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

[1]\fnmShota \surHamanaka

[1]\orgdivDepartment of Mathematics, \orgnameGraduate School of Science, Osaka University, \orgaddress\street1-1 Machikaneyama-cho, \cityToyonaka, \postcode560-0043, \stateOsaka, \countryJapan

Limit theorems for the total scalar curvature

Abstract

We prove that the lower bound of the total scalar curvatures on a closed nn-manifold is preserved under the W1,p(p>n)W^{1,p}~{}(p>n) convergence of the Riemannian metrics provided that each scalar curvature is nonnegative. We also discuss certain weighted version of this type of theorem.

keywords:
Scalar curvature, Ricci flow, Heat flow, Weak notions of the scalar curvature lower bound
pacs:
[

MSC Classification]53C21, 53E20

1 Introduction

Gromov [8] proved the following “C0C^{0}-limit theorem”.

Theorem 1.1 ([8, p.1118] and [2]).

Let MM be a (possibly open) smooth manifold and κ:M\kappa:M\rightarrow\mathbb{R} a continuous function. Consider a sequence of C2C^{2}-Riemannian metrics gig_{i} on MM that converges to a C2C^{2}-Riemannian metric gg in the local C0C^{0}-sense. Assume that for all i=1,2,i=1,2,\cdots the scalar curvature R(gi)R(g_{i}) of gig_{i} satisfies R(gi)κR(g_{i})\geq\kappa everywhere on M.M. Then R(g)κR(g)\geq\kappa everywhere on M.M.

In contrast, let MM be the same as in the above theorem, for given a continuous function σ:M,\sigma:M\rightarrow\mathbb{R}, the set {g|R(g)σonM}\{g\in\mathcal{M}|~{}R(g)\leq\sigma~{}\mathrm{on}~{}M\} is C0C^{0}-dense in the set \mathcal{M} of all smooth Riemannian metrics on MM ([15, Theorem B]). On the other hand, in our forthcoming paper (see the preprint by the author, arXiv:2301.05444v5), we will show that in a fixed conformal class, the upper bound of the total scalar curvature is preserved under C0C^{0}-convergence of metric tensors. Gromov [8] proved the above theorem by using a gluing technique and the resolution of Geroch’s conjecture. Later, Bamler [2] gave an alternative proof of this theorem using the Ricci–DeTurck flow. On the other hand, Lee–Topping recently proved (in their preprint, arXiv:2203.01223v2) that non-negativity of scalar curvature is not preserved in dimension at least four under the topology of uniform convergence of Riemannian distance. For other studies about the behavior of the scalar curvature lower bound under various weak topologies, see, for example, [5, 9, 11].

As we will see below, we can observe that the point-wise version of the Gromov’s theorem (Theorem 1.1) is false in general. That is, we can easily construct an example of C2C^{2}-Riemannian metrics (gi)(g_{i}) on a smooth manifold MM which satisfies the following: gig_{i} converges to a C2C^{2}-Riemannian metric gg on MM in the C0C^{0}-sense as i.i\rightarrow\infty. And there is a point pMp\in M such that for some κ,\kappa\in\mathbb{R},

R(gi)(p)κforalli,R(g_{i})(p)\geq\kappa~{}~{}\mathrm{for~{}all}~{}i\in\mathbb{N},

but

R(g)(p)<κ.R(g)(p)<\kappa.

Indeed, Lohkamp gave an example in [4, Lecture Series 2, Counterexample 2.3.2].

Example 1.1 ([4, Lecture Series 2, Counterexample 2.3.2]. cf. Example 4.2 and 4.3 in this paper).

For each i,i\in\mathbb{N}, we define a smooth metric on n\mathbb{R}^{n} as

gi:={exp(2fi)gEuclonDα,i:={xn||x|gEucl2α/i},gEuclinnD2α,i.g_{i}:=\begin{cases}\exp{\left(2f_{i}\right)}\cdot g_{Eucl}&\mathrm{on}~{}D_{\alpha,i}:=\{x\in\mathbb{R}^{n}~{}|~{}|x|_{g_{Eucl}}^{2}\leq\alpha/i\},\\ g_{Eucl}&\mathrm{in}~{}\mathbb{R}^{n}\setminus D_{2\alpha,i}.\end{cases}

Here, gEuclg_{Eucl} denotes the Euclidean metric on n,\mathbb{R}^{n}, |x|gEucl|x|_{g_{Eucl}} is the Euclidean distance from xx to the origin o,o, the smooth function

fi:=αix12x22xn2f_{i}:=\frac{\alpha}{i}-x_{1}^{2}-x_{2}^{2}-\cdots-x_{n}^{2}

whose support is contained in D2α,i,D_{2\alpha,i}, and α\alpha\in\mathbb{R} is a positive constant. Then, using the following fact, we have R(gi)(o)=C(n)>0.R(g_{i})(o)=C(n)>0. Moreover, gig_{i} converges to gEuclg_{Eucl} in the uniformly C0C^{0}-sense on n.\mathbb{R}^{n}.

Fact 1.1 (Conformal change of the scalar curvature).

For a Riemannian metric gg and a C2C^{2}-function ϕ,\phi, set g¯:=e2ϕg,\bar{g}:=e^{2\phi}g, then

R(g¯)=e2ϕR(g)2(n1)e2ϕΔgϕ(n2)(n1)e2ϕ|dϕ|g2.R(\bar{g})=e^{-2\phi}R(g)-2(n-1)~{}e^{-2\phi}\Delta_{g}\phi-(n-2)(n-1)~{}e^{-2\phi}~{}|d\phi|_{g}^{2}.

The proof of this fact is a straightforward calculation. Note that the scalar curvature lower bounds are guaranteed only at the origin oo in this example. But we can never take a metric whose scalar curvature is bounded from below by some positive constant on a small neighborhood of oo and nonnegative on the whole manifold n,\mathbb{R}^{n}, and which is equal to the Euclidean metric outside a compact subset of n.\mathbb{R}^{n}. Indeed, if such a metric exists, then we can construct a metric on the nn-dimensional torus whose scalar curvature is nonnegative everywhere and positive somewhere. But, this is impossible by the resolution of Geroch’s conjecture (see [7, 16, 17]). Of course, we can apply Theorem 1.1 to this sequence (gi)(g_{i}) and M=n.M=\mathbb{R}^{n}. But, we can only take the lower bound κ\kappa such that supnκ0\sup_{\mathbb{R}^{n}}\kappa\leq 0 in this case because the support of fif_{i} shrinks as i.i\rightarrow\infty. Hence we can only obtain the trivial fact R(gEucl)=0κR(g_{Eucl})=0\geq\kappa even though we use Theorem 1.1.

In this paper we investigate some total scalar curvature versions of Theorem 1.1. More precisely, we will consider the following problem: Let MM be a (possibly non-compact) smooth manifold, {gi}i\{g_{i}\}_{i\in\mathbb{N}} a sequence of C2C^{2}-Riemannian metrics and gg a C2C^{2}-Riemannian metric on M.M. If gig_{i} converges to gg in some sense (with respect to gg) and

MR(gi)𝑑volgiκforalli\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa~{}~{}~{}~{}\mathrm{for~{}all}~{}~{}i

for some constant κ.\kappa\in\mathbb{R}. Here Rgi,dvolgiR_{g_{i}},\,d\mathrm{vol}_{g_{i}} denote respectively the scalar curvature of gig_{i} and the Riemannian volume measure of gi.g_{i}. Then, does

MR(g)𝑑volgκ\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa

hold? Or, more generally, let we consider that a sequence of C2C^{2} metrics gig_{i} on MM converging to a C2C^{2} metric in some sense and a sequence of functions fif_{i} on MM converging to a function ff in some sense that satisfy

MR(gi)efi𝑑volgiκforalli\int_{M}R(g_{i})e^{-f_{i}}\,d\mathrm{vol}_{g_{i}}\geq\kappa~{}~{}~{}~{}\mathrm{for~{}all}~{}~{}i

for some constant κ.\kappa\in\mathbb{R}. Then, we ask whether

MR(g)ef𝑑volgκ\int_{M}R(g)e^{-f}\,d\mathrm{vol}_{g}\geq\kappa

holds or not.

We emphasize that we will only consider the situation that metrics converge to some metric with respect to certain topology which is weaker than C2,C^{2}, each metric is at least C2,C^{2}, and the underlying manifolds we consider are assumed to be smooth.

If M2M^{2} is closed (i.e., compact without boundary) surface, from the Gauss-Bonnet theorem,

M2R(g)𝑑volg=4πχ(M)\int_{M^{2}}R(g)\,d\mathrm{vol}_{g}=4\pi\chi(M)

for each Riemannian metric gg on M.M. Here χ(M)\chi(M) denotes the Euler characteristic of M.M. Hence it is sufficient to consider the above problem (unweighted case) in dimension n3n\geq 3 and, unless otherwise mentioned, we will assume below that the dimension of manifolds are greater than or equal to three.

On the other hand, if M2nM^{2n} is a closed complex n(real2n)n~{}(\mathrm{real}~{}2n)-manifold (n1n\geq 1) and g,gi(i=1,2,)g,g_{i}~{}(i=1,2,\cdots) are Kähler metrics on M.M. Let ω\omega and ωi\omega_{i} be the Kähler forms of gg and gig_{i} respectively. Assume that gig_{i} converges to gg (hence ωi\omega_{i} converges to ω\omega) in the C0C^{0}-sense as i.i\rightarrow\infty. Then

MR(gi)𝑑volgi=4π(n1)!(c1(M)[ωin1](M))4π(n1)!(c1(M)[ωn1](M))=MR(g)𝑑volg(i),\begin{split}\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=\frac{4\pi}{(n-1)!}(c_{1}(M)\smallsmile[\omega_{i}^{n-1}](M))\\ &\longrightarrow\frac{4\pi}{(n-1)!}(c_{1}(M)\smallsmile[\omega^{n-1}](M))=\int_{M}R(g)\,d\mathrm{vol}_{g}~{}~{}~{}(i\rightarrow\infty),\end{split}

where c1(M)c_{1}(M) denotes the first Chern class of M.M. Note that we assumed here that the limiting metric gg is Kähler metric on MM as well, but, in our main theorem 3, we will not assume that the limiting metric gg is a Ricci soliton. Although it deviates a bit from our subject, the following interesting result about lower bounds of scalar curvature integrals is also known.

Theorem 1.2 ([4, Lecture Series 1, Theorem 4.1]).

Let MM be a compact nn-dimensional manifold (n3)(n\geq 3) carrying a hyperbolic metric g0.g_{0}. There is a neighborhood 𝒰\mathcal{U} of g0g_{0} in the space of all Riemannian metrics with the C2C^{2}-topology such that for any g𝒰,g\in\mathcal{U},

M(R(g))n/2𝑑volgM(R(g0))n/2𝑑volg0\int_{M}\left(R(g)_{-}\right)^{n/2}d\mathrm{vol}_{g}\geq\int_{M}\left(R(g_{0})_{-}\right)^{n/2}\,d\mathrm{vol}_{g_{0}}

and equality if and only if gg is isometric to g0.g_{0}. Here, R(g):=max{R(g),0}.R(g)_{-}:=\max\left\{-R(g),~{}0\right\}.

Our first main result in this paper is the following.

Main Theorem 1.

Let p>n.p>n. Let MnM^{n} be a closed manifold of dimension n3n\geq 3 and gg a C2C^{2} Riemannian metric on M.M. Assume that (gi)(g_{i}) is a sequence of C2C^{2} Riemannian metrics on MM such that gig_{i} converges to gg on MM in the W1,pW^{1,p}-sense as i,i\rightarrow\infty,

MR(gi)𝑑volgiκforsomeconstantκ\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa~{}~{}~{}\mathrm{for~{}some~{}constant}~{}\kappa\in\mathbb{R}

and R(gi)0R(g_{i})\geq 0 on MM for each i.i. Then

MR(g)𝑑volgκ.\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.
Remark 1.1.

From the assumption R(gi)0,R(g_{i})\geq 0, of course the only meaningful case is κ0.\kappa\geq 0. For the case of κ=0,\kappa=0, this claim follows from Theorem 1.1. However, Main Theorem 1 states that this statement still holds for all κ0.\kappa\geq 0.

Here, we said that a sequence of metrics (gi)i(g_{i})_{i} converges to a metric gg on MM in the W1,pW^{1,p}-sense if gig_{i} and all first weak derivatives of it respectively converge to those of gg with respect to the LpL^{p}-norm of g.g. (Since MM is compact, if (gi)(g_{i}) converges in the W1,pW^{1,p}-sense with respect to g,g, then it also converges W1,pW^{1,p}-sense with respect to any fixed reference metric on M.M.) Note that from Morrey’s embedding, there is a continuous embedding: W1,pC0,1npW^{1,p}\hookrightarrow C^{0,1-\frac{n}{p}} if p>n.p>n. Therefore the same statement of Main Theorem 1 still holds even though one replace W1,p(p>n)W^{1,p}~{}(p>n) with C0,αC^{0,\alpha} for all α(0,1].\alpha\in(0,1]. On the other hand, in Main Theorem 1, if we weaken the assumption from W1,pW^{1,p}-convergence to C0C^{0}-convergence, then the same statement (without the assumption that each gig_{i} has nonnegative scalar curvature) no longer holds true in general. Indeed, we will give an example (Example 4.3) on every closed Riemannian nn-manifold for n3.n\geq 3.

As a corollary of Main Theorem 1 and Theorem 1.1, we obtain the following.

Corollary 1.1.

Let p>n,p>n, and let \mathcal{M} be the space of all C2C^{2}-Riemannian metrics on a closed manifold M.M. For any nonnegative continuous function σC0(M,0)\sigma\in C^{0}(M,\mathbb{R}_{\geq 0}) and constant κ,\kappa\in\mathbb{R}, the subspace

{g|MR(g)𝑑volgκ,R(g)σonM}()\left\{g\in\mathcal{M}~{}\middle|~{}\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa,~{}R(g)\geq\sigma~{}\mathrm{on}~{}M\right\}~{}(\subset\mathcal{M})

is closed in \mathcal{M} with respect to the W1,pW^{1,p}-topology.

As our second main result in this paper, we will prove the following theorem in a more general setting.

Main Theorem 2.

Let p>n2/2.p>n^{2}/2. Suppose that MnM^{n} is a closed nn-manifold (n2n\geq 2), gg is a C2C^{2} Riemannian metric on M,M, and (gi)(g_{i}) is a sequence of C2C^{2} Riemannian metrics on MM such that gig_{i} converges to gg on MM in the W1,pW^{1,p}-sense as i.i\rightarrow\infty. Let (fi)(f_{i}) be a family of functions on MM and ff a function on M.M. Assume the following:

  • (1)

    There ia a positive constant Λ>0\Lambda>0 such that ff and fi(i0)f_{i}~{}(i\geq 0) are Λ\Lambda-Lipschitz functions on M,M,

  • (2)

    fiC0ff_{i}\overset{C^{0}}{\longrightarrow}f uniformly on M,M,

  • (3)

    R(gi)0R(g_{i})\geq 0 on MM for all i,i,

  • (4)

    MR(gi)efi𝑑volgiκ(κ).\int_{M}R(g_{i})\,e^{-f_{i}}d\mathrm{vol}_{g_{i}}\geq\kappa~{}~{}(\kappa\in\mathbb{R}).

Then

MR(g)ef𝑑volgκ.\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}\geq\kappa.

This is non-trivial even in the two-dimensional case because fif_{i} is non-constant in general, hence we cannot use the Gauss-Bonnet theorem. For example, (1)(1) and (2)(2) are automatically satisfied in case that efidvolgi=dvolg0=efdvolge^{-f_{i}}d\mathrm{vol}_{g_{i}}=d\mathrm{vol}_{g_{0}}=e^{-f}d\mathrm{vol}_{g} for some C0C^{0} metric g0g_{0}, and giC1g.g_{i}\overset{C^{1}}{\longrightarrow}g. Indeed, since each fif_{i} can be locally written as fi=log(detgi)log(detg0)f_{i}=\log(\sqrt{detg_{i}})-\log(\sqrt{detg_{0}}) (ff is also represented in the same form) and giC1g,g_{i}\overset{C^{1}}{\longrightarrow}g, so the norm of the first derivatives (|fi|g)i(|\nabla f_{i}|_{g})_{i} are uniformly bounded and fiff_{i}\rightarrow f uniformly on M.M. And, we speculate the assumptions R(gi)0(i=1,2,)R(g_{i})\geq 0~{}(i=1,2,\cdots) in Main Theorem 1 and 2 can be not needed. As a corollary of Main Theorem 2, we can obtain the following and from it, we can also define a new weak notion of scalar curvature lower bounds.

Corollary 1.2 (== Corollary 3.4).

Let p>n2/2p>n^{2}/2 and κ\kappa a constant. Suppose that MM is a closed manifold of dimension n2,n\geq 2, gg is a C2C^{2} Riemannian metric on MM and (gi)(g_{i}) is a sequence of C2C^{2} metric on M.M. Assume the following:

  • (1)

    a sequence (ϕi)(\phi_{i}) of nonnegative continuous functions on MM satisfying: for any positive constant a>0a>0 there is a positive constant Λ>0\Lambda>0 such that log(ϕi+a)\log(\phi_{i}+a) is Λ\Lambda-Lipschitz on MM for all i,i,

  • (2)

    (ϕi)(\phi_{i}) converges to some nonnegative continuous function ϕ\phi in the uniformly C0C^{0}-sense on M,M,

  • (3)

    R(gi)0R(g_{i})\geq 0 on MM for each i,i,

  • (4)

    MR(gi)ϕi𝑑volgiκMϕi𝑑volgi,\int_{M}R(g_{i})\phi_{i}\,d\mathrm{vol}_{g_{i}}\geq\kappa\int_{M}\phi_{i}\,d\mathrm{vol}_{g_{i}},

  • (5)

    gig_{i} converges to gg in the W1,pW^{1,p}-sense.

Then

MR(g)ϕ𝑑volgκMϕ𝑑volg.\int_{M}R(g)\phi\,d\mathrm{vol}_{g}\geq\kappa\int_{M}\phi\,d\mathrm{vol}_{g}.
Remark 1.2.

From (3),(5)(3),(5) and Remark 3.3, it is known that MR(g)ψ𝑑volg0\int_{M}R(g)\psi\,d\mathrm{vol}_{g}\geq 0 for all smooth nonnegative function ψ.\psi. Hence it is reasonable to consider case that κ0\kappa\geq 0 in this setting.

We give necessary notions and prove this corollary in Section 3. Based on this type of limit theorem, we can define a new generalized notion of scalar curvature lower bound via the existence of certain type of approximate sequences as follows.

Definition 1.1.

Let MnM^{n} be a smooth closed nn-manifold and κ\kappa a constant. For any W1,p(p>n2/2)W^{1,p}~{}(p>n^{2}/2) metric gg on M,M, gg is of R(g)κR(g)\geq\kappa in the approximate distributional sense if for any nonnegative continuous function ϕ\phi there is an approximate sequence (ϕi)(\phi_{i}) satisfying (1),(2)(1),(2) in Corollary 1.2, and there exists a W1,p(p>n2/2)W^{1,p}~{}(p>n^{2}/2)-approximate sequence of C2C^{2}-metrics (gi)(g_{i}) satisfying (3)(3)-(5)(5) in Corollary 1.2.

Suppose now that a metric gg in Definition 1.1 is actually C2.C^{2}. Then, from Corollary 1.2, R(g)κR(g)\geq\kappa in the approximate distributional sense on MM implies that the same bound R(g)κR(g)\geq\kappa holds in the distributional sense on M.M. As a result, R(g)κR(g)\geq\kappa in the conventional sense from Remark 3.3. Note that R(g)0R(g)\geq 0 in the conventional sense in this case due to (3)(3) of Corollary 1.2 and Theorem 1.1.

For example, if (M,gi)(M,g_{i}) is a complete gradient shrinking or steady Ricci soliton, then R(gi)0R(g_{i})\geq 0 on MM (see [6, Corollary 2.5], [22, Theorem 1.3]). Hence, from Main Theorem 1, if furthermore each total scalar curvature is bounded from below by some (nonnegative) constant, then such a lower bound is preserved under the W1,p(p>n)W^{1,p}~{}(p>n) convergence of metrics. On the other hand, if we assume that each metric is Ricci soliton (with certain additional assumption), then we can obtain a similar statement under weaker C0C^{0} convergence of metrics. Our third main theorem is the following.

Main Theorem 3.

Let MnM^{n} be a closed nn-manifold and gg a C2C^{2} Riemannian metric on M.M. Let (gi)(g_{i}) be a sequence of Ricci solitons on MM (i.e., 2Ric(gi)=Yigi2λigi-2\,\mathrm{Ric}(g_{i})=\mathcal{L}_{Y_{i}}g_{i}-2\lambda_{i}\,g_{i} for some constant λi\lambda_{i}\in\mathbb{R} and a vector field YiΓ(TM)Y_{i}\in\Gamma(TM)) such that gig_{i} converges to gg on MM in the C0C^{0}-sense as i.i\rightarrow\infty. Assume

MR(gi)𝑑volgiκforsomeconstantκ.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa~{}~{}~{}\mathrm{for~{}some~{}constant}~{}\kappa\in\mathbb{R}.

Moreover, assume that λiC+\lambda_{i}\leq C_{+} for all ii and some constant C+C_{+}\in\mathbb{R} if κ0\kappa\geq 0 (resp. λiC\lambda_{i}\geq C_{-} for some CC_{-}\in\mathbb{R} if κ<0\kappa<0). Then

MR(g)𝑑volgκ.\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.

On a closed manifold, every Ricci soliton is a self-similar solution of the Ricci flow equation, and vice versa. This self-similarity is one of the reasons why the assumption of convergence in Main Theorem 3 can be weaker than W1,p.W^{1,p}.

The rest of the paper will be arranged as follows. In Section 2, we will prove our Main theorems after preparing some lemmas. In Secrion 3, we will prove some corollaries of Main Theorems. In Section 4, we will give several (partial) counterexamples to Main theorem 1 and 2.

2 Proof of Main Theorems

Firstly, we need the following stability result for the Ricci-DeTurck flow like what is proved in [2, Lemma 2]. More precisely, we need the following.

Lemma 2.1 ([11, 19]).

Let (M,h)(M,h) be a closed Riemannian manifold endowed with a C2C^{2}-Riemannian metric h.h. Then there are constants τ,ε>0,C<\tau,\varepsilon>0,~{}C<\infty such that the following is true: Consider a C2C^{2}-Riemannian metric gg that is (1+ϵ)(1+\epsilon)-bilipschitz close to h.h. Then there is a continuous family of Riemannian metrics (gt)t[0,τ](g_{t})_{t\in[0,\tau]} on MM such that the following holds:

  • (a)(a)

    For all t[0,τ],t\in[0,\tau], the metric gtg_{t} is 1.11.1-bilipschitz to h.h.

  • (b)(b)

    (gt)(g_{t}) is smooth on M×(0,τ]M\times(0,\tau] and the map [0,τ]C2(M,S2M),tgt[0,\tau]\rightarrow C^{2}(M,S^{2}M),~{}t\mapsto g_{t} is continuous. In particular, tRgtt\mapsto R_{g_{t}} is continuous on [0,τ][0,\tau].

  • (c)(c)

    g0=gg_{0}=g and (gt)t[0,τ](g_{t})_{t\in[0,\tau]} is a solution to the Ricci DeTurck flow equation

    (RDE)tgt=2Ric(gt)Xh(gt)gt,\mathrm{(RDE)}~{}~{}\frac{\partial}{\partial t}g_{t}=-2~{}\mathrm{Ric}(g_{t})-\mathcal{L}_{X_{h}(g_{t})}g_{t},

    where Xh(gt)gt\mathcal{L}_{X_{h}(g_{t})}g_{t} denotes the Lie derivative of gtg_{t} with respect to the time-dependent vector field Xh(gt)X_{h}(g_{t}) defined in Remark 2.1 below.

  • (d)(d)

    For any t(0,τ]t\in(0,\tau] and any m=0,1,2m=0,1,2 we have

    |hmgt|h<Ctm/2|\nabla_{h}^{m}g_{t}|_{h}<\frac{C}{t^{m/2}}

    where |hmgt|h|\nabla_{h}^{m}g_{t}|_{h} denotes the norm with respect to hh of the covariant derivatives of g(t)g(t) by the Levi-Civita connection of h.h.

  • (e)(e)

    If (gi,t)t[0,τ](g_{i,t})_{t\in[0,\tau]} is a sequence of solutions to (RDE) that are continuous on M×[0,τ]M\times[0,\tau] and smooth on M×(0,τ]M\times(0,\tau] and if gi,0g_{i,0} converges to some metric g0g_{0} in the C0C^{0}-sense, then there is a subsequence of (gi,t)(g_{i,t}) that converges to (gt)(g_{t}) in the C0C^{0}-sense on M×[0,τ]M\times[0,\tau] and in the locally smooth sense on M×(0,τ]M\times(0,\tau] with respect to h.h.

Proof.

The items (a),(c)(a),(c) and (d)(d) are the result of [19, Theorem 1.1]. (e)(e) follows from the standard argument using the derivative estimates in [19, Theorem 1.1] and Arzela-Ascoli’s theorem. (b)(b) follows from the same argument as in the proof of [19, Theorem 5.2]. But we use the equations which is obtained from differentiating the equation (66)(66) in [18] and (2.15)(2.15) in [19] twice with respect to h\nabla_{h} instead of [18, (66) Section 2] and [19, (2.15)] respectively. In particular, we use the continuity of the second derivatives of gg to derive the estimate corresponding to (5.2)(5.2) in the proof of Theorem 5.2 in [19]. ∎

Thanks to [11, Theorem 3.11], if we assume gig_{i} converges to gg in the W1,p(p>n)W^{1,p}~{}(p>n)-sense, we obtain the following as well.

Lemma 2.2 (c.f. [19, Theorem 4.3]).

Let (M,g)(M,g) be a closed Riemannian manifold endowed with a C2C^{2}-Riemannian metric g.g. Then there are constants ε,τ>0\varepsilon,~{}\tau>0 and C<C<\infty such that the following is true: Consider a C2C^{2}-Riemannian metric gg^{\prime} on MM which is ε\varepsilon-close to gg in the W1,p(p>n)W^{1,p}~{}(p>n)-sense with respect to gg (i.e., |gg|W1,p(M,g)<ε,|g-g^{\prime}|_{W^{1,p}(M,g)}<\varepsilon, where |gg|W1,p(g)|g-g^{\prime}|_{W^{1,p}(g)} denotes the W1,pW^{1,p}-norm of the (0,2)(0,2)-tensor ggg-g^{\prime} with respect to gg). Then there is a continuous family of Riemannian metrics (gt)t[0,τ](g^{\prime}_{t})_{t\in[0,\tau]} on MM such that (a)(a)-(c)(c) and (e)(e) in Lemma 2.1 hold. Moreover, the following (d)(d^{\prime}) holds instead of (d)(d) in Lemma 2.1.

  • (d)(d^{\prime})

    For any t(0,τ]t\in(0,\tau] we have

    |ggt|g<Ctn/2pand|g2gt|g<Ctn4p+34|\nabla_{g}g^{\prime}_{t}|_{g}<\frac{C}{t^{n/2p}}~{}~{}\mathrm{and}~{}~{}|\nabla_{g}^{2}g^{\prime}_{t}|_{g}<\frac{C}{t^{\frac{n}{4p}+\frac{3}{4}}}
Proof.

The statements except for the item (d)(d^{\prime}) are the results of Lemma 2.1 (a)(a)-(c),(e)(c),(e) respectively. The item (d)(d^{\prime}) is the result of [11, Theorem 3.11]. Indeed, since gg^{\prime} is W1,pW^{1,p}-close to g,g, the W1,pW^{1,p}-bounded assumption in [11, Theorem 3.11]:

M|gg|p𝑑volgA=A(ε,g)\int_{M}|\nabla_{g}g^{\prime}|^{p}\,d\mathrm{vol}_{g}\leq A=A(\varepsilon,g)

is satisfied. Therefore we obtain the assertion from [11, Theorem 3.11]. ∎

Remark 2.1.

Let (gt)t[0,T)(0<T)(g_{t})_{t\in[0,T)}~{}(0<T) be a solution of the Ricci-DeTurck flow equation with g0=g.g_{0}=g. Choose a background metric g¯\bar{g} on MM and define the Bianchi operator

Xg¯i(h)=(g¯+h)ij(g¯+h)pq(pg¯hqj+12jg¯hpq),X^{i}_{\bar{g}}(h)=(\bar{g}+h)^{ij}(\bar{g}+h)^{pq}\left(-\nabla^{\bar{g}}_{p}h_{qj}+\frac{1}{2}\nabla^{\bar{g}}_{j}h_{pq}\right), (1)

which assigns a vector field to every symmetric 2-form hh on M.M. Let (Φt)tI(\Phi_{t})_{t\in I} be the flow (we call this flow the Ricci-DeTurck deiffeomorphism below) generated by the time-dependent family of vector fields Xg¯(gt),X_{\bar{g}}(g_{t}), i.e.,

tΦt=Xg¯(gt)ΦtwithΦ0=idM.\frac{\partial}{\partial t}\Phi_{t}=X_{\bar{g}}(g_{t})\circ\Phi_{t}~{}~{}\mathrm{with}~{}~{}\Phi_{0}=\mathrm{id}_{M}. (2)

Then g~t:=Φtgt\tilde{g}_{t}:=\Phi^{*}_{t}g_{t} satisfies the Ricci flow equation:

tg~t=2Ric(g~t)withg~0=g0=g.\frac{\partial}{\partial t}\tilde{g}_{t}=-2\,\mathrm{Ric}(\tilde{g}_{t})~{}~{}\mathrm{with}~{}~{}\tilde{g}_{0}=g_{0}=g.

For each gig_{i}, from Lemma 2.2, we have a Ricci-DeTurck flow gi(t)g_{i}(t) defined on [0,τ][0,\tau] for some positive time τ\tau, which is independent of ii. We also have the corresponding Ricci flow g~i(t)\tilde{g}_{i}(t) and the Ricci-DeTurck diffeomorphism Φi,t\Phi_{i,t} both defined on the same interval [0,τ][0,\tau]. Indeed, we assume that there is the largest time 0<t0<τ0<t_{0}<\tau so that the Ricci-DeTurck diffeomorphism exists on [0,t0)[0,t_{0}). On the other hand, since the Ricci-DeTurck flow gi(t)g_{i}(t) is defined on the interval [0,τ][0,\tau], from the Shi-type estimates for gi(t)g_{i}(t), we can show that Φi,t\Phi_{i,t} converges to a diffeomorphism Φi,t0\Phi_{i,t_{0}} as tt0t\nearrow t_{0} by the similar argument in the following proof of Lemma 2.3. This contradicts the definition of the time t0t_{0}. Moreover, under the condition of Lemma 2.1 (e)(e), we can see that for each t0(0,τ),t_{0}\in(0,\tau), {g~i,t}t[t0,τ]\{\tilde{g}_{i,t}\}_{t\in[t_{0},\tau]} smoothly subconverges to {g~t}t[t0,τ].\{\tilde{g}_{t}\}_{t\in[t_{0},\tau]}.

Let MM be a smooth manifold and gg a C2C^{2}-Riemannian metric on M.M. Assume a sequence of C2C^{2}-Riemannian metrics (gi)(g_{i}) on MM converges to gg on MM in the W1,p(p>n2/2)W^{1,p}~{}(p>n^{2}/2)-sense. Then, from Lemma 2.2, there are a positive time τ=τ(M,g)>0\tau=\tau(M,g)>0 and a positive constant C=C(M,g)<C=C(M,g)<\infty such that for sufficiently large i,i, Lemma 2.2 holds for gg and gi.g_{i}. Thus, let (Φi,t)t[0,τ)(\Phi_{i,t})_{t\in[0,\tau)} be the corresponding Ricci-DeTurck diffeomorphism of gig_{i} with background metric gg defined as in (2). The next lemma will be used in the proof of Main Theorem 2. In particular, the assumption p>n2/2p>n^{2}/2 is needed for this lemma.

Lemma 2.3 (Subconvergence of the Ricci-DeTurck diffeomorphisms).

In the above setting, there is a subsequence of (Φi,t)t[0,τ/2](\Phi_{i,t})_{t\in[0,\tau/2]} that converges to a time-dependent C1C^{1}-diffeomorphism (Φt)t[0,τ/2](\Phi_{t})_{t\in[0,\tau/2]} (i.e., for all t[0,τ/2],Φtt\in[0,\tau/2],~{}\Phi_{t} is a homeomorphism, Φt,Φt1\Phi_{t},\Phi_{t}^{-1} are continuously differentiable and dΦt\mathrm{d}\Phi_{t} and dΦt1\mathrm{d}\Phi^{-1}_{t} are mutually inverse. And, there is a subsequence (Φik,t)t[0,τ/2](\Phi_{i_{k},t})_{t\in[0,\tau/2]} such that for all t[0,τ/2],Φik,t,Φik,t1,dΦik,tt\in[0,\tau/2],~{}\Phi_{i_{k},t},\Phi_{i_{k},t}^{-1},\mathrm{d}\Phi_{i_{k},t} and dΦik,t1\mathrm{d}\Phi_{i_{k},t}^{-1} are converges to Φt,Φt1,dΦt\Phi_{t},\Phi_{t}^{-1},\mathrm{d}\Phi_{t} and dΦt1\mathrm{d}\Phi_{t}^{-1} respectively) with Φ0=idM,\Phi_{0}=\mathrm{id}_{M}, where idM:MM\mathrm{id}_{M}:M\rightarrow M denotes the identity map.

Proof.

Step 1(C0C^{0}-convergence): From Lemma 2.2 (d)(d^{\prime}) and the definition (1), the norm (with respect to gg) of the time-dependent vector field defined in (1) is bounded by some positive constant C=C(M,g).C=C(M,g). Thus, applying Gronwall’s lemma to (2), by the same argument in the proof of Lemma 2.1 in [5], one can obtain that

dg(Φi,t(p),Φi,s(p))C|t1n2ps1n2p|,t,s[0,τ/2],pM.d_{g}(\Phi_{i,t}(p),\Phi_{i,s}(p))\leq C|t^{1-\frac{n}{2p}}-s^{1-\frac{n}{2p}}|,~{}~{}~{}~{}t,s\in[0,\tau/2],~{}p\in M. (3)

Here, CC depends only on M,gM,g and τ.\tau. Unlike that in [5, Lemma 2.1], one can get the above type of estimate (in particular, the right-hand side is not |ts||\sqrt{t}-\sqrt{s}| but |t1n2ps1n2p||t^{1-\frac{n}{2p}}-s^{1-\frac{n}{2p}}|). This follows from the first-derivative estimate (Lemma 2.2 (d)(d^{\prime}) and hence |Xg(gi,t)|g|X_{g}(g_{i,t})|_{g} is bounded by Ctn2pCt^{\frac{n}{2p}} where C=C(M,g)C=C(M,g) is a positive constant. On the other hand, taking the derivative of both sides of (2), and using the estimate of the second derivatives (Lemma 2.2 (d)(d^{\prime})), we obtain that

t|dΦi,t|gCtn4p+34|dΦi,t|g,t(0,τ/2],\frac{\partial}{\partial t}\,|\mathrm{d}\Phi_{i,t}|_{g}\leq\frac{C}{t^{\frac{n}{4p}+\frac{3}{4}}}|\mathrm{d}\Phi_{i,t}|_{g},~{}~{}~{}~{}t\in(0,\tau/2],

where |dΦi,t|g|\mathrm{d}\Phi_{i,t}|_{g} denotes the maximum of the operator norm of dΦi,t:TMTM\mathrm{d}\Phi_{i,t}:TM\rightarrow TM with respect to gg on M.M. Hence, for all t(0,τ/2],t\in(0,\tau/2], we have

ddt(u(t)v(t))0,\frac{d}{dt}\left(\frac{u(t)}{v(t)}\right)\leq 0,

where u(t)=|dΦi,t|gu(t)=|\mathrm{d}\Phi_{i,t}|_{g} and v(t)=eCt1n4p34v(t)=e^{Ct^{1-\frac{n}{4p}-\frac{3}{4}}}. Then, since u(t)u(t) is continuous on [0,τ/2],[0,\tau/2], from the mean value theorem and this estimate of time-derivative, we have

|dΦi,t|gCexpt1n4p34,t[0,τ/2].|\mathrm{d}\Phi_{i,t}|_{g}\leq C\exp{t^{1-\frac{n}{4p}-\frac{3}{4}}},~{}~{}~{}~{}t\in[0,\tau/2]. (4)

Since the pullback metric Φi,tgi,t\Phi^{*}_{i,t}\,g_{i,t} satisfies the Ricci flow equation:

t(Φi,tgi,t)=2Ric(Φi,tgi,t),\frac{\partial}{\partial t}(\Phi^{*}_{i,t}\,g_{i,t})=-2\,\mathrm{Ric}(\Phi^{*}_{i,t}\,g_{i,t}),

by Lemma 2.2 (d)(d^{\prime}) and the above estimate, there is a constant C=C(M,g,τ)C=C(M,g,\tau) such that for all points (p,t)M×(0,τ/2](p,t)\in M\times(0,\tau/2] and all vectors vTpM,v\in T_{p}M,

|ddt|v|Φi,tgi,t2|Ctn4p+34|v|Φi,tgi,t2.\left|\frac{d}{dt}|v|^{2}_{\Phi^{*}_{i,t}g_{i,t}}\right|\leq\frac{C}{t^{\frac{n}{4p}+\frac{3}{4}}}|v|^{2}_{\Phi^{*}_{i,t}g_{i,t}}.

From this estimate, Lemma 2.2 (b)(b) and the mean value theorem, we obtain

eCτ1n4p34|v|gi2|v|Φi,tgi,t2eCτ1n4p34|v|gi2.e^{-C\tau^{1-\frac{n}{4p}-\frac{3}{4}}}|v|^{2}_{g_{i}}\leq|v|^{2}_{\Phi^{*}_{i,t}g_{i,t}}\leq e^{C\tau^{1-\frac{n}{4p}-\frac{3}{4}}}|v|^{2}_{g_{i}}.

Since gig_{i} converges to gg in the C0C^{0}-sense, for all sufficiently large i,i, we have

eC~τ1n4p34|v|g2|v|Φi,tgi,t2eC~τ1n4p34|v|g2e^{-\tilde{C}\tau^{1-\frac{n}{4p}-\frac{3}{4}}}|v|^{2}_{g}\leq|v|^{2}_{\Phi^{*}_{i,t}g_{i,t}}\leq e^{\tilde{C}\tau^{1-\frac{n}{4p}-\frac{3}{4}}}|v|^{2}_{g}

for some constant C~=C~(M,g,τ).\tilde{C}=\tilde{C}(M,g,\tau). Therefore, from Lemma 2.2 (a),(a), the C0C^{0}-closedness of gig_{i} and g,g, and the previous estimate, there is a positive constant C=C(M,g,τ)C=C(M,g,\tau) such that for all sufficiently large i,i,

dg(Φi,t(p),Φi,t(q))Cdgi,t(Φi,t(p),Φi,t(q))Cdg(p,q).d_{g}(\Phi_{i,t}(p),\Phi_{i,t}(q))\leq Cd_{g_{i,t}}(\Phi_{i,t}(p),\Phi_{i,t}(q))\leq C\,d_{g}(p,q). (5)

Combining the inequalities (3) and (5), we have

dg(Φi,t(p),Φi,s(q))C(|t1n2ps1n2p|+dg(p,q)),t,s[0,τ/2],p,qM.d_{g}(\Phi_{i,t}(p),\Phi_{i,s}(q))\leq C\left(|t^{1-\frac{n}{2p}}-s^{1-\frac{n}{2p}}|+d_{g}(p,q)\right),~{}~{}~{}t,s\in[0,\tau/2],~{}p,q\in M.

Then, by Arzela-Ascoli’s theorem, a subsequence (Φik,t)t[0,τ/2](\Phi_{i_{k},t})_{t\in[0,\tau/2]} of (Φi,t)t[0,τ/2](\Phi_{i,t})_{t\in[0,\tau/2]} converges to a time-dependent map Φt:MM(t[0,τ/2])\Phi_{t}:M\rightarrow M~{}(t\in[0,\tau/2]) as ik.i_{k}\rightarrow\infty. In exactly the same way, one can prove that there is a subsequence (Φikl,t1)t[0,τ/2](\Phi^{-1}_{i_{k_{l}},t})_{t\in[0,\tau/2]} of (Φik,t1)t[0,τ/2](\Phi^{-1}_{i_{k},t})_{t\in[0,\tau/2]} that converges to some time-dependent map Φ~t:MM(t[0,τ/2])\tilde{\Phi}_{t}:M\rightarrow M~{}(t\in[0,\tau/2]) as ikl.i_{k_{l}}\rightarrow\infty. But, since Φikl,t1Φikl,t=idM,\Phi^{-1}_{i_{k_{l}},t}\circ\Phi_{i_{k_{l}},t}=\mathrm{id}_{M}, Φt1\Phi^{-1}_{t} exists and Φt1=Φ~t\Phi^{-1}_{t}=\tilde{\Phi}_{t} for each t[0,τ/2].t\in[0,\tau/2]. To prevent complicated in expression, we will simply write this converging subsequence (Φikl,t)t[0,τ/2](\Phi_{i_{k_{l}},t})_{t\in[0,\tau/2]} as (Φi,t)t[0,τ/2].(\Phi_{i,t})_{t\in[0,\tau/2]}.

Step 2(C1C^{1}-convergence): Next, we will show that the first derivatives of Φi,t\Phi_{i,t} subconverges as i.i\rightarrow\infty. Let 0<a<1.0<a<1. Since ggt\nabla_{g}g^{\prime}_{t} satisfies a parabolic type PDE (see [18, Section 4, Equation (4)]), from the derivative estimate Lemma 2.2 we obtain that

t(ggtta)Δg(ggtta)aCta1n2p+Ctan2p+Ctan2pn4p34+Cta3n2p.\frac{\partial}{\partial t}\left(\nabla_{g}g^{\prime}_{t}\cdot t^{a}\right)-\Delta_{g}\left(\nabla_{g}g^{\prime}_{t}\cdot t^{a}\right)\leq aCt^{a-1-\frac{n}{2p}}+Ct^{a-\frac{n}{2p}}+Ct^{a-\frac{n}{2p}-\frac{n}{4p}-\frac{3}{4}}+Ct^{a-\frac{3n}{2p}}.

Then, by the parabolic Wq2,1W^{2,1}_{q}-estimate ([12, Ch. 4, Section 3, Theorem 7]), we have

ggttaWq2,1(M×(0,τ/2],g)C||\nabla_{g}g^{\prime}_{t}\cdot t^{a}||_{W^{2,1}_{q}(M\times(0,\tau/2],g)}\leq C

for all qq satisfying (1+n2pa)q<1\left(1+\frac{n}{2p}-a\right)q<1 and q2.q\geq 2. Note that if (1+n2pa)q<1\left(1+\frac{n}{2p}-a\right)q<1 then it holds that

(3n+3p4pa)q,(n2pa)q,(3n2pa)q<1.\left(\frac{3n+3p}{4p}-a\right)q,~{}\left(\frac{n}{2p}-a\right)q,~{}\left(\frac{3n}{2p}-a\right)q<1.

(Recall our assumption p>n2/2np>n^{2}/2\geq n.) In particular, we have

ggtWq2(M,g)Cta||\nabla_{g}g^{\prime}_{t}||_{W^{2}_{q}(M,g)}\leq Ct^{-a}

for all such a,qa,q and t(0,τ/2].t\in(0,\tau/2]. Since, p>n2/2,p>n^{2}/2, we can choose 0<a<10<a<1 sufficiently close to 11 so that

n<11+n2pa.n<\frac{1}{1+\frac{n}{2p}-a}. (6)

Hence, as noted above, the weaker relations:

n<13n+3p4pa,1n2pa,13n2pan<\frac{1}{\frac{3n+3p}{4p}-a},~{}\frac{1}{\frac{n}{2p}-a},~{}\frac{1}{\frac{3n}{2p}-a}

are satisfied as well under the above condition (6). Thus, from Morrey’s embedding theorem,

g2gtCα(M,g)Cta,t(0,τ/2]||\nabla^{2}_{g}g^{\prime}_{t}||_{C^{\alpha}(M,g)}\leq Ct^{-a},~{}~{}~{}~{}t\in(0,\tau/2]

for some 0<α<1.0<\alpha<1. Therefore, by the same arguments that derived (4) above, we obtain that

|dΦi,t|Cα(M,g)Cexpt1a,t[0,τ/2].|\mathrm{d}\Phi_{i,t}|_{C^{\alpha}(M,g)}\leq C\exp{t^{1-a}},~{}~{}~{}~{}t\in[0,\tau/2]. (7)

Then, by the inequalities (4) and (7), we can apply the Arzela-Ascoli’s theorem and obtain that there is a subsequence (Φik,t)t[0,τ/2](\Phi_{i_{k},t})_{t\in[0,\tau/2]} such that as ik,i_{k}\rightarrow\infty, (Φik,t)t[0,τ/2](Φt)t[0,τ/2](\Phi_{i_{k},t})_{t\in[0,\tau/2]}\rightarrow(\Phi_{t})_{t\in[0,\tau/2]} for some time-dependent map (Φt)t[0,τ/2].(\Phi_{t})_{t\in[0,\tau/2]}. Moreover, (Φt)t[0,τ/2](\Phi_{t})_{t\in[0,\tau/2]} are differentiable on MM and (dΦik,t)t[0,τ/2](dΦt)t[0,τ/2](\mathrm{d}\Phi_{i_{k},t})_{t\in[0,\tau/2]}\rightarrow(\mathrm{d}\Phi_{t})_{t\in[0,\tau/2]} as ik.i_{k}\rightarrow\infty. Similarly, there is a subsequence (Φikl,t)t[0,τ/2](\Phi_{i_{k_{l}},t})_{t\in[0,\tau/2]} of (Φik,t)t[0,τ/2](\Phi_{i_{k},t})_{t\in[0,\tau/2]} such that

(Φikl,t1)t[0,τ/2](Φ~t)t[0,τ/2]and(dΦikl,t1)t[0,τ/2](dΦ~t)t[0,τ/2](\Phi^{-1}_{i_{k_{l}},t})_{t\in[0,\tau/2]}\rightarrow(\tilde{\Phi}_{t})_{t\in[0,\tau/2]}~{}~{}\mathrm{and}~{}~{}(\mathrm{d}\Phi^{-1}_{i_{k_{l}},t})_{t\in[0,\tau/2]}\rightarrow(\mathrm{d}\tilde{\Phi}_{t})_{t\in[0,\tau/2]}

as ikl.i_{k_{l}}\rightarrow\infty. Since Φikl,tΦikl,t1=Φikl,t1Φikl,t=idM\Phi_{i_{k_{l}},t}\circ\Phi^{-1}_{i_{k_{l}},t}=\Phi^{-1}_{i_{k_{l}},t}\circ\Phi_{i_{k_{l}},t}=\mathrm{id}_{M} and dΦikl,tdΦikl,t1=dΦikl,t1dΦikl,t=idTM,\mathrm{d}\Phi_{i_{k_{l}},t}\circ\mathrm{d}\Phi^{-1}_{i_{k_{l}},t}=\mathrm{d}\Phi^{-1}_{i_{k_{l}},t}\circ\mathrm{d}\Phi_{i_{k_{l}},t}=\mathrm{id}_{TM}, dΦt\mathrm{d}\Phi_{t} are invertible and dΦt1=dΦ~t\mathrm{d}\Phi_{t}^{-1}=\mathrm{d}\tilde{\Phi}_{t} for all t[0,τ/2].t\in[0,\tau/2]. Finally, Φ0=idM\Phi_{0}=\mathrm{id}_{M} easily follows from the definition (2) and the above construction of (Φt)t[0,τ/2].(\Phi_{t})_{t\in[0,\tau/2]}.

Remark 2.2.

In Lemma 2.3, it is not known that the limit (Φt)t[0,τ/2](\Phi_{t})_{t\in[0,\tau/2]} is the Ricci-DeTurck diffeomorphism of the Ricci flow starting at gg with background metric g.g. Therefore, let (gt)t[0.τ/2](g_{t})_{t\in[0.\tau/2]} be the Ricci-DeTurck flow starting at gg with background metric g,g, then we don’t know whether or not (Φtgt)t[0,τ/2](\Phi_{t}^{*}g_{t})_{t\in[0,\tau/2]} is the solution of the Ricci flow equation starting at g.g.

Next, we need the following stability of the heat flow with the Ricci flow background which is proved by Lee and Tam [14, Theorem 3.1].

Lemma 2.4 ([14, Theorem 3.1]).

Let (Mn,g0)(M^{n},g_{0}) be a closed manifold of dimension n2n\geq 2 and {fi}\{f_{i}\} a family of functions on MM satisfying the assumptions (1)(1) and (2)(2) in Main Theorem 2. Suppose g(t)(t[0,τ])g(t)~{}(t\in[0,\tau]) be a solution of the Ricci flow equation starting at g0g_{0} such that |Rm(g(t))|at1|\mathrm{Rm}(g(t))|\leq a\cdot t^{-1} for some a>0a>0 on (0,τ].(0,\tau]. Then for all i0,i_{0}\in\mathbb{N}, there are positive constants τ0=τ0(n,Λ,i0)>0\tau_{0}=\tau_{0}(n,\Lambda,i_{0})>0 and C0=C0(n,a,Λ,i0)>0C_{0}=C_{0}(n,a,\Lambda,i_{0})>0 such that the following holds. For all ii0,i\geq i_{0}, there exists Fi(t)C(M)(t(0,min{τ,τ0})F_{i}(t)\in C^{\infty}(M)~{}(t\in(0,\min\{\tau,\tau_{0}\}) satisfies the heat flow equation:

tFi(t)=ΔgtFi(t)\frac{\partial}{\partial t}F_{i}(t)=\Delta_{g_{t}}F_{i}(t)

such that

  • (A)

    (Λ22(n1)t)(Fi)gEuclg(t),\left(\Lambda^{-2}-2(n-1)t\right)(F_{i})^{*}g_{Eucl}\leq g(t),

  • (B)

    supxMdgEucl(Fi(x,t),fi(x))C0t.\sup_{x\in M}d_{g_{Eucl}}(F_{i}(x,t),f_{i}(x))\leq C_{0}\sqrt{t}.

Moreover, for any integer l0,l\geq 0, there is a constant C=C(n,l,a,Λ,i0)>0C=C(n,l,a,\Lambda,i_{0})>0 such that for all t(0,min{τ,τ0}],t\in(0,\min\{\tau,\tau_{0}\}],

|ldFi|Ctl/2.|\nabla^{l}dF_{i}|\leq C\cdot t^{-l/2}.

Here, gEuclg_{Eucl} denotes the Euclidean metric on .\mathbb{R}.

Proof.

Since MM is compact, the image of ff is compact in the target space .\mathbb{R}. From the assumption (2)(2) in Main Theorem 2, there is a compact neighbourhood N()N(\subset\mathbb{R}) of the image of ff such that the image of fif_{i} is contained in NN for all ii0.i\geq i_{0}. Then, from the assumption (1)(1) of Main Theorem 2, one can apply the proof of Lemma 3.1 and Theorem 3.1 in [14] to (M,g(t)),(N,h:=gEucl|N)(M,g(t)),~{}(N,h:=g_{Eucl}|_{N}) and fi:MN(forallii0).f_{i}:M\rightarrow N\subset\mathbb{R}~{}(\mathrm{for~{}all}~{}i\geq i_{0}). Hence we obtain the desired assertions from Theoem 3.1 in [14]. ∎

The following lemma is a key to prove our main theorems. Note that we only need p>np>n for this lemma.

Lemma 2.5 (cf. [2, Lemma 4]).

Let MnM^{n} be a closed nn-manifold (n2n\geq 2) and gg a C2C^{2}-Riemannian metric on M.M. Suppose f:Mf:M\rightarrow\mathbb{R} be a Λ\Lambda-Lipschitz function for some Λ>0.\Lambda>0. Let gg^{\prime} be a C2C^{2}-Riemannian metric which is sufficiently W1,p(p>n)W^{1,p}~{}(p>n)-close to gg as in Lemma 2.2. Then for any given positive constant δ>0,\delta>0, there is a constant τ=τ(M,g,|gg|W1,p(M,g),δ,Λ)>0\tau=\tau\left(M,g,|g-g^{\prime}|_{W^{1,p}(M,g)},\delta,\Lambda\right)>0 such that the following holds: Assume that MR(g)𝑑m>a,\int_{M}R(g^{\prime})\,dm>a, where dm:=efdvolgdm:=e^{-f}d\mathrm{vol}_{g^{\prime}} for some aa\in\mathbb{R} and R(g)0R(g^{\prime})\geq 0 on M.M. Then there is a solution (gt)t[0,τ](g^{\prime}_{t})_{t\in[0,\tau]} to the Ricci flow equation with initial metric gg^{\prime} and a solution (ft)t(0,τ](f_{t})_{t\in(0,\tau]} of the heat flow equation with the Ricci flow background such that MR(gt)eft𝑑volg(t)>aδ\int_{M}R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}(t)}>a-\delta for all t[0,τ].t\in[0,\tau].

Proof.

From Lemma 2.2 and 2.4, there is a sufficiently small 1>τ>01>\tau^{\prime}>0 such that there is a Ricci flow (g(t))t[0,τ](g^{\prime}(t))_{t\in[0,\tau^{\prime}]} starting at gg^{\prime} and there is a heat flow (ft)t(0,τ](f_{t})_{t\in(0,\tau^{\prime}]} with ftff_{t}\rightarrow f as t0t\searrow 0 uniformly. Along these (gt)(g^{\prime}_{t}) and (ft),(f_{t}), for all t(0,τ],t\in(0,\tau^{\prime}],

ddt(MR(gt)eft𝑑volgt)=M(Δg(t)R(gt)+2|Ric(gt)|gt2R(gt)2)eft𝑑volgt+M(teft)R(gt)𝑑volgt=M(2|Ric(gt)|gt2R(gt)2)eft𝑑volgt2M(Δgtft|ft|2)R(gt)eft𝑑volgtM(2nR(gt)2R(gt)2)eft𝑑volgt2M(Δgtft|ft|2)R(gt)eft𝑑volgt=(2n1)MR(gt)2eft𝑑volgt2M(Δgtft|ft|2)R(gt)eft𝑑volgtCtn4p34MR(gt)eft𝑑volgt.\begin{split}&\frac{d}{dt}\left(\int_{M}R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}\right)\\ &=\int_{M}\left(\Delta_{g(t)}R(g^{\prime}_{t})+2|\mathrm{Ric}(g^{\prime}_{t})|^{2}_{g^{\prime}_{t}}-R(g^{\prime}_{t})^{2}\right)\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}+\int_{M}\left(\frac{\partial}{\partial t}e^{-f_{t}}\right)R(g^{\prime}_{t})\,d\mathrm{vol}_{g^{\prime}_{t}}\\ &=\int_{M}\left(2|\mathrm{Ric}(g^{\prime}_{t})|^{2}_{g^{\prime}_{t}}-R(g^{\prime}_{t})^{2}\right)\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}-2\int_{M}\left(\Delta_{g^{\prime}_{t}}f_{t}-|\nabla f_{t}|^{2}\right)R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}\\ &\geq\int_{M}\left(\frac{2}{n}R(g^{\prime}_{t})^{2}-R(g^{\prime}_{t})^{2}\right)\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}-2\int_{M}\left(\Delta_{g^{\prime}_{t}}f_{t}-|\nabla f_{t}|^{2}\right)R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}\\ &=\left(\frac{2}{n}-1\right)\int_{M}R(g^{\prime}_{t})^{2}\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}-2\int_{M}\left(\Delta_{g^{\prime}_{t}}f_{t}-|\nabla f_{t}|^{2}\right)R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}\\ &\geq-Ct^{-\frac{n}{4p}-\frac{3}{4}}\int_{M}R(g^{\prime}_{t})\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}_{t}}.\end{split}

Here, CC is a positive constant depending on M,ΛM,\Lambda and |gg|W1,p(M,g).|g-g^{\prime}|_{W^{1,p}(M,g)}. The first equality follows from the evolution of the scalar curvature and the volume form under the Ricci flow, and tft=Δgtft(t(0,τ]).\partial_{t}f_{t}=\Delta_{g_{t}}f_{t}~{}(t\in(0,\tau^{\prime}]). The second equality follows from applying the divergence formula to the term MΔg(t)R(g(t))eft𝑑volg(t).\int_{M}\Delta_{g^{\prime}(t)}R(g^{\prime}(t))\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}(t)}. The first inequality follows from the Cauchy-Schwarz inequality n|Ric(gt)|2R(gt)2.n|\mathrm{Ric}(g^{\prime}_{t})|^{2}\geq R(g^{\prime}_{t})^{2}. Finally, we have obtained the last inequality as follows. From the assumption R(g)0,R(g^{\prime})\geq 0, by the maximum principle under the Ricci flow, we have R(gt)0.R(g^{\prime}_{t})\geq 0. Since the scalar curvature is invariant under the pullback action by a diffeomorphism, from the C1C^{1}-closedness assumption, one can apply the same derivative estimate (d)(d^{\prime}) in Lemma 2.2 to one R(gt)R(g^{\prime}_{t}) in the integrand of the left-hand side of the inequality. Moreover, applying the derivative estimate for the heat flow as in Lemma 2.4 to the second integrand of the left-hand side of the last inequality, we obtain the desired estimate.

Set u(t):=MR(gt)eft𝑑volgtu(t):=\int_{M}R(g^{\prime}_{t})e^{-f_{t}}\,d\mathrm{vol}_{g^{\prime}_{t}} and v(t):=exp(C(1n4p34)1t1n4p34).v(t):=\exp\left(C\left(1-\frac{n}{4p}-\frac{3}{4}\right)^{-1}t^{1-\frac{n}{4p}-\frac{3}{4}}\right). Then, by the above calculation, we have

ddt(u(t)v(t))=u(t)v(t)u(t)v(t)v2(t)Ctn4p34u(t)v(t)u(t)(Ctn4p34)v(t)v2(t)=0\begin{split}\frac{d}{dt}\left(\frac{u(t)}{v(t)}\right)&=\frac{u^{\prime}(t)v(t)-u(t)v^{\prime}(t)}{v^{2}(t)}\\ &\geq\frac{-Ct^{-\frac{n}{4p}-\frac{3}{4}}u(t)v(t)-u(t)\left(-Ct^{-\frac{n}{4p}-\frac{3}{4}}\right)v(t)}{v^{2}(t)}=0\end{split}

for all t(0,τ].t\in(0,\tau^{\prime}]. Since u(t)u(t) is continuous on [0,τ][0,\tau^{\prime}] from Lemma 2.1 (b)(b) and Lemma 2.4 (B),(B), by the mean value theorem and the previous derivative estimate, we finally obtain that

MR(g(t))eft𝑑volg(t)v(t)MR(g)ef𝑑volg,t[0,τ].\int_{M}R(g^{\prime}(t))\,e^{-f_{t}}d\mathrm{vol}_{g^{\prime}(t)}\geq v(t)\cdot\int_{M}R(g^{\prime})\,e^{-f}d\mathrm{vol}_{g^{\prime}},~{}~{}~{}~{}t\in[0,\tau^{\prime}].

Therefore, if we take sufficiently small 0<τ<<τ,0<\tau<<\tau^{\prime}, the desired assertion holds for such a τ.\tau.

We prove Main Theorems as follows. First, we prove Main Theorem 2 because we will use almost the same method in the proof of Main theorems 1, 3. The idea of proof here is the same as that of Bamler [2]. We will show the assertion by contradiction. In order to do it, we suppose that the weighted total scalar curvature of the limiting metric is less than κ.\kappa. Then, we can also suppose that it is less than or equal to κδ\kappa-\delta for a positive small constant δ.\delta. On the other hand, from Lemma 2.2, we can flow each metric gig_{i} and potential function fif_{i} respectively along the Ricci–DeTurck flow and the heat flow with the Ricci flow background, up to a uniform (i.e., independent of ii) positive time τ\tau. From Lemma 2.5, if we retake τ\tau sufficiently small (independent of ii), the weighted total scalar curvature is bounded from below by κδ/2\kappa-\delta/2 along the Ricci flow and the heat flow for each i.i. Thus, combining Lemma 2.2, 2.3 and 2.4, we can deduce that the weighted total scalar curvature of the limiting metric is greater than or equal to κδ/2.\kappa-\delta/2. But this contradicts the supposition that the weighted total scalar curvature of the limiting metric is less than κ.\kappa. We prescribe these arguments more precisely below.

Proof of Main Theorem 2.

We show the assertion by contradiction. Suppose that

MR(g)ef𝑑volg<κ.\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}<\kappa.

Then there is a positive constant δ>0\delta>0 such that

MR(g)ef𝑑volgκδ<κ.\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}\leq\kappa-\delta<\kappa.

On the other hand, since MR(gi)efi𝑑volgiκ\int_{M}R(g_{i})\,e^{-f_{i}}\,d\mathrm{vol}_{g_{i}}\geq\kappa, from Lemma 2.2, Lemma 2.5 and Lemma 2.4, there are a τ=τ(δ),\tau=\tau(\delta), a Ricci flow (g~i,t)t[0,τ](\tilde{g}_{i,t})_{t\in[0,\tau]} and a heat flow (fi,t)t(0,τ](f_{i,t})_{t\in(0,\tau]} such that

MR(g~i,t)efi,t𝑑volg~i,tκ12δ\int_{M}R(\tilde{g}_{i,t})\,e^{-f_{i,t}}d\mathrm{vol}_{\tilde{g}_{i,t}}\geq\kappa-\frac{1}{2}\delta

and there also exists a heat flow (ft)t(0,τ](f_{t})_{t\in(0,\tau]} with ftff_{t}\rightarrow f as t0t\searrow 0 uniformly on M.M. By Lemma 2.3, as i,i\rightarrow\infty, we have a solution of the Ricci–DeTurck flow equation (gt)t[0,τ](g_{t})_{t\in[0,\tau]} (retake a sufficiently small τ\tau if necessary) starting at gg and a time-dependent C1C^{1}-diffeomorphism (Φt)t[0,τ](\Phi_{t})_{t\in[0,\tau]} with Φ0=idM\Phi_{0}=\mathrm{id}_{M} such that

MR(Φtgt)eft𝑑volΦtgt=MR(gt)eftΦt1𝑑volgtκ12δ\int_{M}R(\Phi_{t}^{*}g_{t})\,e^{-f_{t}}d\mathrm{vol}_{\Phi_{t}^{*}g_{t}}=\int_{M}R(g_{t})\,e^{-f_{t}\circ\Phi^{-1}_{t}}d\mathrm{vol}_{g_{t}}\geq\kappa-\frac{1}{2}\delta

for all t(0,τ].t\in(0,\tau]. On the other hand, by Lemma 2.2 (b),Φ0=idM(b),~{}\Phi_{0}=\mathrm{id}_{M} and ftf(t0),f_{t}\rightarrow f~{}(t\searrow 0), we have

MR(g)ef𝑑volg=limt0MR(gt)eftΦt1𝑑volgtκ12δ>κδ.\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}=\lim_{t\rightarrow 0}\int_{M}R(g_{t})\,e^{-f_{t}\circ\Phi^{-1}_{t}}d\mathrm{vol}_{g_{t}}\geq\kappa-\frac{1}{2}\delta>\kappa-\delta.

This contradicts our supposition

MR(g)ef𝑑volgκδ\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}\leq\kappa-\delta

and concludes the proof. ∎

Next, we give a proof of Main Theorem 1. This is simpler than the proof of Main Theorem 2 because the unweighted total scalar curvature MR(g)𝑑volg\int_{M}R(g)\,d\mathrm{vol}_{g} is diffeomorphism invariant, hence it does not need to use Lemma 2.3.

Proof of Main Theorem 1.

We firstly prove the first half of the theorem. Then we can prove this part in the same way as in the proof of Main Theorem 2 since we can use Lemma 2.5 with f0f\equiv 0 (hence ft0f_{t}\equiv 0). Note that since the total scalar curvature is diffeomorphism invariant, from Remark 2.1, we have

MR(g~i,t)𝑑volg~i,t=MR(gi,t)𝑑volgi,t,\int_{M}R(\tilde{g}_{i,t})\,d\mathrm{vol}_{\tilde{g}_{i,t}}=\int_{M}R(g_{i,t})\,d\mathrm{vol}_{g_{i,t}},

where g~i,t,gi,t\tilde{g}_{i,t},g_{i,t} are the Ricci-DeTurck flow and the Ricci flow starting at gig_{i} respectively. Note also that the condition p>np>n is sufficient to apply Lemma 2.5. Therefore, under the assumption of Main Theorem 1, arguing in the same way in the proof of Main Theorem 2 (and using the same notations), we obtain

MR(gt)𝑑volgtκ12δ\int_{M}R(g_{t})\,d\mathrm{vol}_{g_{t}}\geq\kappa-\frac{1}{2}\delta

for all t(0,τ],t\in(0,\tau], where gtg_{t} is the Ricci-DeTurck flow starting at g.g. Then we can deduce a contradiction as t0t\rightarrow 0 as in the same way in the proof of Main Theorem 2.

Next, we prove the theorem for n=3n=3 without assuming R(gi)0.R(g_{i})\geq 0. This statement follows from the following more general claim.

Claim 2.1 (cf. [4, Lecture Series 2, Proposition 2.3.1]).

Let (Mn,g)(M^{n},g) be a closed Riemannian manifold of dimension n2.n\geq 2. Suppose that a sequence of C2C^{2}-Riemannian metrics (gi)(g_{i}) on MM converges to gg in the W1,2W^{1,2}-sense. Suppose also that MM is parallelizable i.e., the tangent bundle of MM is trivial. Then

MR(gi)𝑑volgiMR(g)𝑑volgasi.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\rightarrow\int_{M}R(g)\,d\mathrm{vol}_{g}~{}~{}~{}~{}\mathrm{as}~{}~{}i\rightarrow\infty.
Proof of Claim 2.1.

We will use the similar technique in the proof of [4, Lecture Series 2, Proposition 2.3.1]. Since MM is parallelizable, we can take a global section {v1(x),,vn(x)}xM\{v_{1}(x),\cdots,v_{n}(x)\}_{x\in M} of the orthonomal frame bundle of M.M. Define the 1-form ωk():=g(vk,)(k=1,2,,n)\omega_{k}(\cdot):=g(v_{k},\cdot)~{}(k=1,2,\cdots,n) and the vector field vki(i=1,2,,k=1,2,,n)v^{i}_{k}~{}(i=1,2,\cdots,~{}k=1,2,\cdots,n) be the dual of ωk\omega_{k} with respect to gi.g_{i}. Next, we will use the Bochner identity for 1-forms:

Δg=gg+Ric(g)\Delta_{g}=\nabla_{g}^{*}\nabla_{g}+\mathrm{Ric}(g)
(resp.Δgi=gigi+Ric(gi)).(\mathrm{resp.}~{}~{}\Delta_{g_{i}}=\nabla_{g_{i}}^{*}\nabla_{g_{i}}+\mathrm{Ric}(g_{i})).

Here Δg=DgDg(resp.Δgi=DgiDgi)\Delta_{g}=D^{*}_{g}D_{g}~{}(\mathrm{resp.}~{}~{}\Delta_{g_{i}}=D^{*}_{g_{i}}D_{g_{i}}) for the Dirac operator Dg(resp.Dgi)D_{g}~{}(\mathrm{resp.}~{}~{}D_{g_{i}}) related to the deRham complex. Thus, we have for gig_{i} and vkiv^{i}_{k} using this Bochner identity and the integration by parts,

M(Dgiωkigi2giωkigi2)𝑑volgi=MRic(gi)(vki,vki)𝑑volgi.\int_{M}\left(||D_{g_{i}}\omega^{i}_{k}||_{g_{i}}^{2}-||\nabla_{g_{i}}\omega^{i}_{k}||^{2}_{g_{i}}\right)d\mathrm{vol}_{g_{i}}=\int_{M}\mathrm{Ric}(g_{i})(v^{i}_{k},v^{i}_{k})\,d\mathrm{vol}_{g_{i}}.

Snice gigg_{i}\rightarrow g in the W1,2W^{1,2}-sense on M,M, the left-hand side of the above equality converges to the following quantity

M(Dgωkg2gωkg2)𝑑volg=MRic(g)(vk,vk)𝑑volg.\int_{M}\left(||D_{g}\omega_{k}||_{g}^{2}-||\nabla_{g}\omega_{k}||^{2}_{g}\right)d\mathrm{vol}_{g}=\int_{M}\mathrm{Ric}(g)(v_{k},v_{k})\,d\mathrm{vol}_{g}.

Taking the sum from 1 to nn for k,k, we obtain that

MR(gi)𝑑volgiMR(g)𝑑volgasi.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\rightarrow\int_{M}R(g)\,d\mathrm{vol}_{g}~{}~{}~{}~{}\mathrm{as}~{}~{}i\rightarrow\infty.

This completes the proof of the claim. ∎

Since every oriented closed three manifold is parallelizable, using this claim we have

M3R(gi)𝑑volgiM3R(g)𝑑volgasi.\int_{M^{3}}R(g_{i})\,d\mathrm{vol}_{g_{i}}\rightarrow\int_{M^{3}}R(g)\,d\mathrm{vol}_{g}~{}~{}\mathrm{as}~{}i\rightarrow\infty.

(Of course, if necessary, we discuss similarly after taking its orientable double cover.) In particular, if M3R(gi)𝑑volgiκ\int_{M^{3}}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa for all i,i, then M3R(g)𝑑volgκ\int_{M^{3}}R(g)\,d\mathrm{vol}_{g}\geq\kappa. This completes the proof of Main Theorem 1. ∎

Example 2.1.

For example, the following manifolds are known to be parallelizable:

  • (1)

    Every orientable closed three manifold.

  • (2)

    nn-dimensional sphere SnS^{n} where n=0,1,3n=0,1,3 or 7.7.

  • (4)

    Every Lie group.

  • (3)

    The product of parallelizable manifolds.

Question 2.1.
  • In Claim 2.1, is the parallelizability of MM necessary?

  • What is the relation between parallelizability of a closed manifold and W1,2W^{1,2}-convergence of metric tensors on it?

Next, we give a proof of Main Theorem 3.

Proof of Main Theorem 3.

The proof is similar to the one of Main Theorem 2. However, since each metric is Ricci soliton, the Ricci flow starting from such a metric is homothetic. Hence Lemma 2.1, 2.3 and 2.5 are hold “more explicitly” in this situation. That is, the solution (g~i,t)t[0,τ)(\tilde{g}_{i,t})_{t\in[0,\tau)} of the Ricci flow equation starting from gig_{i} constructed in Lemma 2.1 is

g~i,t=(12λit)Φi,t(gi)t[0,τ),\tilde{g}_{i,t}=(1-2\lambda_{i}t)\Phi^{*}_{i,t}(g_{i})~{}~{}~{}~{}t\in[0,\tau),

where Φi,t\Phi_{i,t} is the family of diffeomorphisms generated by the time-dependent vector field Xi(t):=(12λit)1YiX_{i}(t):=(1-2\lambda_{i}t)^{-1}Y_{i} with the initial condition Φi,0=idM,\Phi_{i,0}=\mathrm{id}_{M}, and τ>0\tau>0 is the uniform existence time of the flows guaranteed in Lemma 2.1. Thus, we have

MR(g~i,t)𝑑volg~i,t=(12λit)n/21MR(gi)𝑑volgi.\int_{M}R(\tilde{g}_{i,t})\,d\mathrm{vol}_{\tilde{g}_{i,t}}=(1-2\lambda_{i}t)^{n/2-1}\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}. (8)

for all t[0,τ).t\in[0,\tau). We will firstly consider the case of κ0.\kappa\geq 0. From (8), for any δ>0,\delta>0, there is a positive time τ=τ(δ,C+)>0\tau=\tau(\delta,C_{+})>0 such that for any diffeomorphism Φ:MM\Phi:M\rightarrow M and t[0,τ],t\in[0,\tau],

MR(Φg~i,t)𝑑volΦgi,t=MR(g~i,t)𝑑volg~i,tκδ.\int_{M}R(\Phi^{*}\tilde{g}_{i,t})\,d\mathrm{vol}_{\Phi^{*}g_{i,t}}=\int_{M}R(\tilde{g}_{i,t})\,d\mathrm{vol}_{\tilde{g}_{i,t}}\geq\kappa-\delta.

Similarly, when κ<0,\kappa<0, we see that for any δ>0,\delta>0, there is a positive time τ=τ(δ,C)>0\tau=\tau(\delta,C_{-})>0 such that for any diffeomorphism Φ:MM\Phi:M\rightarrow M and t[0,τ],t\in[0,\tau],

MR(Φg~i,t)𝑑volΦg~i,t=MR(g~i,t)𝑑volg~i,tκδ.\int_{M}R(\Phi^{*}\tilde{g}_{i,t})\,d\mathrm{vol}_{\Phi^{*}\tilde{g}_{i,t}}=\int_{M}R(\tilde{g}_{i,t})\,d\mathrm{vol}_{\tilde{g}_{i,t}}\geq\kappa-\delta.

In particular, we take the inverse of the Ricci-DeTurck diffeomorphism (see Remark 2.1) of g~i,t\tilde{g}_{i,t} as Φ\Phi here, then we have

MR(gi,t)𝑑volgi,tκδforallt[0,τ],\int_{M}R(g_{i,t})\,d\mathrm{vol}_{g_{i,t}}\geq\kappa-\delta~{}~{}~{}\mathrm{for~{}all}~{}t\in[0,\tau],

where gi,tg_{i,t} is the Ricci-DeTurck flow starting at gi.g_{i}. Hence we can use this claim instead of Lemma 2.5 in the proof of Main Theorem 2. Therefore, using Lemma 2.1 instead of Lemma 2.2, we can prove the assertion in the same way as in the proof of Main Theorem 2. ∎

Remark 2.3.

If gig_{i} is a shrinking Ricci soliton (i.e., λi>0\lambda_{i}>0 in the above situation), then the maximal existence time of the corresponding Ricci flow is (2λi)1.(2\lambda_{i})^{-1}. But, since we assume gig_{i} converges to gg in the C0C^{0}-sense, Lemma 2.1 implicitly prevents λi\lambda_{i} being λi\lambda_{i}\rightarrow\infty as i.i\rightarrow\infty.

3 Some corollaries

In the assumption in Main Theorem 1, the nonnegativity of each scalar curvature can be replaced with general lower bound of each scalar curvature provided that a suitable condition about the volumes is additionally assumed.

Corollary 3.1 (for Main Theorem 1).

Let MnM^{n} be a closed nn-manifold and gg a C2C^{2}-Riemannian metric on M.M. Let (gi)(g_{i}) be a sequence of C2C^{2}-Riemannian metrics on MM that converges to gg on MM in the W1,pW^{1,p}-sense (p>n)(p>n). Assume that there are a constant κ\kappa\in\mathbb{R} and a continuous function σC0(M)\sigma\in C^{0}(M) such that for all i,i,

  • MR(gi)𝑑volgiκ,\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa,

  • R(gi)σR(g_{i})\geq\sigma on MM, and

  • Vol(M,gi)Vol(M,g),\mathrm{Vol}(M,g_{i})\geq\mathrm{Vol}(M,g),

Here Vol(M,g)\mathrm{Vol}(M,g) is the volume of MM with respect to g.g. Then

MR(g)𝑑volgκ.\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.
Remark 3.1.

If the limiting metric gg here is a hyperbolic metric (i.e., its sectional curvature is constant 1-1) and each metric gig_{i} has its volume entropy h(gi)n1(n3),h(g_{i})\leq n-1~{}(n\geq 3), then Vol(M,gi)Vol(M,g)\mathrm{Vol}(M,g_{i})\geq\mathrm{Vol}(M,g) for all i.i. This is the result of [3]. Here, the volume entropy of a closed Riemannian manifold (M,G)(M,G) is given by the limit

h(G)=limrlog(Vol(BG~(p0,r),G~))r,h(G)=\lim_{r\rightarrow\infty}\frac{\log\left(\mathrm{Vol}\left(B_{\tilde{G}}(p_{0},r),\tilde{G}\right)\right)}{r},

where Vol(BG~(p0,r),G~)\mathrm{Vol}\left(B_{\tilde{G}}(p_{0},r),\tilde{G}\right) is the volume of a ball of radius rr in the universal cover (M~,G~).(\tilde{M},\tilde{G}). (In particular, Bishop-Gromov’s volume comparison implies that a manifold with Ricci curvature lower bound RicG(n1)G\mathrm{Ric}_{G}\geq-(n-1)G satisfies h(G)n1.h(G)\leq n-1.) Moreover, as a deep corollary of Perelman’s resolution of the Geometrization conjecture of W. Thurston, a very strong generalization of this result [3] in dimension 3 holds (see [1]): if (M,g0)(M,g_{0}) is a closed hyperbolic 3-manifold, for any metric gg on M,M,

ifR(g)6thenVol(M,g)Vol(M,g0).\mathrm{if}~{}~{}R(g)\geq-6~{}~{}\mathrm{then}~{}~{}\mathrm{Vol}(M,g)\geq\mathrm{Vol}(M,g_{0}).
Proof of Corollary 3.1.

There is a sufficient large natural number nn\in\mathbb{N} such that the Riemannian product manifold (N:=M×Sn,giN:=gi×gstd)(N:=M\times S^{n},g^{N}_{i}:=g_{i}\times g_{\mathrm{std}}) satisfies R(giN)0,R(g^{N}_{i})\geq 0, where (Sn,gstd)(S^{n},g_{\mathrm{std}}) is the standard Riemannian nn-sphere of constant scalar curvature n(n1).n(n-1). Indeed, since

R(giN)=R(gi)+R(gstd)=R(gi)+n(n1)minMσ+n(n1),R(g^{N}_{i})=R(g_{i})+R(g_{\mathrm{std}})=R(g_{i})+n(n-1)\geq\min_{M}\sigma+n(n-1),

we see that R(giN)0R(g^{N}_{i})\geq 0 for sufficiently large n.n. Therefore we have

NR(giN)𝑑volgiN=M×Sn(R(gi)+R(gstd))𝑑volgi×gstd=Vol(Sn,gstd)MR(gi)𝑑volgi+Vol(M,gi)Vol(Sn,gstd)n(n1).\begin{split}\int_{N}R(g^{N}_{i})\,d\mathrm{vol}_{g^{N}_{i}}&=\int_{M\times S^{n}}\left(R(g_{i})+R(g_{\mathrm{std}})\right)\,d\mathrm{vol}_{g_{i}\times g_{\mathrm{std}}}\\ &=\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\\ &~{}~{}~{}~{}+\mathrm{Vol}(M,g_{i})\,\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,n(n-1).\end{split}

On the other hand, from our assumption, giN=gi×gstdg_{i}^{N}=g_{i}\times g_{\mathrm{std}} converges to g×gstdg\times g_{\mathrm{std}} in the W1,pW^{1,p}-sense and

NR(giN)𝑑volgiNVol(Sn,gstd)κ+Vol(M,g)Vol(Sn,gstd)n(n1).\int_{N}R(g^{N}_{i})\,d\mathrm{vol}_{g^{N}_{i}}\geq\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,\kappa+\mathrm{Vol}(M,g)\,\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,n(n-1).

Therefore, from Main Theorem 1,

Vol(Sn,gstd)MR(g)𝑑volg+Vol(M,g)Vol(Sn,gstd)n(n1)=NR(g×gstd)𝑑volg×gstdVol(Sn,gstd)κ+Vol(M,g)Vol(Sn,gstd)n(n1).\begin{split}\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,&\int_{M}R(g)\,d\mathrm{vol}_{g}+\mathrm{Vol}(M,g)\,\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,n(n-1)\\ &=\int_{N}R(g\times g_{\mathrm{std}})\,d\mathrm{vol}_{g\times g_{\mathrm{std}}}\\ &\geq\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,\kappa+\mathrm{Vol}(M,g)\,\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,n(n-1).\end{split}

Then, subtracting Vol(M,g)Vol(Sn,gstd)n(n1)\mathrm{Vol}(M,g)\,\mathrm{Vol}(S^{n},g_{\mathrm{std}})\,n(n-1) from both sides and dividing both sides by Vol(Sn,gstd)>0,\mathrm{Vol}(S^{n},g_{\mathrm{std}})>0, we obtain MR(g)𝑑volgκ.\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.

We also present another corollary of Corollary 3.1 here. In order to do this, we need to recall the definition of the Yamabe constant:

Definition 3.1 (Yamabe constant).

The Yamabe constant Y(M,g)Y(M,g) of a closed Riemannian manifold (M,g)(M,g) is defined as

Y(M,g):=inf{MR(g~)𝑑volg~|g~[g]andVol(M,g~)=1},Y(M,g):=\inf\left\{\int_{M}R(\tilde{g})\,d\mathrm{vol}_{\tilde{g}}~{}\middle|~{}\tilde{g}\in[g]~{}\mathrm{and}~{}\mathrm{Vol}(M,\tilde{g})=1\right\},

where [g]:={g~=u4n2g|uC(M),u>0onM}[g]:=\left\{\tilde{g}=u^{\frac{4}{n-2}}g~{}|~{}u\in C^{\infty}(M),~{}u>0~{}\mathrm{on}~{}M\right\} is the conformal class of the metric g.g. By the definition, Y(M,g)Y(M,g) depends only on the conformal class [g][g] of g.g. A Riemannian metric g~[g]\tilde{g}\in[g] with Vol(M,g~)=1\mathrm{Vol}(M,\tilde{g})=1 is called Yamabe metric of [g][g] if

Y(M,g)=MR(g~)𝑑volg~.Y(M,g)=\int_{M}R(\tilde{g})\,d\mathrm{vol}_{\tilde{g}}.
Corollary 3.2 (for Corollary 3.1).

Let MnM^{n} be a closed nn-manifold and gg a C2C^{2}-Riemannian metric on M.M. Let (gi)(g_{i}) be a sequence of C2C^{2}-Riemannian metrics on MM that converges to gg on MM in the W1,pW^{1,p}-sense (p>n)(p>n), and Vol(M,gi)=1.\mathrm{Vol}(M,g_{i})=1. Assume that gg is a Yamabe metric of [g],[g], and there are a constant κ\kappa\in\mathbb{R} and a continuous function σC0(M)\sigma\in C^{0}(M) such that for all i,i, Y(M,gi)κY(M,g_{i})\geq\kappa and R(gi)σR(g_{i})\geq\sigma on M.M. Then Y(M,g)κ.Y(M,g)\geq\kappa.

Proof.

From the definition of Y(M,gi)Y(M,g_{i}) and Vol(M,gi)=1,\mathrm{Vol}(M,g_{i})=1, we have

MR(gi)𝑑volgiκ.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq\kappa.

Since gigg_{i}\rightarrow g in the W1,pW^{1,p}-sense (p>n)(p>n) and Vol(M,gi)=1,\mathrm{Vol}(M,g_{i})=1, we also have Vol(M,g)=1.\mathrm{Vol}(M,g)=1. Hence, from Corollary 3.1, we have

MR(g)𝑑volgκ.\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.

Therefore, since gg is a Yamabe metric of [g][g] and Vol(M,g)=1,\mathrm{Vol}(M,g)=1, we obtain

Y(M,g)=MR(g)𝑑volgκ.Y(M,g)=\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\kappa.

Remark 3.2.

If σκ,\sigma\geq\kappa, this corollary directly follows from Gromov’s theorem (Theorem 1.1 in this paper). Indeed, since gigg_{i}\rightarrow g in the W1,pW^{1,p}-sense for p>np>n (in particular, in the C0C^{0}-sense), from Theorem 1.1, we have R(g)σ.R(g)\geq\sigma. On the other hand, since gg is a Yamabe metric of [g][g] (hence, its scalar curvature R(g)R(g) is constant) and Vol(M,g)=1,\mathrm{Vol}(M,g)=1, we have

Y(M,g)=MR(g)𝑑volgσκ.Y(M,g)=\int_{M}R(g)\,d\mathrm{vol}_{g}\geq\sigma\geq\kappa.

However, if σκ,\sigma\leq\kappa, this corollary does not follow from Gromov’s theorem.

Up to this point, we have only dealt with case that the underlying manifold is closed. Under a special situation, we can prove a similar statement as Claim 2.1 for a non-compact manifold.

Proposition 3.1 (Conformal deformations on an open manifold).

Let (Mn,g)(M^{n},g) be a non-compact Riemannian nn-manifold (n3n\geq 3) with MR(g)𝑑volg<+\int_{M}R(g)\,d\mathrm{vol}_{g}<+\infty and ui:Mu_{i}:M\rightarrow\mathbb{R} a sequence of positive C2C^{2}-functions. Assume that each uiu_{i} is equal to 1 outside a compact set and

ui1uniformlyC1senseonM.u_{i}\rightarrow 1~{}\mathrm{uniformly}~{}C^{1}\mathrm{-sense~{}on}~{}M.

Set gi:=ui4n2g.g_{i}:=u_{i}^{\frac{4}{n-2}}g. Then

MR(gi)𝑑volgiiMR(g)𝑑volg.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\overset{i\rightarrow\infty}{\longrightarrow}\int_{M}R(g)\,d\mathrm{vol}_{g}.
Proof.

From the formula for the scalar curvature and the volume form under this conformal change:

R(gi)=4n1n2uin+2n2Δgui+ui4n2R(g),volgi=ui2nn2volg,R(g_{i})=-4\frac{n-1}{n-2}u_{i}^{-\frac{n+2}{n-2}}\Delta_{g}u_{i}+u_{i}^{-\frac{4}{n-2}}R(g),~{}~{}~{}~{}\mathrm{vol}_{g_{i}}=u_{i}^{\frac{2n}{n-2}}\mathrm{vol}_{g},

we have

MR(gi)𝑑volgi=4n1n2Muin+2n2Δgui(ui2nn2dvolg)+Mui2n4n2R(g)𝑑volg=4n1n2MuiΔgui𝑑volg+Mui2n4n2R(g)𝑑volg=4n1n2M|gui|g2𝑑volg+Mui2n4n2R(g)𝑑volg.\begin{split}\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=-4\frac{n-1}{n-2}\int_{M}u_{i}^{-\frac{n+2}{n-2}}\Delta_{g}u_{i}\left(u_{i}^{\frac{2n}{n-2}}\,d\mathrm{vol}_{g}\right)+\int_{M}u_{i}^{\frac{2n-4}{n-2}}R(g)\,d\mathrm{vol}_{g}\\ &=-4\frac{n-1}{n-2}\int_{M}u_{i}\Delta_{g}u_{i}\,d\mathrm{vol}_{g}+\int_{M}u_{i}^{\frac{2n-4}{n-2}}R(g)\,d\mathrm{vol}_{g}\\ &=4\frac{n-1}{n-2}\int_{M}|\nabla_{g}u_{i}|_{g}^{2}d\mathrm{vol}_{g}+\int_{M}u_{i}^{\frac{2n-4}{n-2}}R(g)\,d\mathrm{vol}_{g}.\end{split}

We have used the divergence formula and the fact that uiu_{i} is equal to 1 outside a compact set in the third equality. Since uii1u_{i}\overset{i\rightarrow\infty}{\longrightarrow}1 C1C^{1}-uniformly on M,M, we have

M|gui|g2𝑑volg0asi\int_{M}|\nabla_{g}u_{i}|_{g}^{2}d\mathrm{vol}_{g}\rightarrow 0~{}~{}\mathrm{as}~{}i\rightarrow\infty

and

Mui2n4n2R(g)𝑑volgMR(g)𝑑volgasi\int_{M}u_{i}^{\frac{2n-4}{n-2}}R(g)\,d\mathrm{vol}_{g}\rightarrow\int_{M}R(g)\,d\mathrm{vol}_{g}~{}~{}\mathrm{as}~{}i\rightarrow\infty

Therefore we obtain the desired assertion from the above equality. ∎

The proof of Proposition 3.1 also provides the following.

Corollary 3.3.

In Main Theorem 1, if p=,p=\infty, and if for each i,i, gi=uigg_{i}=u_{i}\,g for some positive C2C^{2}-functions ui:M+,u_{i}:M\rightarrow\mathbb{R}_{+}, then the assumption R(gi)0(i=1,2,)R(g_{i})\geq 0~{}(i=1,2,\cdots) is not needed.

Lee and LeFloch [13] defined a notion of distributional scalar curvature for smooth manifolds that have a metric tensor which has only certain lower regularity.

Definition 3.2 (Distributional scalar curvature ([13, Definition 2.1], [11, Section 2])).

Let MM be a smooth manifold endowed with a smooth background metric h.h. Given any Riemannian metric gg with Lloc(M)Wloc1,2(M)L^{\infty}_{loc}(M)\cap W^{1,2}_{loc}(M) regularity and locally bounded inverse g1Lloc(M),g^{-1}\in L^{\infty}_{loc}(M), the scalar curvature distribution RgR_{g} is defined, for every compactly supported smooth test function u:Mu:M\rightarrow\mathbb{R} by

Rg,u:=M(V¯(udvolgdvolh)+Fudvolgdvolh)𝑑volh,\langle R_{g},u\rangle:=\int_{M}\left(-V\cdot\overline{\nabla}\left(u\frac{d\mathrm{vol}_{g}}{d\mathrm{vol}_{h}}\right)+Fu\frac{d\mathrm{vol}_{g}}{d\mathrm{vol}_{h}}\right)\,d\mathrm{vol}_{h},

where V=(Vk)Γ(M)V=(V^{k})\in\Gamma(M) is given by Vk:=gijΓijkgikΓjij,V^{k}:=g^{ij}\Gamma^{k}_{ij}-g^{ik}\Gamma^{j}_{ji}, FF is a function as

F:=Rh¯kgijΓijk+¯kgikΓjij+gij(ΓklkΓijlΓjlkΓikl)F:=R_{h}-\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 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). Here, ¯\overline{\nabla} denotes the Levi-Civita connection of h.h.

Let κ\kappa be a continuous function on M.M. We say that RgκR_{g}\geq\kappa in the distributional sense if Rg,uMκu𝑑volg0\langle R_{g},u\rangle-\int_{M}\kappa u\,d\mathrm{vol}_{g}\geq 0 for any nonnegative compactly supported test function uC+(M)C0(M).u\in C^{\infty}_{+}(M)\cap C^{\infty}_{0}(M).

Remark 3.3.

If a metric gg is C2,C^{2}, then the scalar curvature distribution Rg,u\langle R_{g},u\rangle coincides with MR(g)u𝑑volg.\int_{M}R(g)u\,d\mathrm{vol}_{g}.

For more details about the distributional scalar curvature and related results, see [11, 13, 20, 21]. From Gromov’s C0C^{0}-limit theorem (Theorem 1.1), there has already been a definition of scalar curvature lower bounds for C0C^{0} metrics (see [14, Definition 1.2] for example). Namely, a C0C^{0} metric gg on a smooth manifold MM is of R(g)κR(g)\geq\kappa on MM in the Gromov’s sense if and only if there exists a sequence of C2C^{2} metrics (gi)(g_{i}) such that gig_{i} converge C0C^{0}-locally to gg and satisfy R(gi)κR(g_{i})\geq\kappa on M.M. Note that Burkhardt-Guim [5] pointed out that her definition (via the Ricci–DeTurck flow) and this Gromov’s definition are actually equivalent on a closed manifold. For example, on tori, there is no metric gg which is of R(g)κ>0R(g)\geq\kappa>0 in the Gromov’s sense (or equivalently in the sense of [5, Definition 1.2]) from the resolution of Geroch’s conjecture [7, 16, 17]. In contrast, a metric gg which is of R(g)κ>0R(g)\geq\kappa>0 in the sense of the definition 3.2 might exist on a torus. At least on a manifold whose Yamabe invariant is nonpositive, the question of how different these definitions are is related to Schoen’s conjecture (cf. [11], see also the preprint, arXiv:2111.05582v2 by Lee and Tam). As a corollary of Main Theorem 2, we can obtain the following. This is the same as Corollary 1.2 in Section 1.

Corollary 3.4.

Let p>n2/2p>n^{2}/2 and κ\kappa a constant. Suppose that MM is a closed manifold of dimension n2,n\geq 2, gg is a C2C^{2} Riemannian metric on MM and (gi)(g_{i}) is a sequence of C2C^{2} metric on M.M. Assume the following:

  • (1)

    a sequence (ϕi)(\phi_{i}) of nonnegative smooth functions on MM satisfying: for any positive constant a>0a>0 there is a positive constant Λ>0\Lambda>0 such that log(ϕi+a)\log(\phi_{i}+a) is Λ\Lambda-Lipschitz on MM for all i,i,

  • (2)

    (ϕi)(\phi_{i}) converges to some nonnegative continuous function ϕ\phi in the uniformly C0C^{0}-sense on M,M,

  • (3)

    R(gi)0R(g_{i})\geq 0 on MM for each i,i,

  • (4)

    MR(gi)ϕi𝑑volgiκMϕi𝑑volgi,\int_{M}R(g_{i})\phi_{i}\,d\mathrm{vol}_{g_{i}}\geq\kappa\int_{M}\phi_{i}\,d\mathrm{vol}_{g_{i}},

  • (5)

    gig_{i} converges to gg in the W1,pW^{1,p}-sense.

Then

MR(g)ϕ𝑑volgκMϕ𝑑volg.\int_{M}R(g)\phi\,d\mathrm{vol}_{g}\geq\kappa\int_{M}\phi\,d\mathrm{vol}_{g}.
Proof.

From (3)(3) and (4),(4), for any (small) positive constant a>0,a>0,

MR(gi)(ϕi+a)𝑑volgiκMϕi𝑑volgi=κMϕ𝑑volg+κ(Mϕi𝑑volgiMϕ𝑑volg).\begin{split}\int_{M}R(g_{i})(\phi_{i}+a)\,d\mathrm{vol}_{g_{i}}&\geq\kappa\int_{M}\phi_{i}\,d\mathrm{vol}_{g_{i}}\\ &=\kappa\int_{M}\phi\,d\mathrm{vol}_{g}+\kappa\left(\int_{M}\phi_{i}\,d\mathrm{vol}_{g_{i}}-\int_{M}\phi\,d\mathrm{vol}_{g}\right).\end{split}

Then applying Main Theorem 2 to fi=log(ϕi+a)f_{i}=-\log(\phi_{i}+a) and f=log(ϕ+a),f=-\log(\phi+a), we obtain that

MR(g)(ϕ+a)𝑑volgiκMϕ𝑑volg+κδ\int_{M}R(g)(\phi+a)\,d\mathrm{vol}_{g_{i}}\geq\kappa\int_{M}\phi\,d\mathrm{vol}_{g}+\kappa\delta

for all 0<δ<<1.0<\delta<<1. Here a,δ>0a,\delta>0 can be taken to be arbitrarily small, hence we obtain the desired inequality:

MR(g)ϕ𝑑volgiκMϕ𝑑volg.\int_{M}R(g)\phi\,d\mathrm{vol}_{g_{i}}\geq\kappa\int_{M}\phi\,d\mathrm{vol}_{g}.

Question 3.1.

Can we replace the condition (3)(3) and (4)(4) with “R(gi)κR(g_{i})\geq\kappa in the distributional sense for some nonnegative constant κ0\kappa\geq 0”?

If we make the regularity of convergence much stronger W1,p(p>n2/2)W^{1,p}~{}(p>n^{2}/2) in [11, Theorem 3.2 (1)], then it can be proven that the above question is positively true.

4 Counterexamples

If MM is a compact smooth manifold and C2C^{2}-Riemannian metrics {gi}\{g_{i}\} converges to a C2C^{2}-Riemannian metric gg on MM in the C2C^{2} sense as ii\rightarrow\infty, then maxMR(gi)maxMR(g).\max_{M}R(g_{i})\rightarrow\max_{M}R(g). So, by Lebesgue’s dominated convergence theorem, we have

MR(gi)𝑑volgiMR(g)𝑑volgasi.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\rightarrow\int_{M}R(g)\,d\mathrm{vol}_{g}~{}~{}~{}~{}\mathrm{as}~{}~{}i\rightarrow\infty.

However, if MM is non-compact, it is not known that there is a Lebesgue integrable function f:Mf:M\rightarrow\mathbb{R} such that |R(gi)|f|R(g_{i})|\leq f a.e. on MM for all i.i. Hence in this situation, Lebesgue’s dominated convergence theorem cannot be applied in general. Indeed, the following example implies that Main Theorem 1 (without the assumption that each gig_{i} has nonnegative scalar curvature) does not hold in general if MM is non-compact.

Example 4.1 (CC^{\infty} locally uniformly convergence and incomplete limiting metric).

The limiting metric of the first example constructed below is incomplete on n(n3)\mathbb{R}^{n}~{}(n\geq 3). Consider the smooth positive function ui:nu_{i}:\mathbb{R}^{n}\rightarrow\mathbb{R} defined as

ui=ϕ(i1+leir2)+1u_{i}=\phi\left(i^{-1+l}e^{-ir^{2}}\right)+1

and (n,gi:=ui4n2gEucl)(n3).\left(\mathbb{R}^{n},~{}g_{i}:=u_{i}^{\frac{4}{n-2}}\cdot g_{Eucl}\right)~{}(n\geq 3). Here ϕ:n[0,1]\phi:\mathbb{R}^{n}\rightarrow[0,1] is a smooth cut-off function such that ϕ1\phi\equiv 1 on the closed ball Br0¯:={xn|r(x)r0}\overline{B_{r_{0}}}:=\{x\in\mathbb{R}^{n}|~{}r(x)\leq r_{0}\} and ϕ0\phi\equiv 0 outside of the ε/2\varepsilon/2-neighbourhood (Br0¯)ε/2\left(\overline{B_{r_{0}}}\right)_{\varepsilon/2} of Br0¯\overline{B_{r_{0}}} where r0>0r_{0}>0 is an arbitrarily fixed positive constant. Here, gEuclg_{Eucl} denotes the Euclidean metric on n,r:n0\mathbb{R}^{n},~{}r:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0} is the Euclidean distance function from the origin ono\in\mathbb{R}^{n}, and

l:=n+24when{n=2m+1(m1),n=2m(m2).l:=\frac{n+2}{4}~{}\mathrm{when}\begin{cases}n=2m+1~{}(m\geq 1),\\ n=2m~{}(m\geq 2).\end{cases}

Then for each i,i, (n,gi)(\mathbb{R}^{n},g_{i}) is a non-compact smooth Riemannian manifold with

R(gi)=uin+2n2(4n1n2ΔgEuclui+R(gEucl)ui)=uin+2n2(4n1n2ΔgEuclui).\begin{split}R(g_{i})&=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}+R(g_{Eucl})u_{i}\right)\\ &=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}\right).\end{split} (9)

Moreover, gig_{i} converges to gEculg_{Ecul} in the locally CC^{\infty}-sense in n{o}\mathbb{R}^{n}\setminus\{o\} but not in the C0C^{0}-sense on n.\mathbb{R}^{n}. Note that (n{o},gEucl)(\mathbb{R}^{n}\setminus\{o\},g_{Eucl}) is incomplete. On Br0¯,\overline{B_{r_{0}}},

|ui|2=j=1n|xji1+leir2|2=j=1n|rxjri1+leir2|2=j=1n|xjr(2)ilreir2|2=4i2lr2e2ir2.\begin{split}|\nabla u_{i}|^{2}=\sum^{n}_{j=1}\left|\frac{\partial}{\partial x^{j}}i^{-1+l}e^{-ir^{2}}\right|^{2}&=\sum^{n}_{j=1}\left|\frac{\partial r}{\partial x^{j}}\frac{\partial}{\partial r}i^{-1+l}e^{-ir^{2}}\right|^{2}\\ &=\sum^{n}_{j=1}\left|\frac{x^{j}}{r}(-2)i^{l}re^{-ir^{2}}\right|^{2}\\ &=4i^{2l}r^{2}e^{-2ir^{2}}.\end{split}

When ii\rightarrow\infty,

R(gi){0onBr0¯{o},ato.R(g_{i})\rightarrow\begin{cases}0&\mathrm{on}~{}\overline{B_{r_{0}}}\setminus\{o\},\\ \infty&\mathrm{at}~{}o.\end{cases}

Indeed, we can observe such a behavior from the form of the scalar curvature on Br0¯\overline{B_{r_{0}}} as follows.

R(gi)=4(n1)n2uin+2n2(2iln+4il+1r2)eir2R(g_{i})=-\frac{4(n-1)}{n-2}u_{i}^{-\frac{n+2}{n-2}}\left(-2i^{l}n+4i^{l+1}r^{2}\right)e^{-ir^{2}} (10)

From (9) and the divergence formula, we have

nR(gi)𝑑volgi=4n1n2nuin+2n2ΔgEuclui(ui2nn2dvolgEucl)=4n1n2nuiΔgEuclui𝑑volgEucl=4n1n2(Br0¯)ε|ui|2𝑑volgEucl16n1n2Br0¯i2lr2e2ir2𝑑volgEucl=16n1n2Vol(Sn1)0r0i2lr2e2ir2rn1𝑑r,\begin{split}\int_{\mathbb{R}^{n}}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}^{-\frac{n+2}{n-2}}\Delta_{g_{Eucl}}u_{i}\left(u_{i}^{\frac{2n}{n-2}}\,d\mathrm{vol}_{g_{Eucl}}\right)\\ &=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}\Delta_{g_{Eucl}}u_{i}\,d\mathrm{vol}_{g_{Eucl}}\\ &=4\frac{n-1}{n-2}\int_{\left(\overline{B_{r_{0}}}\right)_{\varepsilon}}|\nabla u_{i}|^{2}d\mathrm{vol}_{g_{Eucl}}\\ &\geq 16\frac{n-1}{n-2}\int_{\overline{B_{r_{0}}}}i^{2l}r^{2}e^{-2ir^{2}}d\mathrm{vol}_{g_{Eucl}}\\ &=16\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\int^{r_{0}}_{0}i^{2l}r^{2}e^{-2ir^{2}}r^{n-1}\,dr,\end{split} (11)

where Vol(Sn1)\mathrm{Vol}(S^{n-1}) denotes the volume of (n1)(n-1)-sphere with respect to the standard metric. Here,

0r0rn+1e2ir2𝑑r=[12irne2ir2]0r0+n2i0r0rn1e2ir2𝑑r=12ir0ne2ir02+n2i0r0rn1e2ir2𝑑r\begin{split}\int^{r_{0}}_{0}r^{n+1}e^{-2ir^{2}}dr&=\left[-\frac{1}{2i}r^{n}e^{-2ir^{2}}\right]^{r_{0}}_{0}+\frac{n}{2i}\int^{r_{0}}_{0}r^{n-1}e^{-2ir^{2}}dr\\ &=-\frac{1}{2i}r_{0}^{n}e^{-2ir_{0}^{2}}+\frac{n}{2i}\int^{r_{0}}_{0}r^{n-1}e^{-2ir^{2}}dr\end{split}

Set the left hand side of this equation as In+1:=0r0rn+1e2ir2𝑑r.I_{n+1}:=\int^{r_{0}}_{0}r^{n+1}e^{-2ir^{2}}dr. Then we have

{In+1=e2ir022ik=0ms=0k(n2s2i)+s=0m(2s+12i)I0ifn=2m+1(m1),In+1=e2ir022ik=0m1s=0k(n2s2i)+s=1m(2s2i)I1ifn=2m(m2).\begin{cases}I_{n+1}=-\frac{e^{-2ir_{0}^{2}}}{2i}\sum^{m}_{k=0}\prod^{k}_{s=0}\left(\frac{n-2s}{2i}\right)+\prod^{m}_{s=0}\left(\frac{2s+1}{2i}\right)\,I_{0}&\mathrm{if}~{}n=2m+1~{}(m\geq 1),\\ I_{n+1}=-\frac{e^{-2ir_{0}^{2}}}{2i}\sum^{m-1}_{k=0}\prod^{k}_{s=0}\left(\frac{n-2s}{2i}\right)+\prod^{m}_{s=1}\left(\frac{2s}{2i}\right)\,I_{1}&\mathrm{if}~{}n=2m~{}(m\geq 2).\end{cases}

Moreover,

I0=0r0e2ir2𝑑r(0r00π2e2ir2r𝑑r𝑑θ)1/2=π2(12ie2ir022i),\begin{split}I_{0}=\int^{r_{0}}_{0}e^{-2ir^{2}}\,dr&\geq\left(\int^{r_{0}}_{0}\int^{\frac{\pi}{2}}_{0}e^{-2ir^{2}}rdrd\theta\right)^{1/2}\\ &=\sqrt{\frac{\pi}{2}\left(\frac{1}{2i}-\frac{e^{-2ir_{0}^{2}}}{2i}\right)},\end{split}

and

I1=0r0re2ir2𝑑r=[12ie2ir2]0r0=12i12ie2ir02.I_{1}=\int^{r_{0}}_{0}re^{-2ir^{2}}\,dr=\left[-\frac{1}{2i}e^{-2ir^{2}}\right]^{r_{0}}_{0}=\frac{1}{2i}-\frac{1}{2i}e^{-2ir^{2}_{0}}.

Combining these, as i,i\rightarrow\infty, we can see that the rightmost term of (11) converges to

{16n1n2Vol(Sn1)π2s=0m(2s+12)(>0)ifn=2m+1(m1),8n1n2Vol(Sn1)s=1ms(>0)ifn=2m(m2).\begin{cases}16\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\frac{\sqrt{\pi}}{2}\prod_{s=0}^{m}\left(\frac{2s+1}{2}\right)\,(>0)&\mathrm{if}~{}n=2m+1~{}(m\geq 1),\\ 8\,\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\prod^{m}_{s=1}s\,(>0)&\mathrm{if}~{}n=2m~{}(m\geq 2).\end{cases}

Hence, for all sufficiently large i,i,

nR(gi)dvolgi{8n1n2Vol(Sn1)π2s=0m(2s+12)(>0)ifn=2m+1(m1),4n1n2Vol(Sn1)s=1ms(>0)ifn=2m(m2).\begin{split}\int_{\mathbb{R}^{n}}&R(g_{i})\,d\mathrm{vol}_{g_{i}}\\ &\geq\begin{cases}8\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\frac{\sqrt{\pi}}{2}\prod_{s=0}^{m}\left(\frac{2s+1}{2}\right)\,(>0)&\mathrm{if}~{}n=2m+1~{}(m\geq 1),\\ 4\,\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\prod^{m}_{s=1}s\,(>0)&\mathrm{if}~{}n=2m~{}(m\geq 2).\end{cases}\end{split}

Note that R(gi)R(g_{i}) cannot be non-negative on n\mathbb{R}^{n} by the positive mass theorem or the resolution of Geroch’s conjecture on tori [7, 16, 17]). Indeed, from (10),

R(gi)(o)=8n(n1)n2il(i1+l+1)n+2n2>0,R(g_{i})(o)=8\frac{n(n-1)}{n-2}i^{l}\left(i^{-1+l}+1\right)^{-\frac{n+2}{n-2}}~{}~{}>0,

and

R(gi)(x)=4n1n2il(i1+leir024+1)n+2n2(2n+ir02)eir024<0R(g_{i})(x)=-4\frac{n-1}{n-2}i^{l}\left(i^{-1+l}e^{-\frac{ir^{2}_{0}}{4}}+1\right)^{-\frac{n+2}{n-2}}\left(-2n+ir^{2}_{0}\right)\,e^{-\frac{ir^{2}_{0}}{4}}~{}~{}<0

for any point x{xn|r(x)=r02}x\in\left\{x\in\mathbb{R}^{n}|~{}r(x)=\frac{r_{0}}{2}\right\} and sufficiently large i.i.

Next, we will construct a counterexample (to Main Theorem 1 without the assumption that each gig_{i} has nonnegative scalar curvature), in which each gig_{i} is complete.

Example 4.2 (Not C1C^{1} but C0C^{0}).

Consider (n,gi:=ui4n2gEucl)(n3,i=2,3,).\left(\mathbb{R}^{n},~{}g_{i}:=u_{i}^{\frac{4}{n-2}}\cdot g_{Eucl}\right)~{}(n\geq 3,~{}i=2,3,\cdots). Here the smooth positive function ui:nu_{i}:\mathbb{R}^{n}\rightarrow\mathbb{R} has been defined as

ui=ϕ(i1sin(ir2))+1.u_{i}=\phi\left(i^{-1}\sin(ir^{2})\right)+1.

Here, ϕ:n[0,1]\phi:\mathbb{R}^{n}\rightarrow[0,1] is a smooth cut-off function such that ϕ1\phi\equiv 1 on Br0¯:={xn|r(x)r0}\overline{B_{r_{0}}}:=\{x\in\mathbb{R}^{n}|~{}r(x)\leq r_{0}\} and ϕ0\phi\equiv 0 outside of the ε/2\varepsilon/2-neighbourhood Br0¯.\overline{B_{r_{0}}}. where r0>0r_{0}>0 is an arbitrarily fixed positive constant. Here, r:n0r:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0} is the Euclidean distance function from the origin on.o\in\mathbb{R}^{n}. Then, for each i,i, (n,gi)(\mathbb{R}^{n},g_{i}) is a non-compact smooth Riemannian manifold with

R(gi)=uin+2n2(4n1n2ΔgEuclui+R(gEucl)ui)=uin+2n2(4n1n2ΔgEuclui),\begin{split}R(g_{i})&=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}+R(g_{Eucl})u_{i}\right)\\ &=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}\right),\end{split} (12)

and gigEculg_{i}\rightarrow g_{Ecul} on n\mathbb{R}^{n} in the uniformly C0C^{0} but not in the C1C^{1}-sense. On Br0¯,\overline{B_{r_{0}}},

|ui|2=j=1n|xji1sin(ir2)|2=j=1n|rxjri1sin(ir2)|2=j=1n|xjr(2r)cos(ir2)|2=4r2cos2(ir2)=2r2(1+cos(2ir2)).\begin{split}|\nabla u_{i}|^{2}=\sum^{n}_{j=1}\left|\frac{\partial}{\partial x^{j}}i^{-1}\sin(ir^{2})\right|^{2}&=\sum^{n}_{j=1}\left|\frac{\partial r}{\partial x^{j}}\frac{\partial}{\partial r}i^{-1}\sin(ir^{2})\right|^{2}\\ &=\sum^{n}_{j=1}\left|\frac{x^{j}}{r}(2r)\cos(ir^{2})\right|^{2}\\ &=4r^{2}\cos^{2}(ir^{2})\\ &=2r^{2}(1+\cos(2ir^{2})).\end{split}

Note that R(gi)R(g_{i}) is not non-negative on n.\mathbb{R}^{n}. Indeed, for sufficiently large ii and kk\in\mathbb{Z} such that Br0¯{xn|r(x)=(2k1)2i},\overline{B_{r_{0}}}\cap\left\{x\in\mathbb{R}^{n}|~{}r(x)=\sqrt{\frac{(2k-1)}{2i}}\right\}\neq\emptyset, we can take a point xiBr0¯{xn|r(x)=(2k1)2i}.x_{i}\in\overline{B_{r_{0}}}\cap\left\{x\in\mathbb{R}^{n}|~{}r(x)=\sqrt{\frac{(2k-1)}{2i}}\right\}. Then

R(gi)(xi){8(n1)(2k1)πn2ifkisodd,8(n1)(2k1)πn2ifkiseven.R(g_{i})(x_{i})\rightarrow\begin{cases}\frac{8(n-1)(2k-1)\pi}{n-2}&\mathrm{if}~{}k~{}\mathrm{is~{}odd},\\ -\frac{8(n-1)(2k-1)\pi}{n-2}&\mathrm{if}~{}k~{}\mathrm{is~{}even}.\end{cases}

This is checked as follows. For sufficiently large i,i, such a point xix_{i} is contained in Br0¯.\overline{B_{r_{0}}}. Hence, from the above formula and the choice of the point xi,x_{i}, we have

R(gi)(xi)=uin+2n2(4n1n2ΔgEuclui)=4n1n2(i1+1)n+2n2(2ncos(ir(xi)2)4ir2sin(ir(xi)2))=(1)k+18(n1)(2k1)πn2(i1+1)n+2n2.\begin{split}R(g_{i})(x_{i})&=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}\right)\\ &=-4\frac{n-1}{n-2}\left(i^{-1}+1\right)^{-\frac{n+2}{n-2}}\left(2n\cos(ir(x_{i})^{2})-4ir^{2}\sin(ir(x_{i})^{2})\right)\\ &=(-1)^{k+1}\frac{8(n-1)(2k-1)\pi}{n-2}\left(i^{-1}+1\right)^{-\frac{n+2}{n-2}}.\end{split}

Since (i1+1)n+2n21(i),\left(i^{-1}+1\right)^{-\frac{n+2}{n-2}}\rightarrow 1~{}(i\rightarrow\infty), we can observe the desired behavior of the scalar curvature as above. Moreover, from (12) and the divergence formula, we have

nR(gi)𝑑volgi=4n1n2nuin+2n2ΔgEuclui(ui2nn2dvolgEucl)=4n1n2nuiΔgEuclui𝑑volgEucl=4n1n2(Br0¯)ε|ui|2𝑑volgEucl8n1n2Br0¯r2(1+cos(2ir2))𝑑volgEucl=8n1n2Vol(Sn1)0r0r2(1+cos(2ir2))rn1𝑑r.\begin{split}\int_{\mathbb{R}^{n}}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}^{-\frac{n+2}{n-2}}\Delta_{g_{Eucl}}u_{i}\left(u_{i}^{\frac{2n}{n-2}}\,d\mathrm{vol}_{g_{Eucl}}\right)\\ &=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}\Delta_{g_{Eucl}}u_{i}\,d\mathrm{vol}_{g_{Eucl}}\\ &=4\frac{n-1}{n-2}\int_{\left(\overline{B_{r_{0}}}\right)_{\varepsilon}}|\nabla u_{i}|^{2}d\mathrm{vol}_{g_{Eucl}}\\ &\geq 8\frac{n-1}{n-2}\int_{\overline{B_{r_{0}}}}r^{2}(1+\cos(2ir^{2}))d\mathrm{vol}_{g_{Eucl}}\\ &=8\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\int_{0}^{r_{0}}r^{2}(1+\cos(2ir^{2}))r^{n-1}\,dr.\end{split}

Here,

0r0r2(1+cos(2ir2))rn1dr=[1n+2rn+2]0r0+0r0rn+1cos(2ir2)𝑑r=1n+2r0n+2+[rni14sin(2ir2)]0r0+ni216[rn2cos(2ir2)]0r0n(n2)i2160r0rn3cos(2ir2)𝑑r1n+2r0n+2+[rni14sin(2ir2)]0r0+ni216[rn2cos(2ir2)]0r0n(n2)i1160r0rn3𝑑r=1n+2r0n+2+r0ni14sin(2ir02)+ni216r0n2cos(2ir02)ni216r0n2.\begin{split}\int_{0}^{r_{0}}&r^{2}(1+\cos(2ir^{2}))r^{n-1}dr=\left[\frac{1}{n+2}r^{n+2}\right]_{0}^{r_{0}}+\int_{0}^{r_{0}}r^{n+1}\cos(2ir^{2})\,dr\\ &=\frac{1}{n+2}r^{n+2}_{0}+\left[\frac{r^{n}i^{-1}}{4}\sin(2ir^{2})\right]^{r_{0}}_{0}+\frac{ni^{-2}}{16}\left[r^{n-2}\cos(2ir^{2})\right]^{r_{0}}_{0}\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-\frac{n(n-2)i^{-2}}{16}\int^{r_{0}}_{0}r^{n-3}\cos(2ir^{2})\,dr\\ &\geq\frac{1}{n+2}r^{n+2}_{0}+\left[\frac{r^{n}i^{-1}}{4}\sin(2ir^{2})\right]^{r_{0}}_{0}+\frac{ni^{-2}}{16}\left[r^{n-2}\cos(2ir^{2})\right]^{r_{0}}_{0}\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-\frac{n(n-2)i^{-1}}{16}\int^{r_{0}}_{0}r^{n-3}\,dr\\ &=\frac{1}{n+2}r^{n+2}_{0}+\frac{r_{0}^{n}i^{-1}}{4}\sin(2ir_{0}^{2})+\frac{ni^{-2}}{16}\,r_{0}^{n-2}\cos(2ir_{0}^{2})-\frac{ni^{-2}}{16}r_{0}^{n-2}.\end{split}

Since

r0ni14sin(2ir02)+ni216r0n2cos(2ir02)ni216r0n20asi,\frac{r_{0}^{n}i^{-1}}{4}\sin(2ir_{0}^{2})+\frac{ni^{-2}}{16}\,r_{0}^{n-2}\cos(2ir_{0}^{2})-\frac{ni^{-2}}{16}r_{0}^{n-2}\rightarrow 0~{}~{}\mathrm{as}~{}i\rightarrow\infty,

there is a sufficiently large i0=i0(n,r0)i_{0}=i_{0}(n,r_{0}) such that for all ii0,i\geq i_{0},

r0ni14sin(2ir02)+ni216r0n2cos(2ir02)ni216r0n2>12(n+2)r0n+2.\frac{r_{0}^{n}i^{-1}}{4}\sin(2ir_{0}^{2})+\frac{ni^{-2}}{16}\,r_{0}^{n-2}\cos(2ir_{0}^{2})-\frac{ni^{-2}}{16}r_{0}^{n-2}>-\frac{1}{2(n+2)}r^{n+2}_{0}.

Hence, for all ii0,i\geq i_{0},

nR(gi)𝑑volgi>4n1(n+2)(n2)Vol(Sn1)r0n+2>0.\int_{\mathbb{R}^{n}}R(g_{i})\,d\mathrm{vol}_{g_{i}}>4\frac{n-1}{(n+2)(n-2)}\mathrm{Vol}(S^{n-1})\,r_{0}^{n+2}>0.

From the Morrey embedding, we have

C1W1,pC0,1npC0ifp>n.C^{1}\hookrightarrow W^{1,p}\hookrightarrow C^{0,1-\frac{n}{p}}\hookrightarrow C^{0}~{}~{}~{}\mathrm{if}~{}p>n.

Therefore the same statement of Main Theorem 1 still holds even though one replace W1,p(p>n)W^{1,p}~{}(p>n) with C0,αC^{0,\alpha} for all α(0,1].\alpha\in(0,1]. On the other hand, in Main Theorem 1, if we weaken the assumption from W1,pW^{1,p} to C0,C^{0}, then the same statement (without the assumption R(gi)0R(g_{i})\geq 0) does not hold in general. Indeed, using the same local construction as in the previous example in dimension 3,\geq 3, we can also construct a counterexample on a closed manifold to Main Theorem 1 (without the assumption that each gig_{i} has nonnegative scalar curvature) as follows. Note that all metrics gig_{i} in each such example has sign-changing scalar curvature, i.e., for each i,i, there are some points xi,yiMx_{i},y_{i}\in M s.t. R(gi)(xi)<0<R(gi)(yi).R(g_{i})(x_{i})<0<R(g_{i})(y_{i}).

Example 4.3 (On every closed manifold).

Consider (Mn,gi:=ui4n2g0)\left(M^{n},~{}g_{i}:=u_{i}^{\frac{4}{n-2}}\cdot g_{0}\right) (n3,i=2,3,),(n\geq 3,~{}i=2,3,\cdots), where MnM^{n} is a closed nn-manifold and g0g_{0} is a Riemannian metric on M.M. Here the smooth positive function ui:Mu_{i}:M\rightarrow\mathbb{R} has been defined as

ui=ϕ(i1sin(ih2))+1.u_{i}=\phi\left(i^{-1}\sin(ih^{2})\right)+1.

Here, ϕ:M[0,1]\phi:M\rightarrow[0,1] is a smooth cut-off function such that ϕ1\phi\equiv 1 on Br0¯(p):={xM|dg0(p,x)r0}\overline{B_{r_{0}}}(p):=\{x\in M|~{}d_{g_{0}}(p,x)\leq r_{0}\} and ϕ0\phi\equiv 0 outside of the ε/2\varepsilon/2-neighbourhood Br0¯\overline{B_{r_{0}}} for some point pMp\in M where 0<r0<inj(M,g0)0<r_{0}<\mathrm{inj}(M,g_{0}) is a sufficiently small positive constant. Here, h:=dg0(,p):M0h:=d_{g_{0}}(\cdot,p):M\rightarrow\mathbb{R}_{\geq 0} is the distance function of g0g_{0} from the point pp and inj(M,g0)\mathrm{inj}(M,g_{0}) is the injectivity radius of (M,g0).(M,g_{0}). Then, for each i,i, (M,gi)(M,g_{i}) is a smooth Riemannian manifold with

R(gi)=uin+2n2(4n1n2Δg0ui+R(g0)ui)R(g_{i})=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{0}}u_{i}+R(g_{0})u_{i}\right)

and gig_{i} converges to g0g_{0} on MM in the C0C^{0} but not in the C1C^{1}-sense. In the same calculation as in the previous example, we have

MR(gi)dvolgi=4n1n2Muin+2n2(Δg0ui+Rg0ui)ui2nn2𝑑volg0=4n1n2M(uiΔg0ui+Rg0ui2)𝑑volg0=4n1n2(Br0¯)ε|ui|g02𝑑volg0+MR(g0)ui2𝑑volg08n1n2Br0¯h2(1+cos(2ih2))𝑑volg0+MR(g0)ui2𝑑volg08n1n2C~0r0h2(1+cos(2ih2))hn1𝑑h+MR(g0)ui2𝑑volg0.\begin{split}\int_{M}&R(g_{i})\,d\mathrm{vol}_{g_{i}}\\ &=-4\frac{n-1}{n-2}\int_{M}u_{i}^{-\frac{n+2}{n-2}}\left(\Delta_{g_{0}}u_{i}+R_{g_{0}}u_{i}\right)\,u_{i}^{\frac{2n}{n-2}}\,d\mathrm{vol}_{g_{0}}\\ &=-4\frac{n-1}{n-2}\int_{M}\left(u_{i}\Delta_{g_{0}}u_{i}+R_{g_{0}}u_{i}^{2}\right)\,d\mathrm{vol}_{g_{0}}\\ &=4\frac{n-1}{n-2}\int_{\left(\overline{B_{r_{0}}}\right)_{\varepsilon}}|\nabla u_{i}|_{g_{0}}^{2}d\mathrm{vol}_{g_{0}}+\int_{M}R(g_{0})u_{i}^{2}\,d\mathrm{vol}_{g_{0}}\\ &\geq 8\frac{n-1}{n-2}\int_{\overline{B_{r_{0}}}}h^{2}(1+\cos(2ih^{2}))d\mathrm{vol}_{g_{0}}+\int_{M}R(g_{0})u_{i}^{2}\,d\mathrm{vol}_{g_{0}}\\ &\geq 8\frac{n-1}{n-2}\widetilde{C}\int_{0}^{r_{0}}h^{2}(1+\cos(2ih^{2}))h^{n-1}\,dh+\int_{M}R(g_{0})u_{i}^{2}\,d\mathrm{vol}_{g_{0}}.\end{split}

Here, the constant C~\widetilde{C} depends only on nn and g0.g_{0}. Thus, from the observation as in the previous example, there is i0i_{0}\in\mathbb{N} and a positive constant C=C(n,g0,r0)>0C=C(n,g_{0},r_{0})>0 such that for all ii0,i\geq i_{0},

8n1n2C~0r0h2(1+cos(2ih2))hn1𝑑hC>0.8\frac{n-1}{n-2}\widetilde{C}\int_{0}^{r_{0}}h^{2}(1+\cos(2ih^{2}))h^{n-1}\,dh\geq C>0.

Therefore, for all ii0,i\geq i_{0},

MR(gi)𝑑volgiC+MR(g0)ui2𝑑volg0.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq C+\int_{M}R(g_{0})u_{i}^{2}\,d\mathrm{vol}_{g_{0}}.

Moreover, by the definition of ui,u_{i},

|MR(g0)ui2𝑑volg0MR(g0)𝑑volg0|M|R(g0)|(2i1+i2)𝑑volg0.\left|\int_{M}R(g_{0})u^{2}_{i}\,d\mathrm{vol}_{g_{0}}-\int_{M}R(g_{0})\,d\mathrm{vol}_{g_{0}}\right|\leq\int_{M}\left|R(g_{0})\right|\left(2i^{-1}+i^{-2}\right)\,d\mathrm{vol}_{g_{0}}.

Hence there is a sufficiently large i1i_{1} such that for all ii1,i\geq i_{1},

|MR(g0)ui2𝑑volg0MR(g0)𝑑volg0|C2.\left|\int_{M}R(g_{0})u^{2}_{i}\,d\mathrm{vol}_{g_{0}}-\int_{M}R(g_{0})\,d\mathrm{vol}_{g_{0}}\right|\leq\frac{C}{2}.

Thus, for all imax{i0,i1},i\geq\max\{i_{0},i_{1}\}, we have

MR(gi)𝑑volgiC+MR(g0)ui2𝑑volg0>MR(g0)𝑑volg0.\int_{M}R(g_{i})\,d\mathrm{vol}_{g_{i}}\geq C+\int_{M}R(g_{0})u_{i}^{2}\,d\mathrm{vol}_{g_{0}}>\int_{M}R(g_{0})\,d\mathrm{vol}_{g_{0}}.

Here, we have a question about the regularity of convergence in the assumption of Main Theorem 2.

Question 4.1.

Are there any C2C^{2}-metrics (gi)(g_{i}) and Λ\Lambda-Lipschitz (Λ>0\Lambda>0) functions (fi)(f_{i}) on a closed nn-manifold Mn(n3)M^{n}~{}(n\geq 3) satisfying the followings ?

  • gig_{i} converges to a C2C^{2}-metric gg in the W1,n22W^{1,\frac{n^{2}}{2}}-sense,

  • fif_{i} converges to a Λ\Lambda-Lipschitz function ff in the uniformly C0C^{0}-sense,

  • there is a constant κ\kappa such that MR(gi)efi𝑑volgiκ>MR(g)ef𝑑volg.\int_{M}R(g_{i})\,e^{-f_{i}}d\mathrm{vol}_{{g}_{i}}\geq\kappa>\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g}.

Or, additionally,

  • there is a point piMp_{i}\in M for each ii such that R(gi)(pi)R(g_{i})(p_{i})\rightarrow-\infty as i.i\rightarrow\infty.

In the following Example 4.4, we give another counterexample which is similar to the one in Example 4.2 (n3n\geq 3). However, in the following example of dimension 3,\geq 3, the support of ui1u_{i}-1 ((n,dgEucl)\subset(\mathbb{R}^{n},d_{g_{Eucl}})) with the origin ono\in\mathbb{R}^{n} converges to (n,gEucl,o)(\mathbb{R}^{n},g_{Eucl},o) in the pointed Gromov-Hausdorff sense as i.i\rightarrow\infty. (In Example 4.2, the support of ui1u_{i}-1 has been contained in a fixed compact subset.) Hence, unfortunately it is not possible to localize this construction directly and construct such a counterexample on a closed manifold as in Example 4.3. On the other hand, in the two-dimensional example of Example 4.4, the support of eui1e^{u_{i}}-1 is not compact for each ii (see the last half of Example 4.4).

Example 4.4 (Not C2C^{2} but C1C^{1}).

Consider (n,gi:=ui4n2gEucl)(n3,i=2,3,).\left(\mathbb{R}^{n},~{}g_{i}:=u_{i}^{\frac{4}{n-2}}\cdot g_{Eucl}\right)~{}(n\geq 3,~{}i=2,3,\cdots). Here the smooth positive function ui:nu_{i}:\mathbb{R}^{n}\rightarrow\mathbb{R} has been defined as

ui=ϕi(i2sin(ir2))+1.u_{i}=\phi_{i}\left(i^{-2}\sin(ir^{2})\right)+1.

Here, ϕi:n[0,1]\phi_{i}:\mathbb{R}^{n}\rightarrow[0,1] is a smooth cut-off function such that ϕi1\phi_{i}\equiv 1 on Bri¯:={xn|r(x)ri}\overline{B_{r_{i}}}:=\{x\in\mathbb{R}^{n}|~{}r(x)\leq r_{i}\} and ϕi0\phi_{i}\equiv 0 outside of the ε/2\varepsilon/2-neighbourhood Bri¯.\overline{B_{r_{i}}}. where ri:=i2n+2.r_{i}:=i^{\frac{2}{n+2}}. Note that rir_{i}\rightarrow\infty as i.i\rightarrow\infty. Here, r:n0r:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0} is the Euclidean distance function from the origin o.o. Then, for each i,i, (n,gi)(\mathbb{R}^{n},g_{i}) is a non-compact smooth Riemannian manifold with

R(gi)=uin+2n2(4n1n2ΔgEuclui+R(gEucl)ui)=uin+2n2(4n1n2ΔgEuclui).\begin{split}R(g_{i})&=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}+R(g_{Eucl})u_{i}\right)\\ &=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}\right).\end{split}

Moreover, gig_{i} converges to gEculg_{Ecul} on n\mathbb{R}^{n} in the uniformly C1C^{1} but not in the C2C^{2}-sense. On Bri¯,\overline{B_{r_{i}}},

|ui|2=j=1n|xji2sin(ir2)|2=j=1n|rxjri2sin(ir2)|2=j=1n|xjr(2i1r)cos(ir2)|2=4i2r2cos2(ir2)=2i2r2(1+cos(2ir2)).\begin{split}|\nabla u_{i}|^{2}=\sum^{n}_{j=1}\left|\frac{\partial}{\partial x^{j}}i^{-2}\sin(ir^{2})\right|^{2}&=\sum^{n}_{j=1}\left|\frac{\partial r}{\partial x^{j}}\frac{\partial}{\partial r}i^{-2}\sin(ir^{2})\right|^{2}\\ &=\sum^{n}_{j=1}\left|\frac{x^{j}}{r}(2i^{-1}r)\cos(ir^{2})\right|^{2}\\ &=4i^{-2}r^{2}\cos^{2}(ir^{2})\\ &=2i^{-2}r^{2}(1+\cos(2ir^{2})).\end{split}

Note that R(gi)R(g_{i}) is not non-negative on n.\mathbb{R}^{n}. Indeed, when i,R(x)i\rightarrow\infty,~{}R(x) oscillates for each x{xn|r(x)0}.x\in\{x\in\mathbb{R}^{n}|~{}r(x)\neq 0\}. Indeed, for sufficiently large i,i, such a point xx is contained in Bri¯.\overline{B_{r_{i}}}. Hence, from the above formula,

R(gi)(x)=uin+2n2(4n1n2ΔgEuclui)=4(n1n2)2ni1cos(ir(x)2)4r(x)2sin(ir(x)2)(i2sin(ir(x)2)+1)n+2n2.\begin{split}R(g_{i})(x)&=u_{i}^{-\frac{n+2}{n-2}}\left(-4\frac{n-1}{n-2}\Delta_{g_{Eucl}}u_{i}\right)\\ &=-4\left(\frac{n-1}{n-2}\right)\frac{2ni^{-1}\cos(ir(x)^{2})-4r(x)^{2}\sin(ir(x)^{2})}{\left(i^{-2}\sin(ir(x)^{2})+1\right)^{\frac{n+2}{n-2}}}.\end{split}

Since (i1+1)n+2n21\left(i^{-1}+1\right)^{-\frac{n+2}{n-2}}\rightarrow 1 and 2ni1cos(ir(x)2)02ni^{-1}\cos(ir(x)^{2})\rightarrow 0 as i,i\rightarrow\infty, we can easily observe the desired behavior of the scalar curvature. Moreover, by the divergence formula,

nR(gi)𝑑volgi=4n1n2nuin+2n2ΔgEuclui(ui2nn2dvolgEucl)=4n1n2nuiΔgEuclui𝑑volgEucl=4n1n2(Bri¯)ε|ui|2𝑑volgEucl8n1n2Bri¯r2i2(1+cos(2ir2))𝑑volgEucl=8n1n2Vol(Sn1)0rir2i2(1+cos(2ir2))rn1𝑑r.\begin{split}\int_{\mathbb{R}^{n}}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}^{-\frac{n+2}{n-2}}\Delta_{g_{Eucl}}u_{i}\left(u_{i}^{\frac{2n}{n-2}}\,d\mathrm{vol}_{g_{Eucl}}\right)\\ &=-4\frac{n-1}{n-2}\int_{\mathbb{R}^{n}}u_{i}\Delta_{g_{Eucl}}u_{i}\,d\mathrm{vol}_{g_{Eucl}}\\ &=4\frac{n-1}{n-2}\int_{\left(\overline{B_{r_{i}}}\right)_{\varepsilon}}|\nabla u_{i}|^{2}d\mathrm{vol}_{g_{Eucl}}\\ &\geq 8\frac{n-1}{n-2}\int_{\overline{B_{r_{i}}}}r^{2}i^{-2}(1+\cos(2ir^{2}))d\mathrm{vol}_{g_{Eucl}}\\ &=8\frac{n-1}{n-2}\mathrm{Vol}(S^{n-1})\int_{0}^{r_{i}}r^{2}i^{-2}(1+\cos(2ir^{2}))r^{n-1}\,dr.\end{split}

Here,

0rii2r2(1+cos(2ir2))rn1dr=[i2n+2rn+2]0ri+0rii2rn+1cos(2ir2)𝑑r=1n+2+[rni34sin(2ir2)]0ri+ni416[rn2cos(2ir2)]0rin(n2)i4160rirn3cos(2ir2)𝑑r1n+2+[rni34sin(2ir2)]0ri+ni416[rn2cos(2ir2)]0rin(n2)i4160rirn3𝑑r=1n+2+rini34sin(2iri2)+ni416rin2cos(2iri2)ni416rin21n+2+i2nn+2i34sin(2i1+4n+2)+ni416i2(n2)n+2cos(2i1+4n+2)ni416i2(n2)n+2.\begin{split}\int_{0}^{r_{i}}&i^{-2}r^{2}(1+\cos(2ir^{2}))r^{n-1}dr=\left[\frac{i^{-2}}{n+2}r^{n+2}\right]_{0}^{r_{i}}+\int_{0}^{r_{i}}i^{-2}r^{n+1}\cos(2ir^{2})\,dr\\ &=\frac{1}{n+2}+\left[\frac{r^{n}i^{-3}}{4}\sin(2ir^{2})\right]^{r_{i}}_{0}+\frac{ni^{-4}}{16}\left[r^{n-2}\cos(2ir^{2})\right]^{r_{i}}_{0}\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-\frac{n(n-2)i^{-4}}{16}\int^{r_{i}}_{0}r^{n-3}\cos(2ir^{2})\,dr\\ &\geq\frac{1}{n+2}+\left[\frac{r^{n}i^{-3}}{4}\sin(2ir^{2})\right]^{r_{i}}_{0}+\frac{ni^{-4}}{16}\left[r^{n-2}\cos(2ir^{2})\right]^{r_{i}}_{0}\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-\frac{n(n-2)i^{-4}}{16}\int^{r_{i}}_{0}r^{n-3}\,dr\\ &=\frac{1}{n+2}+\frac{r_{i}^{n}i^{-3}}{4}\sin(2ir_{i}^{2})+\frac{ni^{-4}}{16}\,r_{i}^{n-2}\cos(2ir_{i}^{2})-\frac{ni^{-4}}{16}r_{i}^{n-2}\\ &\geq\frac{1}{n+2}+\frac{i^{\frac{2n}{n+2}}i^{-3}}{4}\sin(2i^{1+\frac{4}{n+2}})+\frac{ni^{-4}}{16}\,i^{\frac{2(n-2)}{n+2}}\cos(2i^{1+\frac{4}{n+2}})-\frac{ni^{-4}}{16}i^{\frac{2(n-2)}{n+2}}.\end{split}

Since

i2nn+2i34sin(2i1+4n+2)+ni416i2(n2)n+2cos(2i1+4n+2)ni416i2(n2)n+20asi,\frac{i^{\frac{2n}{n+2}}i^{-3}}{4}\sin(2i^{1+\frac{4}{n+2}})+\frac{ni^{-4}}{16}\,i^{\frac{2(n-2)}{n+2}}\cos(2i^{1+\frac{4}{n+2}})-\frac{ni^{-4}}{16}i^{\frac{2(n-2)}{n+2}}\rightarrow 0~{}~{}\mathrm{as}~{}i\rightarrow\infty,

there is a sufficiently large i0=i0(n)i_{0}=i_{0}(n) such that for all ii0,i\geq i_{0},

i2nn+2i34sin(2i1+4n+2)+ni416i2(n2)n+2cos(2i1+4n+2)ni416i2(n)n+2>12(n+2).\frac{i^{\frac{2n}{n+2}}i^{-3}}{4}\sin(2i^{1+\frac{4}{n+2}})+\frac{ni^{-4}}{16}\,i^{\frac{2(n-2)}{n+2}}\cos(2i^{1+\frac{4}{n+2}})-\frac{ni^{-4}}{16}i^{\frac{2(n-)}{n+2}}>-\frac{1}{2(n+2)}.

Hence for all ii0,i\geq i_{0},

nR(gi)𝑑volgi>4n1(n+2)(n2)Vol(Sn1)>0.\int_{\mathbb{R}^{n}}R(g_{i})\,d\mathrm{vol}_{g_{i}}>4\frac{n-1}{(n+2)(n-2)}\mathrm{Vol}(S^{n-1})>0.

Next, we will construct a two-dimensional example. Consider the smooth function uiu_{i} on 2\mathbb{R}^{2} defined by

ui:=eir2sin(i2r2)(i=1,2,),u_{i}:=e^{-ir^{2}}\sin\left(-\frac{i}{2}r^{2}\right)~{}~{}(i=1,2,\cdots),

where r():=|o|r(\cdot):=|o-\cdot| denotes the Euclidean distance function from the origin o2.o\in\mathbb{R}^{2}. Then uiu_{i} uniformly converges to the constant function 0 in the C1C^{1} topology on 2\mathbb{R}^{2}, but uiu_{i} does not converge to 0 in the C2C^{2} topology on 2\mathbb{R}^{2}. Hence the sequence of complete metrics (gi:=euigEucl)(g_{i}:=e^{u_{i}}g_{Eucl}) on 2\mathbb{R}^{2} uniformly converges to gEuclg_{Eucl} in the C1C^{1} sense on 2,\mathbb{R}^{2}, but gig_{i} does not converge to gEuclg_{Eucl} in the C2C^{2} sense on 2.\mathbb{R}^{2}. Set a:=i,b:=12a=i2.a:=-i,b:=\frac{1}{2}a=-\frac{i}{2}. Then we can check that

Δui=(4a+4a2r2)ear2sin(br2)+(4bcosr24b2r2sinbr2)ear2+8abr2ear2cosbr2,\begin{split}\Delta u_{i}&=(4a+4a^{2}r^{2})e^{ar^{2}}\sin(br^{2})+(4b\cos r^{2}-4b^{2}r^{2}\sin br^{2})e^{ar^{2}}\\ &~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+8abr^{2}e^{ar^{2}}\cos br^{2},\end{split}

and

2R(gi)𝑑volgi=2ΔgEuclui𝑑volgEcul=02π0rΔgEuclui𝑑r𝑑θ.\int_{\mathbb{R}^{2}}R(g_{i})\,d\mathrm{vol}_{g_{i}}=-\int_{\mathbb{R}^{2}}\Delta_{g_{Eucl}}u_{i}\,d\mathrm{vol}_{g_{Ecul}}=-\int^{2\pi}_{0}\int_{0}^{\infty}r\Delta_{g_{Eucl}}u_{i}\,dr\,d\theta.

Moreover,

  • I:=0rear2cosbr2dr=a2(a2+b2),I:=\int^{\infty}_{0}re^{ar^{2}}\cos br^{2}\,dr=-\frac{a}{2(a^{2}+b^{2})},

  • J:=0rear2sinbr2dr=b2(a2+b2),J:=\int^{\infty}_{0}re^{ar^{2}}\sin br^{2}\,dr=-\frac{b}{2(a^{2}+b^{2})},

  • K:=0r3ear2cosbr2dr=aI+bJa2+b2,K:=\int^{\infty}_{0}r^{3}e^{ar^{2}}\cos br^{2}\,dr=-\frac{aI+bJ}{a^{2}+b^{2}},

  • L:=0r3ear2sinbr2dr=aJbIa2+b2.L:=\int^{\infty}_{0}r^{3}e^{ar^{2}}\sin br^{2}\,dr=-\frac{aJ-bI}{a^{2}+b^{2}}.

Combining these, we obtain that

2R(gi)𝑑volgi=2π(4bI+4aJ4(a2b2)aJbIa2+b28abbJaIa2+b2)=2π8a3b(a2+b2)2=128π25>0=2R(gEucl)𝑑volgEcul.\begin{split}\int_{\mathbb{R}^{2}}R(g_{i})\,d\mathrm{vol}_{g_{i}}&=-2\pi\left(4bI+4aJ-4(a^{2}-b^{2})\frac{aJ-bI}{a^{2}+b^{2}}-8ab\frac{bJ-aI}{a^{2}+b^{2}}\right)\\ &=2\pi\frac{8a^{3}b}{(a^{2}+b^{2})^{2}}\\ &=\frac{128\pi}{25}>0=\int_{\mathbb{R}^{2}}R(g_{Eucl})\,d\mathrm{vol}_{g_{Ecul}}.\end{split}

Note that we have used b=12ab=\frac{1}{2}a in the third equality.

Question 4.2.

As we have seen in the above examples, in Main Theorem 1 (without the assumption that the scalar curvature of each gig_{i} is nonnegative), we cannot weaken the assumptions that the manifold is closed and the convergence is in the sense of W1,p(p>n)W^{1,p}~{}(p>n) to that the manifold is open and the convergence is in the sense of C0,C^{0}, respectively. Then, can we weaken the assumptions in Main Theorems in any sense?

Remark 4.1.

In the above examples, we have constructed these counterexamples by deforming the Euclidean metric locally in a conformal direction. Then, due to the factors from changes of the volume forms, the total scalar curvatures are uniformly bounded from below by a positive constant. On the other hand, if we try to investigate the similar examples for the weighted total scalar curvature MR(g)ef𝑑volg\int_{M}R(g)\,e^{-f}d\mathrm{vol}_{g} (i.e., counterexamples to Main Theorem 2), we cannot use the same method in the above examples since there is no contribution from the factors associated with the conformal changes of the volume forms in this situation.


Acknowledgements The author thanks Prof. Boris Botvinnik for suggesting the problem related to Main Theorem 2. The author also thanks Prof. Kazuo Akutagawa for giving him the opportunity to visit the University of Oregon from September 25 to October 10, 2022.

Author Contributions SH has written the manuscript.

Funding The author had been supported by the Foundation of Research Fellows of Mathematical Society of Japan (from April 2022 to March 2023), and by Mitsubishi Electric Corporation Advanced Technology R&D Center (from May 2022 to March 2024). The author was supported by JSPS KAKENHI Grant Number 24KJ0153.

Data availability Not applicable.

Declarations

Conflict of interest The author declares that there is no conflict of interest.

Ethics approval and consent to participate Not applicable.

Consent for publication The author declares the consent for publication.

References

  • [1] M. T. Anderson, Canonical metrics on 3-manifolds and 4-manifolds, Asian J. Math. 10, 127–163 (2006).
  • [2] R. H. Bamler, A Ricci flow proof of a result by Gromov on lower bounds for scalar curvature, Math. Res. Letters 23, 325–337 (2016).
  • [3] G. Besson, G. Courtois and S. Gallot, Entropies et rigidités des espaces localement symétriques de courbure strictement négative, GAFA 5, 731–799 (1995).
  • [4] G. Besson, J. Lohkamp, P. Pansu and P. Petersen, Riemannian geometry, Fields institute monographs 4, American Mathematical Society (1996).
  • [5] P. Burkhardt-Guim, Pointwise lower scalar curvature bounds for C0C^{0} metrics via regularizing Ricci flow, Geom. Funct. Anal. 29, 1703–1772 (2019).
  • [6] B.-L. Chen, Strong uniqueness of the Ricci flow, J. Differ Geom. 82, 363–382 (2009).
  • [7] M. Gromov and H. Blaine Lawson Jr., Spin and scalar curvature in the presence of a fundamental group. I, Ann. of Math. 111, 209–230 (1980).
  • [8] M. Gromov, Dirac and Plateau billiards in domains with corners, Cent. Eur. J. Math. 12, 1109–1156 (2014).
  • [9] Y. Huang and M.-C. Lee, Scalar curvature lower bound under integral convergence, Math. Z. 303(1), 2 (2023).
  • [10] S. Huang and L.-F. Tam, Short-Time Existence for Harmonic Map Heat Flow with Time-Dependent Metrics, J. Geom. Anal. 32, 1–32 (2022).
  • [11] W. Jiang, W. Sheng and H. Zhang, Weak scalar curvature lower bounds along Ricci flow, Science China Mathematics 66(6), 1141–1160 (2023).
  • [12] N. V. Krylov, Lectures on elliptic and parabolic equations in Sobolev spaces, Graduate Studies in Mathematics 96, American Mathematical Society (2008).
  • [13] D. A. Lee and P. G. LeFloch, The positive mass theorem for manifolds with distributional curvature, Comm. Math. Phys. 339, 99-120 (2015).
  • [14] M.-C. Lee and L.-F. Tam, Rigidity of Lipschitz map using harmonic map heat flow, arXiv preprint arXiv:2207.11017 (2022), to appear in Amer. J. Math.
  • [15] J. Lohkamp, Curvature h-principles, Ann. of Math. 142, 457–498 (1995).
  • [16] R. Schoen and S.-T. Yau, Existence of incompressible minimal surfaces and the topology of three dimensional manifolds with non-negative scalar curvature, Ann. of Math. 110, 127–142 (1979).
  • [17] R. Schoen and S.-T. Yau, On the structure of manifolds with positive scalar curvature, Manuscripta Math. 28, 159–183 (1979).
  • [18] W.-X. Shi, Deforming the metric on complete Riemannian manifolds, J. Differential Geom. 30, 223–301 (1989).
  • [19] M. Simon, Deformation of C0C^{0} Riemannian metrics in the direction of their Ricci curvature, Commun. Anal. Geom. 10, 1033–1074 (2002).
  • [20] C. Sormani, W. Tian and C. Wang, An extreme limit with nonnegative scalar curvature, Nonlinear Anal. 239, 113427 (2024).
  • [21] W. Tian and C. Wang, Compactness of sequences of warped product circles over spheres with nonnegative scalar curvature, Math. Ann. 390, 1–57 (2024).
  • [22] Z. H. Zhang, On the completeness of gradient Ricci solitons, Proc. Am. Math. Soc. 137, 2755–2759 (2009).