5 blows up near , however, because we are integrating with respect to the metric , the volume near is small. Also, the difference is small. So the blowing up of the Ricci curvature will not cause an issue.
3.3. Barrier construction with non-isometric tangent cones
Since , the manifold is topologically a cone near and it is a point at . According to the assumptions of item (3) of Theorem 1.2, at also locally resembles a cone, that is,
(3.22) |
where is a parameter, is a metric on a two dimensional disk and is small compare to . In other words, the tangent cone at is a cone with the metric .
Now we can also identify near as and as a function on . Let , we see that as a function on only depends on . So we view as a function on . Since on , we have that . Now we discuss the case that on .
Lemma 3.7.
If on , then . That is, and have isometric tangent cones at .
Proof.
Since , so we can rescale and by the same scale to obtain a cone but with two different metrics and . For , set . Since the metric comparison, the mean curvature and the scalar curvature comparison are preserved by rescaling, so , the scalar curvature and the mean curvature of at satisfies .
Since both , are warped product metrics, the comparison reduces to Gaussian curvature comparison of and by a direct computation of scalar curvature (or Gauss equation). Let be the geodesic curvatures of with respect to . By direct calculation, the second fundamental form of in the direction vanishes with respect to both metrics and the second fundamental form of in with respect to agree. It then follows from and (2.14) that .
To summarize, we have comparisons on that , and along . By Gauss-Bonnet theorem, on and it follows that . Therefore, and have isometric tangent cones at . ∎
By the above lemma, the case is the case which implies isometric tangent cones of and at . This is the case we have already addressed in Subsection 3.2. Without loss of generality, we assume that .
We first consider the difference of of the perturbation for . We now represent at and its value at the graphical perturbations of by to avoid notational confusion. By the first variation of the mean curvatures,
(3.23) | ||||
(3.24) |
where is the Laplacian with respect to the metric .
Remark 3.8.
We have converges to as by the metric (3.22) near , and to indicate that the limit carries the metric , we use instead of only.
Lemma 3.9.
We have that
Proof.
Since converges to a truncated radial cone and converges to the section of the radial cone with unit distance to , so the section has second fundamental form and by rescaling,
as .
At a point , the value of is given by where is the projection of to the second coordinate. Since as a function on only depends on , we see that the value of the function at the graphical perturbation of is given by . Since , so
Hence
which proves the lemma. ∎
Let which is a function on the limit , so
(3.25) |
where is the Laplacian of . Recall that , so
Let be the dihedral angles formed by and , and be the angles formed by and the graphical perturbation of .
Lemma 3.10.
The dihedral angles formed by and approach as .
Proof.
Since converges to a truncated radial cone, converges to the section of the radial cone with unit distance to , and this section is orthogonal to the radial direction in the limit, so the intersection angles of and approaches as . ∎
Lemma 3.11.
We have that .
Proof.
The lemma can be deduced from that is approximately the radial direction as , the scaling property of and the following lemma. ∎
Lemma 3.12.
Let be a 2-surface with boundary and be the cone over . Then the second fundamental form of in in the direction vanishes.
Proof.
Let be a tangent vector field over , then by direct calculation . So since on the metric is . Due to the same reason, the unit normal vector of in is tangent to , so the claim is proved. ∎
We are interested in the difference between and the value of which to avoid confusion we denote by () at (the graphical perturbation of) . Using the relation of and , . By the expansion of angles (see (3.5)), we see
And
Since each is stable capillary minimal surface under the metric , so we know that
Based on the above asymptotic analysis and Lemmas 3.10 and 3.11, we see
(3.26) |
on where is the outward normal of in . By the elliptic strong maximum principle, the operator
is an isomorphism since in due to Lemma 3.9 and . In other words, we can specify the limits
(3.27) | |||
(3.28) |
by choosing a suitable .
We have these facts: by Lemma 3.10, both and tend to as , so is a function on ; ;
(3.29) |
for small with a remainder term depending on . Hence, we can specify a function to counter-effect the remainder term in and make the remainder term in (3.29) strictly negative. That is, we can specify a function such that
(3.30) | ||||
(3.31) |
for some negative function . Recall the definitions of , , , and by continuity, there exists a surface satisfying
This surface is a lower barrier in the sense of Definition 1.5.
Proof of item (3) of Theorem 1.2.
Assume that and do not have isometric tangent cone at , then we can construct a barrier such that in and the angle along . But due to item (4) of Theorem 1.2, this is not possible. So and have isometric tangent cones at , then by the construction of the foliation in Theorem 3.3, again we have a barrier near , but the barrier condition is now not strict. We can extend the rigidity in item (4) of Theorem 1.2 beyond the barrier and to all of . ∎
Remark 3.13.
By considering only the mean curvature, this provides an alternative proof of Theorem 1.1 in dimension 3. Moreover, we allow conical metrics of at two antipodal points.
Remark 3.14.
During the construction of barriers in the case of non-isometric cones, the Gauss-Bonnet theorem is only used in Lemma 3.7.
4. Construction of barriers (II)
In this section, we prove items (2) and (1) of Theorem 1.2. Our method is similar to the previous work [CW23].
4.1. Proof of item (2) of Theorem 1.2
For convenience, we set . We will construct a lower barrier near . As before, for any , we set to be the -level set of and to be the -sublevel set, that is, all points of which lie below . We see from the assumption on the tangent cone at that the sequence converges to some right circular cone in equipped with a flat metric as . Then converges to the same cone but with a different constant metric . The cone can be represented as a cone in the Euclidean space .
We have the existence of a barrier if and have non-isometric tangent cones at .
Lemma 4.1.
Proof.
First, we note that the mean curvature comparison and the metric comparison (we only need boundary metric comparison) are preserved in the limits. Because that the tangent cone of at is a round (solid) circular cone and that converges to its axial section as , it follows from the angle comparison of [CW23, Proposition 4.9] that there exists a plane in such that the the dihedral angles formed by and in the metric are everywhere larger than .
We gain a lot of freedom to construct the barrier from the strict comparison of angles. The rest of the argument is analogous to [CW23, Proposition 4.10]. ∎
Remark 4.2.
Note that the scalar curvature comparison is not needed here.
Proof of item (2) of Theorem 1.2.
First, the tangent cones of and at must be isometric. Indeed, by Lemma 4.1 and item (4) of Theorem 1.2, the barrier constructed in Lemma 4.1 cannot have in and hold strictly along .
By following Subsection 3.2, we can construct graphical perturbations of which satisfy Proposition 3.3. For every sufficiently small , is a barrier in the sense of Definition 1.5, we conclude that for the region bounded by and for every from item (4) of Theorem 1.2. Hence, we finish the proof of item (2) of Theorem 1.2. ∎
4.2. Proof of item (1) of Theorem 1.2
This part is a slightly extension of the argument in Section 5 in our previous paper [CW23] and Ko-Yao’s paper [KY24]. So we only sketch the key steps here and refer to the above papers for more details.
Again, we set and we construct a lower barrier near . Suppose is given by
near , where is a smooth function such that and is positive definite at under metric . Note that we need to assume , otherwise, the manifold will have a cusp at point . For simplicity, we assume , where is determined by (1.5).
To better illustrate the situation, we can choose the coordinate on such that the expansion of metric in (1.5) at is given by
where .
For simplicity, we denote as the linearised part of at . After a suitable rotation, we can write for some constant metric defined as
(4.1) |
where the matrix
is positive definite and satisfies .
We assume the manifold can be written as
where is a smooth function with . Here, we have used the fact that is positive definite at under metric .
We write as the inverse matrix of , and define several constants as
and consider the function defined by
and the surface is defined by
We use an ellipse to parameterize where is given by
Then, the surface can be written as a map such that
where satisfies
We also use for short. We have the following result by the argument in [CW23].
Proposition 4.3.
Suppose the metric can be written as where is the constant metric defined in (4.1) and is a bounded symmetric two-tensor. Then, we have
for any . Here, is a bounded term (not related to and ) which is also odd symmetric with respect to , is a bounded term (not related to ) relying on linearly.
Sketch of the proof.
We use the same argument for Proposition 5.1 in [KY24] and also track the term to get the following expansions
where we assume . Together with the remaining computation for Proposition 5.1 in [KY24] and the Corollary 5.5 in [CW23] (see the proof for Corollary 5.17 in [CW23]), we can establish the result. ∎
As a corollary of Proposition 4.3, we can easily establish the following results for ,
(4.2) | ||||
(4.3) |
and the following proposition.
Proposition 4.4.
Suppose the conditions in Proposition 4.3 hold. Then, for any , we can find (might rely on ) such that for any , we have
(4.4) |
for any .
We need to analyze the asymptotic behavior of mean curvature. We define the following mean curvatures:
Using the same computation for Corollary 5.2 in [KY24], we have
Proposition 4.5.
Suppose the metric can be written as where is a constant metric defined in (4.1), and is a bounded symmetric two-tensor. Then, we have the following formula for the behavior of mean curvature
(4.5) |
for any . Here, we write and for short.
We have two subcases to consider.
If , then we can use the continuation of with respect to and , together with Proposition 4.4, we can show the following results (cf. Proposition 5.10 in [CW23]).
Proposition 4.6.
Suppose the metric can be written as where is a constant metric defined in (4.1), and is a bounded symmetric two-tensor. If , we can choose some small such that for any and for each .
Now, we focus on the case . In particular, it implies and .
Then, we need to construct a foliation near . We define the vector field . Given where , we can define the perturbation surface by
where is the unit normal vector field of .
We write . Using the variational formula for mean curvatures and contact angles, we have
where denotes the Laplacian-Beltrami operator on under the metric , and is the unit normal vector field of under the metric . Here, we have used (4.3).
By using the same argument for Proposition 5.27 in [CW23], together with the asymptotic behavior of mean curvature, for each sufficiently small, we can find such that the mean curvature is where is a function only depends on , the contact angle , and satisfies the following
for any . A finer analysis of will give for sufficiently small (cf. Proposition 5.28 in [CW23]), and it leads to the following.
Proposition 4.7.
We can construct a surface near such that the mean curvature of is not greater than and it has prescribed contact angle with .
References
- [Amb15] Lucas C. Ambrozio. Rigidity of area-minimizing free boundary surfaces in mean convex three-manifolds. J. Geom. Anal., 25(2):1001–1017, 2015.
- [BBHW24] Christian Bär, Simon Brendle, Bernhard Hanke, and Yipeng Wang. Scalar curvature rigidity of warped product metrics. SIGMA Symmetry Integrability Geom. Methods Appl., 20:Paper No. 035, 26, 2024.
- [BBI01] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in Metric Geometry, volume 33 of Graduate Studies in Mathematics. American Mathematical Society, Providence, Rhode Island, 2001.
- [BBN10] Hubert Bray, Simon Brendle, and Andre Neves. Rigidity of area-minimizing two-spheres in three-manifolds. Comm. Anal. Geom., 18(4):821–830, 2010.
- [BM11] Simon Brendle and Fernando C. Marcques. Scalar Curvature Rigidity of Geodesic Balls in . Journal of Differential Geometry, 88(3):379–394, 2011.
- [CLZ24] Jianchun Chu, Man-Chun Lee, and Jintian Zhu. Llarull’s theorem on punctured sphere with metric. arXiv: 2405.19724, 2024.
- [CW23] Xiaoxiang Chai and Gaoming Wang. Scalar curvature comparison of rotationally symmetric sets. arXiv/2304.13152, 2023.
- [CW24] Xiaoxiang Chai and Xueyuan Wan. Scalar curvature rigidity of domains in a warped product. arXiv:2407.10212 [math], 2024.
- [CWXZ24] Simone Cecchini, Jinmin Wang, Zhizhang Xie, and Bo Zhu. Scalar curvature rigidity of the four-dimensional sphere. arXiv: 2402.12633v2, 2024.
- [CZ24] Simone Cecchini and Rudolf Zeidler. Scalar and mean curvature comparison via the Dirac operator. Geom. Topol., 28(3):1167–1212, 2024.
- [DPM15] G. De Philippis and F. Maggi. Regularity of free boundaries in anisotropic capillarity problems and the validity of Young’s law. Arch. Ration. Mech. Anal., 216(2):473–568, 2015.
- [FCS80] Doris Fischer-Colbrie and Richard Schoen. The structure of complete stable minimal surfaces in 3-manifolds of non-negative scalar curvature. Communications on Pure and Applied Mathematics, 33(2):199–211, 1980.
- [Gro21] Misha Gromov. Four Lectures on Scalar Curvature. arXiv:1908.10612 [math], 2021.
- [HKKZ] Sven Hirsch, Demetre Kazaras, Marcus Khuri, and Yiyue Zhang. Rigid comparison geometry for riemannian bands and open incomplete manifolds. (arXiv:2209.12857).
- [HLS23] Yuhao Hu, Peng Liu, and Yuguang Shi. Rigidity of 3D spherical caps via -bubbles. Pacific J. Math., 323(1):89–114, 2023.
- [KY24] Dongyeong Ko and Xuan Yao. Scalar curvature comparison and rigidity of -dimensional weakly convex domains. arXiv:2410.20548, 2024.
- [Li20] Chao Li. A polyhedron comparison theorem for 3-manifolds with positive scalar curvature. Invent. Math., 219(1):1–37, 2020.
- [Lis10] Mario Listing. Scalar curvature on compact symmetric spaces. arXiv:1007.1832 [math], 2010.
- [Lla98] Marcelo Llarull. Sharp estimates and the Dirac operator. Math. Ann., 310(1):55–71, 1998.
- [Lot21] John Lott. Index theory for scalar curvature on manifolds with boundary. Proc. Amer. Math. Soc., 149(10):4451–4459, 2021.
- [MO89] Maung Min-Oo. Scalar curvature rigidity of asymptotically hyperbolic spin manifolds. Math. Ann., 285(4):527–539, 1989.
- [MP21] Pengzi Miao and Annachiara Piubello. Mass and Riemannian Polyhedra. arXiv:2101.02693 [gr-qc], 2021.
- [RS97] Antonio Ros and Rabah Souam. On stability of capillary surfaces in a ball. Pacific J. Math., 178(2):345–361, 1997.
- [SY79a] R. Schoen and Shing Tung Yau. Existence of incompressible minimal surfaces and the topology of three-dimensional manifolds with nonnegative scalar curvature. Ann. of Math. (2), 110(1):127–142, 1979.
- [SY79b] Richard Schoen and Shing Tung Yau. On the proof of the positive mass conjecture in general relativity. Comm. Math. Phys., 65(1):45–76, 1979.
- [WWZ24] Jinmin Wang, Zhichao Wang, and Bo Zhu. Scalar-mean rigidity theorem for compact manifolds with boundary. arXiv: 2409.14503, 2024.
- [WX23a] Jinmin Wang and Zhizhang Xie. Dihedral rigidity for submanifolds of warped product manifolds. arXiv: 2303.13492, 2023.
- [WX23b] Jinmin Wang and Zhizhang Xie. Scalar curvature rigidity of degenerate warped product spaces. arXiv: 2306.05413, 2023.
- [Ye91] Rugang Ye. Foliation by constant mean curvature spheres. Pacific Journal of Mathematics, 147(2):381–396, 1991.