Compactness of sequences of warped product circles over spheres with nonnegative scalar curvature
Abstract.
Gromov and Sormani conjectured that a sequence of three dimensional Riemannian manifolds with nonnegative scalar curvature and some additional uniform geometric bounds should have a subsequence which converges in some sense to a limit space with some generalized notion of nonnegative scalar curvature. In this paper, we study the pre-compactness of a sequence of three dimensional warped product manifolds with warped circles over standard that have nonnegative scalar curvature, a uniform upper bound on the volume, and a positive uniform lower bound on the , which is the minimum area of closed minimal surfaces in the manifold. We prove that such a sequence has a subsequence converging to a Riemannian metric for all , and that the limit metric has nonnegative scalar curvature in the distributional sense as defined by Lee-LeFloch.
1. Introduction
In [8] and [7], Gromov conjectured that a sequence of Riemannian manifolds with nonnegative scalar curvature, , should have a subsequence which converges in some weak sense to a limit space with some generalized notion of “nonnegative scalar curvature”. In light of the examples constructed by Basilio, Dodziuk, and Sormani in [2], the condition in (2) below was added to prevent collapsing happening, and the conjecture was made more precise at an IAS Emerging Topics Workshop co-organized by Gromov and Sormani as follows [18]:
Conjecture 1.1.
Let be a sequence of closed oriented three dimensional Riemannian manifolds without boundary satisfying
(1) |
(2) |
Then there exists a subsequence which is still denoted as that converges in the volume preserving intrinsic flat sense to a three dimensional rectifiable limit space . Furthermore, is a connected geodesic metric space, that has Euclidean tangent cones almost everywhere, and has nonnegative generalized scalar curvature.
In a joint work with Jiewon Park [15], the authors confirmed Conjecture 1.1 for sequences of rotationally symmetric Riemannian manifolds . In our proof the condition provides a uniform lower bound for the warping functions in the closed region between any two minimal surfaces. As a result, we can prevent counter examples like the sequence of round spheres shrinking to a point, and we can also prevent the formation of thin tunnels between two non-collapsed regions. The regularity of the limit metric is high, and the convergence of the sequence of warping functions is strong. In particular, in [15] we proved that the limit warping function is Lipschitz and that the sequence of warping functions converges to the limit function in the norm in closed regions away from the two poles.
In this paper, we study the warped product case of the Conjecture 1.1. We consider the following:
Definition 1.2.
Let be a sequence of Riemannian manifold such that
(3) |
where and are the standard metrics on and respectively, and the function is smooth for each . Here and are the geodesic polar coordinate for . We also use the notation to denote .
We consider the convergence of the warping function and prove the sharp regularity of the limit warping function in the following theorem:
Theorem 1.3.
Let be a sequence of warped product Riemannian manifolds such that each has non-negative scalar curvature. If we assume that
(4) |
then we have the following:
-
(i)
After passing to a subsequence if needed, the sequence of warping functions converges to some limit function in for all .
-
(ii)
The limit function is in , for all such that .
-
(iii)
The essential infimum of is strictly positive, i.e. .
-
(iv)
If we allow as a limit, then the limit
(5) exists for every . Moreover, is lower semi-continuous and strictly positive everywhere on , and a.e. on .
The definition of essential infimum is given in Definition 4.6. In the proof of convergence properties in items (i) and (ii) in Theorem 1.3, we only need nonnegative scalar curvature condition and volume uniform upper bound condition. In the proof of part (iii) of Theorem 1.3, we make essential use of condition combined with the spherical mean inequality [Proposition 2.4], Min-Max minimal surface theory and a covering argument. This is an interesting new way of applying the condition to prevent collapsing. Then the part (iv) follows from (iii) and an interesting ball average monotonicity property [Proposition 2.6]. The ball average monotonicity is obtained from spherical mean inequality by using the trick as in the proof of Bishop-Gromov volume comparison theorem.
Remark 1.4.
The extreme example constructed by Sormani and authors in [19] shows that the regularity for is sharp for the limit warping function .
Proposition 1.5.
Let be a sequence of warped product manifolds such that each has non-negative scalar curvature, and the sequence satisfies conditions in . Then there exists such that , for all and , where obtained in Theorem 1.3.
As an application of Proposition 1.5, we have:
Corollary 1.6.
Let be a sequence of warped product manifolds such that each has non-negative scalar curvature, and the sequence satisfies conditions in . Then the systoles of , for all , have a uniform positive lower bound given by , where obtained in Theorem 1.3.
The systole of a Riemannian manifold is defined to be the length of the shortest closed geodesic in the manifold [Definition 4.16]. In order to estimate systole of warped product manifolds: , in Lemma 4.18 we establish an interesting dichotomy property for closed geodesics in a general warped product manifold with as a typical fiber, with metric tensor as , where is a -dimensional complete Riemannian manifold without boundary and is a positive smooth function on . The dichotomy property in Lemma 4.18 has its own interests independently, and shall be useful in other studies of closed geodesics in such warped product manifolds.
The convergence of the warping functions in Theorem 1.3 leads to the convergence of the Riemannian metrics, we prove the following:
Theorem 1.7.
Let be a sequence of warped product Riemannian manifolds such that each has non-negative scalar curvature. If we assume that
(6) |
Then there exists a subsequence and a (weak) warped product Riemannian metric for such that
(7) |
Theorem 1.7 is proved in §5.1. The definition of a (weak) warped product Riemannian metric is given in Definition 5.1, and the spaces and are defined in Definition 5.3. The condition is used to prevent converging to a non-metric tensor in , with the help of the non-collapsing property of in the item (iii) in Theorem 1.3.
In the limit space we calculate the scalar curvature as a distribution using the definition by Lee and LeFloch [10], and we prove the following:
Theorem 1.8.
Theorem 1.8 is proved in §5.2. In general, it is still an interesting and difficult problem to formulate suitable notions of generalized (or weak) nonnegative scalar curvature in Conjecture 1.1. A natural candidate is the volume-limit notion of nonnegative scalar curvature. But recently Kazara and Xu constructed a sequence of warped product metrics on whose limit space does not have nonnegative scalar curvature in the sense of volume-limit in Theorem 1.3 in [9]. There are other candidates, like Gromov’s polyhedron comparison notion [7, 12] and Burkhardt-Guim’s Ricci flow notion [4] of nonnegative scalar curvature for -metrics. However, as mentioned in Remark 1.4, the regularity, for , is the best regularity for our limit metrics, and in general our limit metrics are not continuous. Lee and Lefloch [10] defined the scalar curvature distribution for -metrics. Our limit metric obtained in Theorem 1.7 does not satisfy the regularity requirement in [10], but when we add up different terms in the integrand, the divergent terms cancel with each other and the scalar curvature is still well defined as a distribution. This is discussed in detail in Remark 5.18. Interestingly, we obtain the continuity of distributional total scalar curvature in Theorem 1.8. More importantly, the scalar curvature distribution of Lee-LeFloch enables us to see the concentration of scalar curvature on the singular set, see §4.4 in [19].
In Appendix A, we study pre-compactness of the sequence of warped product spheres over circle , that is, are diffeomorphic to with warped product metric tensors
(8) |
The study of this case is similar to the rotationally symmetric case studied in [15]. The key is to obtain a uniform bound for the norm of gradient of from nonnegative scalar curvature condition [Lemma A.4]. By combining this with uniform diameter upper bound and the condition, we prove that a subsequence of converges in and sense to a bounded positive Lipschitz function [Theorem A.1]. Moreover, we prove that the limit Riemannian metric has nonnegative distributional scalar curvature in the sense of Lee-LeFloch [Theorem A.2].
The proof of Theorem A.1 is similar to that of Theorems 4.1 and 4.8 in [15]. We include it here to show the difference with the rotationally symmetric case and the difference with Theorem 1.3 and Theorem 1.7.
The proof of Theorem A.2 shows that in this case the regularity requirement in Lee-LeFloch [10] is essential for the definition of the scalar curvature as a distribution. This provides an interesting contrast with the proof of Theorem 1.8.
The article is organized as follows: in Section 2, we derive several analysis properties of warping functions from the uniform geometric bounds of metric as in (3). In particular, we show that metrics in (3) have nonnegative scalar curvature if and only if the warping functions satisfy the differential inequality [Lemma 2.1]:
(9) |
where is the Lapacian on the standard round sphere , taken to be the trace of the Hessian. Moreover, a positive number is a uniform upper bound of volumes of metrics in (3) if and only if satisfy [Lemma 2.2]
(10) |
It is well-known that the spherical mean property of (sub, sup)-harmonic functions plays important roles in the study of these functions. Inspired by this, we prove a spherical mean inequality for functions satisfying the differential inequality (9) [Proposition 2.4]. It turns out that the spherical mean inequality is very important in the proof of non-collapsing property in Section 4, in particular, in the proof of Proposition 4.10. Furthermore, by employing the trick in the proof of Bishop-Gromov volume comparison theorem, we prove a ball average monotonicity property for [Proposition 2.6], which helps us to obtain lower semi-continuity of the limit warping function in Proposition 3.7.
In Section 3, we study the convergence of a sequence of positive functions on satisfying (9) and (10). We prove that there exists a subsequence of such sequence and a function such that the subsequence converges to in for any [Proposition 3.5]. The proof of this convergence result is very different from that in cases of warped product metrics as in [15] and in (8). Because warping functions in [15] and in (8) have one variable, whereas in (3) have two variables, it is more difficult to obtain sub-convergence of , and we make use of the Moser-Trudinger inequality in (25) in [14]. The regularity of the limit function is weaker than . The extreme example constructed by Sormani and authors in [19] shows that the regularity for is sharp for .
In Section 4, we use the condition to show that the limit function has positive essential infimum [Theorem 4.13] and that the warping functions have a positive uniform lower bound [Proposition 4.15]. This enables us to define weak warped product Riemnnian metric on in Definition 5.1, and is crucial in the study of geometric convergence of warped product circles over sphere with metric tensor as in (3). Moreover, as a consequence of Proposition 4.15, we obtain a positive uniform lower bound for the systole of the warped product manifolds [Proposition 4.20].
The condition can be viewed as a noncollapsing condition. As shown in [15] and in Lemma A.6 below, it is not difficult to see this in cases of metric tensors as in [15] and (8). In the case of metric tensors as in (3), however, the implication of the condition is much more complicated. We need to use the Min-Max minimal surface theory of Marques and Neves (see e.g. [13]), the maximum principle for weak solutions (Theorem 8.19 in [6]), and the spherical mean inequality obtained in Proposition 2.4, in order to obtain noncollapsing from the condition.
In Section 5, we prove that a subsequence of , with as in (3) having nonnegative scalar curvatures and uniform upper bounded volumes and satisfying condition, converges to a weak metric tensor in the sense of for all [Theorem 5.5]. Moreover, we prove that the limit metric has nonnegative distributional scalar curvature in the sense of Lee-LeFloch [Theorem 5.11].
Note that in the case of metric tensors as in [15] and (8), we need the diameter uniform upper bound condition in addition to nonnegative scalar curvature condition and the condition for getting convergence [Theorem 1.3 in [15] and Theorem A.1], whereas in the case of metric tensors as in (3), we need the volume uniform upper bound condition instead of the diameter uniform upper bound condition [Theorem 5.5].
Acknowledgements:
The authors would like to thank the Fields Institute for hosting the Summer School on Geometric Analysis in July 2017 where we met Professor Christina Sormani and she started to guide us working on the project concerning compactness of manifolds with nonnegative scalar curvatures. We are grateful to Professor Sormani for her constant encouragement and inspiring discussions. In particular, Professor Sormani suggested us the method of spherical means, and it turns out to be very useful in the study of warping functions in Theorem 1.3. We thank Brian Allen for discussions and interest in this work. Wenchuan Tian was partially supported by the AMS Simons Travel Grant. Changliang Wang was partially supported by the Fundamental Research Funds for the Central Universities and Shanghai Pilot Program for Basic Research.
2. Consequences of the geometric hypotheses on
In this section we prove several consequences of the uniform geometric bounds. In Subsection 2.1, we derive the differential inequality satisfied by the warping function and prove that the uniform volume bounds on sequence of Riemannian manifolds implies the uniform norm of the warping function.
In Subsection 2.2, we prove the spherical mean inequality for the warping function [Proposition 2.6], which is our main analytic tool. In Subsection 2.3, we prove a ball average monotonicity property for the warping function [Proposition 2.4].
The implication of the condition is more complicated we discuss that in Section 4.
2.1. Basic consequences of the hypotheses
Lemma 2.1 (Non-negative scalar curvature condition).
The scalar curvature of warped product manifolds are given by
(11) |
where is the Laplacian on with respect to the standard metric , taken to be the trace of the Hessian (without the negative sign).
Thus have nonnegative scalar curvature if and only if
(12) |
Proof.
By using the Ricci curvature formula for warped product metrics as in Proposition 9.106 of [3], we can easily obtain the scalar curvature of as Then the second claim directly follows, since . ∎
Lemma 2.2 (Volume upper bound condition).
The warped product manifolds have volume if and only if
(13) |
Proof.
The Riemannian volume measure of is given by
(14) |
Thus the volume of is given by
(15) |
Then the claim directly follows. ∎
2.2. Spherical mean inequality
In this subsection, we prove a spherical mean inequality [Proposition 2.4] for the smooth functions on satisfying the differential inequality . By Lemma 2.1, this is equivalent to studying the warping function of warped product manifolds with nonnegative scalar curvature. The spherical mean inequality plays an important role in the proof of Proposition 4.10.
The derivation of the spherical mean value inequality is similar to that of the mean value property of harmonic functions. We start with the following lemma.
Lemma 2.3.
Let be a smooth function on . Consider the spherical mean given by
(16) |
where is the geodesic ball in the standard with center and radius . The derivative of satisfies
(17) |
Proof.
Using the geodesic polar coordinate on centered at , one can write as
(18) |
Then taking derivative with respective to gives
(19) | |||||
(20) | |||||
(21) | |||||
(22) | |||||
(23) |
∎
Now we use Lemma 2.3 to prove the spherical mean inequality.
Proposition 2.4.
Let be a smooth function on satisfying . Then for any fixed and , one has
(24) |
where is the geodesic ball in the with center and radius .
Moreover, by taking limit as , one has
(25) |
for any .
Proof.
By Lemma 2.3 and the assumption , one has
(26) |
Integrating this differential inequality for from to gives
(27) | |||||
(28) | |||||
(29) | |||||
(30) | |||||
(31) | |||||
(32) |
∎
2.3. Ball average monotonicity
In this subsection, we further derive a ball average monotonicity [Proposition 2.6] for a smooth function on satisfying . The proof uses the spherical mean inequality [Proposition 2.4] and the trick as in the proof of Bishop-Gromov volume comparison theorem. This ball average monotonicity is used in Proposition 3.7 to prove that the ball average limit as exists everywhere for the limit function.
Lemma 2.5.
Let be a smooth function on satisfying and , where is a positive constant. For any fixed , the spherical mean
(33) |
is a non-increasing function in for
Proof.
The spherical mean inequality in Proposition 2.4 says that for any and ,
(34) |
By rearranging this inequality, we obtain that for any fixed ,
(35) |
This completes the proof. ∎
Combine this spherical mean monotonicity with the trick as in the proof of Bishop-Gromov volume comparison theorem, we obtain the following ball average monotonicity.
Proposition 2.6.
Let be a smooth function on satisfying and , then ,
(36) |
where is the distance between and in the standard .
3. limit of warping function for
In this section, we study the pre-compactness of a sequence of positive smooth functions satisfying the inequalities
(56) |
Here is a positive constant. By Lemmas 2.1 and 2.2, the inequlities in (56) are equivalent to the requirements that the Riemannian manifolds have nonnegative scalar curvature and uniform volume upper bound.
In Subsection 3.1, we prove that a sequence of positive smooth functions on satisfying requirements in (56) has a convergent subsequence in for any , and that the limit function is in for any [Proposition 3.5].
In Subsection 3.2, we apply the ball average monotonicity property obtained in Proposition 2.6 to prove that the limit function has a lower semi-continuous representative [Proposition 3.7, Remark 3.8].
3.1. limit function for
We first derive the gradient estimate for the sequence of function in Lemma 3.1, which is used to obtain estimate for by using Moser-Trudinger inequality in Lemma 3.2.
Lemma 3.1.
Let be a sequence of positive functions on satisfying
(57) |
We have
(58) |
Proof.
Note that
(59) |
By equation (59) and the assumption, we have
(60) |
Integrating it over , and using Stokes’ theorem, we get
(61) |
∎
Lemma 3.2.
Let be a sequence of positive functions on satisfying
(62) |
Then we have
(63) |
for all and .
Proof.
By the Moser-Trudinger inequality (inequality (25) in [14]), for any smooth function we have
(64) |
Here is the Levi-Civita connection of the standard metric and is the volume form on with respect to the standard metric . Take , then we have
(65) | |||||
(66) | |||||
(67) |
By the fact that , we have
(68) |
On the hand, by Lemma 3.1 we have
(69) |
This completes the proof. ∎
Next, we show that such sequence of function is uniformly bounded in for .
Lemma 3.3.
Let be a sequence of positive functions on satisfying
(70) |
Then the sequence is uniformly bounded in for , i.e. for each , there exists a constant such that
(71) |
Proof.
We use the uniform bound to prove convergence in the following lemma.
Lemma 3.4.
Let be a sequence of positive functions on satisfying
(78) |
Then for each fixed , there exists a subsequence and such that
(79) |
for each .
Moreover, for any ,
as , where is the weak gradient of .
Proof.
For each fixed , by using Rellich-Kondrachov compactness theorem, the uniform estimate of Sobolev norms in Lemma 3.3 implies that there exists a subsequence of , which is still denoted by , converging to in for . Then by the weak compactness in space (see, e.g. Theorem 1.42 in [5]), we can obtain that . Indeed, for all implies that and are both uniformly bounded. Then the weak compactness in space implies that there exist a further subsequence, denoted by , and such that
(80) |
i.e.
(81) |
On the other hand,
(82) |
since in . Thus,
(83) |
Therefore, is the gradient of in the sense of distribution, and so . For any , by taking in (81), we obtain
(84) |
∎
Now we use Lemma 3.4 and diagonal argument to find a subsequence converging in for all and prove the following proposition:
Proposition 3.5.
Let be a sequence of positive functions on satisfying
(85) |
Then there exists a subsequence and for all , such that
(86) |
Moreover, for any ,
(87) |
as , where is the weak gradient of .
Proof.
The proof is a diagonal argument. We apply Lemma 3.4 for .
For , by applying Lemma 3.4 to and , we obtain a subsequence, denoted by , and such that
(88) |
For , by applying Lemma 3.4 to the subsequence and , we obtain a subsequence, , and such that
(89) |
Then by repeating this process for , we can obtain a family of decreasing subsequence and for all , such that for each fixed
(90) |
Now we take the diagonal subsequence . By the construction of and as , we have that is a Cauchy sequence in for all . Thus there exists such that
(91) |
Then by the uniqueness of limit, in for all . Furthermore, because and as , we see that the norm of the weak derivative of is bounded for any . Thus for all .
3.2. Lower semi-continuous representative of the limit function
For the limit function obtained in Proposition 3.5, Lebesgue-Besicovitch differential theorem implies that
(92) |
holds for a.e. with respect to the volume measure . In Proposition 3.7, by applying the ball average monotonicity property in Proposition 2.6, we will show that the limit of ball average in (92) actually exists for all , and that the limit produces a lower semi-continuous function.
Proposition 3.7.
Let be a sequence of smooth positive functions on satisfying
(93) |
Then the limit function, , obtained in Proposition 3.5, has the following properties.
-
(i)
For each fixed , the ball average
(94) is non-increasing in , where is a positive real number such that . Note that the existence of such is guaranteed by Lemma 3.2.
-
(ii)
Consequently, the limit
(95) exists, allowing as a limit, for every . Moreover, is a lower semi-continuous function on .
Proof.
By Lemma 3.2, there exists such that
(96) |
Then by applying Proposition 2.6 to functions , we obtain that for any fixed
(97) |
holds for any and all .
By Proposition 3.5 in . Then for any fixed , and any fixed , by taking the limit as , we obtain
(98) |
So for each fixed , the ball average
(99) |
is non-increasing for . Therefore, for any the limit
(100) |
exists as a finite number or .
On the other hand, by direct calculation
(101) |
as . Thus the limit
(102) |
exists for all .
For each fixed , we have that is a continuous function of , since , , and for all . Then by the monotonicity in (98), we have
(103) |
In other words, is the supremum of a sequence of continuous function. Thus is lower semi-continuous. ∎
Remark 3.8.
Recall that by (92), hold for a.e. , thus holds for a.e. . So as a function, has a lower semi-continuous representative .
4. Positivity of the limit warping functions
In this section, we prove that the limit warping function has a positive essential infimum, provided that the Riemannian manifold satisfies both requirements in (56) and the condition [Theorem 4.13]. The main tools we use in the proof of Theorem 4.13 include the maximum principle, the Min-Max minimal surface theory of Marques and Neves, and the spherical mean inequality we obtained in Proposition 2.4.
The maximum principle for weak solutions (Theorem 8.19 in [6]) requires regularity, but in general we only have for [Remark 3.6]. To overcome this difficulty, in Subsection 4.1, we consider the truncation of warping functions as defined in Definition 4.1, and obtain a limit function for the sequence of truncated function [Lemma 4.4]. This enables us to apply maximum principle for weak solutions (Theorem 8.19 in [6]) to , and prove that either or on [Proposition 4.7].
In Subsection 4.3, we use Min-Max minimal surface theory of Marques and Neves and the spherical mean inequality in Proposition 2.4 to obtain an upper bound for in terms of norm of the warping function , provided that the norm of is sufficiently small [Proposition 4.10].
In Subsection 4.4, we use Proposition 4.7 and Proposition 4.10 to prove Theorem 4.13. Moreover, as an application of Theorem 4.13, we obtain a positive uniform lower bound for warping functions , if the warped product manifolds satisfy requirements in (56) and the condition [Proposition 4.15].
4.1. regularity of limit of truncated warping functions
We define the truncation of a function firstly:
Definition 4.1.
Let be a positive smooth function. Let be a real number, for each , we define
(104) |
Then is a positive continuous function on with the maximal value not greater than .
From the definition we can prove the following lemma:
Lemma 4.2.
Let be a positive smooth function, and let be a regular value of the function . If
(105) |
then for all such that we have
(106) |
Proof.
By Theorem 4.4 from [5], we have for all
(107) |
As a result we have
(108) |
Here, since is a regular value of , from the Regular Level Set Theorem we know that the level set is am embedded submanifold of dimension in . Hence we can apply Stokes’ theorem to get the last step. Moreover, since is the outer unit normal vector on the boundary of the set , we have
(109) |
Hence we can drop the boundary term to get the inequality
(110) |
Since
(111) |
we have
(112) |
This finishes the proof. ∎
We can prove similar results for a sequence of functions:
Lemma 4.3.
Let be a sequence of smooth positive function defined on . If
(113) |
then there exists such that for all with we have
(114) |
Moreover, we can choose as large as we want.
Proof.
Note that if for some then we have . On the other hand, if for some then . Either way the inequality (114) holds.
In general, by Sard’s theorem, for each function , the critical values of has measure zero, and the union of all the critical sets for each of the function also has measure zero. As a result, there exists such that for each either is a regular value or . By Lemma 4.2 we get inequality (114). Moreover, we can choose as large as we want. This finishes the proof. ∎
Next we prove similar results for the limit function, but before that we need to consider the regularity of the limit function:
Lemma 4.4.
Let be a real number. Let be a sequence of positive smooth functions on satisfying
(115) |
Then the sequence is uniformly bounded in :
(116) |
As a result, there exists such that converges to in , and that converges to weakly in .
Proof.
Now we prove the following proposition concerning the limit function:
Lemma 4.5.
Proof.
By Lemma 4.4 we know that converges to in , and that converges to weakly in . As a result, for any we have that
(122) |
and that
(123) |
As a result, by (114) we have for all such that
(124) |
Hence by Theorem 8.19 in [6], we have that either the essential infimum of is bounded away from zero or is the zero function. This finishes the proof. ∎
We need the definition of essential infimum of a function:
Definition 4.6.
Consider the standard and use to denote the standard volume measure in . Let be an open subset of . Let be measurable. Define the set
(125) |
We use to denote the essential infimum of in and define
(126) |
Finally, we apply the maximum principle for weak solution to prove the following property for the essential infimum of .
Proposition 4.7.
Let be a sequence of positive smooth functions on satisfying
(127) |
If we further assume that in for some , then either the essential infimum of is bounded away from zero or a.e. on .
Proof.
Since as , choose a subsequence if needed, then we have poiintwise almost everywhere in . Let be a real number that satisfies the requirement in Lemma 4.3. Construct a truncated sequence as in Definition 4.1. By Lemma 4.4, choose a subsequence if needed, there exists such that converges to in norm. As a result, choose a subsequence if needed we have pointwise almost everywhere in .
It suffices to show that if the essential infimum then in . We assume that . Since for each we have , we have . This implies that for any , we have
(128) |
and
(129) |
Let be the north pole of , and be the south pole. and are upper and lower hemispheres respectively. Then either
(130) |
or
(131) |
Without loss of generality we assume that . Since in , for any , and such that we have
(132) |
Now by Lemma 4.5, satisfies
(133) |
on in the weak sense. Hence by the strong maximum principle for weak solutions (see Theorem 8.19 in [6]), the equality in (132) implies that is constant on . This is true for any , thus on . Moreover, since , for almost every we have,
(134) |
and hence a.e. on . This finishes the proof. ∎
4.2. A -sweepout of the warped product manifold
Because we will apply the Min-Max minimal surface theory to get an upper bound for in §4.3, in this subsection we briefly recall some basic notions in geometric measure theory following Marques and Neves [13], and construct a -sweepout for , which will be used in the proof in Lemma 4.11. For an excellent survey and more details about these materials we refer to [13] and references therein.
A -current on is a continuous linear functional on the space of compactly supported smooth -forms: . Its boundary is a -current that is defined as for . A -current is said to be an integer multiplicity -current if it can be written as
(135) |
where is a -measurable countable -rectifiable set, that is with and is an embedded -dimensional -submanifold for all , is a -integrable -valued function, and is a -form such that is a volume form for at where a -dimensional tangent space is well-defined. Note that this tangent space is well-defined for -a.e. , provided for every compact set . Also note that the form give an orientation for . The mass of an integer multiplicity -current is defined as
(136) |
where is the pointwise maximal norm of a form .
In particular, a -dimensional embedded smooth submanifold of can be viewed as an integer multiplicity -current by integrating a -form over it. Its current boundary is given by its usual boundary, and its mass is the -dimensional volume of the submanifold.
Let be a manifold embedded in . The space of integral -currents on , denoted by , is defined to be the space of -current such that both and are integer multiplicity currents with finite mass and support contained in . The space of -cycles, denoted by , is defined to be the space of those so that for some .
A rectifiable -varifold V is defined to be a certain Radon measure on , where is the Grassmannian of -planes in . An integral -current given as in naturally associates a rectifiable -varifold, denoted by , as
(137) |
Here is the natural projection map from to , and is rank- tangent bundle of consisting of at where its -dimensional tangent plane can be well defined. Note that: in the varifold expression of , we forget the orientation of determined by the -form in the current expression of .
The space can be endowed with various metrics and have different induced topologies. Given , the flat metric is defined by
and induces the flat topology on . We also denote and have
(138) |
For , the F-metric is defined by Pitts in [16] as:
(139) |
where is the F-metric on the associated varifolds defined on page 66 in [16] as:
Recall that (see page 66 in [16])
(140) |
and hence
(141) |
For the Min-Max theory for minimal surfaces, the space of mod integral -currents and mod -cycles are also needed. They are denoted by and , respectively, and defined by an equivalence relation: if for . The notions of boundary, mass and metrics defined above for can be extended to . For a -dimensional manifold , the Constancy Theorem (Theorem 26.27 in [17]) says that if has , then either or .
Then we recall some basic facts about the topology of , that is endowed with flat metric. Their proofs can be found in [13], also see [1]. Let be the dimension of the manifold . Then is contractible and the continuous map
(142) |
is a -fold covering map. The homotopy groups are:
(143) |
For the calculation of the fundamental group, one notes that the map
(144) | |||||
(145) |
is an isomorphism. Here is a loop in with , and is the unique lift to with . Then by applying Hurewicz Theorem, one can obtain:
(146) |
The the action of the fundamental cohomology class on a homology class induced by a loop is nonzero if and only if the loop is homotopically non-trivial.
We take the following definition of -sweepout from [13].
Definition 4.8.
A continuous map is called a -sweepout if .
Here is the space endowed with the F-metric given in .
Now we return back our warped product manifold , that is with Riemannian metric
(147) |
For each fixed , we construct a -sweepout of consisting of tori , where denotes the geodesic ball on centered at with radius . In other words, we consider the map
(148) | ||||
Lemma 4.9.
The map given in provides a -sweepout of as in Definition 4.8.
Proof.
Clearly, , and hence can be viewed as a map from to by identifying the end points of the interval .
Now we show the continuity of the map on . This is clear for , since varies smoothly for . Then the continuity at follows from the inequality in and the estimate:
(149) |
as , since the warping function is smooth on . The continuity at follows similarly, since as .
Because by the definition flat metric is less than or equal to F-metric, is also continuous if we endow the flat metric on . So is a loop in , and represents a non-trivial element:
(150) |
This is because by the definition of the map we have that the unique lift of with is given by
(151) | ||||
and has . Consequently, , and so is a -sweepout. ∎
4.3. Bound from above by -norm of warping function
In this subsection, we derive an upper bound for in terms of , provided that is small relative to .
Proposition 4.10.
Let be a warped product Riemannian manifolds with metric tensor as in (3) that has nonnegative scalar curvature and . If , then we have .
Recall that is the infimum of areas of closed embedded minimal surfaces in . Proposition 4.10 is crucial in the proof of Theorem 4.13 below. In order to prove Proposition 4.10, we first prove the following two lemmas.
First of all, we use the Min-Max minimal surface theory of Marques and Neves to bound from above by areas of some tori in .
Lemma 4.11.
Let be a warped product Riemannian manifold with metric tensor as in (3). For each , there exists a torus , , whose area is not less than , i.e.
(152) |
where is the geodesic ball in the standard centered at with radius .
Proof.
We will use Min-Max minimal surface theory of Marques and Neves to prove the lemma.
For each fixed point , by Lemma 4.9, the map in gives a -sweepout of as in Definition 4.8. For , the image are tori in with mass:
(153) |
Clearly, is a continuous function of on with . Thus there exist such that
(154) |
Let be the homotopy class of the -sweepout , which consists of all continuous maps with such that and are homotopic to each other in the flat topology. By Lemma 2.2.6 in [13], the width
(155) |
since is a -sweepout and so is a non-trivial homotopy class. Then Min-Max Theorem of Marques-Neves (see Theorem 2.2.7 in [13]) implies that there exists a smooth embedded minimal surface in achieving the width, i.e. .
Finally, by the definitions of the width in (155) and , and by the choice of , we have
(156) |
Because is an arbitrary point on , this completes the proof. ∎
Next, we apply Lemma 4.11 and the spherical mean inequality from Proposition 2.4 to prove the following lemma.
Lemma 4.12.
Let be a warped product Riemannian manifold with metric tensors as in (3) that have non-negative scalar curvatures and . If , then there exists a set satisfying that for each there exists such that
-
(i)
,
-
(ii)
and
(157) holds for all .
Proof.
For any point , we denote its antipodal point by . By Lemma 4.11, for any , there exists such that the torus in has area
(158) |
Since , we have
(159) |
Thus, we have
(160) |
Now if , then we include the point in the set , and if , then we include its antipodal point in the set , and we set . Then we still have
(161) |
since .
By the construction of the set , contains at least one of any pair of antipodal points on , and for any , there exists such that
(162) |
Then we have that the area of the open set is at least half of the area of the whole sphere , i.e.
(163) |
Indeed, otherwise, we have
(164) |
On the other hand, because for each either or is contained in , we have
(165) |
So
(166) | |||||
(167) | |||||
(168) |
This gives a contradiction. So we have .
Because has non-negative scalar curvature, by Lemma 2.1, we have . Then by the spherical mean inequality in Proposition 2.4, for any and any we have that
(169) |
since and . By rearrange the inequality, we obtain that for any and any ,
(170) | |||||
(171) | |||||
(172) | |||||
(173) |
∎
Proof of Proposition 4.10.
By Lemma 4.12, there exists a set such that
(174) |
and for any , there exists such that
(175) |
holds for all .
By the Vitali covering theorem, there exists a countable sequence of points such that the collection of balls are disjoint with each other, and that
(176) |
By Lemma 4.12 we have
(177) |
As a result, we have
(178) |
Integrating this inequality from to gives
(179) | |||||
(180) | |||||
(181) |
Then by summing the above inequalities for together, we obtain
(182) |
since are disjoint balls. In the standard we have
(183) |
As a result, we have
(184) | |||||
(185) | |||||
(186) | |||||
(187) | |||||
(188) |
This completes the proof. ∎
4.4. Positivity of the limit of warping functions
In this subsection, we use Proposition 4.7 and Proposition 4.10 to prove Theorem 1.3, we restate it here for the convenience of the reader
Theorem 4.13.
Let be a sequence of warped product manifolds such that each has non-negative scalar curvature. If we assume that
(189) |
then we have the following:
-
After passing to a subsequence if needed, the sequence of warping functions converges to some limit function in for all .
-
The limit function is in , for all such that .
-
The essential infimum of is strictly positive, i.e. .
-
If we allow as a limit, then the limit
(190) exists for every . Moreover, is lower semi-continuous and strictly positive everywhere on , and a.e. on .
Proof.
By Lemma 2.1 and Lemma 2.2, the nonnegative scalar curvature condition and imply that the sequence of warping functions satisfies the hypothesis in Proposition 3.5. By applying Proposition 3.5, we get the desired convergence.
By applying Proposition 3.5 we get that , for all .
We prove by contradiction. Recall that is the essential infimum of as defined in Definition 4.6. First note that , since . Assume that , then by Proposition 4.7 we have almost everywhere in and hence
(191) |
Therefore, for all sufficiently large , we have . Then by Proposition 4.10, we have for all sufficiently large . This contradicts with that in as in (191). This finishes the proof of part .
Because warping functions satisfy the requirements in Proposition 3.7, the existence of the limit
(192) |
the lower semi-continuity of and a.e. on directly follow from Proposition 3.7.
Thus we only need to prove that for all . Let
(193) |
By the continuity of the distance funciton , there exists such that for all we have
(194) |
As a result, we have
(195) |
Then because in Proposition 3.7 we proved that for each fixed the ball average is non-increasing in , and
(196) |
we have that for each fixed ,
(197) | |||||
(198) | |||||
(199) | |||||
(200) |
This completes the proof of theorem. ∎
Remark 4.14.
Theorem 4.13 implies that the limit function has a everywhere positive lower semi-continuous representative as a function in for . For the rest of paper, will always denote this everywhere positive lower semi-continuous representative.
We end this section with Proposition 4.15 below. The proof of Proposition 4.15 uses Theorem 4.13 and the spherical mean inequality from Proposition 2.4. The positive uniform lower bound for warping functions obtained in Proposition 4.15 is important in proving geometric convergences of the sequence of warped product manifolds in our next paper.
Proposition 4.15.
Let be a sequence of warped product manifolds with metric tensors as in (3) that have non-negative scalar curvature and satisfy
(201) |
Let . Then there exists such that , for all and all .
Proof.
By Lemma 2.1, the non-negativity of scalar curvature of implies that
(202) |
Therefore, by the spherical mean inequality in Proposition 2.4, we have
(203) |
Then multiplying the inequality by gives us
(204) |
for all and . Let
(205) |
and let denote the essential infimum of the limit function which is strictly positive by Theorem 4.13.
Now integrating the inequality (204) with respect to from to gives us
(206) | |||||
(208) | |||||
(210) | |||||
Then by dividing the inequality by we obtain
(211) |
for all and . By Lemma 3.2 we have , and by direct calculation we have that
(212) |
we can choose such that
(213) |
Moreover, because in , we can choose such that
(214) |
4.5. Uniform systole positive lower bound
In this subsection, as an application of non-collapsing of warping functions obtained in Proposition 4.15, we derive a uniform positive lower bound for the systole of the sequence of warped product manifolds satisfying assumptions in Proposition 4.15.
Definition 4.16 (Systole).
The systole of a Riemannian manifold , which is denoted by is defined to be the length of the shortest closed geodesic in .
Remark 4.17.
People may usually consider so-called -systole that is the length of a shortest non-contractible closed geodesic. But in the study of compactness problem of manifolds with nonnegative scalar curvature, we also need to take into account contractible closed geodesic, for example, in a dumbell, which is diffeomorphic to , we may have a short contractible closed geodesic.
First of all we derive an interesting dichotomy property for closed geodesics in warped product manifolds: , that is, the product manifold endowed with the metric , where is a -dimensional (either compact or completep non-compact) Riemannian manifold without boundary, and is a positive smooth function on .
Lemma 4.18.
There is a dichotomy for closed geodesics in , that is, a closed geodesic in either wraps around the fiber , or is a geodesic in the base .
Proof.
Let is a coordinate on the fiber . The warped product metric then can be written as
(215) |
Let
(216) |
be a closed geodesic in , and without loss of generality, we assume . We have two possible cases as following:
Case 1: . In this case, clearly, the geodesic wraps around the fiber .
Case 2: . In this case, we show that by a proof by contradiction, and then clearly, is a closed geodesic on base . Otherwise, we have
(217) |
Moreover, there exists such that , since due to the closeness of the geodesic . Consequently, is a critical point of the function , i.e. . As a result, the tangent vector of the geodesic at , , is tangent to . On the other hand, there is a geodesic contained in that passes through the point and is tangent to at this point. Then by the uniqueness of the geodesic with given tangent vector at a point, and the fact that base is totally geodesic in the warped product manifold , which can be seen easily by Koszul’s formula, or see Proposition 9.104 in [3], we can obtain , and this contradicts with . ∎
By the dichotomy of closed geodesics in Lemma 4.18, we can obtain a lower bound estimate for the systole of .
Lemma 4.19.
The systole of the warped product Riemannian manifold is greater than or equal to .
Proof.
Let is a closed geodesic in . By Lemma 4.18, either wraps around the fiber , or is a closed geodesic in the base manifold .
If wraps around the fiber , then , and so the length of :
(218) | |||||
(219) | |||||
(220) |
If is a closed geodesic in the base , then by the definition of systole, the length of is greater than or equal to .
These estimates of length of closed geodesics imply the lower bound of systole in the conclusion. ∎
By combining the lower bound estimate of systole in Lemma 4.19 and Proposition 4.15, we immediately have the following uniform lower bound for systoles.
Proposition 4.20.
Let be a sequence of warped product manifolds with metric tensors as in (3) that have non-negative scalar curvature and satisfy
(221) |
Let . Then the systoles of , for all , have a uniform positive lower bound given by .
Proof.
First note that the base manifold of the sequence of the warped product manifolds is the standard -sphere, and its systole is equal to , since the image of a closed geodesic in is always a great circle.
5. Nonnegative distributional scalar curvature of limit metric
Now we use the positive limit function obtained in Theorem 4.13 to define a weak warped product metrics:
Definition 5.1.
Let be a function defined on such that it is almost everywhere positive and finite on . We further assume that for . Define
(222) |
to be a (weak) warped product Riemannian metric on in the sense of defining an inner product on the tangent space at (almost) every point of .
Remark 5.2.
In general, is only defined almost everywhere in with respect to the standard product volume measure , since may have value as on a measure zero set in . Note that we allow as ball average limit in Proposition 3.7. For example, in the extreme example constructed by Christina Sormani and authors in [19], the limit warping function equal to at two poles of .
In Subsection 5.1, we show regularity of the weak metric tensor defined in Definition 5.1 for [Proposition 5.4], and prove that the warped product metrics converge to in the sense for any [Theorem 5.5].
In Subsection 5.2, we show that the limit weak metric has nonnegative distributional scalar curvature in the sense of Lee-LeFloch [Theorem 5.11].
5.1. limit Riemannian metric
we prove the regularity of the metric tensor. Before that we need the following definition:
Definition 5.3.
We define as the set of all tensors defined almost everywhere on such that its norm measured in terms of is finite where is the isometric product metric
(223) |
We define as the set of all tensors, , defined almost everywhere on such that both the norm of and the norm of measured in terms of are finite where is the connection corresponding to the metric .
Now we prove the regularity of the metric tensor defined in Definition 5.1:
Proposition 5.4 (Regularity of the metric tensor).
Proof.
Using the background metric, , we have
(225) | |||||
(226) | |||||
(227) |
is finite, since by the assumption, for any , and Sobolev embedding theorem, we have for any .
Now for the gradient estimate, we fix an arbitrary . We use to denote the connection of the background metric . Clearly, we have
(228) |
and
(229) |
Moreover, since we have
(230) |
where is the gradient of on . As a result, we have
(231) | |||||
(232) |
where is chosen so that , and . Then again by Sobolev embedding theorem we have for any , thus we obtain that is finite for any . This completes the proof. ∎
Then we apply Proposition 3.5 to prove Theorem 1.7 which concerns the pre-compactness of warped product circles over sphere with non-negative scalar curvature. We restate Theorem 1.7 as follows:
Theorem 5.5.
Proof.
By Lemma 2.1 and Lemma 2.2, the assumptions in (4) for implies that the warping functions satisfy the assumptions in Proposition 3.5. Thus, by applying Proposition 3.5, we have that there exists a subsequence of warping functions and for all , such that
(234) |
Let . Then by Proposition 5.4, we have
(235) |
Moreover, because
(236) |
we have that for any ,
(237) | |||||
(238) | |||||
(239) | |||||
(240) | |||||
(241) |
since in for any . ∎
5.2. Nonnegative distributional scalar curvature of
Building upon work of Mardare-LeFloch [11], Dan Lee and Philippe LeFloch defined a notion of distributional scalar curvature for smooth manifolds that have a metric tensor which is only . See Definition 2.1 of [10] which we review below in Definition 5.7.
In Theorem 5.5 we proved that if a sequence of smooth warped product circles over the sphere with non-negative scalar curvature have uniform bounded volumes, then a subsequence of the smooth warped product metric converges to a weak warped product metric in the sense of for any . For the rest of this section, we use to denote such limit metric. We use as a background metric .
In Theorem 5.11, we prove that this limit (weak) metric has nonnegative distributional scalar curvature in the sense of Lee-LeFloch . In Remarks 5.9-5.10, we discuss how the metric tensors studied by Lee and LeFloch have stronger regularity than the regularity of but their definition of distributional scalar curvature is still valid in our case.
First we recall Definition 2.1 in the work of Lee-LeFloch [10]. In their paper, they assume that
Definition 5.7 (Lee-LeFloch).
Let be a smooth manifold endowed with a smooth background metric, . Let be a metric tensor defined on with regularity and locally bounded inverse .
The scalar curvature distribution is defined as a distributions in such that for every test function
(242) |
where the dot product is taken using the metric , is the Levi-Civita connection of , and are volume measure with respect to and respectively, is a vector field given by
(243) |
where
(244) |
(245) |
and
(246) |
The Riemannian metric has nonnegative distributional scalar curvature, if for every nonnegative test function in the integral in (242).
Definition 5.8 (Distributional total scalar curvature).
Note that for a -metric, the distributional total scalar curvature is exactly the usual total scalar curvature.
Remark 5.9.
By the regularity assumption for the Riemannian metric in the work of Lee-LeFloch [10], one has the regularity , , and the density of volume measure with respect to is
(247) |
Thus
(248) |
and
(249) |
are both finite.
Remark 5.10.
We are ready to prove Theorem 1.8. We restate it as follows:
Theorem 5.11.
The limit metric obtained in Theorem 5.5 has nonnegative distributional scalar curvature on in the sense of Lee-LeFloch as in Definition 5.7. In particular, (242) is finite and nonnegative for any nonnegative test function, . Moreover, the total scalar curvatures of converge to the distributional total scalar curvature of .
The proof of Theorem 5.11 consists of straightforward but technical calculations. For the convenience of readers, we provide some details of the calculations in the following lemmas.
We use as background metric, and use coordinate on , where is a polar coordinate on and is a coordinate on . The corresponding local frame of the tangent bundle is . In this coordinate system, both and are diagonal and given as
(250) |
First of all, by the formula of Christoffel symbols:
(251) |
one can easily obtain the following lemma:
Lemma 5.12.
The Christoffel symbols of the Levi-Civita connection of the background metric , in the coordinate , all vanish except
(252) |
and
(253) |
Then by Lemma 5.12, the formula
(254) |
and the diagonal expression of in (250), one can obtain the following lemma:
Lemma 5.13.
For the limit metric, , with the background metric, , the Christoffel symbols defined by Lee-LeFloch as in (244), in the coordinate , all vanish except
(255) |
and
(256) |
Note also that
Lemma 5.14.
Note that the volume forms are:
(257) |
and
(258) |
which are both defined almost everywhere. In particular,
(259) |
is in for .
Proof.
The first claim holds away from and by the definition of volume form, and the second claim holds almost everywhere on . So almost everywhere which gives us the third claim. The rest follows from Proposition 3.5. ∎
Now we are ready to compute the vector field and the function defined by Lee-LeFloch as in (243) and .
Lemma 5.15.
For the limit metric with the background metric , the vector field defined in (243), in the local frame , is given by
(260) |
Furthermore
(261) |
Proof.
Lemma 5.16.
For the limit metric with the background metric , the function defined in (245) is given by
(271) |
Furthermore,
(272) |
Here is the norm of weak gradient of with respect to the standard metric .
Proof.
Lemma 5.17.
Proof.
Remark 5.18.
As explained in Remark 3.6, for any , which is obtained in in Proposition 3.5, is the best regularity for in general, and we cannot expect is in . So the integral appearing in both (298) and (300) may be divergent (c.f. Lemma 4.16 in [19]). But if we sum the integrants in (298) and (300) firstly and then integrate, then this possible divergent integrant terms cancel out and we obtain a finite integral as in the following lemma.
Lemma 5.19.
For the limit metric , the scalar curvature distribution defined in Definition 5.7 can be expressed, for every test function , as the integral
(302) |
and this is finite for any test function . Here is defined as in (350), and are (weak) gradients of functions and on standard sphere respectively, and is the Riemannian metric on .
Proof.
We now apply these lemmas to prove Theorem 5.11:
Proof.
By the expression (11) of the scalar curvature of , we have that for any test function ,
(303) | |||||
(304) | |||||
(305) |
where . Then, by using the nonnegative scalar curvature condition , Proposition 3.5 and Lemma 5.19, possibly after passing to a subsequence, we obtain for any nonnegative test function ,
(306) | |||||
(307) | |||||
(308) | |||||
(309) |
Thus, for all nonnegative test function . By setting in equations (306)-(309), we obtain the convergence of distributional total scalar curvature. ∎
Appendix A convergence in case
In this appendix, we will derive convergence in the case of warped product spheres over circle with nonnegative scalar curvature, and show that the limit metric has nonnegative distributional scalar curvature in the sense of Lee-LeFloch. Specifically, we will prove the following two theorems.
Theorem A.1.
Let be a family of warped Riemannian manifolds with metric tensors as in (8) satisfying
(310) |
and
(311) |
for all , where is the scalar curvature of . Then there is a subsequence of warping functions that converges in to a Lipschitz function , which has Lipschitz constant 1 and satisfies
(312) |
Moreover, let , then is a Lipschitz continuous Riemannian metric tensor on , and a subsequence of converges in to .
Here, as before, we still use as a background metric. Then we can compute the scalar curvature distribution of Lee-LeFloch and have the following property.
Theorem A.2.
The study of this case is similar as the case of rotationally symmetric metrics on sphere, which was studied by authors with Jiewon Park in [15]. But there are some difference between these two cases. For example, in the rotationally symmetric metrics on sphere, in general condition may not be able to prevent collapsing happening near two poles [Lemma 4.3 in [15]], however, in the case of , condition can provide a positive uniform lower bound for [Lemma A.6] and hence prevent collapsing happening.
The key ingredient is a uniform gradient estimate obtained by using nonnegative scalar curvature condition [Lemma A.4]. Moreover, for the minimal value of warping function , we obtain a uniform upper bound from uniform upper bounded diameter condition [Lemma A.3] and a uniform lower bound from condition [Lemma A.6]. Then we combine these estimates to prove Theorem A.1 at the end of Subsection A.1. Finally, in Subsection A.2, we will prove Theorem A.2.
A.1. Convergence of a subsequence
Lemma A.3.
Let be a family of warped product Riemannian manifolds with metric tensors as in (8), having uniformly upper bounded diameters, i.e. , then we have
Proof.
Let be the minimum point of the function . Then clearly the distance between antipodal points on the sphere is . So we have , and the claim follows. ∎
Lemma A.4.
Let be a family of warped product Riemannian manifolds with metric tensors as in (8). The scalar curvature of the warped product metric is given by
(313) |
Here the Laplace is the trace of the Hessian.
Moreover, if , then we have on .
Proof.
First, by using the formula of Ricci curvature for warped product metrics as in 9.106 in [3], one can easily obtain that the scalar curvature of is given as in (313).
Now we prove the second claim by contradiction. Assume for some , at some point, let’s say . Take a unit vector field on such that is in the same direction as at the point . Let be the first point such that while moving from the point on in the opposite direction of the unit vector field . Then let be the integral curve of the vector field with the initial point . Let such that . Set . Then (at least) for ,
(314) |
and
(315) |
By the Mean Value Theorem, there exists such that
(316) |
since and .
On the other hand, because , by using the scalar curvature , one has
(317) |
So
(318) |
since by the choice of . This produces a contradiction, and so on .
∎
Lemma A.5.
Let be a family of warped product Riemannian manifolds with metric tensors as in (8). If for some then there is a minimal surface in .
Proof.
Define . Then for all , is an embedded submanifold with mean curvature
(319) |
∎
Lemma A.6.
Let be a family of warped product Riemannian manifolds with metric tensors as in (8) satisfying . Then we have .
Proof.
By applying Lemma A.5, we have that there exists a minimal surface on at the minimal value point of . The area of is given by
(320) |
Thus by the condition, , and the conclusion follows. ∎
Now we will use above lemmas to prove Theorem A.1:
Proof.
We complete the proof in the following three steps.
Step 1. Uniform convergence of warping functions. By applying Lemma A.3 and Lemma A.4 we immediately obtain the uniform upper bound
(321) |
By combining this uniform upper bound with the uniform lower bound obtained in Lemma A.6, we have that the warping functions are uniformly bounded, i.e.
(322) |
Moreover, Lemma A.4 implies function are equicontinuous. Thus by applying Arzelà-Ascoli theorem we obtain that are uniformly convergent a continuous function satisfying
(323) |
Meanwhile, the uniform gradient estimate obtained in Lemma A.4 also implies that the limit function is Lipschitz with Lipschitz constant . Because a Lipschitz function is , we actually have .
Step 2. convergence of warping functions. We will estimate the bounded variation norm of warping functions. First note that
(324) |
Thus,
(325) |
furthermore,
(326) | |||||
(327) | |||||
(328) |
Then by the expression of the scalar curvature in Lemma A.4, the nonnegative scalar curvature condition implies
(329) |
The last inequality here follows from Lemma A.6. Lemma A.4 also tells us that on for all . Consequently, we have
(330) | |||||
(331) | |||||
(332) |
As a result, by Theorem 5.5 in [5] we have that a subsequence, which is still denoted by , converges to some in norm, and it is easy to see that in the weak sense. Moreover, since and , we have in norm. Indeed, note that by the Hölder inequality,
(333) |
As a result, in .
Step 3. convergence of metrics. Note that
(334) |
and
(335) |
Therefore, by applying the uniform bound , and convergence of to , we can obtain that converges to in . ∎
A.2. Nonnegative distributional scalar curvature of the limit metric
In this subsection, we compute the distributional scalar curvature of the limit metric tensor obtained in Theorem A.1 with the background metric in the sense of Lee-LeFloch, and prove Theorem A.2. Throughout this subsection, always denotes the limit metric obtained in Theorem A.1.
By the definition of in Definition 5.7 and the Christofell symbols in Lemma 5.12, one can obtain the following lemma:
Lemma A.7.
For the limit metric, , with the background metric, , the Christoffel symbols defined by Lee-LeFloch as in (244), in the coordinate , all vanish except
(336) |
(337) |
and
(338) |
Note also that
Lemma A.8.
Note that the volume forms are:
(339) |
and
(340) |
which are both defined everywhere away from and . In particular,
(341) |
is in for all .
Proof.
The first claim holds away from and by the definition of volume form, and the second claim holds almost everywhere on . So almost everywhere which gives us the third claim. The rest follows from Proposition 3.5. ∎
Now we are ready to compute the vector field and the function defined by Lee-LeFloch as in (243) and .
Lemma A.9.
For the limit metric with the background metric , the vector field defined in (243), in the local frame , is given by
(342) |
Furthermore
(343) |
Lemma A.10.
For the limit metric with the background metric , the function defined in (245) is given by
(344) |
Furthermore,
(345) |
Lemma A.11.
Proof.
Remark A.12.
Lemma A.13.
Proof.
We now apply these lemmas to prove Theorem A.2:
Proof.
By the expression (313) of the scalar curvature of , we have that for any test function ,
(352) | |||||
(353) | |||||
(354) |
Then, by using the nonnegative scalar curvature condition , and convergence property of in Theorem A.1, possibly after passing to a subsequence, we obtain for any nonnegative test function ,
(356) | |||||
(357) | |||||
(358) | |||||
(359) |
Thus, for all nonnegative test function . ∎
References
- [1] Jr. Almgren, Frederick Justin. The homotopy groups of the integral cycle groups. Topology, 1:257–299, 1962.
- [2] J. Basilio, J. Dodziuk, and C. Sormani. Sewing Riemannian manifolds with positive scalar curvature. J. Geom. Anal., 28(4):3553–3602, 2018.
- [3] Arthur L. Besse. Einstein Manifolds. Springer, Berlin, 1987.
- [4] Paula Burkhardt-Guim. Pointwise lower scalar curvature bounds for metrics via regularizing Ricci flow. Geom. Funct. Anal., 29:1703–1772, 2019.
- [5] Lawrence C. Evans and Ronald F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
- [6] David Gilbag and Neil S. Trudinger. Elliptic partial differential equations of second order. Springer-Verlag, Berlin, second edition edition, 1983.
- [7] Misha Gromov. Dirac and Plateau billiards in domains with corners. Cent. Eur. J. Math., 12(8):1109–1156, 2014.
- [8] Misha Gromov. Plateau-Stein manifolds. Cent. Eur. J. Math., 12(7):923–951, 2014.
- [9] Demetre Kazaras and Kai Xu. Drawstrings and flexibility in the georch conjecture. arXiv:2309.03756, 2023.
- [10] Dan A. Lee and Philippe G. LeFloch. The positive mass theorem for manifolds with distributional curvature. Comm. Math. Phys., 339(1):99–120, 2015.
- [11] Philippe G. LeFloch and Cristinel Mardare. Definition and stability of Lorentzian manifolds with distributional curvature. Port. Math. (N.S.), 64(4):535–573, 2007.
- [12] Chao Li. A polyhedron comparison theorem for 3-manifolds with positive scalar curvature. Invent. Math., 219(1):1–37, 2020.
- [13] Fernando C. Marques and André Neves. Applications of min-max methods to geometry. Lecture Notes in Math., 2263:41–77, Springer, Cham, 2020.
- [14] Enrico Onofri. On the positivity of the effective action in a theory of random surfaces. Comm. Math. Phys., 86(3):321–326, 1982.
- [15] Jiewon Park, Wenchuan Tian, and Changliang Wang. A compactness theorem for rotationally symmetric Riemannian manifolds with positive scalar curvature. Pure Appl. Math. Q., 14(3-4):529–561, 2018.
- [16] Jon T. Pitts. Existence and regularity of minimal surfaces on Riemannian manifolds, volume 27 of Mathematical Notes. Princeton University Press, 1981.
- [17] Leon Simon. Lectures on geometric measure theory. Proceedings of the Centre of Mathematical Analysis. Australian National University, Canberra, 1983.
- [18] Christina Sormani. Scalar curvature and intrinsic flat convergence. In Measure theory in non-smooth spaces, Partial Differ. Equ. Meas. Theory, pages 288–338. De Gruyter Open, Warsaw, 2017.
- [19] Christina Sormani, Wenchuan Tian, and Changliang Wang. An extreme limit with nonnegative scalar. arXiv preprint arXiv:, 2023.