[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 -manifold is preserved under the 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 boundpacs:
[MSC Classification]53C21, 53E20
1 Introduction
Gromov [8] proved the following “-limit theorem”.
Theorem 1.1 ([8, p.1118] and [2]).
Let be a (possibly open) smooth manifold and a continuous function. Consider a sequence of -Riemannian metrics on that converges to a -Riemannian metric in the local -sense. Assume that for all the scalar curvature of satisfies everywhere on Then everywhere on
In contrast, let be the same as in the above theorem, for given a continuous function the set is -dense in the set of all smooth Riemannian metrics on ([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 -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 -Riemannian metrics on a smooth manifold which satisfies the following: converges to a -Riemannian metric on in the -sense as And there is a point such that for some
but
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 we define a smooth metric on as
Here, denotes the Euclidean metric on is the Euclidean distance from to the origin the smooth function
whose support is contained in and is a positive constant. Then, using the following fact, we have Moreover, converges to in the uniformly -sense on
Fact 1.1 (Conformal change of the scalar curvature).
For a Riemannian metric and a -function set then
The proof of this fact is a straightforward calculation. Note that the scalar curvature lower bounds are guaranteed only at the origin 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 and nonnegative on the whole manifold and which is equal to the Euclidean metric outside a compact subset of Indeed, if such a metric exists, then we can construct a metric on the -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 and But, we can only take the lower bound such that in this case because the support of shrinks as Hence we can only obtain the trivial fact 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 be a (possibly non-compact) smooth manifold, a sequence of -Riemannian metrics and a -Riemannian metric on If converges to in some sense (with respect to ) and
for some constant Here denote respectively the scalar curvature of and the Riemannian volume measure of Then, does
hold? Or, more generally, let we consider that a sequence of metrics on converging to a metric in some sense and a sequence of functions on converging to a function in some sense that satisfy
for some constant Then, we ask whether
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 each metric is at least and the underlying manifolds we consider are assumed to be smooth.
If is closed (i.e., compact without boundary) surface, from the Gauss-Bonnet theorem,
for each Riemannian metric on Here denotes the Euler characteristic of Hence it is sufficient to consider the above problem (unweighted case) in dimension 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 is a closed complex -manifold () and are Kähler metrics on Let and be the Kähler forms of and respectively. Assume that converges to (hence converges to ) in the -sense as Then
where denotes the first Chern class of Note that we assumed here that the limiting metric is Kähler metric on as well, but, in our main theorem 3, we will not assume that the limiting metric 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 be a compact -dimensional manifold carrying a hyperbolic metric There is a neighborhood of in the space of all Riemannian metrics with the -topology such that for any
and equality if and only if is isometric to Here,
Our first main result in this paper is the following.
Main Theorem 1.
Let Let be a closed manifold of dimension and a Riemannian metric on Assume that is a sequence of Riemannian metrics on such that converges to on in the -sense as
and on for each Then
Remark 1.1.
Here, we said that a sequence of metrics converges to a metric on in the -sense if and all first weak derivatives of it respectively converge to those of with respect to the -norm of (Since is compact, if converges in the -sense with respect to then it also converges -sense with respect to any fixed reference metric on ) Note that from Morrey’s embedding, there is a continuous embedding: if Therefore the same statement of Main Theorem 1 still holds even though one replace with for all On the other hand, in Main Theorem 1, if we weaken the assumption from -convergence to -convergence, then the same statement (without the assumption that each has nonnegative scalar curvature) no longer holds true in general. Indeed, we will give an example (Example 4.3) on every closed Riemannian -manifold for
Corollary 1.1.
Let and let be the space of all -Riemannian metrics on a closed manifold For any nonnegative continuous function and constant the subspace
is closed in with respect to the -topology.
As our second main result in this paper, we will prove the following theorem in a more general setting.
Main Theorem 2.
Let Suppose that is a closed -manifold (), is a Riemannian metric on and is a sequence of Riemannian metrics on such that converges to on in the -sense as Let be a family of functions on and a function on Assume the following:
-
(1)
There ia a positive constant such that and are -Lipschitz functions on
-
(2)
uniformly on
-
(3)
on for all
-
(4)
Then
This is non-trivial even in the two-dimensional case because is non-constant in general, hence we cannot use the Gauss-Bonnet theorem. For example, and are automatically satisfied in case that for some metric , and Indeed, since each can be locally written as ( is also represented in the same form) and so the norm of the first derivatives are uniformly bounded and uniformly on And, we speculate the assumptions 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 and a constant. Suppose that is a closed manifold of dimension is a Riemannian metric on and is a sequence of metric on Assume the following:
-
(1)
a sequence of nonnegative continuous functions on satisfying: for any positive constant there is a positive constant such that is -Lipschitz on for all
-
(2)
converges to some nonnegative continuous function in the uniformly -sense on
-
(3)
on for each
-
(4)
-
(5)
converges to in the -sense.
Then
Remark 1.2.
From and Remark 3.3, it is known that for all smooth nonnegative function Hence it is reasonable to consider case that 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 be a smooth closed -manifold and a constant. For any metric on is of in the approximate distributional sense if for any nonnegative continuous function there is an approximate sequence satisfying in Corollary 1.2, and there exists a -approximate sequence of -metrics satisfying - in Corollary 1.2.
Suppose now that a metric in Definition 1.1 is actually Then, from Corollary 1.2, in the approximate distributional sense on implies that the same bound holds in the distributional sense on As a result, in the conventional sense from Remark 3.3. Note that in the conventional sense in this case due to of Corollary 1.2 and Theorem 1.1.
For example, if is a complete gradient shrinking or steady Ricci soliton, then on (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 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 convergence of metrics. Our third main theorem is the following.
Main Theorem 3.
Let be a closed -manifold and a Riemannian metric on Let be a sequence of Ricci solitons on (i.e., for some constant and a vector field ) such that converges to on in the -sense as Assume
Moreover, assume that for all and some constant if (resp. for some if ). Then
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
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 be a closed Riemannian manifold endowed with a -Riemannian metric Then there are constants such that the following is true: Consider a -Riemannian metric that is -bilipschitz close to Then there is a continuous family of Riemannian metrics on such that the following holds:
-
For all the metric is -bilipschitz to
-
is smooth on and the map is continuous. In particular, is continuous on .
-
and is a solution to the Ricci DeTurck flow equation
where denotes the Lie derivative of with respect to the time-dependent vector field defined in Remark 2.1 below.
-
For any and any we have
where denotes the norm with respect to of the covariant derivatives of by the Levi-Civita connection of
-
If is a sequence of solutions to (RDE) that are continuous on and smooth on and if converges to some metric in the -sense, then there is a subsequence of that converges to in the -sense on and in the locally smooth sense on with respect to
Proof.
The items and are the result of [19, Theorem 1.1]. follows from the standard argument using the derivative estimates in [19, Theorem 1.1] and Arzela-Ascoli’s theorem. 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 in [18] and in [19] twice with respect to instead of [18, (66) Section 2] and [19, (2.15)] respectively. In particular, we use the continuity of the second derivatives of to derive the estimate corresponding to in the proof of Theorem 5.2 in [19]. ∎
Thanks to [11, Theorem 3.11], if we assume converges to in the -sense, we obtain the following as well.
Lemma 2.2 (c.f. [19, Theorem 4.3]).
Let be a closed Riemannian manifold endowed with a -Riemannian metric Then there are constants and such that the following is true: Consider a -Riemannian metric on which is -close to in the -sense with respect to (i.e., where denotes the -norm of the -tensor with respect to ). Then there is a continuous family of Riemannian metrics on such that - and in Lemma 2.1 hold. Moreover, the following holds instead of in Lemma 2.1.
-
For any we have
Proof.
Remark 2.1.
Let be a solution of the Ricci-DeTurck flow equation with Choose a background metric on and define the Bianchi operator
(1) |
which assigns a vector field to every symmetric 2-form on Let be the flow (we call this flow the Ricci-DeTurck deiffeomorphism below) generated by the time-dependent family of vector fields i.e.,
(2) |
Then satisfies the Ricci flow equation:
For each , from Lemma 2.2, we have a Ricci-DeTurck flow defined on for some positive time , which is independent of . We also have the corresponding Ricci flow and the Ricci-DeTurck diffeomorphism both defined on the same interval . Indeed, we assume that there is the largest time so that the Ricci-DeTurck diffeomorphism exists on . On the other hand, since the Ricci-DeTurck flow is defined on the interval , from the Shi-type estimates for , we can show that converges to a diffeomorphism as by the similar argument in the following proof of Lemma 2.3. This contradicts the definition of the time . Moreover, under the condition of Lemma 2.1 , we can see that for each smoothly subconverges to
Let be a smooth manifold and a -Riemannian metric on Assume a sequence of -Riemannian metrics on converges to on in the -sense. Then, from Lemma 2.2, there are a positive time and a positive constant such that for sufficiently large Lemma 2.2 holds for and Thus, let be the corresponding Ricci-DeTurck diffeomorphism of with background metric defined as in (2). The next lemma will be used in the proof of Main Theorem 2. In particular, the assumption is needed for this lemma.
Lemma 2.3 (Subconvergence of the Ricci-DeTurck diffeomorphisms).
In the above setting, there is a subsequence of that converges to a time-dependent -diffeomorphism (i.e., for all is a homeomorphism, are continuously differentiable and and are mutually inverse. And, there is a subsequence such that for all and are converges to and respectively) with where denotes the identity map.
Proof.
Step 1(-convergence): From Lemma 2.2 and the definition (1), the norm (with respect to ) of the time-dependent vector field defined in (1) is bounded by some positive constant Thus, applying Gronwall’s lemma to (2), by the same argument in the proof of Lemma 2.1 in [5], one can obtain that
(3) |
Here, depends only on and Unlike that in [5, Lemma 2.1], one can get the above type of estimate (in particular, the right-hand side is not but ). This follows from the first-derivative estimate (Lemma 2.2 and hence is bounded by where 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 ), we obtain that
where denotes the maximum of the operator norm of with respect to on Hence, for all we have
where and . Then, since is continuous on from the mean value theorem and this estimate of time-derivative, we have
(4) |
Since the pullback metric satisfies the Ricci flow equation:
by Lemma 2.2 and the above estimate, there is a constant such that for all points and all vectors
From this estimate, Lemma 2.2 and the mean value theorem, we obtain
Since converges to in the -sense, for all sufficiently large we have
for some constant Therefore, from Lemma 2.2 the -closedness of and and the previous estimate, there is a positive constant such that for all sufficiently large
(5) |
Combining the inequalities (3) and (5), we have
Then, by Arzela-Ascoli’s theorem, a subsequence of converges to a time-dependent map as In exactly the same way, one can prove that there is a subsequence of that converges to some time-dependent map as But, since exists and for each To prevent complicated in expression, we will simply write this converging subsequence as
Step 2(-convergence): Next, we will show that the first derivatives of subconverges as Let Since satisfies a parabolic type PDE (see [18, Section 4, Equation (4)]), from the derivative estimate Lemma 2.2 we obtain that
Then, by the parabolic -estimate ([12, Ch. 4, Section 3, Theorem 7]), we have
for all satisfying and Note that if then it holds that
(Recall our assumption .) In particular, we have
for all such and Since, we can choose sufficiently close to so that
(6) |
Hence, as noted above, the weaker relations:
are satisfied as well under the above condition (6). Thus, from Morrey’s embedding theorem,
for some Therefore, by the same arguments that derived (4) above, we obtain that
(7) |
Then, by the inequalities (4) and (7), we can apply the Arzela-Ascoli’s theorem and obtain that there is a subsequence such that as for some time-dependent map Moreover, are differentiable on and as Similarly, there is a subsequence of such that
as Since and are invertible and for all Finally, easily follows from the definition (2) and the above construction of ∎
Remark 2.2.
In Lemma 2.3, it is not known that the limit is the Ricci-DeTurck diffeomorphism of the Ricci flow starting at with background metric Therefore, let be the Ricci-DeTurck flow starting at with background metric then we don’t know whether or not is the solution of the Ricci flow equation starting at
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 be a closed manifold of dimension and a family of functions on satisfying the assumptions and in Main Theorem 2. Suppose be a solution of the Ricci flow equation starting at such that for some on Then for all there are positive constants and such that the following holds. For all there exists satisfies the heat flow equation:
such that
-
(A)
-
(B)
Moreover, for any integer there is a constant such that for all
Here, denotes the Euclidean metric on
Proof.
Since is compact, the image of is compact in the target space From the assumption in Main Theorem 2, there is a compact neighbourhood of the image of such that the image of is contained in for all Then, from the assumption of Main Theorem 2, one can apply the proof of Lemma 3.1 and Theorem 3.1 in [14] to and 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 for this lemma.
Lemma 2.5 (cf. [2, Lemma 4]).
Let be a closed -manifold () and a -Riemannian metric on Suppose be a -Lipschitz function for some Let be a -Riemannian metric which is sufficiently -close to as in Lemma 2.2. Then for any given positive constant there is a constant such that the following holds: Assume that where for some and on Then there is a solution to the Ricci flow equation with initial metric and a solution of the heat flow equation with the Ricci flow background such that for all
Proof.
From Lemma 2.2 and 2.4, there is a sufficiently small such that there is a Ricci flow starting at and there is a heat flow with as uniformly. Along these and for all
Here, is a positive constant depending on and The first equality follows from the evolution of the scalar curvature and the volume form under the Ricci flow, and The second equality follows from applying the divergence formula to the term The first inequality follows from the Cauchy-Schwarz inequality Finally, we have obtained the last inequality as follows. From the assumption by the maximum principle under the Ricci flow, we have Since the scalar curvature is invariant under the pullback action by a diffeomorphism, from the -closedness assumption, one can apply the same derivative estimate in Lemma 2.2 to one 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.
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 Then, we can also suppose that it is less than or equal to for a positive small constant On the other hand, from Lemma 2.2, we can flow each metric and potential function respectively along the Ricci–DeTurck flow and the heat flow with the Ricci flow background, up to a uniform (i.e., independent of ) positive time . From Lemma 2.5, if we retake sufficiently small (independent of ), the weighted total scalar curvature is bounded from below by along the Ricci flow and the heat flow for each 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 But this contradicts the supposition that the weighted total scalar curvature of the limiting metric is less than We prescribe these arguments more precisely below.
Proof of Main Theorem 2.
We show the assertion by contradiction. Suppose that
Then there is a positive constant such that
On the other hand, since , from Lemma 2.2, Lemma 2.5 and Lemma 2.4, there are a a Ricci flow and a heat flow such that
and there also exists a heat flow with as uniformly on By Lemma 2.3, as we have a solution of the Ricci–DeTurck flow equation (retake a sufficiently small if necessary) starting at and a time-dependent -diffeomorphism with such that
for all On the other hand, by Lemma 2.2 and we have
This contradicts our supposition
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 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 (hence ). Note that since the total scalar curvature is diffeomorphism invariant, from Remark 2.1, we have
where are the Ricci-DeTurck flow and the Ricci flow starting at respectively. Note also that the condition 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
for all where is the Ricci-DeTurck flow starting at Then we can deduce a contradiction as as in the same way in the proof of Main Theorem 2.
Next, we prove the theorem for without assuming This statement follows from the following more general claim.
Claim 2.1 (cf. [4, Lecture Series 2, Proposition 2.3.1]).
Let be a closed Riemannian manifold of dimension Suppose that a sequence of -Riemannian metrics on converges to in the -sense. Suppose also that is parallelizable i.e., the tangent bundle of is trivial. Then
Proof of Claim 2.1.
We will use the similar technique in the proof of [4, Lecture Series 2, Proposition 2.3.1]. Since is parallelizable, we can take a global section of the orthonomal frame bundle of Define the 1-form and the vector field be the dual of with respect to Next, we will use the Bochner identity for 1-forms:
Here for the Dirac operator related to the deRham complex. Thus, we have for and using this Bochner identity and the integration by parts,
Snice in the -sense on the left-hand side of the above equality converges to the following quantity
Taking the sum from 1 to for we obtain that
This completes the proof of the claim. ∎
Since every oriented closed three manifold is parallelizable, using this claim we have
(Of course, if necessary, we discuss similarly after taking its orientable double cover.) In particular, if for all then . 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)
-dimensional sphere where or
-
(4)
Every Lie group.
-
(3)
The product of parallelizable manifolds.
Question 2.1.
-
•
In Claim 2.1, is the parallelizability of necessary?
-
•
What is the relation between parallelizability of a closed manifold and -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 of the Ricci flow equation starting from constructed in Lemma 2.1 is
where is the family of diffeomorphisms generated by the time-dependent vector field with the initial condition and is the uniform existence time of the flows guaranteed in Lemma 2.1. Thus, we have
(8) |
for all We will firstly consider the case of From (8), for any there is a positive time such that for any diffeomorphism and
Similarly, when we see that for any there is a positive time such that for any diffeomorphism and
In particular, we take the inverse of the Ricci-DeTurck diffeomorphism (see Remark 2.1) of as here, then we have
where is the Ricci-DeTurck flow starting at 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 is a shrinking Ricci soliton (i.e., in the above situation), then the maximal existence time of the corresponding Ricci flow is But, since we assume converges to in the -sense, Lemma 2.1 implicitly prevents being as
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 be a closed -manifold and a -Riemannian metric on Let be a sequence of -Riemannian metrics on that converges to on in the -sense . Assume that there are a constant and a continuous function such that for all
-
•
-
•
on , and
-
•
Here is the volume of with respect to Then
Remark 3.1.
If the limiting metric here is a hyperbolic metric (i.e., its sectional curvature is constant ) and each metric has its volume entropy then for all This is the result of [3]. Here, the volume entropy of a closed Riemannian manifold is given by the limit
where is the volume of a ball of radius in the universal cover (In particular, Bishop-Gromov’s volume comparison implies that a manifold with Ricci curvature lower bound satisfies ) 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 is a closed hyperbolic 3-manifold, for any metric on
Proof of Corollary 3.1.
There is a sufficient large natural number such that the Riemannian product manifold satisfies where is the standard Riemannian -sphere of constant scalar curvature Indeed, since
we see that for sufficiently large Therefore we have
On the other hand, from our assumption, converges to in the -sense and
Therefore, from Main Theorem 1,
Then, subtracting from both sides and dividing both sides by we obtain ∎
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 of a closed Riemannian manifold is defined as
where is the conformal class of the metric By the definition, depends only on the conformal class of A Riemannian metric with is called Yamabe metric of if
Corollary 3.2 (for Corollary 3.1).
Let be a closed -manifold and a -Riemannian metric on Let be a sequence of -Riemannian metrics on that converges to on in the -sense , and Assume that is a Yamabe metric of and there are a constant and a continuous function such that for all and on Then
Proof.
From the definition of and we have
Since in the -sense and we also have Hence, from Corollary 3.1, we have
Therefore, since is a Yamabe metric of and we obtain
∎
Remark 3.2.
If this corollary directly follows from Gromov’s theorem (Theorem 1.1 in this paper). Indeed, since in the -sense for (in particular, in the -sense), from Theorem 1.1, we have On the other hand, since is a Yamabe metric of (hence, its scalar curvature is constant) and we have
However, if 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 be a non-compact Riemannian -manifold () with and a sequence of positive -functions. Assume that each is equal to 1 outside a compact set and
Set Then
Proof.
From the formula for the scalar curvature and the volume form under this conformal change:
we have
We have used the divergence formula and the fact that is equal to 1 outside a compact set in the third equality. Since -uniformly on we have
and
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 and if for each for some positive -functions then the assumption 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 be a smooth manifold endowed with a smooth background metric Given any Riemannian metric with regularity and locally bounded inverse the scalar curvature distribution is defined, for every compactly supported smooth test function by
where is given by is a function as
and Here, denotes the Levi-Civita connection of
Let be a continuous function on We say that in the distributional sense if for any nonnegative compactly supported test function
Remark 3.3.
If a metric is then the scalar curvature distribution coincides with
For more details about the distributional scalar curvature and related results, see [11, 13, 20, 21]. From Gromov’s -limit theorem (Theorem 1.1), there has already been a definition of scalar curvature lower bounds for metrics (see [14, Definition 1.2] for example). Namely, a metric on a smooth manifold is of on in the Gromov’s sense if and only if there exists a sequence of metrics such that converge -locally to and satisfy on 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 which is of 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 which is of 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 and a constant. Suppose that is a closed manifold of dimension is a Riemannian metric on and is a sequence of metric on Assume the following:
-
(1)
a sequence of nonnegative smooth functions on satisfying: for any positive constant there is a positive constant such that is -Lipschitz on for all
-
(2)
converges to some nonnegative continuous function in the uniformly -sense on
-
(3)
on for each
-
(4)
-
(5)
converges to in the -sense.
Then
Proof.
From and for any (small) positive constant
Then applying Main Theorem 2 to and we obtain that
for all Here can be taken to be arbitrarily small, hence we obtain the desired inequality:
∎
Question 3.1.
Can we replace the condition and with “ in the distributional sense for some nonnegative constant ”?
If we make the regularity of convergence much stronger in [11, Theorem 3.2 (1)], then it can be proven that the above question is positively true.
4 Counterexamples
If is a compact smooth manifold and -Riemannian metrics converges to a -Riemannian metric on in the sense as , then So, by Lebesgue’s dominated convergence theorem, we have
However, if is non-compact, it is not known that there is a Lebesgue integrable function such that a.e. on for all 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 has nonnegative scalar curvature) does not hold in general if is non-compact.
Example 4.1 ( locally uniformly convergence and incomplete limiting metric).
The limiting metric of the first example constructed below is incomplete on . Consider the smooth positive function defined as
and Here is a smooth cut-off function such that on the closed ball and outside of the -neighbourhood of where is an arbitrarily fixed positive constant. Here, denotes the Euclidean metric on is the Euclidean distance function from the origin , and
Then for each is a non-compact smooth Riemannian manifold with
(9) |
Moreover, converges to in the locally -sense in but not in the -sense on Note that is incomplete. On
When ,
Indeed, we can observe such a behavior from the form of the scalar curvature on as follows.
(10) |
From (9) and the divergence formula, we have
(11) |
where denotes the volume of -sphere with respect to the standard metric. Here,
Set the left hand side of this equation as Then we have
Moreover,
and
Combining these, as we can see that the rightmost term of (11) converges to
Hence, for all sufficiently large
Next, we will construct a counterexample (to Main Theorem 1 without the assumption that each has nonnegative scalar curvature), in which each is complete.
Example 4.2 (Not but ).
Consider Here the smooth positive function has been defined as
Here, is a smooth cut-off function such that on and outside of the -neighbourhood where is an arbitrarily fixed positive constant. Here, is the Euclidean distance function from the origin Then, for each is a non-compact smooth Riemannian manifold with
(12) |
and on in the uniformly but not in the -sense. On
Note that is not non-negative on Indeed, for sufficiently large and such that we can take a point Then
This is checked as follows. For sufficiently large such a point is contained in Hence, from the above formula and the choice of the point we have
Since we can observe the desired behavior of the scalar curvature as above. Moreover, from (12) and the divergence formula, we have
Here,
Since
there is a sufficiently large such that for all
Hence, for all
From the Morrey embedding, we have
Therefore the same statement of Main Theorem 1 still holds even though one replace with for all On the other hand, in Main Theorem 1, if we weaken the assumption from to then the same statement (without the assumption ) does not hold in general. Indeed, using the same local construction as in the previous example in dimension we can also construct a counterexample on a closed manifold to Main Theorem 1 (without the assumption that each has nonnegative scalar curvature) as follows. Note that all metrics in each such example has sign-changing scalar curvature, i.e., for each there are some points s.t.
Example 4.3 (On every closed manifold).
Consider where is a closed -manifold and is a Riemannian metric on Here the smooth positive function has been defined as
Here, is a smooth cut-off function such that on and outside of the -neighbourhood for some point where is a sufficiently small positive constant. Here, is the distance function of from the point and is the injectivity radius of Then, for each is a smooth Riemannian manifold with
and converges to on in the but not in the -sense. In the same calculation as in the previous example, we have
Here, the constant depends only on and Thus, from the observation as in the previous example, there is and a positive constant such that for all
Therefore, for all
Moreover, by the definition of
Hence there is a sufficiently large such that for all
Thus, for all we have
Here, we have a question about the regularity of convergence in the assumption of Main Theorem 2.
Question 4.1.
Are there any -metrics and -Lipschitz () functions on a closed -manifold satisfying the followings ?
-
•
converges to a -metric in the -sense,
-
•
converges to a -Lipschitz function in the uniformly -sense,
-
•
there is a constant such that
Or, additionally,
-
•
there is a point for each such that as
In the following Example 4.4, we give another counterexample which is similar to the one in Example 4.2 (). However, in the following example of dimension the support of () with the origin converges to in the pointed Gromov-Hausdorff sense as (In Example 4.2, the support of 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 is not compact for each (see the last half of Example 4.4).
Example 4.4 (Not but ).
Consider Here the smooth positive function has been defined as
Here, is a smooth cut-off function such that on and outside of the -neighbourhood where Note that as Here, is the Euclidean distance function from the origin Then, for each is a non-compact smooth Riemannian manifold with
Moreover, converges to on in the uniformly but not in the -sense. On
Note that is not non-negative on Indeed, when oscillates for each Indeed, for sufficiently large such a point is contained in Hence, from the above formula,
Since and as we can easily observe the desired behavior of the scalar curvature. Moreover, by the divergence formula,
Here,
Since
there is a sufficiently large such that for all
Hence for all
Next, we will construct a two-dimensional example. Consider the smooth function on defined by
where denotes the Euclidean distance function from the origin Then uniformly converges to the constant function in the topology on , but does not converge to in the topology on . Hence the sequence of complete metrics on uniformly converges to in the sense on but does not converge to in the sense on Set Then we can check that
and
Moreover,
-
•
-
•
-
•
-
•
Combining these, we obtain that
Note that we have used 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 is nonnegative), we cannot weaken the assumptions that the manifold is closed and the convergence is in the sense of to that the manifold is open and the convergence is in the sense of 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 (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 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 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).