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

Compactness of sequences of warped product circles over spheres with nonnegative scalar curvature

Wenchuan Tian Department of Mathematics, University of California, Santa Barbara, CA93106-3080 [email protected]  and  Changliang Wang School of Mathematical Sciences and Institute for Advanced Study, Tongji University, Shanghai 200092, China [email protected]
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 𝕊2{\mathbb{S}}^{2} that have nonnegative scalar curvature, a uniform upper bound on the volume, and a positive uniform lower bound on the MinA\operatorname{MinA}, which is the minimum area of closed minimal surfaces in the manifold. We prove that such a sequence has a subsequence converging to a W1,pW^{1,p} Riemannian metric for all p<2p<2, 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, Scalar0\operatorname{Scalar}\geq 0, 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 MinA\operatorname{MinA} 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 {Mj3}j=1\{M_{j}^{3}\}_{j=1}^{\infty} be a sequence of closed oriented three dimensional Riemannian manifolds without boundary satisfying

(1) Scalarj0,Vol(Mj)V,Diam(Mj)D,\operatorname{Scalar}_{j}\geq 0,\ \ {\rm Vol}(M_{j})\leq V,\ \ \operatorname{Diam}(M_{j})\leq D,
(2) MinA(Mj3)=inf{Area(Σ):Σ closed min surf in Mj3}A0>0.\operatorname{MinA}(M^{3}_{j})=\inf\{\operatorname{Area}(\Sigma)\,:\,\Sigma\textrm{ closed min surf in }M_{j}^{3}\,\}\geq A_{0}>0.

Then there exists a subsequence which is still denoted as {Mj}j=1\{M_{j}\}_{j=1}^{\infty} that converges in the volume preserving intrinsic flat sense to a three dimensional rectifiable limit space MM_{\infty}. Furthermore, MM_{\infty} 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 (Mj3,gj)(M^{3}_{j},g_{j}). In our proof the MinA\operatorname{MinA} 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 W1,2W^{1,2} norm in closed regions away from the two poles.

In this paper, we study the 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} warped product case of the Conjecture 1.1. We consider the following:

Definition 1.2.

Let {(𝕊2×𝕊1,gj)}j=1\{({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{j})\}_{j=1}^{\infty} be a sequence of Riemannian manifold such that

(3) gj=g𝕊2+fj2g𝕊1=dr2+sin(r)2dθ2+fj2dφ2, for j=1,2,3,g_{j}=g_{{\mathbb{S}}^{2}}+f_{j}^{2}g_{{\mathbb{S}}^{1}}=dr^{2}+\sin(r)^{2}d\theta^{2}+f_{j}^{2}d\varphi^{2},\text{ for }j=1,2,3,...

where g𝕊2g_{{\mathbb{S}}^{2}} and g𝕊1g_{{\mathbb{S}}^{1}} are the standard metrics on 𝕊2{\mathbb{S}}^{2} and 𝕊1{\mathbb{S}}^{1} respectively, and the function fj:𝕊2(0,)f_{j}:{\mathbb{S}}^{2}\to(0,\infty) is smooth for each jj. Here rr and θ\theta are the geodesic polar coordinate for 𝕊2{\mathbb{S}}^{2}. We also use the notation 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} to denote (𝕊2×𝕊1,gj)({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{j}).

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 {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product Riemannian manifolds such that each 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has non-negative scalar curvature. If we assume that

(4) Vol(𝕊2×fj𝕊1)V and MinA(𝕊2×fj𝕊1)A>0,j,{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V\text{ and }\operatorname{MinA}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\geq A>0,\ \ \forall j\in\mathbb{N},

then we have the following:

  1. (i)

    After passing to a subsequence if needed, the sequence of warping functions {fj}j=1\{f_{j}\}_{j=1}^{\infty} converges to some limit function ff_{\infty} in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for all q[1,)q\in[1,\infty).

  2. (ii)

    The limit function ff_{\infty} is in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}), for all pp such that 1p<21\leq p<2.

  3. (iii)

    The essential infimum of ff_{\infty} is strictly positive, i.e. inf𝕊2f>0\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0.

  4. (iv)

    If we allow ++\infty as a limit, then the limit

    (5) f¯(x):=limr0Br(x)f\overline{f_{\infty}}(x):=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}

    exists for every x𝕊2x\in{\mathbb{S}}^{2}. Moreover, f¯\overline{f_{\infty}} is lower semi-continuous and strictly positive everywhere on 𝕊2{\mathbb{S}}^{2}, and f¯=f\overline{f_{\infty}}=f_{\infty} a.e. on 𝕊2{\mathbb{S}}^{2}.

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 MinA\operatorname{MinA} 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 MinA\operatorname{MinA} 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 W1,pW^{1,p} regularity for 1p<21\leq p<2 is sharp for the limit warping function ff_{\infty}.

By applying Theorem 1.3 and the spherical mean inequality [Proposition 2.4], we obtain:.

Proposition 1.5.

Let {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product manifolds such that each 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has non-negative scalar curvature, and the sequence satisfies conditions in (4)(\ref{eqn-Intro-thm-main-condition}). Then there exists j0j_{0}\in\mathbb{N} such that fj(x)e4>0f_{j}(x)\geq\frac{e_{\infty}}{4}>0, for all jj0j\geq j_{0} and x𝕊2x\in{\mathbb{S}}^{2}, where e=inf𝕊2f>0e_{\infty}=\inf_{{\mathbb{S}}^{2}}f_{\infty}>0 obtained in Theorem 1.3.

As an application of Proposition 1.5, we have:

Corollary 1.6.

Let {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product manifolds such that each 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has non-negative scalar curvature, and the sequence satisfies conditions in (4)(\ref{eqn-Intro-thm-main-condition}). Then the systoles of 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}, for all jj\in\mathbb{N}, have a uniform positive lower bound given by min{2π,e2π}\min\left\{2\pi,\frac{e_{\infty}}{2}\pi\right\}, where e:=inf𝕊2f>0e_{\infty}:=\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0 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: 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}, in Lemma 4.18 we establish an interesting dichotomy property for closed geodesics in a general warped product manifold N×f𝕊1N\times_{f}{\mathbb{S}}^{1} with 𝕊1{\mathbb{S}}^{1} as a typical fiber, with metric tensor as g=gN+f2g𝕊1g=g_{N}+f^{2}g_{{\mathbb{S}}^{1}}, where (N,gN)(N,g_{N}) is a nn-dimensional complete Riemannian manifold without boundary and ff is a positive smooth function on NN. 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 {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product Riemannian manifolds such that each 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has non-negative scalar curvature. If we assume that

(6) Vol(𝕊2×fj𝕊1)V and MinA(𝕊2×fj𝕊1)A>0,j,{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V\text{ and }\operatorname{MinA}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\geq A>0,\ \ \forall j\in\mathbb{N},

Then there exists a subsequence gjkg_{j_{k}} and a (weak) warped product Riemannian metric gW1,p(𝕊2×𝕊1,g0)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for p[1,2)p\in[1,2) such that

(7) gjkg inLq(𝕊2×𝕊1,g0),q[1,).g_{j_{k}}\rightarrow g_{\infty}\ \ \text{ in}\ \ L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}),\ \ \forall q\in[1,\infty).

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 Lq(𝕊2×𝕊1,g0)L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) and W1,p(𝕊2×𝕊1,g0)W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) are defined in Definition 5.3. The MinA\operatorname{MinA} condition is used to prevent gjkg_{j_{k}} converging to a non-metric tensor in W1,p(𝕊2×𝕊1,g0)W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}), with the help of the non-collapsing property of ff_{\infty} 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.

The limit metric gg_{\infty} obtained in Theorem 1.7 has nonnegative distributional scalar curvature on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} in the sense of Lee-LeFloch. [10]. Moreover, the total scalar curvatures of gjg_{j} converge to the distributional total scalar curvature of gg_{\infty}.

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 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} 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 C0C^{0}-metrics. However, as mentioned in Remark 1.4, the W1,pW^{1,p} regularity, for 1p<21\leq p<2, 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 Wloc1,2W^{1,2}_{loc}-metrics. Our limit metric gg_{\infty} 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 (Mj3,gj)(M^{3}_{j},g_{j}), that is, Mj3M^{3}_{j} are diffeomorphic to 𝕊1×𝕊2{\mathbb{S}}^{1}\times{\mathbb{S}}^{2} with warped product metric tensors

(8) gj=g𝕊1+hj2g𝕊2,wherehj:𝕊1(0,).g_{j}=g_{{\mathbb{S}}^{1}}+h_{j}^{2}g_{{\mathbb{S}}^{2}},\ \ \textrm{where}\ \ h_{j}:{\mathbb{S}}^{1}\rightarrow(0,\infty).

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 hjh_{j} from nonnegative scalar curvature condition [Lemma A.4]. By combining this with uniform diameter upper bound and the MinA\operatorname{MinA} condition, we prove that a subsequence of {hj}j=1\{h_{j}\}^{\infty}_{j=1} converges in C0C^{0} and W1,2W^{1,2} sense to a bounded positive Lipschitz function h:𝕊1(0,)h_{\infty}:{\mathbb{S}}^{1}\rightarrow(0,\infty) [Theorem A.1]. Moreover, we prove that the limit W1,2W^{1,2} Riemannian metric g=g𝕊1+h2g𝕊2g_{\infty}=g_{{\mathbb{S}}^{1}}+h^{2}_{\infty}g_{{\mathbb{S}}^{2}} 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 fjf_{j} from the uniform geometric bounds of metric gjg_{j} as in (3). In particular, we show that metrics gjg_{j} in (3) have nonnegative scalar curvature if and only if the warping functions fjf_{j} satisfy the differential inequality [Lemma 2.1]:

(9) Δfjfj,on𝕊2,\Delta f_{j}\leq f_{j},\ \ \textrm{on}\ \ {\mathbb{S}}^{2},

where Δ\Delta is the Lapacian on the standard round sphere 𝕊2{\mathbb{S}}^{2}, taken to be the trace of the Hessian. Moreover, a positive number VV is a uniform upper bound of volumes of metrics gjg_{j} in (3) if and only if fjf_{j} satisfy [Lemma 2.2]

(10) 𝕊2fj𝑑volg𝕊2V2π.\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\leq\frac{V}{2\pi}.

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 fjf_{j} 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 fjf_{j} [Proposition 2.6], which helps us to obtain lower semi-continuity of the limit warping function ff_{\infty} in Proposition 3.7.

In Section 3, we study the convergence of a sequence {fj}j=1\{f_{j}\}^{\infty}_{j=1} of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying (9) and (10). We prove that there exists a subsequence of such sequence {fj}\{f_{j}\} and a function fW1,p(𝕊2)(1p<2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2})\,(1\leq p<2) such that the subsequence converges to ff_{\infty} in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for any q1q\geq 1 [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 hjh_{j} in [15] and in (8) have one variable, whereas fjf_{j} in (3) have two variables, it is more difficult to obtain sub-convergence of {fj}\{f_{j}\}, and we make use of the Moser-Trudinger inequality in (25) in [14]. The regularity of the limit function ff_{\infty} is weaker than hh_{\infty}. The extreme example constructed by Sormani and authors in [19] shows that the W1,pW^{1,p} regularity for 1p<21\leq p<2 is sharp for ff_{\infty}.

In Section 4, we use the MinA\operatorname{MinA} condition to show that the limit function ff_{\infty} has positive essential infimum [Theorem 4.13] and that the warping functions fjf_{j} have a positive uniform lower bound [Proposition 4.15]. This enables us to define weak warped product Riemnnian metric gg_{\infty} on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} 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 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} [Proposition 4.20].

The MinA\operatorname{MinA} 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 MinA\operatorname{MinA} 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 MinA\operatorname{MinA} condition.

In Section 5, we prove that a subsequence of {gj}j=1\{g_{j}\}^{\infty}_{j=1}, with gjg_{j} as in (3) having nonnegative scalar curvatures and uniform upper bounded volumes and satisfying MinA\operatorname{MinA} condition, converges to a weak metric tensor gW1,p(𝕊2×𝕊1,g0)(1p<2)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})\,(1\leq p<2) in the sense of Lq(𝕊2×𝕊1,g0)L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for all q1q\geq 1 [Theorem 5.5]. Moreover, we prove that the limit metric gg_{\infty} 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 MinA\operatorname{MinA} 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 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}

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 fjf_{j} and prove that the uniform volume bounds on sequence of Riemannian manifolds implies the uniform L1L^{1} norm of the warping function.

In Subsection 2.2, we prove the spherical mean inequality for the warping function ff [Proposition 2.6], which is our main analytic tool. In Subsection 2.3, we prove a ball average monotonicity property for the warping function ff [Proposition 2.4].

The implication of the MinA\operatorname{MinA} 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 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} are given by

(11) Scalarj=22Δfjfj,\operatorname{Scalar}_{j}=2-2\frac{\Delta f_{j}}{f_{j}},

where Δ\Delta is the Laplacian on 𝕊2{\mathbb{S}}^{2} with respect to the standard metric g𝕊2g_{{\mathbb{S}}^{2}}, taken to be the trace of the Hessian (without the negative sign).

Thus 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} have nonnegative scalar curvature if and only if

(12) Δfjfj.\Delta f_{j}\leq f_{j}.
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 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} as Scalarj=22Δfjfj.\operatorname{Scalar}_{j}=2-2\frac{\Delta f_{j}}{f_{j}}. Then the second claim directly follows, since fj>0f_{j}>0. ∎

Lemma 2.2 (Volume upper bound condition).

The warped product manifolds 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} have volume Vol(𝕊2×fj𝕊1)V{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V if and only if

(13) 𝕊2fj𝑑vol𝕊2V2π.\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi}.
Proof.

The Riemannian volume measure of gjg_{j} is given by

(14) dvolgj=fjdvolg𝕊2dvolg𝕊1.d{\rm vol}_{g_{j}}=f_{j}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}d{\rm vol}_{g_{{\mathbb{S}}^{1}}}.

Thus the volume of 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} is given by

(15) Vol(𝕊2×fj𝕊1)=𝕊2×𝕊1fj𝑑volg𝕊2𝑑vol𝕊1=2π𝕊2fj𝑑volg𝕊2.{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})=\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}f_{j}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}d{\rm vol}_{{\mathbb{S}}^{1}}=2\pi\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}.

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 ff on 𝕊2{\mathbb{S}}^{2} satisfying the differential inequality Δff\Delta f\leq f. By Lemma 2.1, this is equivalent to studying the warping function of warped product manifolds 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} 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 ff be a smooth function on 𝕊2{\mathbb{S}}^{2}. Consider the spherical mean given by

(16) ϕ(r):=Br(p)f𝑑s,\phi(r):=\fint_{\partial B_{r}(p)}fds,

where Br(p)B_{r}(p) is the geodesic ball in the standard 𝕊2\mathbb{S}^{2} with center pp and radius rr. The derivative of ϕ(r)\phi(r) satisfies

(17) ddrϕ(r)=12πsinrBr(p)Δf𝑑vol𝕊2.\frac{d}{dr}\phi(r)=\frac{1}{2\pi\sin r}\int_{B_{r}(p)}\Delta fd{\rm vol}_{{\mathbb{S}}^{2}}.
Proof.

Using the geodesic polar coordinate (r,θ)(r,\theta) on 𝕊2\mathbb{S}^{2} centered at pp, one can write ϕ(r)\phi(r) as

(18) ϕ(r)=02πf(r,θ)sinrdθ2πsinr=02πf(r,θ)𝑑θ2π.\phi(r)=\frac{\int^{2\pi}_{0}f(r,\theta)\sin rd\theta}{2\pi\sin r}=\frac{\int^{2\pi}_{0}f(r,\theta)d\theta}{2\pi}.

Then taking derivative with respective to rr gives

(19) ϕ(r)\displaystyle\phi^{\prime}(r) =\displaystyle= 12π02πfr𝑑θ\displaystyle\frac{1}{2\pi}\int^{2\pi}_{0}\frac{\partial f}{\partial r}d\theta
(20) =\displaystyle= 12π02πf,r\displaystyle\frac{1}{2\pi}\int^{2\pi}_{0}\langle\nabla f,\partial_{r}\rangle
(21) =\displaystyle= 12πsinr02πf,rsinrdθ\displaystyle\frac{1}{2\pi\sin r}\int^{2\pi}_{0}\langle\nabla f,\partial_{r}\rangle\sin rd\theta
(22) =\displaystyle= 12πsinrBr(p)f,r𝑑s\displaystyle\frac{1}{2\pi\sin r}\int_{\partial B_{r}(p)}\langle\nabla f,\partial_{r}\rangle ds
(23) =Stokes\displaystyle\overset{Stokes}{=} 12πsinrBr(p)Δf𝑑vol𝕊2.\displaystyle\frac{1}{2\pi\sin r}\int_{B_{r}(p)}\Delta fd{\rm vol}_{\mathbb{S}^{2}}.

Now we use Lemma 2.3 to prove the spherical mean inequality.

Proposition 2.4.

Let ff be a smooth function on 𝕊2{\mathbb{S}}^{2} satisfying Δff\Delta f\leq f. Then for any fixed p𝕊2p\in\mathbb{S}^{2} and 0<r0<r1π20<r_{0}<r_{1}\leq\frac{\pi}{2}, one has

(24) Br1(p)f𝑑sBr0(p)f𝑑sfL2(𝕊2)2π(r1r0),\fint_{\partial B_{r_{1}}(p)}fds-\fint_{\partial B_{r_{0}}(p)}fds\leq\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}(r_{1}-r_{0}),

where Br(p)B_{r}(p) is the geodesic ball in the 𝕊2\mathbb{S}^{2} with center pp and radius rr.

Moreover, by taking limit as r00r_{0}{\rightarrow}0, one has

(25) Br(p)f𝑑sf(p)fL2(𝕊2)2πr,\fint_{\partial B_{r}(p)}fds-f(p)\leq\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}r,

for any 0<rπ20<r\leq\frac{\pi}{2}.

Proof.

By Lemma 2.3 and the assumption Δff\Delta f\leq f, one has

(26) ϕ(r)12πsinrBr(p)f𝑑vol𝕊2.\phi^{\prime}(r)\leq\frac{1}{2\pi\sin r}\int_{B_{r}(p)}fd{\rm vol}_{\mathbb{S}^{2}}.

Integrating this differential inequality for rr from r0r_{0} to r1r_{1} gives

(27) ϕ(r1)ϕ(r0)\displaystyle\phi(r_{1})-\phi(r_{0}) \displaystyle\leq r0r1(12πsinrBr(p)f𝑑vol𝕊2)𝑑r\displaystyle\int^{r_{1}}_{r_{0}}\left(\frac{1}{2\pi\sin r}\int_{B_{r}(p)}fd{\rm vol}_{\mathbb{S}^{2}}\right)dr
(28) \displaystyle\leq r0r1(12πsinrfL2(𝕊2)Area(Br(p)))𝑑r\displaystyle\int^{r_{1}}_{r_{0}}\left(\frac{1}{2\pi\sin r}\|f\|_{L^{2}(\mathbb{S}^{2})}\sqrt{{\rm Area}(B_{r}(p))}\right)dr
(29) =\displaystyle= fL2(𝕊2)2πr0r11cosrsinr𝑑r\displaystyle\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}\int^{r_{1}}_{r_{0}}\frac{\sqrt{1-\cos r}}{\sin r}dr
(30) =\displaystyle= fL2(𝕊2)2πr0r111+cosr𝑑r\displaystyle\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}\int^{r_{1}}_{r_{0}}\frac{1}{\sqrt{1+\cos r}}dr
(31) \displaystyle\leq fL2(𝕊2)2πr0r11𝑑r(0<r0<r1π2)\displaystyle\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}\int^{r_{1}}_{r_{0}}1dr\qquad\left(0<r_{0}<r_{1}\leq\frac{\pi}{2}\right)
(32) =\displaystyle= fL2(𝕊2)2π(r1r0).\displaystyle\frac{\|f\|_{L^{2}(\mathbb{S}^{2})}}{\sqrt{2\pi}}(r_{1}-r_{0}).

2.3. Ball average monotonicity

In this subsection, we further derive a ball average monotonicity [Proposition 2.6] for a smooth function on 𝕊2{\mathbb{S}}^{2} satisfying Δff\Delta f\leq f. 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 r0r\to 0 exists everywhere for the limit function.

Lemma 2.5.

Let ff be a smooth function on 𝕊2{\mathbb{S}}^{2} satisfying Δff\Delta f\leq f and fL2(𝕊2)C2π\|f\|_{L^{2}({\mathbb{S}}^{2})}\leq C\sqrt{2\pi}, where CC is a positive constant. For any fixed x𝕊2x\in{\mathbb{S}}^{2}, the spherical mean

(33) Br(x)(fCr)=Br(x)(fCr)2πsinr\fint_{\partial B_{r}(x)}\left(f-Cr\right)=\frac{\int_{\partial B_{r}(x)}(f-Cr)}{2\pi\sin r}

is a non-increasing function in rr for r(0,π2]r\in(0,\frac{\pi}{2}]

Proof.

The spherical mean inequality in Proposition 2.4 says that for any x𝕊2x\in{\mathbb{S}}^{2} and 0<r0<r1π20<r_{0}<r_{1}\leq\frac{\pi}{2},

(34) B1(x)fBr0(x)ffL2(𝕊2)2π(r1r0)C(r1r0).\fint_{\partial B_{1}(x)}f-\fint_{\partial B_{r_{0}}(x)}f\leq\frac{\|f\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}(r_{1}-r_{0})\leq C(r_{1}-r_{0}).

By rearranging this inequality, we obtain that for any fixed x𝕊2x\in{\mathbb{S}}^{2},

(35) Br1(x)(fCr1)Br0(x)(fCr0),0<r0r1π2.\fint_{\partial B_{r_{1}}(x)}(f-Cr_{1})\leq\fint_{\partial B_{r_{0}}(x)}(f-Cr_{0}),\quad\forall 0<r_{0}\leq r_{1}\leq\frac{\pi}{2}.

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 ff be a smooth function on 𝕊2{\mathbb{S}}^{2} satisfying Δff\Delta f\leq f and fL2(𝕊2)C2π\|f\|_{L^{2}({\mathbb{S}}^{2})}\leq C\sqrt{2\pi}, then 0<r<Rπ2\forall 0<r<R\leq\frac{\pi}{2},

(36) BR(x)(f(y)Cd(y,x))𝑑vol(y)Br(x)(f(y)Cd(y,x))𝑑vol(y),\fint_{B_{R}(x)}\left(f(y)-Cd(y,x)\right)d{\rm vol}(y)\leq\fint_{B_{r}(x)}\left(f(y)-Cd(y,x)\right)d{\rm vol}(y),

where d(y,x)d(y,x) is the distance between yy and xx in the standard 𝕊2{\mathbb{S}}^{2}.

Proof.

Step 1.

(37) Br(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle\int_{B_{r}(x)}\left(f(y)-Cd(y,x)\right)d{\rm vol}(y)
(38) =\displaystyle= 0r(Bs(x)(fCs))𝑑s\displaystyle\int^{r}_{0}\left(\int_{\partial B_{s}(x)}(f-Cs)\right)ds
(39) =\displaystyle= 0r(2πsins)(Bs(x)(fCs))𝑑s\displaystyle\int^{r}_{0}(2\pi\sin s)\left(\fint_{\partial B_{s}(x)}(f-Cs)\right)ds
(40) \displaystyle\geq Br(x)(fCr)0r2πsinsds(by(35)andsr)\displaystyle\fint_{\partial B_{r}(x)}(f-Cr)\cdot\int^{r}_{0}2\pi\sin sds\quad(\text{by}\ \ (\ref{eqn: spherical mean non-increasing})\ \ \text{and}\ \ s\leq r)
(41) =\displaystyle= Vol(Br(x))Br(x)(fCr).\displaystyle{\rm Vol}(B_{r}(x))\fint_{\partial B_{r}(x)}(f-Cr).

So

(42) Br(x)(f(y)Cd(y,x))𝑑vol(y)Br(x)(f(y)Cr)\fint_{B_{r}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)\geq\fint_{\partial B_{r}(x)}(f(y)-Cr)

Step 2. Let Ar,R(x)=BR(x)Br(x)A_{r,R}(x)=B_{R}(x)\setminus B_{r}(x). Similar as in step 1, we have

(43) Ar,R(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle\int_{A_{r,R}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)
(44) =\displaystyle= rR(Bs(x)(fCs)𝑑σ)𝑑s\displaystyle\int^{R}_{r}\left(\int_{\partial B_{s}(x)}(f-Cs)d\sigma\right)ds
(45) =\displaystyle= rR(2πsins)(Bs(x)(fCs)𝑑σ)𝑑s\displaystyle\int^{R}_{r}(2\pi\sin s)\left(\fint_{\partial B_{s}(x)}(f-Cs)d\sigma\right)ds
(46) \displaystyle\leq Br(x)(fCr)𝑑σrR(2πsins)𝑑s(by(35)andsr)\displaystyle\fint_{\partial B_{r}(x)}(f-Cr)d\sigma\cdot\int^{R}_{r}(2\pi\sin s)ds\quad(\text{by}\ \ (\ref{eqn: spherical mean non-increasing})\ \ \text{and}\ \ s\geq r)
(47) =\displaystyle= vol(Ar,R(x))Br(x)(fCr)𝑑σ\displaystyle{\rm vol}(A_{r,R}(x))\fint_{\partial B_{r}(x)}(f-Cr)d\sigma

So

(48) Ar,R(x)(f(y)Cd(y,x))𝑑vol(y)Br(x)(fCr)𝑑σ.\fint_{A_{r,R}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)\leq\fint_{\partial B_{r}(x)}(f-Cr)d\sigma.

Step 3. By combining (42) and (48), we obtain that for 0<r<Rπ20<r<R\leq\frac{\pi}{2}

(49) Ar,R(x)(f(y)Cd(y,x))𝑑vol(y)Br(x)(f(y)Cd(y,x))𝑑vol(y).\fint_{A_{r,R}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)\leq\fint_{B_{r}(x)}(f(y)-Cd(y,x))d{\rm vol}(y).

Step 4.

(50) BR(x)(fCd(y,x))𝑑vol(y)\displaystyle\int_{B_{R}(x)}(f-Cd(y,x))d{\rm vol}(y)
(51) =\displaystyle= Br(x)(fCd(y,x))𝑑vol(y)+Ar,R(x)(fCd(y,x))𝑑vol(y)\displaystyle\int_{B_{r}(x)}(f-Cd(y,x))d{\rm vol}(y)+\int_{A_{r,R}(x)}(f-Cd(y,x))d{\rm vol}(y)
(53) \displaystyle\leq Br(x)(fCd(y,x))𝑑vol(y)\displaystyle\int_{B_{r}(x)}(f-Cd(y,x))d{\rm vol}(y)
+Vol(Ar,R(x))Br(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle+{\rm Vol}(A_{r,R}(x))\cdot\fint_{B_{r}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)
(54) =\displaystyle= (Vol(Br(x))+vol(Ar,R(x)))Br(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle\left({\rm Vol}(B_{r}(x))+{\rm vol}(A_{r,R}(x))\right)\fint_{B_{r}(x)}(f(y)-Cd(y,x))d{\rm vol}(y)
(55) =\displaystyle= Vol(BR(x))Br(x)(f(y)Cd(y,x))𝑑vol(y).\displaystyle{\rm Vol}(B_{R}(x))\fint_{B_{r}(x)}(f(y)-Cd(y,x))d{\rm vol}(y).

This completes the proof. ∎

3. W1,pW^{1,p} limit of warping function for 1p<21\leq p<2

In this section, we study the LqL^{q} pre-compactness of a sequence of positive smooth functions fjf_{j} satisfying the inequalities

(56) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Here VV is a positive constant. By Lemmas 2.1 and 2.2, the inequlities in (56) are equivalent to the requirements that the Riemannian manifolds 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} have nonnegative scalar curvature and uniform volume upper bound.

In Subsection 3.1, we prove that a sequence of positive smooth functions fjf_{j} on 𝕊2{\mathbb{S}}^{2} satisfying requirements in (56) has a convergent subsequence in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for any 1q<+1\leq q<+\infty, and that the limit function is in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}) for any 1p<21\leq p<2 [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. W1,pW^{1,p} limit function for p<2p<2

We first derive the gradient estimate for the sequence of function lnfj\ln f_{j} in Lemma 3.1, which is used to obtain LpL^{p} estimate for fif_{i} by using Moser-Trudinger inequality in Lemma 3.2.

Lemma 3.1.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(57) Δfjfj,j.\Delta f_{j}\leq f_{j},\quad\forall j\in\mathbb{N}.

We have

(58) lnfjL2(𝕊2)2Vol(𝕊2),j.\|\nabla\ln f_{j}\|_{L^{2}({\mathbb{S}}^{2})}^{2}\leq{\rm Vol}({\mathbb{S}}^{2}),\quad\forall j\in\mathbb{N}.
Proof.

Note that

(59) Δlnfj=Δfjfj|fj|2fj2.\Delta\ln f_{j}=\frac{\Delta f_{j}}{f_{j}}-\frac{|\nabla f_{j}|^{2}}{f_{j}^{2}}.

By equation (59) and the assumption, we have

(60) |lnfj|2=|fj|2fj2=ΔfjfjΔlnfj1Δlnfj.|\nabla\ln f_{j}|^{2}=\frac{|\nabla f_{j}|^{2}}{f_{j}^{2}}=\frac{\Delta f_{j}}{f_{j}}-\Delta\ln f_{j}\leq 1-\Delta\ln f_{j}.

Integrating it over 𝕊2{\mathbb{S}}^{2}, and using Stokes’ theorem, we get

(61) lnfjL2(𝕊2)2=𝕊2|lnfj|2Vol(𝕊2).\|\nabla\ln f_{j}\|_{L^{2}({\mathbb{S}}^{2})}^{2}=\int_{{\mathbb{S}}^{2}}|\nabla\ln f_{j}|^{2}\leq{\rm Vol}({\mathbb{S}}^{2}).

Lemma 3.2.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(62) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Then we have

(63) fjLp(𝕊2)p4πexp(Vp8π2+p24),\|f_{j}\|_{L^{p}({\mathbb{S}}^{2})}^{p}\leq 4\pi\exp\left(\frac{Vp}{8\pi^{2}}+\frac{p^{2}}{4}\right),

for all jj\in\mathbb{N} and p[1,+)p\in[1,+\infty).

Proof.

By the Moser-Trudinger inequality (inequality (25) in [14]), for any smooth function ψ:𝕊2\psi:{\mathbb{S}}^{2}\to\mathbb{R} we have

(64) 𝕊2eψ𝑑vol𝕊24πexp(14π𝕊2(ψ+14|ψ|2)𝑑vol𝕊2).\int_{{\mathbb{S}}^{2}}e^{\psi}d{\rm vol}_{{\mathbb{S}}^{2}}\leq 4\pi\exp\left(\frac{1}{4\pi}\int_{{\mathbb{S}}^{2}}\left(\psi+\frac{1}{4}|\nabla\psi|^{2}\right)d{\rm vol}_{{\mathbb{S}}^{2}}\right).

Here \nabla is the Levi-Civita connection of the standard metric g𝕊2g_{{\mathbb{S}}^{2}} and dvol𝕊2d{\rm vol}_{{\mathbb{S}}^{2}} is the volume form on 𝕊2{\mathbb{S}}^{2} with respect to the standard metric g𝕊2g_{{\mathbb{S}}^{2}}. Take ψ=plnfj\psi=p\ln f_{j}, then we have

(65) fjLp(𝕊2)p\displaystyle\|f_{j}\|_{L^{p}({\mathbb{S}}^{2})}^{p} =\displaystyle= 𝕊2fjp𝑑vol𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}f_{j}^{p}d{\rm vol}_{{\mathbb{S}}^{2}}
(66) =\displaystyle= 𝕊2eplnfj𝑑vol𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}e^{p\ln f_{j}}d{\rm vol}_{{\mathbb{S}}^{2}}
(67) \displaystyle\leq 4πexp(14π𝕊2(plnfj+p24|lnfj|2)𝑑vol𝕊2).\displaystyle 4\pi\exp\left(\frac{1}{4\pi}\int_{{\mathbb{S}}^{2}}\left(p\ln f_{j}+\frac{p^{2}}{4}|\nabla\ln f_{j}|^{2}\right)d{\rm vol}_{{\mathbb{S}}^{2}}\right).

By the fact that lnxx,x>0\ln x\leq x,\forall x>0, we have

(68) 𝕊2lnfj𝕊2fjV2π.\int_{{\mathbb{S}}^{2}}\ln f_{j}\leq\int_{{\mathbb{S}}^{2}}f_{j}\leq\frac{V}{2\pi}.

On the hand, by Lemma 3.1 we have

(69) 𝕊2|lnfj|2vol(𝕊2)=4π.\int_{{\mathbb{S}}^{2}}|\nabla\ln f_{j}|^{2}\leq{\rm vol}({\mathbb{S}}^{2})=4\pi.

This completes the proof. ∎

Next, we show that such sequence of function is uniformly bounded in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}) for p[1,2)p\in[1,2).

Lemma 3.3.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(70) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Then the sequence is uniformly bounded in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}) for p[1,2)p\in[1,2), i.e. for each p[1,2)p\in[1,2), there exists a constant C(p)C(p) such that

(71) fjW1,p(𝕊2)C(p),j.\|f_{j}\|_{W^{1,p}({\mathbb{S}}^{2})}\leq C(p),\quad\forall j\in\mathbb{N}.
Proof.

For any 1p<21\leq p<2,

(72) |fj|p=|lnfj|p|fj|p.|\nabla f_{j}|^{p}=|\nabla\ln f_{j}|^{p}\cdot|f_{j}|^{p}.

The Cauchy-Schwarz inequality implies that

(73) fjLp(𝕊2)\displaystyle\|\nabla f_{j}\|_{L^{p}(\mathbb{S}^{2})}
(74) =\displaystyle= (𝕊2|lnfj|p|fj|p)1p\displaystyle\left(\int_{\mathbb{S}^{2}}|\nabla\ln f_{j}|^{p}\cdot|f_{j}|^{p}\right)^{\frac{1}{p}}
(75) \displaystyle\leq lnfjL2(𝕊2)fjL2p2p(𝕊2)\displaystyle\|\nabla\ln f_{j}\|_{L^{2}(\mathbb{S}^{2})}\cdot\|f_{j}\|_{L^{\frac{2p}{2-p}}(\mathbb{S}^{2})}
(76) \displaystyle\leq lnfjL2(𝕊2)(fjL2p2p(𝕊2)+Vol(𝕊2))\displaystyle\|\nabla\ln f_{j}\|_{L^{2}(\mathbb{S}^{2})}\cdot\left(\|f_{j}\|_{L^{\frac{2p}{2-p}}(\mathbb{S}^{2})}+{\rm Vol}({\mathbb{S}}^{2})\right)
(77) \displaystyle\leq (vol(𝕊2))12((4π)2p2pexp(V8π2+p2(2p))+Vol(𝕊2)).\displaystyle\left({\rm vol}(\mathbb{S}^{2})\right)^{\frac{1}{2}}\left((4\pi)^{\frac{2-p}{2p}}\exp\left(\frac{V}{8\pi^{2}}+\frac{p}{2(2-p)}\right)+{\rm Vol}({\mathbb{S}}^{2})\right).

Here in the last step, we used Lemma 3.1 and Lemma 3.2. Moreover, by Lemma 3.2 again, for each p[1,2)p\in[1,2), fjLp(𝕊2)\|f_{j}\|_{L^{p}({\mathbb{S}}^{2})} is uniformly bounded for all jj\in\mathbb{N}. Hence for each p[1,2)p\in[1,2), fjW1,p(𝕊2)\|f_{j}\|_{W^{1,p}({\mathbb{S}}^{2})} is uniformly bounded for all jj\in\mathbb{N}. ∎

We use the uniform W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}) bound to prove convergence in the following lemma.

Lemma 3.4.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(78) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Then for each fixed p[1,2)p\in[1,2), there exists a subsequence {fjk(p)}k=1\{f_{j^{(p)}_{k}}\}_{k=1}^{\infty} and f,pW1,p(𝕊2)f_{\infty,p}\in W^{1,p}(\mathbb{S}^{2}) such that

(79) fjk(p)f,p,inLq(𝕊2),f_{j^{(p)}_{k}}\rightarrow f_{\infty,p},\quad\text{in}\ \ L^{q}({\mathbb{S}}^{2}),

for each 1q<2p2p1\leq q<\frac{2p}{2-p}.

Moreover, for any φC(𝕊2)\varphi\in C^{\infty}({\mathbb{S}}^{2}),

𝕊2(fjk(p)φ+fjk(p),φ)dvolg𝕊2𝕊2(f,pφ+f,p,φ)dvolg𝕊2,\int_{{\mathbb{S}}^{2}}\left(f_{j^{(p)}_{k}}\varphi+\langle\nabla f_{j^{(p)}_{k}},\nabla\varphi\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\rightarrow\int_{{\mathbb{S}}^{2}}\left(f_{\infty,p}\varphi+\langle\nabla f_{\infty,p},\nabla\varphi\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

as jk(p)j^{(p)}_{k}\rightarrow\infty, where f,p\nabla f_{\infty,p} is the weak gradient of f,pf_{\infty,p}.

Proof.

For each fixed p[1,2)p\in[1,2), by using Rellich-Kondrachov compactness theorem, the uniform estimate of Sobolev norms in Lemma 3.3 implies that there exists a subsequence of {fj}\{f_{j}\}, which is still denoted by {fj}\{f_{j}\}, converging to f,pf_{\infty,p} in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for 1q<2p2p1\leq q<\frac{2p}{2-p}. Then by the weak compactness in LpL^{p} space (see, e.g. Theorem 1.42 in [5]), we can obtain that f,pW1,p(𝕊2)f_{\infty,p}\in W^{1,p}({\mathbb{S}}^{2}). Indeed, fjW1,p(𝕊2)C\|f_{j}\|_{W^{1,p}({\mathbb{S}}^{2})}\leq C for all jj\in\mathbb{N} implies that fjLp(𝕊2)\|f_{j}\|_{L^{p}({\mathbb{S}}^{2})} and fjLp(𝕊2)\|\nabla f_{j}\|_{L^{p}({\mathbb{S}}^{2})} are both uniformly bounded. Then the weak compactness in LpL^{p} space implies that there exist a further subsequence, denoted by fjk(p)f_{j^{(p)}_{k}}, and XLp(𝕊2,T𝕊2)X\in L^{p}({\mathbb{S}}^{2},{\rm T}{\mathbb{S}}^{2}) such that

(80) fjk(p)XinLp(𝕊2,T𝕊2),\nabla f_{j^{(p)}_{k}}\rightharpoonup X\quad\text{in}\ \ L^{p}({\mathbb{S}}^{2},{\rm T}{\mathbb{S}}^{2}),

i.e.

(81) 𝕊2fjk(p),Y𝑑volg𝕊2𝕊2X,Y𝑑volg𝕊2,YC(𝕊2,T𝕊2).\int_{{\mathbb{S}}^{2}}\langle\nabla f_{j^{(p)}_{k}},Y\rangle d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\rightarrow\int_{{\mathbb{S}}^{2}}\langle X,Y\rangle d{\rm vol}_{g_{{\mathbb{S}}^{2}}},\quad\forall Y\in C^{\infty}({\mathbb{S}}^{2},{\rm T}{\mathbb{S}}^{2}).

On the other hand,

(82) 𝕊2fjk(p),Y𝑑volg𝕊2=𝕊2fjdivY𝑑volg𝕊2𝕊2f,pdivY𝑑volg𝕊2,\int_{{\mathbb{S}}^{2}}\langle\nabla f_{j^{(p)}_{k}},Y\rangle d{\rm vol}_{g_{{\mathbb{S}}^{2}}}=\int_{{\mathbb{S}}^{2}}f_{j}{\rm div}Yd{\rm vol}_{g_{{\mathbb{S}}^{2}}}\rightarrow\int_{{\mathbb{S}}^{2}}f_{\infty,p}{\rm div}Yd{\rm vol}_{g_{{\mathbb{S}}^{2}}},

since fjk(p)f,pf_{j^{(p)}_{k}}\rightarrow f_{\infty,p} in LpL^{p}. Thus,

(83) 𝕊2f,pdivY𝑑volg𝕊2=𝕊2X,Y𝑑volg𝕊2,YC(𝕊2,T𝕊2).\int_{{\mathbb{S}}^{2}}f_{\infty,p}{\rm div}Yd{\rm vol}_{g_{{\mathbb{S}}^{2}}}=\int_{{\mathbb{S}}^{2}}\langle X,Y\rangle d{\rm vol}_{g_{{\mathbb{S}}^{2}}},\quad\forall Y\in C^{\infty}({\mathbb{S}}^{2},{\rm T}{\mathbb{S}}^{2}).

Therefore, X=f,pX=\nabla f_{\infty,p} is the gradient of f,pf_{\infty,p} in the sense of distribution, and so f,pW1,p(𝕊2,g𝕊2)f_{\infty,p}\in W^{1,p}({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}). For any φC(𝕊2)\varphi\in C^{\infty}({\mathbb{S}}^{2}), by taking Y=φY=\nabla\varphi in (81), we obtain

(84) 𝕊2(fjk(p)φ+fjk(p),φ)dvolg𝕊2𝕊2(f,pφ+f,p,φ)dvolg𝕊2.\int_{{\mathbb{S}}^{2}}\left(f_{j^{(p)}_{k}}\varphi+\langle\nabla f_{j^{(p)}_{k}},\nabla\varphi\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\rightarrow\int_{{\mathbb{S}}^{2}}\left(f_{\infty,p}\varphi+\langle\nabla f_{\infty,p},\nabla\varphi\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}.

Now we use Lemma 3.4 and diagonal argument to find a subsequence converging in LqL^{q} for all q1q\geq 1 and prove the following proposition:

Proposition 3.5.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(85) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Then there exists a subsequence {fjk}k=1\{f_{j_{k}}\}_{k=1}^{\infty} and fW1,p(𝕊2)f_{\infty}\in W^{1,p}(\mathbb{S}^{2}) for all p[1,2)p\in[1,2), such that

(86) fjkf,inLq(𝕊2),q[1,).f_{j_{k}}\rightarrow f_{\infty},\quad\text{in}\ \ L^{q}({\mathbb{S}}^{2}),\ \ \forall q\in[1,\infty).

Moreover, for any φC(𝕊2)\varphi\in C^{\infty}({\mathbb{S}}^{2}),

(87) 𝕊2(fjkφ+fjk,φ)𝑑volg𝕊2𝕊2(fφ+f,φ)𝑑volg𝕊2,\int_{{\mathbb{S}}^{2}}\left(f_{j_{k}}\varphi+\langle\nabla f_{j_{k}},\nabla\varphi\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\rightarrow\int_{{\mathbb{S}}^{2}}\left(f_{\infty}\varphi+\langle\nabla f_{\infty},\nabla\varphi\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

as jkj_{k}\rightarrow\infty, where f\nabla f_{\infty} is the weak gradient of ff_{\infty}.

Proof.

The proof is a diagonal argument. We apply Lemma 3.4 for p=21n+1,n=1,2,3,p=2-\frac{1}{n+1},n=1,2,3,\dots.

For n=1n=1, by applying Lemma 3.4 to {fj}j=1\{f_{j}\}_{j=1}^{\infty} and p=212p=2-\frac{1}{2}, we obtain a subsequence, denoted by fjk(1),1f_{j^{(1)}_{k},1}, and f,1W1,212f_{\infty,1}\in W^{1,2-\frac{1}{2}} such that

(88) fjk(1),1f,1inLq(𝕊2),1q<6,ask.f_{j^{(1)}_{k},1}\rightarrow f_{\infty,1}\ \ \text{in}\ \ L^{q}({\mathbb{S}}^{2}),\ \ \forall 1\leq q<6,\ \ \text{as}\ \ k\rightarrow\infty.

For n=2n=2, by applying Lemma 3.4 to the subsequence {fjk(1),1}k=1\left\{f_{j^{(1)}_{k},1}\right\}_{k=1}^{\infty} and p=213p=2-\frac{1}{3}, we obtain a subsequence, {fjk(2),2}k=1{fjk(1),1}k=1\left\{f_{j^{(2)}_{k},2}\right\}_{k=1}^{\infty}\subset\left\{f_{j^{(1)}_{k},1}\right\}_{k=1}^{\infty}, and f,2W1,213f_{\infty,2}\in W^{1,2-\frac{1}{3}} such that

(89) fjk(2),2f,2inLq(𝕊2),1q<10,ask.f_{j^{(2)}_{k},2}\rightarrow f_{\infty,2}\ \ \text{in}\ \ L^{q}({\mathbb{S}}^{2}),\ \ \forall 1\leq q<10,\ \ \text{as}\ \ k\rightarrow\infty.

Then by repeating this process for n=3,4,5,n=3,4,5,\dots, we can obtain a family of decreasing subsequence {fjk(n),n}k=1{fjk(n1),n1}k=1\left\{f_{j^{(n)}_{k},n}\right\}_{k=1}^{\infty}\subset\left\{f_{j^{(n-1)}_{k},n-1}\right\}_{k=1}^{\infty} and f,nW1,21n+1f_{\infty,n}\in W^{1,2-\frac{1}{n+1}} for all nn\in\mathbb{N}, such that for each fixed nn\in\mathbb{N}

(90) fjk(n),nf,ninLq(𝕊2),1q<4n+2,ask.f_{j^{(n)}_{k},n}\rightarrow f_{\infty,n}\ \ \text{in}\ \ L^{q}({\mathbb{S}}^{2}),\ \ \forall 1\leq q<4n+2,\ \ \text{as}\ \ k\rightarrow\infty.

Now we take the diagonal subsequence {fjk:=ffjk(k),kk}\left\{f_{j_{k}}:=f_{f_{j^{(k)}_{k},k}}\mid k\in\mathbb{N}\right\}. By the construction of fjkf_{j_{k}} and 4k+2+4k+2\rightarrow+\infty as k+k\rightarrow+\infty, we have that {fjk}\{f_{j_{k}}\} is a Cauchy sequence in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for all q[1,)q\in[1,\infty). Thus there exists fLq(𝕊2)f_{\infty}\in L^{q}({\mathbb{S}}^{2}) such that

(91) fjkfinLq(𝕊2),ask,q[1,).f_{j_{k}}\rightarrow f_{\infty}\ \ \text{in}\ \ \in L^{q}({\mathbb{S}}^{2}),\ \ \text{as}\ \ k\rightarrow\infty,\ \ \forall q\in[1,\infty).

Then by the uniqueness of L2L^{2} limit, f=f,nf_{\infty}=f_{\infty,n} in L2(𝕊2)L^{2}({\mathbb{S}}^{2}) for all nn\in\mathbb{N}. Furthermore, because f,nW1,21n+1(𝕊2)f_{\infty,n}\in W^{1,2-\frac{1}{n+1}}({\mathbb{S}}^{2}) and 21n+122-\frac{1}{n+1}\rightarrow 2^{-} as nn\rightarrow\infty, we see that the LpL^{p} norm of the weak derivative of ff_{\infty} is bounded for any p[1,2)p\in[1,2). Thus fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for all p[1,2)p\in[1,2).

Finally, the last claim in (87) follows from that {fjk}k=1{fjk(1),1}k=1\left\{f_{j_{k}}\right\}_{k=1}^{\infty}\subset\left\{f_{j^{(1)}_{k},1}\right\}_{k=1}^{\infty} and the corresponding convergence in Lemma 3.4 for p=212p=2-\frac{1}{2}, in particular for the subsequence {fjk(1),1}k=1\left\{f_{j^{(1)}_{k},1}\right\}_{k=1}^{\infty}. ∎

Remark 3.6.

The extreme example constructed by Christina Sormani and authors in [19] shows that W1,pW^{1,p} regularity for p<2p<2 is the best regularity we can expect for ff_{\infty} in general (see Lemma 3.4 in [19]).

3.2. Lower semi-continuous representative of the limit function

For the limit function ff_{\infty} obtained in Proposition 3.5, Lebesgue-Besicovitch differential theorem implies that

(92) limr0Br(x)f𝑑volg𝕊2=f(x)\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}=f_{\infty}(x)

holds for a.e. x𝕊2x\in{\mathbb{S}}^{2} with respect to the volume measure dvolg𝕊2d{\rm vol}_{g_{{\mathbb{S}}^{2}}}. 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 x𝕊2x\in{\mathbb{S}}^{2}, and that the limit produces a lower semi-continuous function.

Proposition 3.7.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of smooth positive functions on 𝕊2{\mathbb{S}}^{2} satisfying

(93) Δfjfj,𝕊2fj𝑑vol𝕊2V2π,j.\Delta f_{j}\leq f_{j},\quad\int_{{\mathbb{S}}^{2}}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}\leq\frac{V}{2\pi},\quad\forall j\in\mathbb{N}.

Then the limit function, ff_{\infty}, obtained in Proposition 3.5, has the following properties.

  1. (i)

    For each fixed x𝕊2x\in{\mathbb{S}}^{2}, the ball average

    (94) Br(x)(f(y)Cd(y,x))𝑑vol(y)\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)

    is non-increasing in r(0,π2)r\in\left(0,\frac{\pi}{2}\right), where CC is a positive real number such that supjfjL2(𝕊2)C2π\sup_{j\in\mathbb{N}}\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}\leq C\sqrt{2\pi}. Note that the existence of such CC is guaranteed by Lemma 3.2.

  2. (ii)

    Consequently, the limit

    (95) f¯(x):=limr0Br(x)f=limr0Br(x)(f(y)Cd(y,x))𝑑vol(y)\overline{f_{\infty}}(x):=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)

    exists, allowing ++\infty as a limit, for every x𝕊2x\in{\mathbb{S}}^{2}. Moreover, f¯\overline{f_{\infty}} is a lower semi-continuous function on 𝕊2{\mathbb{S}}^{2}.

Proof.

By Lemma 3.2, there exists CC\in\mathbb{R} such that

(96) fjL2(𝕊2)C2π,j.\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}\leq C\sqrt{2\pi},\ \ \forall j\in\mathbb{N}.

Then by applying Proposition 2.6 to functions fjf_{j}, we obtain that for any fixed x𝕊2x\in{\mathbb{S}}^{2}

(97) BR(x)(fj(y)Kd(y,x))𝑑vol(y)Br(x)(fj(y)Cd(y,x))𝑑vol(y)\fint_{B_{R}(x)}(f_{j}(y)-Kd(y,x))d{\rm vol}(y)\leq\fint_{B_{r}(x)}(f_{j}(y)-Cd(y,x))d{\rm vol}(y)

holds for any 0<r<R<π20<r<R<\frac{\pi}{2} and all jj\in\mathbb{N}.

By Proposition 3.5 fjff_{j}\rightarrow f_{\infty} in L1(𝕊2)L^{1}({\mathbb{S}}^{2}). Then for any fixed x𝕊2x\in{\mathbb{S}}^{2}, and any fixed 0<r<R<π20<r<R<\frac{\pi}{2}, by taking the limit as j+j\rightarrow+\infty, we obtain

(98) BR(x)(f(y)Cd(y,x))𝑑vol(y)Br(x)(f(y)Cd(y,x))𝑑vol(y),\fint_{B_{R}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)\leq\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y),

So for each fixed x𝕊2x\in{\mathbb{S}}^{2}, the ball average

(99) Br(x)(f(y)Cd(y,x))𝑑vol(y)\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)

is non-increasing for r(0,π2)r\in\left(0,\frac{\pi}{2}\right). Therefore, for any x𝕊2x\in{\mathbb{S}}^{2} the limit

(100) limr0Br(x)(f(y)Cd(y,x))𝑑vol(y)\lim_{r\rightarrow 0}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)

exists as a finite number or ++\infty.

On the other hand, by direct calculation

(101) Br(x)d(y,x)𝑑vol(y)=0r2πssinsds0r2πsin(s)𝑑s=sinrrcosr1cosr0,\fint_{B_{r}(x)}d(y,x)d{\rm vol}(y)=\frac{\int^{r}_{0}2\pi s\sin sds}{\int^{r}_{0}2\pi\sin(s)ds}=\frac{\sin r-r\cos r}{1-\cos r}\rightarrow 0,

as r0r\rightarrow 0. Thus the limit

(102) f¯(x):=limr0Br(x)f=limr0Br(x)(f(y)Cd(y,x))𝑑vol(y)\overline{f_{\infty}}(x):=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)

exists for all x𝕊2x\in{\mathbb{S}}^{2}.

For each fixed 0<r<π20<r<\frac{\pi}{2}, we have that Br(x)(f(y)Cd(y,x))𝑑vol(y)\fint_{B_{r}(x)}(f_{\infty}(y)-Cd(y,x))d{\rm vol}(y) is a continuous function of x𝕊2x\in{\mathbb{S}}^{2}, since fL2(𝕊2)f_{\infty}\in L^{2}({\mathbb{S}}^{2}), Cd(y,x)CπCd(y,x)\leq C\pi, and Area(Br(x))=2πsinr{\rm Area}(B_{r}(x))=2\pi\sin r for all x𝕊2x\in{\mathbb{S}}^{2}. Then by the monotonicity in (98), we have

(103) f¯(x)=supr>0Br(x)(f(y)Cd(y,x))𝑑vol(y).\overline{f_{\infty}}(x)=\sup_{r>0}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y).

In other words, f¯\overline{f_{\infty}} is the supremum of a sequence of continuous function. Thus f¯\overline{f_{\infty}} is lower semi-continuous. ∎

Remark 3.8.

Recall that by (92), limr0Br(x)f𝑑volg𝕊2=f(x)\lim\limits_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}=f_{\infty}(x) hold for a.e. x𝕊2x\in{\mathbb{S}}^{2}, thus f¯(x)=f(x)\overline{f_{\infty}}(x)=f_{\infty}(x) holds for a.e. x𝕊2x\in{\mathbb{S}}^{2}. So as a W1,pW^{1,p} function, ff_{\infty} has a lower semi-continuous representative f¯\overline{f_{\infty}}.

4. Positivity of the limit warping functions

In this section, we prove that the limit warping function ff_{\infty} has a positive essential infimum, provided that the Riemannian manifold 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} satisfies both requirements in (56) and the MinA\operatorname{MinA} 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 W1,2W^{1,2} regularity, but in general we only have fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for p<2p<2 [Remark 3.6]. To overcome this difficulty, in Subsection 4.1, we consider the truncation of warping functions f¯jK\bar{f}^{K}_{j} as defined in Definition 4.1, and obtain a W1,2(𝕊2)W^{1,2}({\mathbb{S}}^{2}) limit function f¯K\bar{f}^{K}_{\infty} for the sequence of truncated function f¯jK\bar{f}^{K}_{j} [Lemma 4.4]. This enables us to apply maximum principle for weak solutions (Theorem 8.19 in [6]) to f¯K\bar{f}^{K}_{\infty}, and prove that either inff¯K>0\inf\bar{f}^{K}_{\infty}>0 or f¯K0\bar{f}^{K}_{\infty}\equiv 0 on 𝕊2{\mathbb{S}}^{2} [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 MinA(𝕊×f𝕊1)\operatorname{MinA}({\mathbb{S}}\times_{f}{\mathbb{S}}^{1}) in terms of L1L^{1} norm of the warping function ff, provided that the L2L^{2} norm of ff 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 fjf_{j}, if the warped product manifolds 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} satisfy requirements in (56) and the MinA\operatorname{MinA} condition [Proposition 4.15].

4.1. W1,2W^{1,2} regularity of limit of truncated warping functions

We define the truncation of a function firstly:

Definition 4.1.

Let f:𝕊2f:{\mathbb{S}}^{2}\to\mathbb{R} be a positive smooth function. Let K>0K>0 be a real number, for each x𝕊2x\in{\mathbb{S}}^{2}, we define

(104) f¯K(x)={f(x), if f(x)<K,K, if f(x)K.\bar{f}^{K}(x)=\begin{cases}f(x),&\text{ if }\ \ f(x)<K,\\ K,&\text{ if }\ \ f(x)\geq K.\end{cases}

Then f¯K\bar{f}^{K} is a positive continuous function on 𝕊2{\mathbb{S}}^{2} with the maximal value not greater than KK.

From the definition we can prove the following lemma:

Lemma 4.2.

Let f:𝕊2f:{\mathbb{S}}^{2}\to\mathbb{R} be a positive smooth function, and let K>0K>0 be a regular value of the function ff. If

(105) Δff\Delta f\leq f

then for all uW1,2(𝕊2)u\in W^{1,2}({\mathbb{S}}^{2}) such that u0u\geq 0 we have

(106) 𝕊2u,f¯K𝕊2uf¯K.-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}^{K}\rangle\leq\int_{{\mathbb{S}}^{2}}u\bar{f}^{K}.
Proof.

By Theorem 4.4 from [5], we have for all K>0K>0

(107) f¯K={f,a.e. on {f(x)<K},0,a.e. on {f(x)K}.\nabla\bar{f}^{K}=\begin{cases}\nabla f,&\text{a.e. on }\{f(x)<K\},\\ 0,&\text{a.e. on }\{f(x)\geq K\}.\end{cases}

As a result we have

(108) 𝕊2u,f¯K={f<K}u,f={f<K}uΔf{f<K}uνf.\begin{split}-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}^{K}\rangle&=-\int_{\{f<K\}}\langle\nabla u,\nabla f\rangle\\ &=\int_{\{f<K\}}u\Delta f-\int_{\partial\{f<K\}}u\partial_{\nu}f.\end{split}

Here, since KK is a regular value of ff, from the Regular Level Set Theorem we know that the level set {f=K}={f<K}\{f=K\}=\partial\{f<K\} is am embedded submanifold of dimension 11 in 𝕊2{\mathbb{S}}^{2}. Hence we can apply Stokes’ theorem to get the last step. Moreover, since ν\nu is the outer unit normal vector on the boundary of the set {f<K}\{f<K\}, we have

(109) νf0.\partial_{\nu}f\geq 0.

Hence we can drop the boundary term to get the inequality

(110) 𝕊2u,f¯K{f<K}uΔf.-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}^{K}\rangle\leq\int_{\{f<K\}}u\Delta f.

Since

(111) Δff,\Delta f\leq f,

we have

(112) 𝕊2u,f¯K{f<K}uΔf{f<K}uf𝕊2uf¯K.-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}^{K}\rangle\leq\int_{\{f<K\}}u\Delta f\leq\int_{\{f<K\}}uf\leq\int_{{\mathbb{S}}^{2}}u\bar{f}^{K}.

This finishes the proof. ∎

We can prove similar results for a sequence of functions:

Lemma 4.3.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of smooth positive function defined on 𝕊2{\mathbb{S}}^{2}. If

(113) Δfjfj,j,\Delta f_{j}\leq f_{j},\ \ \forall j\in\mathbb{N},

then there exists K>0K>0 such that for all uW1,2(𝕊2)u\in W^{1,2}({\mathbb{S}}^{2}) with u0u\geq 0 we have

(114) 𝕊2u,f¯jK𝕊2uf¯jKj.-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}_{j}^{K}\rangle\leq\int_{{\mathbb{S}}^{2}}u\bar{f}_{j}^{K}\quad\forall j\in\mathbb{N}.

Moreover, we can choose KK as large as we want.

Proof.

Note that if 0<Kinfx𝕊2fj(x)0<K\leq\inf\limits_{x\in{\mathbb{S}}^{2}}f_{j}(x) for some ii then we have f¯jK(x)=K\bar{f}_{j}^{K}(x)=K. On the other hand, if supx𝕊2f(x)K\sup\limits_{x\in{\mathbb{S}}^{2}}f(x)\leq K for some ii then f¯jK(x)=fj(x)\bar{f}_{j}^{K}(x)=f_{j}(x). Either way the inequality (114) holds.

In general, by Sard’s theorem, for each function fjf_{j}, the critical values of fjf_{j} 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 K>0K>0 such that for each fjf_{j} either KK is a regular value or fj1({K})=f_{j}^{-1}(\{K\})=\emptyset. By Lemma 4.2 we get inequality (114). Moreover, we can choose KK 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 K>0K>0 be a real number. Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive smooth functions on 𝕊2{\mathbb{S}}^{2} satisfying

(115) Δfjfj,j.\Delta f_{j}\leq f_{j},\quad\forall j\in\mathbb{N}.

Then the sequence {f¯jK}j=1\{\bar{f}^{K}_{j}\}_{j=1}^{\infty} is uniformly bounded in W1,2(𝕊2)W^{1,2}({\mathbb{S}}^{2}):

(116) f¯jKW1,2(𝕊2)2Kvol(𝕊2).\|\bar{f}^{K}_{j}\|_{W^{1,2}({\mathbb{S}}^{2})}\leq 2K{\rm vol}({\mathbb{S}}^{2}).

As a result, there exists f¯KW1,2(𝕊2)\bar{f}_{\infty}^{K}\in W^{1,2}({\mathbb{S}}^{2}) such that f¯jK\bar{f}^{K}_{j} converges to f¯K\bar{f}^{K}_{\infty} in L2(𝕊2)L^{2}({\mathbb{S}}^{2}), and that f¯jK\bar{f}^{K}_{j} converges to f¯K\bar{f}^{K}_{\infty} weakly in W1,2(𝕊2)W^{1,2}({\mathbb{S}}^{2}).

Proof.

By definition of the cutoff in Definition 4.1, we get

(117) f¯jKL2(𝕊2)Kvol(𝕊2).\|\bar{f}_{j}^{K}\|_{L^{2}({\mathbb{S}}^{2})}\leq K\sqrt{{\rm vol}({\mathbb{S}}^{2})}.

By Theorem 4.4 from [5], we have for all K>0K>0 and for each ii

(118) f¯jK={fj,a.e. on {fj(x)<K},0,a.e. on {fj(x)K}.\nabla\bar{f}^{K}_{j}=\begin{cases}\nabla f_{j},&\text{a.e. on }\{f_{j}(x)<K\},\\ 0,&\text{a.e. on }\{f_{j}(x)\geq K\}.\end{cases}

Hence

(119) f¯jL2(𝕊2)2={fj<K}|fj|2={fi<K}|fj|2|lnfj|2K2{fj<K}|lnfj|2K2lnfj2K2vol(𝕊2),\begin{split}\|\nabla\bar{f}_{j}\|^{2}_{L^{2}({\mathbb{S}}^{2})}&=\int_{\{f_{j}<K\}}|\nabla f_{j}|^{2}\\ &=\int_{\{f_{i}<K\}}|f_{j}|^{2}|\nabla\ln f_{j}|^{2}\\ &\leq K^{2}\int_{\{f_{j}<K\}}|\nabla\ln f_{j}|^{2}\\ &\leq K^{2}\|\nabla\ln f_{j}\|^{2}\\ &\leq K^{2}{\rm vol}({\mathbb{S}}^{2}),\end{split}

where the last step follows from Lemma 3.1. Combine inequalities (117) and (119) then we get the desired results. ∎

Now we prove the following proposition concerning the limit function:

Lemma 4.5.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive smooth functions on 𝕊2{\mathbb{S}}^{2} satisfying

(120) Δfjfj,j.\Delta f_{j}\leq f_{j},\quad\forall j\in\mathbb{N}.

Let K>0K>0 be a real number that satisfies the requirement in Lemma 4.3. Let f¯KW1,2(𝕊2)\bar{f}^{K}_{\infty}\in W^{1,2}({\mathbb{S}}^{2}) be the limit function as in Lemma 4.4. Then f¯K\bar{f}^{K}_{\infty} satisfies the inequality

(121) 𝕊2u,f¯K𝕊2uf¯K,-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}_{\infty}^{K}\rangle\leq\int_{{\mathbb{S}}^{2}}u\bar{f}_{\infty}^{K},

for all uW1,2(𝕊2)u\in W^{1,2}({\mathbb{S}}^{2}) such that u0u\geq 0.

Proof.

By Lemma 4.4 we know that f¯jK\bar{f}^{K}_{j} converges to f¯K\bar{f}^{K}_{\infty} in L2(𝕊2)L^{2}({\mathbb{S}}^{2}), and that f¯jK\bar{f}^{K}_{j} converges to f¯K\bar{f}^{K}_{\infty} weakly in W1,2(𝕊2)W^{1,2}({\mathbb{S}}^{2}). As a result, for any uW1,2(𝕊2)u\in W^{1,2}({\mathbb{S}}^{2}) we have that

(122) 𝕊2uf¯jK𝕊2uf¯K, as j,\int_{{\mathbb{S}}^{2}}u\bar{f}_{j}^{K}\to\int_{{\mathbb{S}}^{2}}u\bar{f}_{\infty}^{K},\ \ \text{ as }\ \ j\to\infty,

and that

(123) 𝕊2u,f¯jK𝕊2u,f¯K, as j.\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}_{j}^{K}\rangle\to\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}_{\infty}^{K}\rangle,\ \ \text{ as }j\to\infty.

As a result, by (114) we have for all uW1,2(𝕊2)u\in W^{1,2}({\mathbb{S}}^{2}) such that u0u\geq 0

(124) 𝕊2u,f¯K𝕊2uf¯K.-\int_{{\mathbb{S}}^{2}}\langle\nabla u,\nabla\bar{f}_{\infty}^{K}\rangle\leq\int_{{\mathbb{S}}^{2}}u\bar{f}_{\infty}^{K}.

Hence by Theorem 8.19 in [6], we have that either the essential infimum of f¯K\bar{f}^{K}_{\infty} is bounded away from zero or f¯K\bar{f}^{K}_{\infty} is the zero function. This finishes the proof. ∎

We need the definition of essential infimum of a function:

Definition 4.6.

Consider the standard 𝕊2{\mathbb{S}}^{2} and use mm to denote the standard volume measure in 𝕊2{\mathbb{S}}^{2}. Let UU be an open subset of 𝕊2{\mathbb{S}}^{2} . Let f:Uf:U\to\mathbb{R} be measurable. Define the set

(125) Ufess={a:m(f1(,a))=0}.U_{f}^{ess}=\{a\in\mathbb{R}:m(f^{-1}(-\infty,a))=0\}.

We use infUf\inf_{U}f to denote the essential infimum of ff in UU and define

(126) infUf=supUfess\inf_{U}f=\sup U^{ess}_{f}

Finally, we apply the maximum principle for weak solution to prove the following property for the essential infimum of ff_{\infty}.

Proposition 4.7.

Let {fj}j=1\{f_{j}\}_{j=1}^{\infty} be a sequence of positive smooth functions on 𝕊2{\mathbb{S}}^{2} satisfying

(127) Δfjfj,j.\Delta f_{j}\leq f_{j},\quad\forall j\in\mathbb{N}.

If we further assume that fjff_{j}\to f_{\infty} in L2(𝕊2)L^{2}({\mathbb{S}}^{2}) for some ff_{\infty}, then either the essential infimum of ff_{\infty} is bounded away from zero or f=0f_{\infty}=0 a.e. on 𝕊2{\mathbb{S}}^{2}.

Proof.

Since fjfL2(𝕊2)0\|f_{j}-f_{\infty}\|_{L^{2}({\mathbb{S}}^{2})}\to 0 as jj\to\infty, choose a subsequence if needed, then we have fjff_{j}\to f_{\infty} poiintwise almost everywhere in 𝕊2{\mathbb{S}}^{2}. Let K>0K>0 be a real number that satisfies the requirement in Lemma 4.3. Construct a truncated sequence {f¯jK}j=1\{\bar{f}^{K}_{j}\}_{j=1}^{\infty} as in Definition 4.1. By Lemma 4.4, choose a subsequence if needed, there exists f¯KW1,2(𝕊2)\bar{f}^{K}_{\infty}\in W^{1,2}({\mathbb{S}}^{2}) such that f¯jK\bar{f}^{K}_{j} converges to f¯K\bar{f}^{K}_{\infty} in L2(𝕊2)L^{2}({\mathbb{S}}^{2}) norm. As a result, choose a subsequence if needed we have f¯jKf¯K\bar{f}_{j}^{K}\to\bar{f}_{\infty}^{K} pointwise almost everywhere in 𝕊2{\mathbb{S}}^{2}.

It suffices to show that if the essential infimum inf𝕊2f=0\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}=0 then f¯K=f=0\bar{f}^{K}_{\infty}=f_{\infty}=0 in 𝕊2{\mathbb{S}}^{2}. We assume that inf𝕊2f=0\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}=0. Since for each jj we have 0<f¯jKfj0<\bar{f}_{j}^{K}\leq f_{j}, we have 0inf𝕊2f¯jKinf𝕊2f=00\leq\inf\limits_{{\mathbb{S}}^{2}}\bar{f}_{j}^{K}\leq\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}=0. This implies that for any δ,δ>0\delta,\delta^{\prime}>0, we have

(128) m((f¯K)1(,δ))>0,m\left(\left(\bar{f}^{K}_{\infty}\right)^{-1}(-\infty,\delta)\right)>0,

and

(129) m((f¯K)1(,δ))=0.m\left(\left(\bar{f}^{K}_{\infty}\right)^{-1}(-\infty,-\delta^{\prime})\right)=0.

Let NN be the north pole of 𝕊2\mathbb{S}^{2}, and SS be the south pole. Bπ2(N)B_{\frac{\pi}{2}}(N) and Bπ2(S)B_{\frac{\pi}{2}}(S) are upper and lower hemispheres respectively. Then either

(130) infBπ2(N)f¯K=0,\inf\limits_{B_{\frac{\pi}{2}}(N)}\bar{f}^{K}_{\infty}=0,

or

(131) infBπ2(S)f¯K=0.\inf\limits_{B_{\frac{\pi}{2}}(S)}\bar{f}^{K}_{\infty}=0.

Without loss of generality we assume that infBπ2(N)f¯K=0\inf\limits_{B_{\frac{\pi}{2}}(N)}\bar{f}^{K}_{\infty}=0. Since f¯K0\bar{f}^{K}_{\infty}\geq 0 in 𝕊2{\mathbb{S}}^{2}, for any r>π2r>\frac{\pi}{2}, and ϵ>0\epsilon>0 such that r+ϵ<πr+\epsilon<\pi we have

(132) infBr(N)f¯K=infBr+ϵ(N)f¯K=0.\inf_{B_{r}(N)}\bar{f}^{K}_{\infty}=\inf_{B_{r+\epsilon}(N)}\bar{f}^{K}_{\infty}=0.

Now by Lemma 4.5, f¯K\bar{f}^{K}_{\infty} satisfies

(133) (Δ1)f¯K0,(\Delta-1)\bar{f}^{K}_{\infty}\leq 0,

on Br+ϵ(N)B_{r+\epsilon}(N) 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 f¯K\bar{f}^{K}_{\infty} is constant on Br(N)B_{r}(N). This is true for any r>π2r>\frac{\pi}{2}, thus f¯K0\bar{f}^{K}_{\infty}\equiv 0 on 𝕊2\mathbb{S}^{2}. Moreover, since K>0K>0, for almost every x𝕊2x\in{\mathbb{S}}^{2} we have,

(134) limjf¯jK=limjfj=0,\lim_{j\to\infty}\bar{f}_{j}^{K}=\lim_{j\to\infty}f_{j}=0,

and hence f=0f_{\infty}=0 a.e. on 𝕊2{\mathbb{S}}^{2}. This finishes the proof. ∎

4.2. A 11-sweepout of the warped product manifold 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}

Because we will apply the Min-Max minimal surface theory to get an upper bound for MinA\operatorname{MinA} in §4.3, in this subsection we briefly recall some basic notions in geometric measure theory following Marques and Neves [13], and construct a 11-sweepout for 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}, 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 kk-current TT on J\mathbb{R}^{J} is a continuous linear functional on the space of compactly supported smooth kk-forms: 𝒟k(J)\mathcal{D}^{k}(\mathbb{R}^{J}). Its boundary T\partial T is a (k1)(k-1)-current that is defined as T(ϕ):=T(dϕ)\partial T(\phi):=T(d\phi) for ϕ𝒟k1(J)\phi\in\mathcal{D}^{k-1}(\mathbb{R}^{J}). A kk-current TT is said to be an integer multiplicity kk-current if it can be written as

(135) T(ϕ)=Sϕ(x),τ(x)θ(x)𝑑k,ϕ𝒟k(J),T(\phi)=\int_{S}\langle\phi(x),\tau(x)\rangle\theta(x)d\mathcal{H}^{k},\quad\phi\in\mathcal{D}^{k}(\mathbb{R}^{J}),

where SS is a k\mathcal{H}^{k}-measurable countable kk-rectifiable set, that is SS0jSjS\subset S_{0}\cup_{j\in\mathbb{N}}S_{j} with k(S0)=0\mathcal{H}^{k}(S_{0})=0 and SjS_{j} is an embedded kk-dimensional C1C^{1}-submanifold for all jj\in\mathbb{N}, θ\theta is a k\mathcal{H}^{k}-integrable \mathbb{N}-valued function, and τ\tau is a kk-form such that τ(x)\tau(x) is a volume form for TxST_{x}S at xx where a kk-dimensional tangent space TxST_{x}S is well-defined. Note that this tangent space TxST_{x}S is well-defined for k\mathcal{H}^{k}-a.e. xSx\in S, provided k(SK)<+\mathcal{H}^{k}(S\cap K)<+\infty for every compact set KJK\subset\mathbb{R}^{J}. Also note that the form τ\tau give an orientation for TxST_{x}S. The mass of an integer multiplicity kk-current TT is defined as

(136) 𝐌(T):=sup{T(ϕ)ϕ𝒟k(J),|ϕ|1},{\bf M}(T):=\sup\{T(\phi)\mid\phi\in\mathcal{D}^{k}(\mathbb{R}^{J}),\ \ |\phi|\leq 1\},

where |ϕ||\phi| is the pointwise maximal norm of a form ϕ\phi.

In particular, a kk-dimensional embedded smooth submanifold of J\mathbb{R}^{J} can be viewed as an integer multiplicity kk-current by integrating a kk-form over it. Its current boundary is given by its usual boundary, and its mass is the kk-dimensional volume of the submanifold.

Let MM be a manifold embedded in J\mathbb{R}^{J}. The space of integral kk-currents on MM, denoted by 𝐈k(M){\bf I}_{k}(M), is defined to be the space of kk-current such that both TT and T\partial T are integer multiplicity currents with finite mass and support contained in MM. The space of kk-cycles, denoted by 𝒵k(M)\mathcal{Z}_{k}(M), is defined to be the space of those T𝐈k(M)T\in{\bf I}_{k}(M) so that T=QT=\partial Q for some Q𝐈k+1(M)Q\in{\bf I}_{k+1}(M).

A rectifiable kk-varifold V is defined to be a certain Radon measure on J×Gk(J)\mathbb{R}^{J}\times G_{k}(\mathbb{R}^{J}), where Gk(J)G_{k}(\mathbb{R}^{J}) is the Grassmannian of kk-planes in J\mathbb{R}^{J}. An integral kk-current T𝐈k(M)T\in{\bf I}_{k}(M) given as in (135)(\ref{eqn-integer-current}) naturally associates a rectifiable kk-varifold, denoted by |T||T|, as

(137) |T|(A)=Sπ(TSA)θ(x)𝑑k.|T|(A)=\int_{S\cap\pi(TS\cap A)}\theta(x)d\mathcal{H}^{k}.

Here π\pi is the natural projection map from J×Gk(J)\mathbb{R}^{J}\times G_{k}(\mathbb{R}^{J}) to J\mathbb{R}^{J}, and TSTS is rank-kk tangent bundle of SS consisting of TxST_{x}S at xSx\in S where its kk-dimensional tangent plane can be well defined. Note that: in the varifold expression (137)(\ref{eqn-varifold}) of |T||T|, we forget the orientation of SS determined by the kk-form τ\tau in the current expression (135)(\ref{eqn-integer-current}) of TT.

The space 𝐈k(M){\bf I}_{k}(M) can be endowed with various metrics and have different induced topologies. Given T,S𝐈k(M)T,S\in{\bf I}_{k}(M), the flat metric is defined by

(T,S):=inf{𝐌(Q)+𝐌(R)TS=Q+R,Q𝐈k(M),R𝐈k+1(M)}\mathcal{F}(T,S):=\inf\left\{{\bf M}(Q)+{\bf M}(R)\mid T-S=Q+\partial R,\ \ Q\in{\bf I}_{k}(M),\ \ R\in{\bf I}_{k+1}(M)\right\}

and induces the flat topology on 𝐈k(M){\bf I}_{k}(M). We also denote (T):=(T,0)\mathcal{F}(T):=\mathcal{F}(T,0) and have

(138) (T)𝐌(T),T𝐈k(M).\mathcal{F}(T)\leq{\bf M}(T),\quad\forall T\in{\bf I}_{k}(M).

For T,S𝐈k(M)T,S\in{\bf I}_{k}(M), the F-metric is defined by Pitts in [16] as:

(139) 𝐅(S,T):=(ST)+𝐅(|S|,|T|),{\bf F}(S,T):=\mathcal{F}(S-T)+{\bf F}(|S|,|T|),

where 𝐅(|S|,|T|){\bf F}(|S|,|T|) is the F-metric on the associated varifolds defined on page 66 in [16] as:

𝐅(|S|,|T|):=sup{|S|(f)|T|(f)fCc(Gk(J)),|f|1,Lip(f)1}.{\bf F}(|S|,|T|):=\sup\left\{|S|(f)-|T|(f)\mid f\in C_{c}(G_{k}(\mathbb{R}^{J})),\ \ |f|\leq 1,\ \ {\rm Lip}(f)\leq 1\right\}.

Recall that (see page 66 in [16])

(140) 𝐅(|S|,|T|)𝐌(ST),{\bf F}(|S|,|T|)\leq{\bf M}(S-T),

and hence

(141) 𝐅(S,T)2𝐌(ST),S,T𝐈k(M).{\bf F}(S,T)\leq 2{\bf M}(S-T),\quad\forall S,T\in{\bf I}_{k}(M).

For the Min-Max theory for minimal surfaces, the space of mod 22 integral kk-currents and mod 22 kk-cycles are also needed. They are denoted by 𝐈k(M;2){\bf I}_{k}(M;\mathbb{Z}_{2}) and 𝒵k(M;2)\mathcal{Z}_{k}(M;\mathbb{Z}_{2}), respectively, and defined by an equivalence relation: TST\equiv S if TS=2QT-S=2Q for T,S,Q𝐈k(M)T,S,Q\in{\bf I}_{k}(M). The notions of boundary, mass and metrics defined above for 𝐈k(M){\bf I}_{k}(M) can be extended to 𝐈k(M;2){\bf I}_{k}(M;\mathbb{Z}_{2}). For a nn-dimensional manifold MM, the Constancy Theorem (Theorem 26.27 in [17]) says that if T𝐈n(M;2)T\in{\bf I}_{n}(M;\mathbb{Z}_{2}) has T=0\partial T=0, then either T=MT=M or T=0T=0.

Then we recall some basic facts about the topology of 𝒵k(M;;2)\mathcal{Z}_{k}(M;\mathcal{F};\mathbb{Z}_{2}), that is 𝒵k(M;2)\mathcal{Z}_{k}(M;\mathbb{Z}_{2}) endowed with flat metric. Their proofs can be found in [13], also see [1]. Let nn be the dimension of the manifold MM. Then 𝐈n(M;;2){\bf I}_{n}(M;\mathcal{F};\mathbb{Z}_{2}) is contractible and the continuous map

(142) :𝐈n(M;;2)𝒵n1(M;;2)\partial:{\bf I}_{n}(M;\mathcal{F};\mathbb{Z}_{2})\rightarrow\mathcal{Z}_{n-1}(M;\mathcal{F};\mathbb{Z}_{2})

is a 22-fold covering map. The homotopy groups are:

(143) πk(𝒵n1(M;;2),0)={0,whenk2,2,whenk=1.\pi_{k}\left(\mathcal{Z}_{n-1}(M;\mathcal{F};\mathbb{Z}_{2}),0\right)=\begin{cases}0,&\text{when}\ \ k\geq 2,\\ \mathbb{Z}_{2},&\text{when}\ \ k=1.\end{cases}

For the calculation of the fundamental group, one notes that the map

(144) P:π1(𝒵n1(M;;2),0)\displaystyle P:\pi_{1}\left(\mathcal{Z}_{n-1}(M;\mathcal{F};\mathbb{Z}_{2}),0\right) \displaystyle\rightarrow {0,M}\displaystyle\{0,M\}
(145) [γ]\displaystyle\left[\gamma\right] \displaystyle\mapsto γ~(1)\displaystyle\tilde{\gamma}(1)

is an isomorphism. Here γ\gamma is a loop in 𝒵n1(M;;2)\mathcal{Z}_{n-1}(M;\mathcal{F};\mathbb{Z}_{2}) with γ(0)=γ(1)=0\gamma(0)=\gamma(1)=0, and γ~\tilde{\gamma} is the unique lift to 𝐈n(M;;2){\bf I}_{n}(M;\mathcal{F};\mathbb{Z}_{2}) with γ~(0)=0\tilde{\gamma}(0)=0. Then by applying Hurewicz Theorem, one can obtain:

(146) H1(𝒵n1(M;;2);2)=2={0,λ¯}.H^{1}\left(\mathcal{Z}_{n-1}(M;\mathcal{F};\mathbb{Z}_{2});\mathbb{Z}_{2}\right)=\mathbb{Z}_{2}=\{0,\bar{\lambda}\}.

The the action of the fundamental cohomology class λ¯\bar{\lambda} 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 11-sweepout from [13].

Definition 4.8.

A continuous map Φ:𝕊1𝒵n1(M;𝐅;2)\Phi:{\mathbb{S}}^{1}\rightarrow\mathcal{Z}_{n-1}(M;{\bf F};\mathbb{Z}_{2}) is called a 11-sweepout if Φ(λ¯)0H1(𝕊1,2)\Phi^{*}(\bar{\lambda})\neq 0\in H^{1}({\mathbb{S}}^{1},\mathbb{Z}_{2}).

Here 𝒵n1(M;𝐅;2)\mathcal{Z}_{n-1}(M;{\bf F};\mathbb{Z}_{2}) is the space 𝒵n1(M;2)\mathcal{Z}_{n-1}(M;\mathbb{Z}_{2}) endowed with the F-metric given in (139)(\ref{eqn-F-metric}).

Now we return back our warped product manifold 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}, that is 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} with Riemannian metric

(147) g=g𝕊2+f2g𝕊1.g=g_{{\mathbb{S}}^{2}}+f^{2}g_{{\mathbb{S}}^{1}}.

For each fixed x𝕊2x\in{\mathbb{S}}^{2}, we construct a 11-sweepout of 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} consisting of tori {Σx,r:=Br(x)×𝕊10rπ}\{\Sigma_{x,r}:=\partial B_{r}(x)\times{\mathbb{S}}^{1}\mid 0\leq r\leq\pi\} , where Br(x)B_{r}(x) denotes the geodesic ball on 𝕊2{\mathbb{S}}^{2} centered at xx with radius rr. In other words, we consider the map

(148) Φ:[0,π]\displaystyle\Phi:[0,\pi] 𝒵2(𝕊2×f𝕊1;𝐅;2),\displaystyle\rightarrow{\mathcal{Z}_{2}(\mathbb{S}^{2}\times_{f}\mathbb{S}^{1};{\bf F};\mathbb{Z}_{2}}),
r\displaystyle r (Br(x)×𝕊1)=Br(x)×𝕊1.\displaystyle\mapsto\partial\left(B_{r}(x)\times{\mathbb{S}}^{1}\right)=\partial B_{r}(x)\times{\mathbb{S}}^{1}.
Lemma 4.9.

The map Φ\Phi given in (148)(\ref{eqn-tori-sweepout}) provides a 11-sweepout of 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} as in Definition 4.8.

Proof.

Clearly, Φ(0)=Φ(π)=0\Phi(0)=\Phi(\pi)=0, and hence Φ\Phi can be viewed as a map from 𝕊1{\mathbb{S}}^{1} to 𝒵2(𝕊2×f𝕊1;𝐅;2)\mathcal{Z}_{2}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1};{\bf F};\mathbb{Z}_{2}) by identifying the end points of the interval [0,π][0,\pi].

Now we show the continuity of the map Φ\Phi on [0,π][0,\pi]. This is clear for r(0,π)r\in(0,\pi), since Br(x)\partial B_{r}(x) varies smoothly for r(0,π)r\in(0,\pi). Then the continuity at t=0t=0 follows from the inequality in (141)(\ref{eqn-F-metric-less-than-mass}) and the estimate:

(149) 𝐌(Φ(r)Φ(0))=𝐌(Φ(r))=𝐌(Br(x)×𝕊1)=f4π2sinr0,{\mathbf{M}}(\Phi(r)-\Phi(0))={\mathbf{M}}(\Phi(r))={\mathbf{M}}\left(\partial B_{r}(x)\times{\mathbb{S}}^{1}\right)=f\cdot 4\pi^{2}\sin r\rightarrow 0,

as r0r\rightarrow 0, since the warping function ff is smooth on 𝕊2{\mathbb{S}}^{2}. The continuity at t=πt=\pi follows similarly, since sinr0\sin r\rightarrow 0 as rπr\rightarrow\pi.

Because by the definition flat metric is less than or equal to F-metric, Φ\Phi is also continuous if we endow the flat metric on 𝒵2(M;2)\mathcal{Z}_{2}(M;\mathbb{Z}_{2}). So Φ\Phi is a loop in 𝒵2(𝕊2×f𝕊1;;2)\mathcal{Z}_{2}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1};\mathcal{F};\mathbb{Z}_{2}), and represents a non-trivial element:

(150) [Φ]0π1(𝒵2(𝕊2×f𝕊1;;2)).\left[\Phi\right]\neq 0\in\pi_{1}\left(\mathcal{Z}_{2}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1};\mathcal{F};\mathbb{Z}_{2})\right).

This is because by the definition of the map Φ\Phi we have that the unique lift Φ~\tilde{\Phi} of Φ\Phi with Φ~(0)=0\tilde{\Phi}(0)=0 is given by

(151) Φ~:[0,π]\displaystyle\tilde{\Phi}:[0,\pi] 𝒵3(𝕊2×f𝕊1;;2),\displaystyle\rightarrow{\mathcal{Z}_{3}(\mathbb{S}^{2}\times_{f}\mathbb{S}^{1};\mathcal{F};\mathbb{Z}_{2}}),
r\displaystyle r Br(x)×𝕊1,\displaystyle\mapsto B_{r}(x)\times{\mathbb{S}}^{1},

and has Φ~(π)=𝕊2×𝕊1\tilde{\Phi}(\pi)={\mathbb{S}}^{2}\times{\mathbb{S}}^{1}. Consequently, Φ(λ¯)0\Phi^{*}(\bar{\lambda})\neq 0, and so Φ\Phi is a 11-sweepout. ∎

4.3. Bound MinA\operatorname{MinA} from above by L1L^{1}-norm of warping function

In this subsection, we derive an upper bound for MinA(𝕊2×f𝕊1)\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}) in terms of fL1(𝕊2)\|f\|_{L^{1}({\mathbb{S}}^{2})}, provided that fL2(𝕊2)\|f\|_{L^{2}({\mathbb{S}}^{2})} is small relative to MinA(𝕊2×f𝕊1)\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}).

Proposition 4.10.

Let 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} be a warped product Riemannian manifolds with metric tensor as in (3) that has nonnegative scalar curvature and MinA(𝕊2×f𝕊1)A>0\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1})\geq A>0. If fL2(𝕊2)<A232π52\|f\|_{L^{2}({\mathbb{S}}^{2})}<\frac{A}{2^{\frac{3}{2}}\pi^{\frac{5}{2}}}, then we have fL1(𝕊1)A100π\|f\|_{L^{1}({\mathbb{S}}^{1})}\geq\frac{A}{100\pi}.

Recall that MinA(𝕊2×f𝕊1)\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}) is the infimum of areas of closed embedded minimal surfaces in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}. 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 MinA(𝕊2×f𝕊1)\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}) from above by areas of some tori in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}.

Lemma 4.11.

Let 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} be a warped product Riemannian manifold with metric tensor as in (3). For each x𝕊2x\in{\mathbb{S}}^{2}, there exists a torus Σx,rx=Brx(x)×𝕊1𝕊2×f𝕊1\Sigma_{x,r_{x}}=\partial B_{r_{x}}(x)\times\mathbb{S}^{1}\subset\mathbb{S}^{2}\times_{f}\mathbb{S}^{1}, 0<rx<π0<r_{x}<\pi, whose area is not less than MinA(𝕊2×f𝕊1)\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}), i.e.

(152) Area(Σx,rx)MinA(𝕊2×f𝕊1),{\rm Area}(\Sigma_{x,r_{x}})\geq\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}),

where Brx(x)B_{r_{x}}(x) is the geodesic ball in the standard 𝕊2{\mathbb{S}}^{2} centered at xx with radius rxr_{x}.

Proof.

We will use Min-Max minimal surface theory of Marques and Neves to prove the lemma.

For each fixed point x𝕊2x\in{\mathbb{S}}^{2}, by Lemma 4.9, the map Φ\Phi in (148)(\ref{eqn-tori-sweepout}) gives a 11-sweepout of 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} as in Definition 4.8. For r[0,π]r\in[0,\pi], the image Φ(r)=Br(x)×𝕊1=:Σx,r\Phi(r)=\partial B_{r}(x)\times{\mathbb{S}}^{1}=:\Sigma_{x,r} are tori in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} with mass:

(153) 𝐌(Φ(r))=Area(Σx,r)=2πBr(x)f𝑑s.{\mathbf{M}}(\Phi(r))={\rm Area}(\Sigma_{x,r})=2\pi\int_{\partial B_{r}(x)}fds.

Clearly, 𝐌(Φ(r)){\mathbf{M}}(\Phi(r)) is a continuous function of rr on [0,π][0,\pi] with 𝐌(Φ(0))=𝐌(Φ(π))=0{\mathbf{M}}(\Phi(0))={\mathbf{M}}(\Phi(\pi))=0. Thus there exist rx(0,π)r_{x}\in(0,\pi) such that

(154) 𝐌(Φ(rx))=max{𝐌(Φ(r))0rπ}.{\mathbf{M}}(\Phi(r_{x}))=\max\{{\mathbf{M}}(\Phi(r))\mid 0\leq r\leq\pi\}.

Let Π\Pi be the homotopy class of the 11-sweepout Φ\Phi, which consists of all continuous maps Φ:[0,π]𝒵2(𝕊2×f𝕊1;𝐅;2)\Phi^{\prime}:[0,\pi]\rightarrow\mathcal{Z}_{2}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1};{\bf F};\mathbb{Z}_{2}) with Φ(0)=Φ(π)\Phi^{\prime}(0)=\Phi^{\prime}(\pi) such that Φ\Phi and Φ\Phi^{\prime} are homotopic to each other in the flat topology. By Lemma 2.2.6 in [13], the width

(155) 𝐋(Π)=infΦΠsupr[0,π]{𝐌(Φ(r))}>0,{\bf L}(\Pi)=\inf_{\Phi^{\prime}\in\Pi}\sup_{r\in[0,\pi]}\{{\bf M}(\Phi^{\prime}(r))\}>0,

since Φ\Phi is a 11-sweepout and so Π\Pi 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 Σ\Sigma in 𝕊2×f𝕊1\mathbb{S}^{2}\times_{f}\mathbb{S}^{1} achieving the width, i.e. Area(Σ)=𝐋(Π)>0{\rm Area}(\Sigma)={\bf L}(\Pi)>0.

Finally, by the definitions of the width in (155) and MinA\operatorname{MinA}, and by the choice of Σx,rx\Sigma_{x,r_{x}}, we have

(156) Area(Σx,rx)𝐋(Π)=Area(Σ)MinA(𝕊2×𝕊1).{\rm Area}(\Sigma_{x,r_{x}})\geq{\bf L}(\Pi)={\rm Area}(\Sigma)\geq\operatorname{MinA}(\mathbb{S}^{2}\times\mathbb{S}^{1}).

Because xx is an arbitrary point on 𝕊2{\mathbb{S}}^{2}, 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 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} be a warped product Riemannian manifold with metric tensors as in (3) that have non-negative scalar curvatures and MinA(𝕊2×f𝕊1)A>0\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1})\geq A>0. If fL2(𝕊2)<A232π52\|f\|_{L^{2}({\mathbb{S}}^{2})}<\frac{A}{2^{\frac{3}{2}}\pi^{\frac{5}{2}}}, then there exists a set 𝕊2\mathcal{H}\subset{\mathbb{S}}^{2} satisfying that for each xx\in\mathcal{H} there exists 0<rxπ20<r_{x}\leq\frac{\pi}{2} such that

  1. (i)

    Area(xBrx10(x))12Area(𝕊2){\rm Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)\geq\frac{1}{2}{\rm Area}({\mathbb{S}}^{2}),

  2. (ii)

    and

    (157) Br(x)f𝑑sA2(2π)2\fint_{\partial B_{r}(x)}fds\geq\frac{A}{2(2\pi)^{2}}

    holds for all r[0,rx]r\in[0,r_{x}].

Proof.

For any point x𝕊2x\in{\mathbb{S}}^{2}, we denote its antipodal point by x¯\bar{x}. By Lemma 4.11, for any x𝕊2x\in{\mathbb{S}}^{2}, there exists 0<rx<π0<r_{x}<\pi such that the torus Σx,rx=Brx(x)×𝕊1\Sigma_{x,r_{x}}=\partial B_{r_{x}}(x)\times{\mathbb{S}}^{1} in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} has area

(158) Area(Σx,rx)MinA(𝕊2×f𝕊1)A.{\rm Area}(\Sigma_{x,r_{x}})\geq\operatorname{MinA}({\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1})\geq A.

Since Area(Σx,rx)=2πBrx(x)f𝑑s{\rm Area}(\Sigma_{x,r_{x}})=2\pi\int_{\partial B_{r_{x}}(x)}fds, we have

(159) 2πBrx(x)f𝑑sA.2\pi\int_{\partial B_{r_{x}}(x)}fds\geq A.

Thus, we have

(160) Brx(x)f𝑑sA2π.\int_{\partial B_{r_{x}}(x)}fds\geq\frac{A}{2\pi}.

Now if 0<rxπ20<r_{x}\leq\frac{\pi}{2}, then we include the point xx in the set \mathcal{H}, and if rx>π2r_{x}>\frac{\pi}{2}, then we include its antipodal point x¯\bar{x} in the set \mathcal{H}, and we set rx¯=πrx<π2r_{\bar{x}}=\pi-r_{x}<\frac{\pi}{2}. Then we still have

(161) Brx¯(x¯)f𝑑s=Brx(x)f𝑑sA2π,\int_{\partial B_{r_{\bar{x}}}(\bar{x})}fds=\int_{\partial B_{r_{x}}(x)}fds\geq\frac{A}{2\pi},

since Brx¯(x¯)=Brx(x)\partial B_{r_{\bar{x}}}(\bar{x})=\partial B_{r_{x}}(x).

By the construction of the set 𝕊2\mathcal{H}\subset{\mathbb{S}}^{2}, \mathcal{H} contains at least one of any pair of antipodal points on 𝕊2{\mathbb{S}}^{2}, and for any xx\in\mathcal{H}, there exists 0<rxπ20<r_{x}\leq\frac{\pi}{2} such that

(162) Brx(x)f𝑑sA2π.\int_{\partial B_{r_{x}}(x)}fds\geq\frac{A}{2\pi}.

Then we have that the area of the open set xBrx10(x)\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x) is at least half of the area of the whole sphere 𝕊2{\mathbb{S}}^{2}, i.e.

(163) Area(xBrx10(x))12Area(𝕊2).\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)\geq\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2}).

Indeed, otherwise, we have

(164) Area(xBrx10(x¯))=Area(xBrx10(x))<12Area(𝕊2).\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(\bar{x})\right)=\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)<\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2}).

On the other hand, because for each x𝕊2x\in{\mathbb{S}}^{2} either xx or x¯\bar{x} is contained in \mathcal{H}, we have

(165) 𝕊2=(xBrx10(x))(xBrx10(x¯)).{\mathbb{S}}^{2}=\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)\cup\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(\bar{x})\right).

So

(166) Area(𝕊2)\displaystyle\operatorname{Area}({\mathbb{S}}^{2}) =\displaystyle= Area((xBrx10(x))(xBrx10(x¯)))\displaystyle\operatorname{Area}\left(\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)\cup\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(\bar{x})\right)\right)
(167) \displaystyle\leq Area(xBrx10(x))+Area(xBrx10(x¯))\displaystyle\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)+\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(\bar{x})\right)
(168) <\displaystyle< 12Area(𝕊2)+12Area(𝕊2)=Area(𝕊2).\displaystyle\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2})+\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2})=\operatorname{Area}({\mathbb{S}}^{2}).

This gives a contradiction. So we have Area(xBrx10(x))12Area(𝕊2)\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)\geq\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2}).

Because 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1} has non-negative scalar curvature, by Lemma 2.1, we have Δff\Delta f\leq f. Then by the spherical mean inequality in Proposition 2.4, for any x𝕊2x\in\mathcal{H}\subset{\mathbb{S}}^{2} and any 0rrx(π2)0\leq r\leq r_{x}(\leq\frac{\pi}{2}) we have that

(169) Brx(x)f𝑑sBr(x)f𝑑sfL2(𝕊2)2π(rxr)A2(2π)2,\fint_{\partial B_{r_{x}}(x)}fds-\fint_{\partial B_{r}(x)}fds\leq\frac{\|f\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}(r_{x}-r)\leq\frac{A}{2(2\pi)^{2}},

since fL2(𝕊2)A232π52\|f\|_{L^{2}({\mathbb{S}}^{2})}\leq\frac{A}{2^{\frac{3}{2}}\pi^{\frac{5}{2}}} and rxrπ2r_{x}-r\leq\frac{\pi}{2}. By rearrange the inequality, we obtain that for any xx\in\mathcal{H} and any 0rrx0\leq r\leq r_{x},

(170) Br(x)f𝑑s\displaystyle\fint_{\partial B_{r}(x)}fds \displaystyle\geq Brx(x)f𝑑sA2(2π)2\displaystyle\fint_{\partial B_{r_{x}}(x)}fds-\frac{A}{2(2\pi)^{2}}
(171) =\displaystyle= 12πsinrxBrx(x)f𝑑sA2(2π)2\displaystyle\frac{1}{2\pi\sin r_{x}}\int_{\partial B_{r_{x}}(x)}fds-\frac{A}{2(2\pi)^{2}}
(172) \displaystyle\geq 12πBrx(x)f𝑑sA2(2π)2\displaystyle\frac{1}{2\pi}\int_{\partial B_{r_{x}}(x)}fds-\frac{A}{2(2\pi)^{2}}
(173) \displaystyle\geq A(2π)2A2(2π)2=A2(2π)2.\displaystyle\frac{A}{(2\pi)^{2}}-\frac{A}{2(2\pi)^{2}}=\frac{A}{2(2\pi)^{2}}.

We now apply Lemma 4.12 and Vitali covering theorem to prove Proposition 4.10:

Proof of Proposition 4.10.

By Lemma 4.12, there exists a set 𝕊2\mathcal{H}\subset{\mathbb{S}}^{2} such that

(174) Area(xBrx10(x))12Area(𝕊2),{\rm Area}(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x))\geq\frac{1}{2}{\rm Area}({\mathbb{S}}^{2}),

and for any xx\in\mathcal{H}, there exists rxπ2r_{x}\leq\frac{\pi}{2} such that

(175) Br(x)fA2(2π)2\fint_{\partial B_{r}(x)}f\geq\frac{A}{2(2\pi)^{2}}

holds for all r[0,rx]r\in[0,r_{x}].

By the Vitali covering theorem, there exists a countable sequence of points {xii}\{x_{i}\mid i\in\mathbb{N}\}\subset\mathcal{H} such that the collection of balls {Brxi10(xi)}\{B_{\frac{r_{x_{i}}}{10}}(x_{i})\} are disjoint with each other, and that

(176) xBrx10(x)iBrxi2(xi).\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\subset\underset{i\in\mathbb{N}}{\cup}B_{\frac{r_{x_{i}}}{2}}(x_{i}).

By Lemma 4.12 we have

(177) A8π2Br(xi)f=12πsinrBr(xi)f𝑑s,r[0,rxi].\frac{A}{8\pi^{2}}\leq\fint_{\partial B_{r}(x_{i})}f=\frac{1}{2\pi\sin r}\int_{\partial B_{r}(x_{i})}fds,\quad\forall r\in[0,r_{x_{i}}].

As a result, we have

(178) A4πsinrBr(xi)f𝑑s,r[0,rxi].\frac{A}{4\pi}\sin r\leq\int_{\partial B_{r}(x_{i})}fds,\quad\forall r\in[0,r_{x_{i}}].

Integrating this inequality from 0 to rxi10\frac{r_{x_{i}}}{10} gives

(179) A8π2Area(Brxi10)\displaystyle\frac{A}{8\pi^{2}}\operatorname{Area}(B_{\frac{r_{x_{i}}}{10}}) =\displaystyle= A8π20rxi102πsinrdr\displaystyle\frac{A}{8\pi^{2}}\int^{\frac{r_{x_{i}}}{10}}_{0}2\pi\sin rdr
(180) \displaystyle\leq 0rxi10(Br(xi)f𝑑s)𝑑r\displaystyle\int^{\frac{r_{x_{i}}}{10}}_{0}\left(\int_{\partial B_{r}(x_{i})}fds\right)dr
(181) =\displaystyle= Brxi10(xi)fvol𝕊2.\displaystyle\int_{B_{\frac{r_{x_{i}}}{10}}(x_{i})}f{\rm vol}_{{\mathbb{S}}^{2}}.

Then by summing the above inequalities for ii\in\mathbb{N} together, we obtain

(182) A8π2i=1+Area(Brxi10)i=1+Brxi10(xi)fvol𝕊2fL1(𝕊2),\frac{A}{8\pi^{2}}\sum^{+\infty}_{i=1}\operatorname{Area}(B_{\frac{r_{x_{i}}}{10}})\leq\sum^{+\infty}_{i=1}\int_{B_{\frac{r_{x_{i}}}{10}}(x_{i})}f{\rm vol}_{{\mathbb{S}}^{2}}\leq\|f\|_{L^{1}({\mathbb{S}}^{2})},

since {Brxi10(xi)i}\{B_{\frac{r_{x_{i}}}{10}}(x_{i})\mid i\in\mathbb{N}\} are disjoint balls. In the standard 𝕊2{\mathbb{S}}^{2} we have

(183) Area(Brxi10(xi))125Area(Brxi2(xi)).\operatorname{Area}\left(B_{\frac{r_{x_{i}}}{10}}(x_{i})\right)\geq\frac{1}{25}\operatorname{Area}\left(B_{\frac{r_{x_{i}}}{2}}(x_{i})\right).

As a result, we have

(184) fL1(𝕊2)\displaystyle\|f\|_{L^{1}({\mathbb{S}}^{2})} \displaystyle\geq A8π2i=1+Area(Brxi10)\displaystyle\frac{A}{8\pi^{2}}\sum^{+\infty}_{i=1}\operatorname{Area}\left(B_{\frac{r_{x_{i}}}{10}}\right)
(185) \displaystyle\geq A200π2i=1+Area(Brxi2(xi))\displaystyle\frac{A}{200\pi^{2}}\sum^{+\infty}_{i=1}\operatorname{Area}\left(B_{\frac{r_{x_{i}}}{2}}(x_{i})\right)
(186) \displaystyle\geq A200π2Area(iBrxi2(xi))\displaystyle\frac{A}{200\pi^{2}}\operatorname{Area}\left(\underset{i\in\mathbb{N}}{\cup}B_{\frac{r_{x_{i}}}{2}}(x_{i})\right)
(187) \displaystyle\geq A200π2Area(xBrx10(x))\displaystyle\frac{A}{200\pi^{2}}\operatorname{Area}\left(\underset{x\in\mathcal{H}}{\cup}B_{\frac{r_{x}}{10}}(x)\right)
(188) \displaystyle\geq A200π212Area(𝕊2)=A100π.\displaystyle\frac{A}{200\pi^{2}}\frac{1}{2}\operatorname{Area}({\mathbb{S}}^{2})=\frac{A}{100\pi}.

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 {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product manifolds such that each 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has non-negative scalar curvature. If we assume that

(189) Vol(𝕊2×fj𝕊1)V and MinA(𝕊2×fj𝕊1)A>0,j,{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V\text{ and }\operatorname{MinA}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\geq A>0,\forall j\in\mathbb{N},

then we have the following:

  1. (i)(i)

    After passing to a subsequence if needed, the sequence of warping functions {fj}j=1\{f_{j}\}_{j=1}^{\infty} converges to some limit function ff_{\infty} in Lq(𝕊2)L^{q}({\mathbb{S}}^{2}) for all q[1,)q\in[1,\infty).

  2. (ii)(ii)

    The limit function ff_{\infty} is in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}), for all pp such that 1p<21\leq p<2.

  3. (iii)(iii)

    The essential infimum of ff_{\infty} is strictly positive, i.e. inf𝕊2f>0\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0.

  4. (iv)(iv)

    If we allow ++\infty as a limit, then the limit

    (190) f¯(x):=limr0Br(x)f\overline{f_{\infty}}(x):=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}

    exists for every x𝕊2x\in{\mathbb{S}}^{2}. Moreover, f¯\overline{f_{\infty}} is lower semi-continuous and strictly positive everywhere on 𝕊2{\mathbb{S}}^{2}, and f¯=f\overline{f_{\infty}}=f_{\infty} a.e. on 𝕊2{\mathbb{S}}^{2}.

Proof.

(i)(i) By Lemma 2.1 and Lemma 2.2, the nonnegative scalar curvature condition and Vol(𝕊2×fj𝕊2)V{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{2})\leq V imply that the sequence of warping functions {fj}j=1\{f_{j}\}_{j=1}^{\infty} satisfies the hypothesis in Proposition 3.5. By applying Proposition 3.5, we get the desired convergence.

(ii)(ii) By applying Proposition 3.5 we get that fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}), for all p[1,2)p\in[1,2).

(iii)(iii) We prove inf𝕊2f>0\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0 by contradiction. Recall that inf𝕊2f\inf\limits_{{\mathbb{S}}^{2}}f_{\infty} is the essential infimum of ff_{\infty} as defined in Definition 4.6. First note that f0f_{\infty}\geq 0, since fj>0,jf_{j}>0,\forall j\in\mathbb{N}. Assume that inf𝕊2f=0\inf\limits_{\mathbb{S}^{2}}f_{\infty}=0, then by Proposition 4.7 we have f=0f_{\infty}=0 almost everywhere in 𝕊2{\mathbb{S}}^{2} and hence

(191) fj0inL2(𝕊2),asj+.f_{j}\rightarrow 0\ \ \text{in}\ \ L^{2}({\mathbb{S}}^{2}),\ \ \text{as}\ \ j\rightarrow+\infty.

Therefore, for all sufficiently large jj, we have fjL2(𝕊2)<A232π52\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}<\frac{A}{2^{\frac{3}{2}}\pi^{\frac{5}{2}}}. Then by Proposition 4.10, we have fjL1(𝕊2)A100π>0\|f_{j}\|_{L^{1}({\mathbb{S}}^{2})}\geq\frac{A}{100\pi}>0 for all sufficiently large jj\in\mathbb{N}. This contradicts with that fj0f_{j}\rightarrow 0 in L2(𝕊2)L^{2}({\mathbb{S}}^{2}) as j+j\rightarrow+\infty in (191). This finishes the proof of part (ii)(ii).

(iv)(iv) Because warping functions fif_{i} satisfy the requirements in Proposition 3.7, the existence of the limit

(192) f¯(x):=limr0Br(x)f,\overline{f_{\infty}}(x):=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty},

the lower semi-continuity of f¯\overline{f_{\infty}} and f¯=f\overline{f_{\infty}}=f_{\infty} a.e. on 𝕊2{\mathbb{S}}^{2} directly follow from Proposition 3.7.

Thus we only need to prove that f¯(x)>0\overline{f_{\infty}}(x)>0 for all x𝕊2x\in{\mathbb{S}}^{2}. Let

(193) e:=inf𝕊2f>0.e_{\infty}:=\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0.

By the continuity of the distance funciton d(y,x)d(y,x), there exists 0<r0<π20<r_{0}<\frac{\pi}{2} such that for all x𝕊2x\in{\mathbb{S}}^{2} we have

(194) f(y)Cd(y,x)>e2, for a.e.yBr0(x).f_{\infty}(y)-Cd(y,x)>\frac{e_{\infty}}{2},\ \ \text{ for a.e.}\ \ y\in B_{r_{0}}(x).

As a result, we have

(195) Br0(x)(f(y)Cd(y,x))𝑑vol(y)>e2,x𝕊2.\fint_{B_{r_{0}}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)>\frac{e_{\infty}}{2},\ \ \forall x\in{\mathbb{S}}^{2}.

Then because in Proposition 3.7 we proved that for each fixed x𝕊2x\in{\mathbb{S}}^{2} the ball average Br0(x)(f(y)Cd(y,x))𝑑vol(y)\fint_{B_{r_{0}}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y) is non-increasing in r(0,π2)r\in\left(0,\frac{\pi}{2}\right), and

(196) limr0Br(x)f=limr0Br(x)(f(y)Cd(y,x))𝑑vol(y),\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}=\lim_{r\rightarrow 0}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y),

we have that for each fixed x𝕊2x\in{\mathbb{S}}^{2},

(197) f¯(x)\displaystyle\overline{f_{\infty}}(x) :=\displaystyle:= limr0Br(x)f\displaystyle\lim_{r\rightarrow 0}\fint_{B_{r}(x)}f_{\infty}
(198) =\displaystyle= sup0<r<π2Br(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle\sup_{0<r<\frac{\pi}{2}}\fint_{B_{r}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)
(199) \displaystyle\geq Br0(x)(f(y)Cd(y,x))𝑑vol(y)\displaystyle\fint_{B_{r_{0}}(x)}\left(f_{\infty}(y)-Cd(y,x)\right)d{\rm vol}(y)
(200) >\displaystyle> e2>0.\displaystyle\frac{e_{\infty}}{2}>0.

This completes the proof of theorem. ∎

Remark 4.14.

Theorem 4.13 implies that the limit function ff_{\infty} has a everywhere positive lower semi-continuous representative f¯\overline{f_{\infty}} as a function in W1,p(𝕊2)W^{1,p}({\mathbb{S}}^{2}) for 1p<21\leq p<2. For the rest of paper, fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) 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 fjf_{j} obtained in Proposition 4.15 is important in proving geometric convergences of the sequence of warped product manifolds {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} in our next paper.

Proposition 4.15.

Let {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product manifolds with metric tensors as in (3) that have non-negative scalar curvature and satisfy

(201) Vol(𝕊2×fj𝕊1)V and MinA(𝕊2×fj𝕊1)A>0,j.{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V\text{ and }\operatorname{MinA}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\geq A>0,\forall j\in\mathbb{N}.

Let e:=inf𝕊2f>0e_{\infty}:=\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0. Then there exists j0j_{0}\in\mathbb{N} such that fj(x)e4>0f_{j}(x)\geq\frac{e_{\infty}}{4}>0, for all jj0j\geq j_{0} and all x𝕊2x\in{\mathbb{S}}^{2}.

Proof.

By Lemma 2.1, the non-negativity of scalar curvature of 𝕊2×fi𝕊1{\mathbb{S}}^{2}\times_{f_{i}}{\mathbb{S}}^{1} implies that

(202) Δfjfj,j.\Delta f_{j}\leq f_{j},\quad\forall j\in\mathbb{N}.

Therefore, by the spherical mean inequality in Proposition 2.4, we have

(203) fj(x)Bs(x)fj𝑑sfjL2(𝕊2)2πs,s(0,π2),x𝕊2,j.f_{j}(x)\geq\fint_{\partial B_{s}(x)}f_{j}ds-\frac{\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}s,\quad\forall s\in\left(0,\frac{\pi}{2}\right),x\in{\mathbb{S}}^{2},j\in\mathbb{N}.

Then multiplying the inequality by Area(Bs(x))=2πsin(s)\operatorname{Area}(\partial B_{s}(x))=2\pi\sin(s) gives us

(204) 2πsin(s)fj(x)Bs(x)fj𝑑sfjL2(𝕊2)2π2πsin(s)s,2\pi\sin(s)f_{j}(x)\geq\int_{\partial B_{s}(x)}f_{j}ds-\frac{\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}2\pi\sin(s)s,

for all s(0,π2),x𝕊2s\in\left(0,\frac{\pi}{2}\right),x\in{\mathbb{S}}^{2} and jj\in\mathbb{N}. Let

(205) V(r):=vol(Br(x))=0r2πsinsds=2π(1cosr),V(r):={\rm vol}(B_{r}(x))=\int^{r}_{0}2\pi\sin sds=2\pi(1-\cos r),

and let e:=inf𝕊2fe_{\infty}:=\inf_{{\mathbb{S}}^{2}}f_{\infty} denote the essential infimum of the limit function ff_{\infty} which is strictly positive by Theorem 4.13.

Now integrating the inequality (204) with respect to ss from 0 to r<π2r<\frac{\pi}{2} gives us

(206) V(r)fj(x)\displaystyle V(r)f_{j}(x) \displaystyle\geq Br(x)fj𝑑vol𝕊2fjL2(𝕊2)2π0r2πssinsds\displaystyle\int_{B_{r}(x)}f_{j}d{\rm vol}_{{\mathbb{S}}^{2}}-\frac{\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}\int^{r}_{0}2\pi s\sin sds
(208) \displaystyle\geq Br(x)f𝑑vol𝕊2ffjL1(𝕊2)\displaystyle\int_{B_{r}(x)}f_{\infty}d{\rm vol}_{{\mathbb{S}}^{2}}-\|f_{\infty}-f_{j}\|_{L^{1}({\mathbb{S}}^{2})}
2πfjL2(𝕊2)(sinrrcosr)\displaystyle-\sqrt{2\pi}\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}(\sin r-r\cos r)
(210) \displaystyle\geq eV(r)ffjL1(𝕊2)\displaystyle e_{\infty}V(r)-\|f_{\infty}-f_{j}\|_{L^{1}({\mathbb{S}}^{2})}
2πfjL2(𝕊2)(sinrrcosr).\displaystyle-\sqrt{2\pi}\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}(\sin r-r\cos r).

Then by dividing the inequality by V(r)V(r) we obtain

(211) fj(x)effjL1(𝕊2)V(r)fjL2(𝕊2)2πsinrrcosr1cosr,f_{j}(x)\geq e_{\infty}-\frac{\|f_{\infty}-f_{j}\|_{L^{1}({\mathbb{S}}^{2})}}{V(r)}-\frac{\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}\frac{\sin r-r\cos r}{1-\cos r},

for all 0<r<π2,x𝕊20<r<\frac{\pi}{2},x\in{\mathbb{S}}^{2} and jj\in\mathbb{N}. By Lemma 3.2 we have supjfjL2(𝕊2)<\sup\limits_{j}\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}<\infty, and by direct calculation we have that

(212) limr0sinrrcosr1cosr=0,\lim_{r\rightarrow 0}\frac{\sin r-r\cos r}{1-\cos r}=0,

we can choose 0<r1<π20<r_{1}<\frac{\pi}{2} such that

(213) |fjL2(𝕊2)2πsinr1r1cosr11cosr1|<e2,j.\left|\frac{\|f_{j}\|_{L^{2}({\mathbb{S}}^{2})}}{\sqrt{2\pi}}\frac{\sin r_{1}-r_{1}\cos r_{1}}{1-\cos r_{1}}\right|<\frac{e_{\infty}}{2},\quad\forall j\in\mathbb{N}.

Moreover, because fjff_{j}\rightarrow f_{\infty} in L1(𝕊2)L^{1}({\mathbb{S}}^{2}), we can choose j0j_{0}\in\mathbb{N} such that

(214) ffjL1(𝕊2)V(r1)e4,jj0.\frac{\|f_{\infty}-f_{j}\|_{L^{1}({\mathbb{S}}^{2})}}{V(r_{1})}\leq\frac{e_{\infty}}{4},\quad\forall j\geq j_{0}.

Finally by combining (211), (213) and (214) together, we conclude that fj(x)e4>0f_{j}(x)\geq\frac{e_{\infty}}{4}>0 for all jj0j\geq j_{0} and x𝕊2x\in{\mathbb{S}}^{2}. ∎

4.5. Uniform systole positive lower bound

In this subsection, as an application of non-collapsing of warping functions fjf_{j} obtained in Proposition 4.15, we derive a uniform positive lower bound for the systole of the sequence of warped product manifolds 𝕊2×fi𝕊1{\mathbb{S}}^{2}\times_{f_{i}}{\mathbb{S}}^{1} satisfying assumptions in Proposition 4.15.

Definition 4.16 (Systole).

The systole of a Riemannian manifold (M,g)(M,g), which is denoted by sys(M,g)sys(M,g) is defined to be the length of the shortest closed geodesic in MM.

Remark 4.17.

People may usually consider so-called π1\pi_{1}-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 𝕊3{\mathbb{S}}^{3}, we may have a short contractible closed geodesic.

First of all we derive an interesting dichotomy property for closed geodesics in warped product manifolds: N×f𝕊1N\times_{f}{\mathbb{S}}^{1}, that is, the product manifold N×𝕊1N\times{\mathbb{S}}^{1} endowed with the metric g=gN+f2g𝕊1g=g_{N}+f^{2}g_{{\mathbb{S}}^{1}}, where (N,gN)(N,g_{N}) is a nn-dimensional (either compact or completep non-compact) Riemannian manifold without boundary, and ff is a positive smooth function on NN.

Lemma 4.18.

There is a dichotomy for closed geodesics in N×f𝕊1N\times_{f}{\mathbb{S}}^{1}, that is, a closed geodesic in N×f𝕊1N\times_{f}{\mathbb{S}}^{1} either wraps around the fiber 𝕊1{\mathbb{S}}^{1}, or is a geodesic in the base NN.

Proof.

Let φ[0,2π]\varphi\in[0,2\pi] is a coordinate on the fiber 𝕊1{\mathbb{S}}^{1}. The warped product metric gg then can be written as

(215) g=gN+f2dφ2.g=g_{N}+f^{2}d\varphi^{2}.

Let

(216) γ(t)=(γN(t),φ(t))t[0,1]\gamma(t)=(\gamma_{N}(t),\varphi(t))\ \ t\in[0,1]

be a closed geodesic in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}, and without loss of generality, we assume φ(0)=0\varphi(0)=0. We have two possible cases as following:

Case 1: φ([0,1])=[0,2π]\varphi([0,1])=[0,2\pi]. In this case, clearly, the geodesic wraps around the fiber 𝕊1{\mathbb{S}}^{1}.

Case 2: φ([0,1])[0,2π]\varphi([0,1])\neq[0,2\pi]. In this case, we show that φ([0,1])={0}\varphi([0,1])=\{0\} by a proof by contradiction, and then clearly, γ\gamma is a closed geodesic on base NN×{φ=0}N\cong N\times\{\varphi=0\}. Otherwise, we have

(217) 0<φ0:=max{φ(t)t[0,1]}<2π.0<\varphi_{0}:=\max\{\varphi(t)\mid t\in[0,1]\}<2\pi.

Moreover, there exists 0<t0<10<t_{0}<1 such that φ(t0)=φ0\varphi(t_{0})=\varphi_{0}, since φ(1)=φ(0)=0\varphi(1)=\varphi(0)=0 due to the closeness of the geodesic γ\gamma. Consequently, t0t_{0} is a critical point of the function φ(t)\varphi(t), i.e. φ(t0)=0\varphi^{\prime}(t_{0})=0. As a result, the tangent vector of the geodesic at t0t_{0}, γ(t0)=(γN(t0),0)\gamma^{\prime}(t_{0})=(\gamma^{\prime}_{N}(t_{0}),0), is tangent to N×{φ=φ0}N\times\{\varphi=\varphi_{0}\}. On the other hand, there is a geodesic contained in N×{φ=φ0}N\times\{\varphi=\varphi_{0}\} that passes through the point (γN(t0),φ0)(\gamma_{N}(t_{0}),\varphi_{0}) and is tangent to (γN(t0),0)(\gamma^{\prime}_{N}(t_{0}),0) at this point. Then by the uniqueness of the geodesic with given tangent vector at a point, and the fact that base NN is totally geodesic in the warped product manifold N×f𝕊1N\times_{f}{\mathbb{S}}^{1}, which can be seen easily by Koszul’s formula, or see Proposition 9.104 in [3], we can obtain φ([0,1])={φ0}\varphi([0,1])=\{\varphi_{0}\}, and this contradicts with φ(0)=0\varphi(0)=0. ∎

By the dichotomy of closed geodesics in Lemma 4.18, we can obtain a lower bound estimate for the systole of N×f𝕊1N\times_{f}{\mathbb{S}}^{1}.

Lemma 4.19.

The systole of the warped product Riemannian manifold N×f𝕊1N\times_{f}{\mathbb{S}}^{1} is greater than or equal to min{sys(N,gN),2πmin𝕊2f}\min\left\{sys(N,g_{N}),2\pi\min\limits_{{\mathbb{S}}^{2}}f\right\}.

Proof.

Let γ(t)=(r(t),θ(t),φ(t)),t[0,1],\gamma(t)=(r(t),\theta(t),\varphi(t)),t\in[0,1], is a closed geodesic in 𝕊2×f𝕊1{\mathbb{S}}^{2}\times_{f}{\mathbb{S}}^{1}. By Lemma 4.18, γ\gamma either wraps around the fiber 𝕊1{\mathbb{S}}^{1}, or γ\gamma is a closed geodesic in the base manifold (N,gN)(N,g_{N}).

If γ\gamma wraps around the fiber 𝕊1{\mathbb{S}}^{1}, then φ([0,1])=[0,2π]\varphi([0,1])=[0,2\pi], and so the length of γ\gamma:

(218) L(γ)=01|γ(t)|g𝑑t\displaystyle L(\gamma)=\int^{1}_{0}|\gamma^{\prime}(t)|_{g}dt \displaystyle\geq 01f(γ(t))|φ(t)|𝑑t\displaystyle\int^{1}_{0}f(\gamma(t))|\varphi^{\prime}(t)|dt
(219) \displaystyle\geq min𝕊2f01|φ(t)|𝑑t\displaystyle\min\limits_{{\mathbb{S}}^{2}}f\int^{1}_{0}|\varphi^{\prime}(t)|dt
(220) \displaystyle\geq 2πmin𝕊2f.\displaystyle 2\pi\min\limits_{{\mathbb{S}}^{2}}f.

If γ\gamma is a closed geodesic in the base (N,gN)(N,g_{N}), then by the definition of systole, the length of γ\gamma is greater than or equal to sys(N,gN)sys(N,g_{N}).

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 {𝕊2×fj𝕊1}j=1\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\}_{j=1}^{\infty} be a sequence of warped product manifolds with metric tensors as in (3) that have non-negative scalar curvature and satisfy

(221) Vol(𝕊2×fj𝕊1)V and MinA(𝕊2×fj𝕊1)A>0,j.{\rm Vol}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\leq V\text{ and }\operatorname{MinA}({\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1})\geq A>0,\forall j\in\mathbb{N}.

Let e:=inf𝕊2f>0e_{\infty}:=\inf\limits_{{\mathbb{S}}^{2}}f_{\infty}>0. Then the systoles of 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}, for all jj\in\mathbb{N}, have a uniform positive lower bound given by min{2π,e2π}\min\left\{2\pi,\frac{e_{\infty}}{2}\pi\right\}.

Proof.

First note that the base manifold of the sequence of the warped product manifolds is the standard 22-sphere, and its systole is equal to 2π2\pi, since the image of a closed geodesic in (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}) is always a great circle.

Then note that e>0e_{\infty}>0 follows from the item (iii)(iii) in Theorem 4.13. For each jj\in\mathbb{N}, by Lemma 4.19, the systole of 𝕊2×fj𝕊1{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1} has a lower bound given by min{2π,2πmin𝕊2fj}\min\left\{2\pi,2\pi\min\limits_{{\mathbb{S}}^{2}}f_{j}\right\}. Then by Proposition 4.15, min𝕊2fje4\min\limits_{{\mathbb{S}}^{2}}f_{j}\geq\frac{e_{\infty}}{4} holds for all jj\in\mathbb{N}. Hence the conclusion follows and we complete the proof. ∎

5. Nonnegative distributional scalar curvature of limit metric

Now we use the positive limit function ff_{\infty} obtained in Theorem 4.13 to define a weak warped product metrics:

Definition 5.1.

Let ff_{\infty} be a function defined on 𝕊2{\mathbb{S}}^{2} such that it is almost everywhere positive and finite on 𝕊2{\mathbb{S}}^{2}. We further assume that fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for 1p<21\leq p<2. Define

(222) g:=g𝕊2+f2g𝕊1,g_{\infty}:=g_{{\mathbb{S}}^{2}}+f_{\infty}^{2}g_{{\mathbb{S}}^{1}},

to be a (weak) warped product Riemannian metric on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} in the sense of defining an inner product on the tangent space at (almost) every point of 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}.

Remark 5.2.

In general, gg_{\infty} is only defined almost everywhere in 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} with respect to the standard product volume measure dvolg𝕊2dvolg𝕊1d{\rm vol}_{g_{{\mathbb{S}}^{2}}}d{\rm vol}_{g_{{\mathbb{S}}^{1}}}, since ff_{\infty} may have value as ++\infty on a measure zero set in 𝕊2{\mathbb{S}}^{2}. Note that we allow ++\infty 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 ++\infty at two poles of 𝕊2{\mathbb{S}}^{2}.

In Subsection 5.1, we show W1,pW^{1,p} regularity of the weak metric tensor gg_{\infty} defined in Definition 5.1 for 1p<21\leq p<2 [Proposition 5.4], and prove that the warped product metrics gj=g𝕊2+fj2g𝕊1g_{j}=g_{{\mathbb{S}}^{2}}+f^{2}_{j}g_{{\mathbb{S}}^{1}} converge to gg_{\infty} in the LqL^{q} sense for any 1q<+1\leq q<+\infty [Theorem 5.5].

In Subsection 5.2, we show that the limit weak metric gg_{\infty} has nonnegative distributional scalar curvature in the sense of Lee-LeFloch [Theorem 5.11].

5.1. W1,pW^{1,p} limit Riemannian metric gg_{\infty}

we prove the regularity of the metric tensor. Before that we need the following definition:

Definition 5.3.

We define Lp(𝕊2×𝕊1,g0)L^{p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) as the set of all tensors defined almost everywhere on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} such that its LpL^{p} norm measured in terms of g0g_{0} is finite where g0g_{0} is the isometric product metric

(223) g0=g𝕊2+g𝕊1 on 𝕊2×𝕊1.g_{0}=g_{{\mathbb{S}}^{2}}+g_{{\mathbb{S}}^{1}}\textrm{ on }{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}.

We define W1,p(𝕊2×𝕊1,g0)W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) as the set of all tensors, hh, defined almost everywhere on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} such that both the LpL^{p} norm of hh and the LpL^{p} norm of ¯h\overline{\nabla}h measured in terms of g0g_{0} are finite where ¯\overline{\nabla} is the connection corresponding to the metric g0g_{0}.

Now we prove the regularity of the metric tensor gg_{\infty} defined in Definition 5.1:

Proposition 5.4 (Regularity of the metric tensor).

The Riemannian metric tensor gg_{\infty} as in Definition 5.1 satisfies

(224) gW1,p(𝕊2×𝕊1,g0)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})

for all p[1,2)p\in[1,2) in the sense of Definition 5.3.

Proof.

Using the background metric, g0g_{0}, we have

(225) gLp(𝕊2×𝕊1,g0)\displaystyle\|g_{\infty}\|_{L^{p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})} =\displaystyle= (2π)1p(2+f4)12Lp(𝕊2)\displaystyle(2\pi)^{\frac{1}{p}}\|(2+f^{4}_{\infty})^{\frac{1}{2}}\|_{L^{p}({\mathbb{S}}^{2})}
(226) \displaystyle\leq (2π)1p2+f2Lp(𝕊2)\displaystyle(2\pi)^{\frac{1}{p}}\|\sqrt{2}+f^{2}_{\infty}\|_{L^{p}({\mathbb{S}}^{2})}
(227) \displaystyle\leq (2π)1p(2(4π)1p+fL2p(𝕊2)2)\displaystyle(2\pi)^{\frac{1}{p}}\left(\sqrt{2}(4\pi)^{\frac{1}{p}}+\|f_{\infty}\|^{2}_{L^{2p}({\mathbb{S}}^{2})}\right)

is finite, since by the assumption, fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for any p[1,2)p\in[1,2), and Sobolev embedding theorem, we have fL2p(𝕊2)f_{\infty}\in L^{2p}({\mathbb{S}}^{2}) for any p[1,)p\in[1,\infty).

Now for the gradient estimate, we fix an arbitrary p[1,2)p\in[1,2). We use ¯\overline{\nabla} to denote the connection of the background metric g0g_{0}. Clearly, we have

(228) ¯g=¯g𝕊2+¯f2g𝕊1+f2¯g𝕊1.\overline{\nabla}g_{\infty}=\overline{\nabla}g_{{\mathbb{S}}^{2}}+\overline{\nabla}f^{2}_{\infty}\otimes g_{{\mathbb{S}}^{1}}+f^{2}_{\infty}\overline{\nabla}g_{{\mathbb{S}}^{1}}.

and

(229) ¯g𝕊2=0, and ¯g𝕊1=0.\overline{\nabla}g_{{\mathbb{S}}^{2}}=0,\text{ and }\overline{\nabla}g_{{\mathbb{S}}^{1}}=0.

Moreover, since ¯f2=2ff\overline{\nabla}f^{2}_{\infty}=2f_{\infty}\nabla f_{\infty} we have

(230) ¯g=2ffg𝕊1,\overline{\nabla}g_{\infty}=2f_{\infty}\nabla f_{\infty}\otimes g_{{\mathbb{S}}^{1}},

where f\nabla f_{\infty} is the gradient of ff_{\infty} on (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}). As a result, we have

(231) ¯gLp(𝕊2×𝕊1,g0)p\displaystyle\|\overline{\nabla}g_{\infty}\|^{p}_{L^{p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})} =\displaystyle= 2π𝕊22pfp|f|p𝑑volg𝕊2\displaystyle 2\pi\int_{{\mathbb{S}}^{2}}2^{p}f^{p}_{\infty}|\nabla f_{\infty}|^{p}d{\rm vol}_{g_{{\mathbb{S}}^{2}}}
(232) =\displaystyle= 2p+1πfLpq(𝕊2,g𝕊2)fLpq(𝕊2),\displaystyle 2^{p+1}\pi\|f_{\infty}\|_{L^{pq^{*}}({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}})}\cdot\|\nabla f_{\infty}\|_{L^{pq}({\mathbb{S}}^{2})},

where q>1q>1 is chosen so that pq<2pq<2, and q=qq1q^{*}=\frac{q}{q-1}. Then again by Sobolev embedding theorem we have fLqf_{\infty}\in L^{q} for any p[1,)p\in[1,\infty), thus we obtain that ¯gLp(𝕊2×𝕊1,g0)\|\overline{\nabla}g_{\infty}\|_{L^{p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})} is finite for any p[1,2)p\in[1,2). This completes the proof. ∎

Then we apply Proposition 3.5 to prove Theorem 1.7 which concerns the LqL^{q} pre-compactness of warped product circles over sphere with non-negative scalar curvature. We restate Theorem 1.7 as follows:

Theorem 5.5.

Let {gj=g𝕊2+fj2g𝕊1j}\{g_{j}=g_{{\mathbb{S}}^{2}}+f^{2}_{j}g_{{\mathbb{S}}^{1}}\mid j\in\mathbb{N}\} be a sequence of warped Riemannian metrics on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} satisfying requirements in (4). Then there exists a subsequence gjkg_{j_{k}} and a (weak) warped Riemannian metric gW1,p(𝕊2×𝕊1,g0)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for p[1,2)p\in[1,2) as in Definition 5.1 such that

(233) gjkg inLq(𝕊2×𝕊1,g0),q[1,).g_{j_{k}}\rightarrow g_{\infty}\ \ \text{ in}\ \ L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}),\ \ \forall q\in[1,\infty).
Proof.

By Lemma 2.1 and Lemma 2.2, the assumptions in (4) for gjg_{j} implies that the warping functions fjf_{j} satisfy the assumptions in Proposition 3.5. Thus, by applying Proposition 3.5, we have that there exists a subsequence fjkf_{j_{k}} of warping functions and fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for all 1p<21\leq p<2, such that

(234) fjkf,inLq(𝕊2),q[1,).f_{j_{k}}\rightarrow f_{\infty},\ \ \text{in}\ \ L^{q}({\mathbb{S}}^{2}),\ \ \forall q\in[1,\infty).

Let g:=g𝕊2+f2g𝕊1g_{\infty}:=g_{{\mathbb{S}}^{2}}+f^{2}_{\infty}g_{{\mathbb{S}}^{1}}. Then by Proposition 5.4, we have

(235) gW1,p(𝕊2×𝕊1,g0)1p<2.g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})\ \ \forall 1\leq p<2.

Moreover, because

(236) gjg=(fj2f2)g𝕊1,g_{j}-g_{\infty}=(f_{j}^{2}-f_{\infty}^{2})g_{{\mathbb{S}}^{1}},

we have that for any q[1,)q\in[1,\infty),

(237) gjkgLq(𝕊2×𝕊1,g0)\displaystyle\|g_{j_{k}}-g_{\infty}\|_{L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})}
(238) =\displaystyle= (2π)1qfjk2f2Lq(𝕊2)\displaystyle(2\pi)^{\frac{1}{q}}\|f^{2}_{j_{k}}-f^{2}_{\infty}\|_{L^{q}({\mathbb{S}}^{2})}
(239) =\displaystyle= (2π)1q(fjkf)(fjk+f)Lq(𝕊2)\displaystyle(2\pi)^{\frac{1}{q}}\|(f_{j_{k}}-f_{\infty})\cdot(f_{j_{k}}+f_{\infty})\|_{L^{q}({\mathbb{S}}^{2})}
(240) \displaystyle\leq (2π)1qfjkfL2q(𝕊2)fjk+fL2q(𝕊2)\displaystyle(2\pi)^{\frac{1}{q}}\|f_{j_{k}}-f_{\infty}\|_{L^{2q}({\mathbb{S}}^{2})}\cdot\|f_{j_{k}}+f_{\infty}\|_{L^{2q}({\mathbb{S}}^{2})}
(241) \displaystyle\rightarrow 0,asjk,\displaystyle 0,\ \ \text{as}\ \ j_{k}\rightarrow\infty,

since fjkff_{j_{k}}\rightarrow f_{\infty} in L2q(𝕊2)L^{2q}({\mathbb{S}}^{2}) for any q[1,)q\in[1,\infty). ∎

Remark 5.6.

As showed by the example constructed by Christina Sormani and authors in [19], gW1,p(𝕊2×𝕊1,g0)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for 1p<21\leq p<2 is the best regularity we can expect in general for the limit weak Riemannian metric gg_{\infty}, see Proposition 3.6 and Remark 3.8 in [19].

5.2. Nonnegative distributional scalar curvature of gg_{\infty}

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 LlocWloc1,2L^{\infty}_{loc}\cap W^{1,2}_{loc}. 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 {𝕊2×fj𝕊1}\{{\mathbb{S}}^{2}\times_{f_{j}}{\mathbb{S}}^{1}\} with non-negative scalar curvature have uniform bounded volumes, then a subsequence of the smooth warped product metric gj=g𝕊2+fj2g𝕊1g_{j}=g_{{\mathbb{S}}^{2}}+f^{2}_{j}g_{{\mathbb{S}}^{1}} converges to a weak warped product metric g=g𝕊2+f2g𝕊1W1,p(𝕊2×𝕊1,g0)(1p<2)g_{\infty}=g_{{\mathbb{S}}^{2}}+f^{2}_{\infty}g_{{\mathbb{S}}^{1}}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0})(1\leq p<2) in the sense of Lq(𝕊2×𝕊1,g0)L^{q}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for any q1q\geq 1. For the rest of this section, we use gg_{\infty} to denote such limit metric. We use g0=g𝕊2+g𝕊1g_{0}=g_{{\mathbb{S}}^{2}}+g_{{\mathbb{S}}^{1}} as a background metric .

In Theorem 5.11, we prove that this limit (weak) metric gg_{\infty} 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 gg_{\infty} 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 MM be a smooth manifold endowed with a smooth background metric, g0g_{0}. Let gg be a metric tensor defined on MM with LlocWloc1,2L^{\infty}_{loc}\cap W^{1,2}_{loc} regularity and locally bounded inverse g1Llocg^{-1}\in L^{\infty}_{loc}.

The scalar curvature distribution Scalarg\operatorname{Scalar}_{g} is defined as a distributions in MM such that for every test function uC0(M)u\in C^{\infty}_{0}(M)

(242) Scalarg,u:=M(V¯(udμgdμg0)+Fudμgdμ0)𝑑μ0,\langle\operatorname{Scalar}_{g},u\rangle:=\int_{M}\left(-V\cdot\overline{\nabla}\left(u\frac{d\mu_{g}}{d\mu_{g_{0}}}\right)+Fu\frac{d\mu_{g}}{\,d\mu_{0}}\right)\,d\mu_{0},

where the dot product is taken using the metric g0g_{0}, ¯\overline{\nabla} is the Levi-Civita connection of g0g_{0}, dμgd\mu_{g} and dμg0d\mu_{g_{0}} are volume measure with respect to gg and g0g_{0} respectively, VV is a vector field given by

(243) Vk:=gijΓijkgikΓjij,V^{k}:=g^{ij}\Gamma^{k}_{ij}-g^{ik}\Gamma^{j}_{ji},

where

(244) Γijk:=12gkl(¯igjl+¯jgil¯lgij),\Gamma^{k}_{ij}:=\frac{1}{2}g^{kl}\left(\overline{\nabla}_{i}g_{jl}+\overline{\nabla}_{j}g_{il}-\overline{\nabla}_{l}g_{ij}\right),
(245) F:=R¯¯kgijΓijk+¯kgikΓjij+gij(ΓklkΓijlΓjlkΓikl),F:=\overline{R}-\overline{\nabla}_{k}g^{ij}\Gamma^{k}_{ij}+\overline{\nabla}_{k}g^{ik}\Gamma^{j}_{ji}+g^{ij}\left(\Gamma^{k}_{kl}\Gamma^{l}_{ij}-\Gamma^{k}_{jl}\Gamma^{l}_{ik}\right),

and

(246) R¯:=gij(kΓ¯ijkiΓ¯kjk+Γ¯ijlΓ¯klkΓ¯kjlΓ¯ilk).\overline{R}:=g^{ij}\left(\partial_{k}\overline{\Gamma}^{k}_{ij}-\partial_{i}\overline{\Gamma}^{k}_{kj}+\overline{\Gamma}^{l}_{ij}\overline{\Gamma}^{k}_{kl}-\overline{\Gamma}^{l}_{kj}\overline{\Gamma}^{k}_{il}\right).

The Riemannian metric gg has nonnegative distributional scalar curvature, if Scalarg,u0\langle\operatorname{Scalar}_{g},u\rangle\geq 0 for every nonnegative test function uu in the integral in (242).

Definition 5.8 (Distributional total scalar curvature).

For a weak metric gg having the regularity as in Definition 5.7, we define the distributional total scalar curvature of gg to be Scalarg,1\langle\operatorname{Scalar}_{g},1\rangle, which is obtained by setting the test function u1u\equiv 1 in the integration in (242).

Note that for a C2C^{2}-metric, the distributional total scalar curvature is exactly the usual total scalar curvature.

Remark 5.9.

By the regularity assumption for the Riemannian metric gg in the work of Lee-LeFloch [10], one has the regularity ΓijkLloc2\Gamma^{k}_{ij}\in L^{2}_{loc}, VLloc2,FLloc1V\in L^{2}_{loc},F\in L^{1}_{loc}, and the density of volume measure dμgd\mu_{g} with respect to dμ0\,d\mu_{0} is

(247) dμgdμ0LlocWloc1,2.\tfrac{d\mu_{g}}{\,d\mu_{0}}\in L^{\infty}_{loc}\cap W^{1,2}_{loc}.

Thus

(248) FirstIntg=M(V¯(udμgdμg0))𝑑μ0FirstInt_{g}=\int_{M}\left(-V\cdot\overline{\nabla}\left(u\frac{d\mu_{g}}{d\mu_{g_{0}}}\right)\right)\,d\mu_{0}

and

(249) SecondIntg=M(Fudμgdμ0)𝑑μ0.SecondInt_{g}=\int_{M}\left(Fu\frac{d\mu_{g}}{\,d\mu_{0}}\right)\,d\mu_{0}.

are both finite.

Remark 5.10.

Our limit metric is less regular than the metrics studied by Lee-LeFloch in [10]. Recall that in Proposition 5.4 we showed gW1,p(𝕊2×𝕊1,g0)g_{\infty}\in W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for 1p<21\leq p<2, and as shown by the extreme example constructed in [19], in general gWloc1,2(𝕊2×𝕊1,g0)g_{\infty}\notin W^{1,2}_{loc}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}), see Proposition 3.6 in [19].

In Remark 5.18 below we show that in genenral both integrals in (248) and (249) may be divergent. However, in Theorem 5.11 below, we show that in our case the sum of (248) and (249) is still well-defined since the singularity cancels out when we add them up.

We are ready to prove Theorem 1.8. We restate it as follows:

Theorem 5.11.

The limit metric gg_{\infty} obtained in Theorem 5.5 has nonnegative distributional scalar curvature on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1} in the sense of Lee-LeFloch as in Definition 5.7. In particular, (242) is finite and nonnegative for any nonnegative test function, uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}). Moreover, the total scalar curvatures of gjg_{j} converge to the distributional total scalar curvature of gg_{\infty}.

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 g0=g𝕊2+g𝕊1g_{0}=g_{{\mathbb{S}}^{2}}+g_{{\mathbb{S}}^{1}} as background metric, and use coordinate {r,θ,φ}\{r,\theta,\varphi\} on 𝕊2×𝕊1{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}, where (r,θ)(r,\theta) is a polar coordinate on 𝕊2{\mathbb{S}}^{2} and φ\varphi is a coordinate on 𝕊1{\mathbb{S}}^{1}. The corresponding local frame of the tangent bundle is {r,θ,φ}\{\partial_{r},\partial_{\theta},\partial_{\varphi}\}. In this coordinate system, both g0g_{0} and gg_{\infty} are diagonal and given as

(250) g0=(1000sin2r0001) and g=(1000sin2r000f2(r,θ)).g_{0}=\begin{pmatrix}1&0&0\\ 0&\sin^{2}r&0\\ 0&0&1\\ \end{pmatrix}\textrm{ and }g_{\infty}=\begin{pmatrix}1&0&0\\ 0&\sin^{2}r&0\\ 0&0&f_{\infty}^{2}(r,\theta)\\ \end{pmatrix}.

First of all, by the formula of Christoffel symbols:

(251) Γ¯jki=12(g0)il((g0)ilxk+(g0)lkxj(g0)jkxl),\overline{\Gamma}^{i}_{jk}=\frac{1}{2}(g_{0})^{il}\left(\frac{\partial(g_{0})_{il}}{\partial x^{k}}+\frac{\partial(g_{0})_{lk}}{\partial x^{j}}-\frac{\partial(g_{0})_{jk}}{\partial x^{l}}\right),

one can easily obtain the following lemma:

Lemma 5.12.

The Christoffel symbols of the Levi-Civita connection ¯\overline{\nabla} of the background metric g0=g𝕊2+g𝕊1g_{0}=g_{{\mathbb{S}}^{2}}+g_{{\mathbb{S}}^{1}}, in the coordinate {r,θ,φ}\{r,\theta,\varphi\}, all vanish except

(252) Γ¯θθr=sinrcosr,\overline{\Gamma}^{r}_{\theta\theta}=-\sin r\cos r,

and

(253) Γ¯rθθ=Γ¯θrθ=cosrsinr.\overline{\Gamma}^{\theta}_{r\theta}=\overline{\Gamma}^{\theta}_{\theta r}=\frac{\cos r}{\sin r}.

Then by Lemma 5.12, the formula

(254) ¯i(g)jl=i((g)jl)Γ¯ijp(g)plΓ¯ilq(g)jq,\overline{\nabla}_{i}(g_{\infty})_{jl}=\partial_{i}\left((g_{\infty})_{jl}\right)-\overline{\Gamma}^{p}_{ij}(g_{\infty})_{pl}-\overline{\Gamma}^{q}_{il}(g_{\infty})_{jq},

and the diagonal expression of gg_{\infty} in (250), one can obtain the following lemma:

Lemma 5.13.

For the limit metric, gg_{\infty}, with the background metric, g0g_{0}, the Christoffel symbols defined by Lee-LeFloch as in (244), in the coordinate {r,θ,φ}\{r,\theta,\varphi\}, all vanish except

(255) Γφφr=frf,Γφφθ=1sin2rfθf,\Gamma^{r}_{\varphi\varphi}=-f_{\infty}\partial_{r}f_{\infty},\quad\Gamma^{\theta}_{\varphi\varphi}=-\frac{1}{\sin^{2}r}f_{\infty}\partial_{\theta}f_{\infty},

and

(256) Γrφφ=Γφrφ=rff,Γθφφ=Γφθφ=θff.\Gamma^{\varphi}_{r\varphi}=\Gamma^{\varphi}_{\varphi r}=\frac{\partial_{r}f_{\infty}}{f_{\infty}},\quad\Gamma^{\varphi}_{\theta\varphi}=\Gamma^{\varphi}_{\varphi\theta}=\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}.

Note also that

Lemma 5.14.

Note that the volume forms are:

(257) dμ0=drsin(r)dθdφd\mu_{0}=\,dr\wedge\sin(r)\,d\theta\wedge\,d\varphi

and

(258) dμ=drsin(r)dθf(r,θ)dφd\mu_{\infty}=dr\wedge\sin(r)\,d\theta\wedge f_{\infty}(r,\theta)\,d\varphi

which are both defined almost everywhere. In particular,

(259) dμdμ0=f(r,θ)\frac{\,d\mu_{\infty}}{\,d\mu_{0}}=f_{\infty}(r,\theta)

is in W1,p(𝕊2×𝕊1,g0)W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for p<2p<2.

Proof.

The first claim holds away from r=0r=0 and r=πr=\pi by the definition of volume form, and the second claim holds almost everywhere on (𝕊2×𝕊1,g0)({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}). So dμ=fdμ0d\mu_{\infty}=f_{\infty}d\mu_{0} almost everywhere which gives us the third claim. The rest follows from Proposition 3.5. ∎

Now we are ready to compute the vector field VV and the function FF defined by Lee-LeFloch as in (243) and (245)(\ref{defn-F}).

Lemma 5.15.

For the limit metric gg_{\infty} with the background metric g0g_{0}, the vector field VV defined in (243), in the local frame {r,θ,φ}\{\partial_{r},\partial_{\theta},\partial_{\varphi}\}, is given by

(260) V=(2rff,2sin2rθff,0).V=\left(-2\frac{\partial_{r}f_{\infty}}{f_{\infty}},-\frac{2}{\sin^{2}r}\frac{\partial_{\theta}f_{\infty}}{f_{\infty}},0\right).

Furthermore

(261) V¯(udμdμ0)=2rffr(uf)+2sin2rθffθ(uf).-V\cdot\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right)=2\frac{\partial_{r}f_{\infty}}{f_{\infty}}\partial_{r}(uf_{\infty})+\frac{2}{\sin^{2}r}\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\partial_{\theta}(uf_{\infty}).
Proof.

By plugging the non-vanishing Christoffel symbols in Lemma 5.13 into

(262) Vk:=gijΓijkgikΓjij,V^{k}:=g^{ij}_{\infty}\Gamma^{k}_{ij}-g^{ik}_{\infty}\Gamma^{j}_{ji},

we get

(263) Vr\displaystyle V^{r} =\displaystyle= gφφΓφφrgrrΓφrφ\displaystyle g^{\varphi\varphi}_{\infty}\Gamma^{r}_{\varphi\varphi}-g^{rr}_{\infty}\Gamma^{\varphi}_{\varphi r}
(264) =\displaystyle= 1(f)2(frf)rff=2rff.\displaystyle\frac{1}{(f_{\infty})^{2}}(-f_{\infty}\partial_{r}f_{\infty})-\frac{\partial_{r}f_{\infty}}{f_{\infty}}=-2\frac{\partial_{r}f_{\infty}}{f_{\infty}}.

Also

(265) Vθ\displaystyle V^{\theta} =\displaystyle= gφφΓφφθgθθΓφθφ\displaystyle g^{\varphi\varphi}_{\infty}\Gamma^{\theta}_{\varphi\varphi}-g^{\theta\theta}_{\infty}\Gamma^{\varphi}_{\varphi\theta}
(266) =\displaystyle= 1f2(1sin2rfθf)1sin2rθff=2sin2rθff.\displaystyle\frac{1}{f^{2}_{\infty}}\left(-\frac{1}{\sin^{2}r}f_{\infty}\partial_{\theta}f_{\infty}\right)-\frac{1}{\sin^{2}r}\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}=-\frac{2}{\sin^{2}r}\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}.
(267) Vφ=gijΓijφgφφΓjφj=0.V^{\varphi}=g^{ij}_{\infty}\Gamma^{\varphi}_{ij}-g^{\varphi\varphi}_{\infty}\Gamma^{j}_{j\varphi}=0.

By Lemma A.8, we now see that,

(268) ¯(udμdμ0)\displaystyle\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right) =\displaystyle= ¯(uf)\displaystyle\overline{\nabla}\left(uf_{\infty}\right)
(269) =\displaystyle= r(uf)r+1sin2rθ(uf)θ+φ(uf)φ\displaystyle\partial_{r}(uf_{\infty})\frac{\partial}{\partial r}+\frac{1}{\sin^{2}r}\partial_{\theta}(uf_{\infty})\frac{\partial}{\partial\theta}+\partial_{\varphi}(uf_{\infty})\frac{\partial}{\partial\varphi}

Thus

(270) V¯(udμdμ0)=2rffr(uf)+2sin2rθffθ(uf)-V\cdot\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right)=2\frac{\partial_{r}f_{\infty}}{f_{\infty}}\partial_{r}(uf_{\infty})+\frac{2}{\sin^{2}r}\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\partial_{\theta}(uf_{\infty})

Lemma 5.16.

For the limit metric gg_{\infty} with the background metric g0g_{0}, the function FF defined in (245) is given by

(271) F=22(rff)22sin2r(θff)2=221(f)2|f|2.F=2-2\left(\frac{\partial_{r}f_{\infty}}{f_{\infty}}\right)^{2}-\frac{2}{\sin^{2}r}\left(\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\right)^{2}=2-2\frac{1}{(f_{\infty})^{2}}|\nabla f_{\infty}|^{2}.

Furthermore,

(272) (Fudμdμ0)=2uf2uf|f|2.\left(Fu\frac{\,d\mu_{\infty}}{d\mu_{0}}\right)=2uf_{\infty}-2\frac{u}{f_{\infty}}|\nabla f_{\infty}|^{2}.

Here |f||\nabla f_{\infty}| is the norm of weak gradient of ff_{\infty} with respect to the standard metric g𝕊2g_{{\mathbb{S}}^{2}}.

Proof.

First note that from the expression of R¯\overline{R} in (246) and the Christofell symbols calculated in Lemma 5.12, one can easily see that

(273) R¯=Rg𝕊2=2.\overline{R}=R_{g_{{\mathbb{S}}^{2}}}=2.

Also recall that

(274) ¯igjl=i(gjl)+Γ¯ipjgpl+Γ¯iqlgjq.\overline{\nabla}_{i}g^{jl}_{\infty}=\partial_{i}(g^{jl}_{\infty})+\overline{\Gamma}^{j}_{ip}g^{pl}_{\infty}+\overline{\Gamma}^{l}_{iq}g^{jq}_{\infty}.

Then by Lemmas 5.12 and 5.13, one has

(275) F\displaystyle F :=\displaystyle:= R¯(¯kgij)Γijk+(¯kgik)Γjij+gij(ΓklkΓijlΓjlkΓikl)\displaystyle\overline{R}-(\overline{\nabla}_{k}g^{ij})\Gamma^{k}_{ij}+(\overline{\nabla}_{k}g^{ik})\Gamma^{j}_{ji}+g^{ij}(\Gamma^{k}_{kl}\Gamma^{l}_{ij}-\Gamma^{k}_{jl}\Gamma^{l}_{ik})
(280) =\displaystyle= 2¯rgφφΓφφr¯θgφφΓφφθ2¯φgrφΓrφφ2¯φgθφΓθφφ\displaystyle 2-\overline{\nabla}_{r}g^{\varphi\varphi}\Gamma^{r}_{\varphi\varphi}-\overline{\nabla}_{\theta}g^{\varphi\varphi}\Gamma^{\theta}_{\varphi\varphi}-2\overline{\nabla}_{\varphi}g^{r\varphi}\Gamma^{\varphi}_{r\varphi}-2\overline{\nabla}_{\varphi}g^{\theta\varphi}\Gamma^{\varphi}_{\theta\varphi}
+¯kgrkΓφrφ+¯kgθkΓφθφ\displaystyle+\overline{\nabla}_{k}g^{rk}\Gamma^{\varphi}_{\varphi r}+\overline{\nabla}_{k}g^{\theta k}\Gamma^{\varphi}_{\varphi\theta}
+gφφΓφrφΓφφr+gφφΓφθφΓφφθ\displaystyle+\cancel{g^{\varphi\varphi}\Gamma^{\varphi}_{\varphi r}\Gamma^{r}_{\varphi\varphi}}+\bcancel{g^{\varphi\varphi}\Gamma^{\varphi}_{\varphi\theta}\Gamma^{\theta}_{\varphi\varphi}}
gφφΓφφrΓrφφgφφΓφφθΓφθφgrrΓrφφΓrφφgφφΓφrφΓφφr\displaystyle-g^{\varphi\varphi}\Gamma^{r}_{\varphi\varphi}\Gamma^{\varphi}_{r\varphi}-\bcancel{g^{\varphi\varphi}\Gamma^{\theta}_{\varphi\varphi}\Gamma^{\varphi}_{\varphi\theta}}-g^{rr}\Gamma^{\varphi}_{r\varphi}\Gamma^{\varphi}_{r\varphi}-\cancel{g^{\varphi\varphi}\Gamma^{\varphi}_{\varphi r}\Gamma^{r}_{\varphi\varphi}}
gθθΓθφφΓθφφgφφΓφθφΓφφθ\displaystyle-g^{\theta\theta}\Gamma^{\varphi}_{\theta\varphi}\Gamma^{\varphi}_{\theta\varphi}-g^{\varphi\varphi}\Gamma^{\varphi}_{\varphi\theta}\Gamma^{\theta}_{\varphi\varphi}
(290) =\displaystyle= 2(r(gφφ)+2Γ¯rφφgφφ)Γφφr(θ(gφφ)+2Γ¯θφφgφφ)Γφφθ\displaystyle 2-\left(\partial_{r}(g^{\varphi\varphi})+2\overline{\Gamma}^{\varphi}_{r\varphi}g^{\varphi\varphi}\right)\Gamma^{r}_{\varphi\varphi}-\left(\partial_{\theta}(g^{\varphi\varphi})+2\overline{\Gamma}^{\varphi}_{\theta\varphi}g^{\varphi\varphi}\right)\Gamma^{\theta}_{\varphi\varphi}
2(φ(grφ)+Γ¯φφrgφφ+Γ¯φrφgrr)Γrφφ\displaystyle-2\left(\partial_{\varphi}(g^{r\varphi})+\overline{\Gamma}^{r}_{\varphi\varphi}g^{\varphi\varphi}+\overline{\Gamma}^{\varphi}_{\varphi r}g^{rr}\right)\Gamma^{\varphi}_{r\varphi}
2(φ(gθφ)+Γ¯φφθgφφ+Γ¯φθφgθθ)Γθφφ\displaystyle-2\left(\partial_{\varphi}(g^{\theta\varphi})+\overline{\Gamma}^{\theta}_{\varphi\varphi}g^{\varphi\varphi}+\overline{\Gamma}^{\varphi}_{\varphi\theta}g^{\theta\theta}\right)\Gamma^{\varphi}_{\theta\varphi}
+(r(grr)+Γ¯rrrgrr+Γ¯rrrgrr)Γφrφ\displaystyle+\left(\partial_{r}(g^{rr})+\overline{\Gamma}^{r}_{rr}g^{rr}+\overline{\Gamma}^{r}_{rr}g^{rr}\right)\Gamma^{\varphi}_{\varphi r}
+(θ(grθ)+Γ¯θθrgθθ+Γ¯θrθgrr)Γφrφ\displaystyle+\left(\partial_{\theta}(g^{r\theta})+\overline{\Gamma}^{r}_{\theta\theta}g^{\theta\theta}+\overline{\Gamma}^{\theta}_{\theta r}g^{rr}\right)\Gamma^{\varphi}_{\varphi r}
+(φ(grφ)+Γ¯φφrgφφ+Γ¯φrφgrr)Γφrφ\displaystyle+\left(\partial_{\varphi}(g^{r\varphi})+\overline{\Gamma}^{r}_{\varphi\varphi}g^{\varphi\varphi}+\overline{\Gamma}^{\varphi}_{\varphi r}g^{rr}\right)\Gamma^{\varphi}_{\varphi r}
+(r(gθr)+Γ¯rrθgrr+Γ¯rθrgθθ)Γφθφ\displaystyle+\left(\partial_{r}(g^{\theta r})+\overline{\Gamma}^{\theta}_{rr}g^{rr}+\overline{\Gamma}^{r}_{r\theta}g^{\theta\theta}\right)\Gamma^{\varphi}_{\varphi\theta}
+(θ(gθθ)+Γ¯θθθgθθ+Γ¯θθθgθθ)Γφθφ\displaystyle+\left(\partial_{\theta}(g^{\theta\theta})+\overline{\Gamma}^{\theta}_{\theta\theta}g^{\theta\theta}+\overline{\Gamma}^{\theta}_{\theta\theta}g^{\theta\theta}\right)\Gamma^{\varphi}_{\varphi\theta}
+(φ(gθφ)+Γ¯φφθgφφ+Γ¯φθφgθθ)Γφθφ\displaystyle+\left(\partial_{\varphi}(g^{\theta\varphi})+\overline{\Gamma}^{\theta}_{\varphi\varphi}g^{\varphi\varphi}+\overline{\Gamma}^{\varphi}_{\varphi\theta}g^{\theta\theta}\right)\Gamma^{\varphi}_{\varphi\theta}
gφφΓφφrΓrφφgφφΓφθφΓφφθgrrΓrφφΓrφφgθθΓφθφΓφθφ\displaystyle-g^{\varphi\varphi}\Gamma^{r}_{\varphi\varphi}\Gamma^{\varphi}_{r\varphi}-g^{\varphi\varphi}\Gamma^{\varphi}_{\varphi\theta}\Gamma^{\theta}_{\varphi\varphi}-g^{rr}\Gamma^{\varphi}_{r\varphi}\Gamma^{\varphi}_{r\varphi}-g^{\theta\theta}\Gamma^{\varphi}_{\varphi\theta}\Gamma^{\varphi}_{\varphi\theta}
(294) =\displaystyle= 2(2)rf(f)3(frf)(2)θf(f)3(1sin2rfθf)\displaystyle 2-(-2)\frac{\partial_{r}f_{\infty}}{(f_{\infty})^{3}}\left(-f_{\infty}\partial_{r}f_{\infty}\right)-(-2)\frac{\partial_{\theta}f_{\infty}}{(f_{\infty})^{3}}\left(-\frac{1}{\sin^{2}r}f_{\infty}\partial_{\theta}f_{\infty}\right)
+(cosrsinr+cosrsinr)Γφrφ1(f)2(frf)(rff)\displaystyle+\left(-\frac{\cos r}{\sin r}+\frac{\cos r}{\sin r}\right)\Gamma^{\varphi}_{\varphi r}-\frac{1}{(f_{\infty})^{2}}(-f_{\infty}\partial_{r}f_{\infty})\left(\frac{\partial_{r}f_{\infty}}{f_{\infty}}\right)
1(f)2(1sin2rfθf)(θff)\displaystyle-\frac{1}{(f_{\infty})^{2}}\left(-\frac{1}{\sin^{2}r}f_{\infty}\partial_{\theta}f_{\infty}\right)\left(\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\right)
(rff)21sin2r(θff)2\displaystyle-\left(\frac{\partial_{r}f_{\infty}}{f_{\infty}}\right)^{2}-\frac{1}{\sin^{2}r}\left(\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\right)^{2}
(295) =\displaystyle= 22(rff)22sin2r(θff)2\displaystyle 2-2\left(\frac{\partial_{r}f_{\infty}}{f_{\infty}}\right)^{2}-\frac{2}{\sin^{2}r}\left(\frac{\partial_{\theta}f_{\infty}}{f_{\infty}}\right)^{2}
(296) =\displaystyle= 221(f)2|f|2.\displaystyle 2-2\frac{1}{(f_{\infty})^{2}}|\nabla f_{\infty}|^{2}.

We immediately obtain our second claim by applying Lemma A.8. ∎

Lemma 5.17.

For gg being our limit metric tensor gg_{\infty} and a smooth nonnegative test function uu, the integrals in (248) and (249) are given by

(297) FirstIntg\displaystyle\quad FirstInt_{g_{\infty}} =\displaystyle= 𝕊2×𝕊1(V¯(udμdμ0))𝑑μ0\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\left(-V\cdot\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right)\right)\,d\mu_{0}
(298) =\displaystyle= 𝕊2(2f,u¯+2u¯f|f|2)𝑑volg𝕊2,\displaystyle\int_{{\mathbb{S}}^{2}}\left(2\langle\nabla f_{\infty},\nabla\bar{u}\rangle+2\frac{\bar{u}}{f_{\infty}}|\nabla f_{\infty}|^{2}\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

and

(299) SecondIntg\displaystyle\quad SecondInt_{g_{\infty}} =\displaystyle= 𝕊2×𝕊1(Fudμdμ0)𝑑μ0\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\left(Fu\frac{\,d\mu_{\infty}}{d\mu_{0}}\right)\,d\mu_{0}
(300) =\displaystyle= 𝕊2(2u¯f2u¯f|f|2)𝑑volg𝕊2,\displaystyle\int_{{\mathbb{S}}^{2}}\left(2\bar{u}f_{\infty}-2\frac{\bar{u}}{f_{\infty}}|\nabla f_{\infty}|^{2}\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

where

(301) u¯(r,θ)=02πu(r,θ,φ)𝑑φ,\bar{u}(r,\theta)=\int^{2\pi}_{0}u(r,\theta,\varphi)d\varphi,

f\nabla f_{\infty} and u¯\nabla\bar{u} are (weak) gradients of functions ff_{\infty} and u¯\bar{u} on standard sphere (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}) respectively, and ,\langle\cdot,\cdot\rangle is the Riemannian metric on (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}).

Proof.

By integrating the formulas in Lemma 5.15 and Lemma 5.16, one can easily obtain the integrals in (298) and (300). ∎

Remark 5.18.

As explained in Remark 3.6, fW1,pf_{\infty}\in W^{1,p} for any 1p<21\leq p<2, which is obtained in in Proposition 3.5, is the best regularity for ff_{\infty} in general, and we cannot expect ff_{\infty} is in Wloc1,2(𝕊2)W^{1,2}_{loc}({\mathbb{S}}^{2}). So the integral 𝕊2u¯f|f|2𝑑volg𝕊2\int_{{\mathbb{S}}^{2}}\frac{\bar{u}}{f_{\infty}}|\nabla f_{\infty}|^{2}d{\rm vol}_{g_{{\mathbb{S}}^{2}}} 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 g=g𝕊2+f2g𝕊1g_{\infty}=g_{{\mathbb{S}}^{2}}+f^{2}_{\infty}g_{{\mathbb{S}}^{1}}, the scalar curvature distribution Scalarg\operatorname{Scalar}_{g_{\infty}} defined in Definition 5.7 can be expressed, for every test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}), as the integral

(302) Scalarg,u=𝕊2(2f,u¯+2fu¯)𝑑volg𝕊2,\langle\operatorname{Scalar}_{g_{\infty}},u\rangle=\int_{{\mathbb{S}}^{2}}\left(2\langle\nabla f_{\infty},\nabla\bar{u}\rangle+2f_{\infty}\bar{u}\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

and this is finite for any test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}). Here u¯\bar{u} is defined as in (350), f\nabla f_{\infty} and u¯\nabla\bar{u} are (weak) gradients of functions ff_{\infty} and u¯\bar{u} on standard sphere (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}) respectively, and ,\langle\cdot,\cdot\rangle is the Riemannian metric on (𝕊2,g𝕊2)({\mathbb{S}}^{2},g_{{\mathbb{S}}^{2}}).

Proof.

The expression in (302) immediately follows from the expressions in (298) and (300) and Definition 5.7. The finiteness of the integral in (302) follows from that fW1,p(𝕊2)f_{\infty}\in W^{1,p}({\mathbb{S}}^{2}) for 1p<21\leq p<2 as proved in Proposition 3.5. ∎

We now apply these lemmas to prove Theorem 5.11:

Proof.

By the expression (11) of the scalar curvature of 𝕊2×fi𝕊1{\mathbb{S}}^{2}\times_{f_{i}}{\mathbb{S}}^{1}, we have that for any test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}),

(303) 𝕊2×𝕊1Scalargjudvolgj\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\operatorname{Scalar}_{g_{j}}ud{\rm vol}_{g_{j}} =\displaystyle= 𝕊2(02π(2fju2Δfju)𝑑φ)𝑑volg𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}\left(\int^{2\pi}_{0}\left(2f_{j}u-2\Delta f_{j}u\right)d\varphi\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}
(304) =\displaystyle= 𝕊2(2fju¯2Δfju¯)𝑑volg𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}\left(2f_{j}\bar{u}-2\Delta f_{j}\bar{u}\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}
(305) =\displaystyle= 𝕊2(2fju¯+2fj,u¯)𝑑volg𝕊2,\displaystyle\int_{{\mathbb{S}}^{2}}\left(2f_{j}\bar{u}+2\langle\nabla f_{j},\nabla\bar{u}\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}},

where u¯(r,θ)=02πu(r,θ,φ)𝑑φ\bar{u}(r,\theta)=\int^{2\pi}_{0}u(r,\theta,\varphi)d\varphi. Then, by using the nonnegative scalar curvature condition Scalargj0\operatorname{Scalar}_{g_{j}}\geq 0, Proposition 3.5 and Lemma 5.19, possibly after passing to a subsequence, we obtain for any nonnegative test function 0uC(𝕊2×𝕊1)0\leq u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}),

(306) 0\displaystyle 0 \displaystyle\leq 𝕊2×𝕊1Scalargjudvolgj\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\operatorname{Scalar}_{g_{j}}ud{\rm vol}_{g_{j}}
(307) =\displaystyle= 𝕊2(2fju¯+2fj,u¯)𝑑volg𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}\left(2f_{j}\bar{u}+2\langle\nabla f_{j},\nabla\bar{u}\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}
(308) \displaystyle\rightarrow 𝕊2(2fu¯+2f,u¯)𝑑volg𝕊2\displaystyle\int_{{\mathbb{S}}^{2}}\left(2f_{\infty}\bar{u}+2\langle\nabla f_{\infty},\nabla\bar{u}\rangle\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}
(309) =\displaystyle= Scalarg,u.\displaystyle\langle\operatorname{Scalar}_{g_{\infty}},u\rangle.

Thus, Scalarg,u0\langle\operatorname{Scalar}_{g_{\infty}},u\rangle\geq 0 for all nonnegative test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}). By setting u1u\equiv 1 in equations (306)-(309), we obtain the convergence of distributional total scalar curvature. ∎

Appendix A W1,2W^{1,2} convergence in 𝕊1×h𝕊2{\mathbb{S}}^{1}\times_{h}{\mathbb{S}}^{2} case

In this appendix, we will derive W1,2W^{1,2} 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 {𝕊1×hj𝕊2}j=1\{{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}\}^{\infty}_{j=1} be a family of warped Riemannian manifolds with metric tensors as in (8) satisfying

(310) Scalarj0,Diam(𝕊1×hj𝕊2)D,\operatorname{Scalar}_{j}\geq 0,\quad\operatorname{Diam}({\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2})\leq D,

and

(311) MinA(𝕊1×hj𝕊2)A>0\operatorname{MinA}({\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2})\geq A>0

for all jj\in\mathbb{N}, where Scalarj\operatorname{Scalar}_{j} is the scalar curvature of 𝕊1×hj𝕊2{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}. Then there is a subsequence of warping functions hjh_{j} that converges in W1,2(𝕊1)W^{1,2}({\mathbb{S}}^{1}) to a Lipschitz function hW1,2(𝕊1)h_{\infty}\in W^{1,2}({\mathbb{S}}^{1}), which has Lipschitz constant 1 and satisfies

(312) A4πhDπ+2π,on𝕊1.\sqrt{\frac{A}{4\pi}}\leq h_{\infty}\leq\frac{D}{\pi}+2\pi,\quad\text{on}\ \ {\mathbb{S}}^{1}.

Moreover, let g:=g𝕊1+h2g𝕊2g_{\infty}:=g_{{\mathbb{S}}^{1}}+h^{2}_{\infty}g_{{\mathbb{S}}^{2}}, then gg_{\infty} is a Lipschitz continuous Riemannian metric tensor on 𝕊1×𝕊2{\mathbb{S}}^{1}\times{\mathbb{S}}^{2}, and a subsequence of {gj=g𝕊1+hj2g𝕊2}j=1\{g_{j}=g_{{\mathbb{S}}^{1}}+h^{2}_{j}g_{{\mathbb{S}}^{2}}\}^{\infty}_{j=1} converges in W1,2(𝕊1×𝕊2,g0)W^{1,2}({\mathbb{S}}^{1}\times{\mathbb{S}}^{2},g_{0}) to gg_{\infty}.

Here, as before, we still use g0=g𝕊1+g𝕊2g_{0}=g_{{\mathbb{S}}^{1}}+g_{{\mathbb{S}}^{2}} as a background metric. Then we can compute the scalar curvature distribution of Lee-LeFloch and have the following property.

Theorem A.2.

The limit metric gg_{\infty} obtained in Theorem A.1 has nonnegative distributional scalar curvature in the sense of Lee-LeFloch as recalled in Definition 5.7.

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 MinA\operatorname{MinA} condition may not be able to prevent collapsing happening near two poles [Lemma 4.3 in [15]], however, in the case of 𝕊1×hj𝕊2{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}, MinA\operatorname{MinA} condition can provide a positive uniform lower bound for hjh_{j} [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 hjh_{j}, we obtain a uniform upper bound from uniform upper bounded diameter condition [Lemma A.3] and a uniform lower bound from MinA\operatorname{MinA} 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 {𝕊1×hj𝕊2}j=1\{{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}\}^{\infty}_{j=1} be a family of warped product Riemannian manifolds with metric tensors as in (8), having uniformly upper bounded diameters, i.e. Diam(𝕊1×hj𝕊2)D\operatorname{Diam}({\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2})\leq D, then we have min𝕊1{hj}Dπ.\min\limits_{{\mathbb{S}}^{1}}\{h_{j}\}\leq\frac{D}{\pi}.

Proof.

Let s0𝕊1s_{0}\in\mathbb{S}^{1} be the minimum point of the function hjh_{j}. Then clearly the distance between antipodal points on the sphere {s0}×𝕊2Mj\{s_{0}\}\times\mathbb{S}^{2}\subset M_{j} is πmin𝕊1{hj}\pi\cdot\min\limits_{{\mathbb{S}}^{1}}\{h_{j}\}. So we have πmin𝕊1{hj}Diam(Mj)D\pi\cdot\min\limits_{{\mathbb{S}}^{1}}\{h_{j}\}\leq\operatorname{Diam}(M_{j})\leq D, and the claim follows. ∎

Lemma A.4.

Let {𝕊1×hj𝕊2}j=1\{{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}\}^{\infty}_{j=1} be a family of warped product Riemannian manifolds with metric tensors as in (8). The scalar curvature of the warped product metric gj=g𝕊1+hj2g𝕊2g_{j}=g_{{\mathbb{S}}^{1}}+h^{2}_{j}g_{{\mathbb{S}}^{2}} is given by

(313) Scalarj=4Δhjhj+21|hj|2hj2.\operatorname{Scalar}_{j}=-4\frac{\Delta h_{j}}{h_{j}}+2\frac{1-|\nabla h_{j}|^{2}}{h^{2}_{j}}.

Here the Laplace is the trace of the Hessian.

Moreover, if Scalarj0\operatorname{Scalar}_{j}\geq 0, then we have |hj|1|\nabla h_{j}|\leq 1 on 𝕊1\mathbb{S}^{1}.

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 Scalarj\operatorname{Scalar}_{j} of 𝕊1×hj𝕊2{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2} is given as in (313).

Now we prove the second claim by contradiction. Assume for some jj, |hj|>1|\nabla h_{j}|>1 at some point, let’s say p𝕊1p\in\mathbb{S}^{1}. Take a unit vector field XX on 𝕊1\mathbb{S}^{1} such that XX is in the same direction as hj\nabla h_{j} at the point pp. Let qq be the first point such that |h|(q)=1|\nabla h|(q)=1 while moving from the point pp on 𝕊1\mathbb{S}^{1} in the opposite direction of the unit vector field XX. Then let γ\gamma be the integral curve of the vector field XX with the initial point γ(0)=q\gamma(0)=q. Let t1>0t_{1}>0 such that γ(t1)=p\gamma(t_{1})=p. Set h~j(t)=hjγ(t)\tilde{h}_{j}(t)=h_{j}\circ\gamma(t). Then (at least) for t[0,t1]t\in[0,t_{1}],

(314) h~j(t)=hj,γ(t)=hj,Xγ(t)=|hj|γ(t),\tilde{h}^{\prime}_{j}(t)=\langle\nabla h_{j},\gamma^{\prime}(t)\rangle=\langle\nabla h_{j},X\rangle\circ\gamma(t)=|\nabla h_{j}|\circ\gamma(t),

and

(315) h~j′′(t)=(Δhj)γ(t).\tilde{h}^{\prime\prime}_{j}(t)=(\Delta h_{j})\circ\gamma(t).

By the Mean Value Theorem, there exists t2(0,t1)t_{2}\in(0,t_{1}) such that

(316) h~j′′(t2)=h~j(t1)h~j(0)t1>0,\tilde{h}^{\prime\prime}_{j}(t_{2})=\frac{\tilde{h}^{\prime}_{j}(t_{1})-\tilde{h}^{\prime}_{j}(0)}{t_{1}}>0,

since h~j(t1)=|hj|(p)>1\tilde{h}^{\prime}_{j}(t_{1})=|\nabla h_{j}|(p)>1 and h~j(0)=|hj|(q)=1\tilde{h}^{\prime}_{j}(0)=|\nabla h_{j}|(q)=1.

On the other hand, because Scalarj0\operatorname{Scalar}_{j}\geq 0, by using the scalar curvature (313)(\ref{scalar-curvature-formula-appendix}), one has

(317) 4h~j′′(t2)h~j(t2)+21(h~j(t2))2(h~j(t2))20-4\frac{\tilde{h}^{\prime\prime}_{j}(t_{2})}{\tilde{h}_{j}(t_{2})}+2\frac{1-(\tilde{h}_{j}(t_{2}))^{2}}{(\tilde{h}_{j}(t_{2}))^{2}}\geq 0

So

(318) h~j′′(t2)1(h~(t2))22h~(t2)<0,\tilde{h}^{\prime\prime}_{j}(t_{2})\leq\frac{1-(\tilde{h}^{\prime}(t_{2}))^{2}}{2\tilde{h}(t_{2})}<0,

since h~j(t2)>1\tilde{h}^{\prime}_{j}(t_{2})>1 by the choice of q=γ(0)q=\gamma(0). This produces a contradiction, and so |hj|1|\nabla h_{j}|\leq 1 on 𝕊1\mathbb{S}^{1}.

Lemma A.5.

Let {𝕊1×hj𝕊2}j=1\{{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}\}^{\infty}_{j=1} be a family of warped product Riemannian manifolds with metric tensors as in (8). If hj(x0)=0\nabla h_{j}(x_{0})=0 for some x0𝕊1x_{0}\in\mathbb{S}^{1} then there is a minimal surface {x0}×𝕊2\{x_{0}\}\times\mathbb{S}^{2} in 𝕊1×hj𝕊2{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}.

Proof.

Define Σx:={x}×𝕊2\Sigma_{x}:=\{x\}\times\mathbb{S}^{2}. Then for all x𝕊1x\in\mathbb{S}^{1}, Σx\Sigma_{x} is an embedded submanifold with mean curvature

(319) Hj=2|hj|(x)hj(x).H_{j}=\frac{2|\nabla h_{j}|(x)}{h_{j}(x)}.

Lemma A.6.

Let {𝕊1×hj𝕊2}j=1\{{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2}\}^{\infty}_{j=1} be a family of warped product Riemannian manifolds with metric tensors as in (8) satisfying MinA(𝕊1×hj𝕊2)A>0\operatorname{MinA}({\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2})\geq A>0. Then we have min𝕊1{hj}A4π>0\min\limits_{{\mathbb{S}}^{1}}\{h_{j}\}\geq\sqrt{\frac{A}{4\pi}}>0.

Proof.

By applying Lemma A.5, we have that there exists a minimal surface Σx0=x0×𝕊2\Sigma_{x_{0}}={x_{0}}\times{\mathbb{S}}^{2} on 𝕊1×hj𝕊2{\mathbb{S}}^{1}\times_{h_{j}}{\mathbb{S}}^{2} at the minimal value point x0x_{0} of hjh_{j}. The area of Σx0\Sigma_{x_{0}} is given by

(320) Area(Σ0)=4πhj2(x0).{\rm Area}(\Sigma_{0})=4\pi h^{2}_{j}(x_{0}).

Thus by the MinA\operatorname{MinA} condition, 4πhj2(x0)A4\pi h^{2}_{j}(x_{0})\geq A, 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) max𝕊1{hj}Dπ+2π,iN.\max_{{\mathbb{S}}^{1}}\{h_{j}\}\leq\frac{D}{\pi}+2\pi,\quad\forall i\in N.

By combining this uniform upper bound with the uniform lower bound obtained in Lemma A.6, we have that the warping functions hjh_{j} are uniformly bounded, i.e.

(322) A4πhjDπ+2πon𝕊1,j.\sqrt{\frac{A}{4\pi}}\leq h_{j}\leq\frac{D}{\pi}+2\pi\quad\text{on}\ \ {\mathbb{S}}^{1},\quad\forall j\in\mathbb{N}.

Moreover, Lemma A.4 implies function hjh_{j} are equicontinuous. Thus by applying Arzelà-Ascoli theorem we obtain that hjh_{j} are uniformly convergent a continuous function ff_{\infty} satisfying

(323) A4πhDπ+2π,on𝕊1.\sqrt{\frac{A}{4\pi}}\leq h_{\infty}\leq\frac{D}{\pi}+2\pi,\quad\text{on}\ \ {\mathbb{S}}^{1}.

Meanwhile, the uniform gradient estimate obtained in Lemma A.4 also implies that the limit function hh_{\infty} is Lipschitz with Lipschitz constant 11. Because a Lipschitz function is W1,W^{1,\infty}, we actually have hW1,(𝕊1)h_{\infty}\in W^{1,\infty}({\mathbb{S}}^{1}).

Step 2. W1,2W^{1,2} convergence of warping functions. We will estimate the bounded variation norm hjBV(𝕊1)\|\nabla h_{j}\|_{BV({\mathbb{S}}^{1})} of warping functions. First note that

(324) 0=𝕊1Δhj={Δhj0}Δhj+{Δhj<0}Δhj.0=\int_{{\mathbb{S}}^{1}}\Delta h_{j}=\int_{\{\Delta h_{j}\geq 0\}}\Delta h_{j}+\int_{\{\Delta h_{j}<0\}}\Delta h_{j}.

Thus,

(325) {Δhj<0}Δhj={Δhj0}Δhj,-\int_{\{\Delta h_{j}<0\}}\Delta h_{j}=\int_{\{\Delta h_{j}\geq 0\}}\Delta h_{j},

furthermore,

(326) hjBV(𝕊1)\displaystyle\|\nabla h_{j}\|_{BV({\mathbb{S}}^{1})} =\displaystyle= 𝕊1|hj|+𝕊1|Δhj|\displaystyle\int_{{\mathbb{S}}^{1}}|\nabla h_{j}|+\int_{{\mathbb{S}}^{1}}|\Delta h_{j}|
(327) =\displaystyle= 𝕊1|hj|+{Δhj0}Δhj{Δhj<0}Δhj\displaystyle\int_{{\mathbb{S}}^{1}}|\nabla h_{j}|+\int_{\{\Delta h_{j}\geq 0\}}\Delta h_{j}-\int_{\{\Delta h_{j}<0\}}\Delta h_{j}
(328) =\displaystyle= 𝕊1|hj|+2{Δhj0}Δhj.\displaystyle\int_{{\mathbb{S}}^{1}}|\nabla h_{j}|+2\int_{\{\Delta h_{j}\geq 0\}}\Delta h_{j}.

Then by the expression of the scalar curvature in Lemma A.4, the nonnegative scalar curvature condition implies

(329) Δhj1|hj|22hj12hjπA,j.\Delta h_{j}\leq\frac{1-|\nabla h_{j}|^{2}}{2h_{j}}\leq\frac{1}{2h_{j}}\leq\sqrt{\frac{\pi}{A}},\quad\forall j\in\mathbb{N}.

The last inequality here follows from Lemma A.6. Lemma A.4 also tells us that |hj|1|\nabla h_{j}|\leq 1 on 𝕊1{\mathbb{S}}^{1} for all jj\in\mathbb{N}. Consequently, we have

(330) hjBV(𝕊1)\displaystyle\|\nabla h_{j}\|_{BV({\mathbb{S}}^{1})} =\displaystyle= 𝕊1|hj|+2{Δhj0}Δhj\displaystyle\int_{{\mathbb{S}}^{1}}|\nabla h_{j}|+2\int_{\{\Delta h_{j}\geq 0\}}\Delta h_{j}
(331) \displaystyle\leq 2π+2Δhj0πA\displaystyle 2\pi+2\int_{\Delta h_{j}\geq 0}\sqrt{\frac{\pi}{A}}
(332) \displaystyle\leq 2π(1+2πA),j.\displaystyle 2\pi\left(1+2\sqrt{\frac{\pi}{A}}\right),\quad\forall j\in\mathbb{N}.

As a result, by Theorem 5.5 in [5] we have that a subsequence, which is still denoted by hj\nabla h_{j}, converges to some ϕBV(𝕊1)\phi\in BV({\mathbb{S}}^{1}) in L1(𝕊1)L^{1}({\mathbb{S}}^{1}) norm, and it is easy to see that ϕ=h\phi=\nabla h_{\infty} in the weak sense. Moreover, since hW1,(𝕊1)h_{\infty}\in W^{1,\infty}({\mathbb{S}}^{1}) and supjhjL(𝕊1)<\sup\limits_{j}\|\nabla h_{j}\|_{L^{\infty}({\mathbb{S}}^{1})}<\infty, we have hjh\nabla h_{j}\rightarrow\nabla h_{\infty} in L2(𝕊1)L^{2}({\mathbb{S}}^{1}) norm. Indeed, note that by the Hölder inequality,

(333) 𝕊1|hjh|2hjhL1(𝕊1)hjhL(𝕊1).\int_{{\mathbb{S}}^{1}}|\nabla h_{j}-\nabla h_{\infty}|^{2}\leq\|\nabla h_{j}-\nabla h_{\infty}\|_{L^{1}({\mathbb{S}}^{1})}\|\nabla h_{j}-\nabla h_{\infty}\|_{L^{\infty}({\mathbb{S}}^{1})}.

As a result, hjhh_{j}\rightarrow h_{\infty} in W1,2(𝕊1)W^{1,2}({\mathbb{S}}^{1}).

Step 3. W1,2W^{1,2} convergence of metrics. Note that

(334) gjg=(hj2h2)g𝕊2,g_{j}-g_{\infty}=(h^{2}_{j}-h^{2}_{\infty})g_{{\mathbb{S}}^{2}},

and

(335) ¯(gjg)=2(hj¯hjh¯h)g𝕊2.\overline{\nabla}(g_{j}-g_{\infty})=2(h_{j}\overline{\nabla}h_{j}-h_{\infty}\overline{\nabla}h_{\infty})\otimes g_{{\mathbb{S}}^{2}}.

Therefore, by applying the uniform bound supjhjL(𝕊1)<\sup\limits_{j}\|\nabla h_{j}\|_{L^{\infty}({\mathbb{S}}^{1})}<\infty, and W1,2W^{1,2} convergence of hjh_{j} to hh_{\infty}, we can obtain that gj=g𝕊1+hj2g𝕊2g_{j}=g_{{\mathbb{S}}^{1}}+h^{2}_{j}g_{{\mathbb{S}}^{2}} converges to gg_{\infty} in W1,2(𝕊1×𝕊2,g0)W^{1,2}({\mathbb{S}}^{1}\times{\mathbb{S}}^{2},g_{0}). ∎

A.2. Nonnegative distributional scalar curvature of the limit metric

In this subsection, we compute the distributional scalar curvature of the limit metric tensor gg_{\infty} obtained in Theorem A.1 with the background metric g0g_{0} in the sense of Lee-LeFloch, and prove Theorem A.2. Throughout this subsection, gg_{\infty} always denotes the limit metric obtained in Theorem A.1.

By the definition of Γijk\Gamma^{k}_{ij} 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, gg_{\infty}, with the background metric, g0g_{0}, the Christoffel symbols defined by Lee-LeFloch as in (244), in the coordinate {φ,r,θ}\{\varphi,r,\theta\}, all vanish except

(336) Γrrφ=hh,Γθθφ=hhsin2r,\Gamma^{\varphi}_{rr}=-h_{\infty}h_{\infty}^{\prime},\quad\Gamma^{\varphi}_{\theta\theta}=-h_{\infty}h^{\prime}_{\infty}\sin^{2}r,
(337) Γφrr=Γrφr=hh,\Gamma^{r}_{\varphi r}=\Gamma^{r}_{r\varphi}=\frac{h^{\prime}_{\infty}}{h_{\infty}},

and

(338) Γφθθ=Γθφθ=hh.\Gamma^{\theta}_{\varphi\theta}=\Gamma^{\theta}_{\theta\varphi}=\frac{h^{\prime}_{\infty}}{h_{\infty}}.

Note also that

Lemma A.8.

Note that the volume forms are:

(339) dμ0=dφdrsin(r)dθ,d\mu_{0}=\,d\varphi\,\wedge dr\wedge\sin(r)\,d\theta,

and

(340) dμ=dφh2drsin(r)dθ,d\mu_{\infty}=d\varphi\wedge h^{2}_{\infty}dr\wedge\sin(r)\,d\theta,

which are both defined everywhere away from r=0r=0 and r=πr=\pi. In particular,

(341) dμdμ0=h2(φ)\frac{\,d\mu_{\infty}}{\,d\mu_{0}}=h^{2}_{\infty}(\varphi)

is in W1,p(𝕊2×𝕊1,g0)W^{1,p}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}) for all p1p\geq 1.

Proof.

The first claim holds away from r=0r=0 and r=πr=\pi by the definition of volume form, and the second claim holds almost everywhere on (𝕊2×𝕊1,g0)({\mathbb{S}}^{2}\times{\mathbb{S}}^{1},g_{0}). So dμ=fdμ0d\mu_{\infty}=f_{\infty}d\mu_{0} almost everywhere which gives us the third claim. The rest follows from Proposition 3.5. ∎

Now we are ready to compute the vector field VV and the function FF defined by Lee-LeFloch as in (243) and (245)(\ref{defn-F}).

Lemma A.9.

For the limit metric gg_{\infty} with the background metric g0g_{0}, the vector field VV defined in (243), in the local frame {φ,r,θ}\{\partial_{\varphi},\partial_{r},\partial_{\theta}\}, is given by

(342) V=(4hh,0,0)=4hhφ.V=\left(-4\frac{h^{\prime}_{\infty}}{h_{\infty}},0,0\right)=-4\frac{h^{\prime}_{\infty}}{h_{\infty}}\frac{\partial}{\partial\varphi}.

Furthermore

(343) V¯(udμdμ0)=4hhφ(uh2).-V\cdot\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right)=4\frac{h^{\prime}_{\infty}}{h_{\infty}}\partial_{\varphi}(uh^{2}_{\infty}).
Lemma A.10.

For the limit metric gg_{\infty} with the background metric g0g_{0}, the function FF defined in (245) is given by

(344) F=2h26(hh)2.F=\frac{2}{h^{2}_{\infty}}-6\left(\frac{h^{\prime}_{\infty}}{h_{\infty}}\right)^{2}.

Furthermore,

(345) (Fudμdμ0)=2u6u(h)2.\left(Fu\frac{\,d\mu_{\infty}}{d\mu_{0}}\right)=2u-6u(h^{\prime}_{\infty})^{2}.
Lemma A.11.

For gg being our limit metric tensor gg_{\infty} and a smooth nonnegative test function uu, the integrals in (248) and (249) are given by

(346) FirstIntg\displaystyle\quad FirstInt_{g_{\infty}} =\displaystyle= 𝕊2×𝕊1(V¯(udμdμ0))𝑑μ0\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\left(-V\cdot\overline{\nabla}\left(u\frac{\,d\mu_{\infty}}{\,d\mu_{0}}\right)\right)\,d\mu_{0}
(347) =\displaystyle= 𝕊1(8(h)2u¯+4hhu¯)𝑑φ,\displaystyle\int_{{\mathbb{S}}^{1}}\left(8(h^{\prime}_{\infty})^{2}\bar{u}+4h^{\prime}_{\infty}h_{\infty}\bar{u}\right)d\varphi,

and

(348) SecondIntg\displaystyle\quad SecondInt_{g_{\infty}} =\displaystyle= 𝕊2×𝕊1(Fudμdμ0)𝑑μ0\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\left(Fu\frac{\,d\mu_{\infty}}{d\mu_{0}}\right)\,d\mu_{0}
(349) =\displaystyle= 𝕊1(2u¯6(h)2u¯)𝑑φ,\displaystyle\int_{{\mathbb{S}}^{1}}\left(2\bar{u}-6(h^{\prime}_{\infty})^{2}\bar{u}\right)d\varphi,

where

(350) u¯(φ)=0π𝑑r02πu(r,θ,φ)𝑑θ.\bar{u}(\varphi)=\int^{\pi}_{0}dr\int^{2\pi}_{0}u(r,\theta,\varphi)d\theta.
Proof.

By integrating the formulas in Lemma A.9 and Lemma A.10, one can easily obtain the integrals in (347) and (349). ∎

Remark A.12.

Here W1,2W^{1,2} regularity of hh_{\infty} implies that the integrals in (347) and (347) are both finite (c.f. Remarks 5.10 and 5.18).

Lemma A.13.

For the limit metric g=g𝕊1+h2g𝕊2g_{\infty}=g_{{\mathbb{S}}^{1}}+h^{2}_{\infty}g_{{\mathbb{S}}^{2}}, the scalar curvature distribution Scalarg\operatorname{Scalar}_{g_{\infty}} defined in Definition 5.7 can be expressed, for every test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}), as the integral

(351) Scalarg,u=𝕊1(2u¯+2(h)2u¯+4hhu¯)𝑑φ,\langle\operatorname{Scalar}_{g_{\infty}},u\rangle=\int_{{\mathbb{S}}^{1}}\left(2\bar{u}+2(h^{\prime}_{\infty})^{2}\bar{u}+4h^{\prime}_{\infty}h_{\infty}\bar{u}\right)d\varphi,

and this is finite for any test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}). Here u¯\bar{u} is defined as in (350).

Proof.

The expression in (351) immediately follows from the expressions in (347) and (349) and Definition 5.7. The finiteness of the integral in (351) follows from that hW1,2(𝕊2)h_{\infty}\in W^{1,2}({\mathbb{S}}^{2}). ∎

We now apply these lemmas to prove Theorem A.2:

Proof.

By the expression (313) of the scalar curvature of 𝕊1×hi𝕊2{\mathbb{S}}^{1}\times_{h_{i}}{\mathbb{S}}^{2}, we have that for any test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}),

(352) 𝕊1×𝕊2Scalargjudvolgj\displaystyle\int_{{\mathbb{S}}^{1}\times{\mathbb{S}}^{2}}\operatorname{Scalar}_{g_{j}}ud{\rm vol}_{g_{j}}
(353) =\displaystyle= 𝕊1(𝕊2(4(Δhj)hju+2u2|hj|2u)𝑑volg𝕊2)𝑑φ\displaystyle\int_{{\mathbb{S}}^{1}}\left(\int_{{\mathbb{S}}^{2}}\left(-4(\Delta h_{j})h_{j}u+2u-2|\nabla h_{j}|^{2}u\right)d{\rm vol}_{g_{{\mathbb{S}}^{2}}}\right)d\varphi
(354) =\displaystyle= 𝕊2(2u¯+2(hj)2u¯+4hjhju¯)𝑑φ.\displaystyle\int_{{\mathbb{S}}^{2}}\left(2\bar{u}+2(h^{\prime}_{j})^{2}\bar{u}+4h^{\prime}_{j}h_{j}\bar{u}\right)d\varphi.

Then, by using the nonnegative scalar curvature condition Scalargj0\operatorname{Scalar}_{g_{j}}\geq 0, and convergence property of hjh_{j} in Theorem A.1, possibly after passing to a subsequence, we obtain for any nonnegative test function 0uC(𝕊2×𝕊1)0\leq u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}),

(356) 0\displaystyle 0 \displaystyle\leq 𝕊2×𝕊1Scalargjudvolgj\displaystyle\int_{{\mathbb{S}}^{2}\times{\mathbb{S}}^{1}}\operatorname{Scalar}_{g_{j}}ud{\rm vol}_{g_{j}}
(357) =\displaystyle= 𝕊1(2u¯+2(hj)2u¯+4hjhju¯)𝑑φ\displaystyle\int_{{\mathbb{S}}^{1}}\left(2\bar{u}+2(h^{\prime}_{j})^{2}\bar{u}+4h^{\prime}_{j}h_{j}\bar{u}\right)d\varphi
(358) \displaystyle\rightarrow 𝕊2(2u¯+2(h)2u¯+4hhu¯)𝑑φ\displaystyle\int_{{\mathbb{S}}^{2}}\left(2\bar{u}+2(h^{\prime}_{\infty})^{2}\bar{u}+4h^{\prime}_{\infty}h_{\infty}\bar{u}\right)d\varphi
(359) =\displaystyle= Scalarg,u.\displaystyle\langle\operatorname{Scalar}_{g_{\infty}},u\rangle.

Thus, Scalarg,u0\langle\operatorname{Scalar}_{g_{\infty}},u\rangle\geq 0 for all nonnegative test function uC(𝕊2×𝕊1)u\in C^{\infty}({\mathbb{S}}^{2}\times{\mathbb{S}}^{1}). ∎

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 C0{C}^{0} 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.