Deformation of Contracting Maps under the Harmonic Map Heat Flow
Abstract.
We investigate the homotopy classes of maps between closed manifolds by studying certain contracting conditions on the singular values of the differential of the map. Building upon the work of Lee and Wan [6], we extend their results on 2-nonnegative maps to a more general class of k-nonnegative maps by exploiting the properties of submatrices and the linearity of the harmonic map heat flow. Our work establishes new rigidity theorems for maps between manifolds with specific curvature bounds and yields new homotopy rigidity results for maps between spheres and complex projective spaces.
1. Introduction
A fundamental question in geometry is to understand the relationship between the geometric properties of a map and its topological properties, such as its homotopy class. In recent years, there has been significant progress in this area, particularly in the study of maps between manifolds with specific curvature conditions.
For instance, Tsui-Wang [11] proved that area-decreasing maps between spheres are homotopically trivial, a result that has been extended to maps between complex projective spaces by Tsai-Tsui-Wang [10]. More recently, Lee-Tam-Wan [5] showed that area-nonincreasing self-maps of higher-dimensional complex projective spaces are either isometric or homotopically trivial. These results often rely on techniques involving mean curvature flow and the analysis of the evolution equations.
Lee-Wan [6] introduced a new class of contracting maps, which lie between area-nonincreasing and distance-nonincreasing maps. They used the harmonic map heat flow to study the rigidity properties of these maps, proving that under certain curvature conditions, such maps are either Riemannian submersions or homotopically trivial. One of their results is as follows:
Theorem 1.1 ([6], Theorem 1.2).
Let and be closed connected manifolds. Suppose
-
(1)
the sectional curvature is positive;
-
(2)
for any unit tangent vectors and in and , respectively.
Then, for every smooth map such that is -nonnegative, is either a Riemannian submersion or homotopically trivial. In particular, is an Einstein manifold in the first case.
Lee and Wan [6] demonstrated that this contracting condition is preserved under the harmonic map heat flow when certain curvature conditions are met. They then characterized the limiting maps by analyzing the evolution inequalities. A crucial step in their work was establishing the preservation of the contracting condition through the linearity of the harmonic map heat flow, which allowed them to apply properties of the curvature submatrix and address problems involving two singular values.
By systematically analyzing how Lee and Wan utilized the properties of submatrices in their work, we extend their approach to encompass contracting conditions involving an arbitrary number of singular values.
Theorem 1.2.
Let and be closed manifolds, . Suppose
-
(1)
the sectional curvature is positive;
-
(2)
for any unit tangent vectors and in and , respectively.
Denote the curvature pinching by . Then there exists a constant such that for every smooth map , if is -nonnegative, then either is homotopically trivial, or and is a Riemannian fiber bundle.
These are new preserved contracting conditions since constants are always greater than . On the other hand, for the contracting condition involving singular values, we have the following result:
Theorem 1.3.
Let and be a closed manifold. Suppose
-
(1)
is an Einstein manifold of positive sectional curvature and curvature pinching ;
-
(2)
for any unit tangent vectors and in and , respectively.
Then, for every smooth map such that , either is homotopically trivial, or is a Riemannian fiber bundle.
Therefore, for a smooth map with , if , then either an isometry or homotopically trivial. On the other hand, for maps between complex projective spaces, we have the following result:
Corollary 1.4.
Let , where . If one of the following holds
-
(i)
for every three singular values ;
-
(ii)
,
then is homotopically trivial.
Acknowledgement: The first named author would like to thank Professor Mu-Tao Wang, Professor Man-Chun Lee, and Jingbo Wan for valuable discussions. J.-L. Hsu and M.-P. Tsui are supported in part by the National Science and Technology Council grants 112-2115-M-002 -015 -MY3.
2. Preliminaries
2.1. A brief introduction to the harmonic map heat flow
For any smooth map between Riemannian manifolds, we define the tension field of to be the vector field along , which can be expressed as
in local coordinates. Then we define a harmonic map heat flow evolving from to be a solution to
When , we say is a harmonic map. When and are closed manifolds, the short-time existence and uniqueness of the harmonic map heat flow have been proved in [1], and we can define the maximally extended harmonic map heat flow evolved from .
By the standard parabolic PDE, we have the following priori estimates:
Proposition 2.1 ([2]).
For a harmonic map heat flow from a closed manifold into a Riemannian manifold with bounded curvature, if is uniformly bounded, then the higher order derivatives are also uniformly bounded.
Thus, the long-time existence and subsequential convergence are guaranteed when is uniformly bounded.
Theorem 2.2.
Let and be closed manifolds, a maximally extended harmonic map heat flow. If is uniformly bounded, then , and there is such that smoothly converges to a harmonic map .
Proof.
By Proposition 2.1, derivatives of any orders of are uniformly bounded. Suppose . Then by Arzelá-Ascoli theorem,
subsequently converges to some harmonic map heat flow
which is compatible to on . However, is maximally extended, and it leads to a contradiction. Thus, .
To prove the subsequently convergence, we consider
are uniformly bounded. By Arzelá-Ascoli theorem, subsequently converges a harmonic map heat flow
of constant energy, so , which is a subsequential limit of , is harmonic. ∎
Whether the subsequential convergence can be enhanced to convergence remains unknown in general. This uniformity question has been studied in [7], [9], and [4]. We also prove this uniformity when the limiting maps are constant.
Proposition 2.3.
Let be a closed manifold and a harmonic map heat flow with uniformly bounded energy density and subsequently smoothly converges to a constant map . Then converges smoothly to .
Proof.
First, we prove that converges uniformly to by applying the maximum principle on , where denotes the distance function to . Let denote the image of . By direct computation, the evolution equation of is
Since at , there is a geodesic ball such that is positive definite on it, and we have
Therefore, is monotone decreasing when , and uniformly converges to . Consequently, the subsequential limit of is unique.
By Proposition 2.1, any sequence always admits a smoothly convergent , so we derive the smooth convergence. ∎
When the limiting map has the same Dirichlet energy as the initial map , the deformation of harmonic map heat flow can also be characterized explicitly.
Proposition 2.4.
Let and be closed connected manifolds, a smooth map. If the harmonic map heat flow subsequently smoothly converges to a harmonic map with , then .
Proof.
Since the harmonic map heat flow is the gradient descent of Dirichlet energy, for any harmonic map heat flow , if , then is time independent.
Now, let be a sequence increasing to such that smoothly converges to . Then is a sequence decreasing to and bounded above by , and thus for all .
According to the case of closed intervals, . Consequently, . ∎
2.2. Evolution equations
We aim to establish time-independent upper bounds for energy density to prove the long-time existence and convergence of harmonic map heat flow. Therefore, we are concerned with the evolution equation of pull-back metric and energy density.
The following evolution equation can be found in [5].
Proposition 2.5.
For a harmonic map heat flow ,
therefore,
To simplify the evolution equation and apply the tensor maximum principle, we need to consider the frame of singular value decomposition (SVD frame), first introduced by M.T. Wang in [12] for studying graphical mean curvature flow.
Formally speaking, for a -map between Riemannian manifolds, we say a singular value decomposition frame (SVD frame) is a pair of orthonormal frames of and such that is diagonalized and . For convenience, we also let . Be aware that a SVD frame is not necessary smooth.
Corollary 2.6.
For a harmonic map heat flow , with respect to a SVD frame,
where the equality holds at when is totally geodesic at it, and
Proof.
It follows from Proposition 2.5 directly. ∎
In this formula, the concept of partial Ricci curvatures naturally arises, which was first introduced by Z.M. Shen in [8] and H. Wu in [13]. We also introduce the positive part of partial Ricci curvature, which will be used later.
Definition 2.7.
For a Riemannian manifold and orthonormal vectors in some tangent space , we define the partial -Ricci curvature to be
and the positive part of partial -Ricci curvature to be
For a subspace , we can also define the partial -Ricci curvature to be a symmetric bilinear form on whose associated quadratic form is defined by
where is an orthonormal basis of .
3. Convergence Results
This section aims to generalize convergence and rigidity results in [6] by describing crucial pointwise curvature conditions for target manifolds. This allows us to generalize their result to various curvature or contracting conditions.
We first show that the -nonnegativity of is preserved through the tensor maximum principle(Proposition 6.1). Let’s look at the evolution equation of the pull-back metric. If the Ricci tensor of is bounded below by a constant , then ; if there is a matrix such that , then ; furthermore, if , these terms can be further simplified to
Therefore, if has sufficiently good properties, we can prove this monotonicity result. Specifically, the properties we need are as follows:
Definition 3.1.
Let be fixed constants, fixed integers.
-
(a)
For any symmetric matrix with non-negative entries, if the following conditions hold, then we say satisfies the condition .
-
•
;
-
•
;
-
•
for every vector with and every principal -submatrix of ,
where denotes the column vector with all entries equal to 1.
-
•
-
(b)
When and the equality in (*) holds if and only if or , then we say satisfies the strong condition . When , and the equality in (*) holds if and only if or , we also say satisfies the strong condition .
-
(c)
For a Riemannian manifold , we say satisfies the (strong) condition if for any set of orthonormal vectors within a tangent space of , there exists a matrix satisfying the (strong) condition such that .
We denote by as an abbreviation.
According to the definition, conditions yield upper bounds for positive part of partial Ricci curvature.
Lemma 3.2.
If is a manifold satisfying the condition , then .
Proof.
For every orthonormal vectors in , let be a matrix satisfying the condition and . Then we have
∎
The following key lemma establishes the preserved condition in our theory.
Lemma 3.3.
Let and be closed connected manifolds. Assume and that satisfies the condition for some . Then is preserved under the harmonic map heat flow.
Proof.
First, consider the SVD frame of on . Then the -nonnegativity of is equivalent to .
The preservation of -nonnegativity of yields an upper bound for energy density, and hence the higher order derivatives are bounded by the regularity theory of harmonic map heat flow.
Theorem 3.4 ([2], Theorem 2.2).
Let be a maximal extended harmonic map heat flow of a closed manifold into another closed manifold . If is uniformly bounded, then and the norms of the higher order derivatives are also uniformly bounded.
Therefore, the long-time existence and convergence follow are proved.
Proposition 3.5.
Let and be closed connected manifolds. Assume and satisfies the condition for some . Then for every smooth map with , the harmonic map heat flow exists for all time, and there is a sequence such that smoothly converges to a harmonic map with .
We can further characterize this limit by using the equality condition in the inequality.
Lemma 3.6.
Let and be closed connected manifolds. Assume and that satisfies the condition for some . Then for every harmonic map with and a SVD frame of at , we have
-
(i)
;
-
(ii)
for all ;
-
(iii)
for every containing indices and is the matrix satisfying the condition such that ;
-
(iv)
for every .
Proof.
Since is harmonic, is a harmonic map heat flow. Consider the SVD frame of on . By Corollary 2.6,
where . Each term in the last two rows is non-positive, so they are all zero by the monotonicity. ∎
Therefore, we can obtain one rigidity result under condition .
Theorem 3.7.
Let and be closed connected manifolds. Assume with strict inequality at some and satisfies the condition for some . Then for every smooth map with , the harmonic map heat flow exists for all time and converges to a constant map smoothly.
The boundary case can be further characterized under the strong condition .
Lemma 3.8.
Let and be closed connected manifolds. Assume and satisfies the strong condition for some . Then for every harmonic map with ,
-
(i)
either and is a totally geodesic Riemannian subimmersion;
-
(ii)
or is a constant map.
Proof.
If is constant, then the proof is done, so we assume is non-constant in the following discussion.
We first show that is a totally geodesic Riemannian subimmersion. By Lemma 3.6, is totally geodesic, so it remains to show that at each point, the first singular values are either all or 0.
Consider the SVD frame of on and used in the condition of . By Lemma 3.6, for any containing elements, . We discuss the case and separately. When , our claim holds by definition. When , since or , or . If , then it’s done. We are going to show the other case is impossible. In this case, since , it follows , . However, evaluating the original evolution equation of energy density yields:
which leads to a contradiction. Thus, always holds. Since , we have .
∎
The following proposition simplifies the discussion concerning Riemannian submersions between closed manifolds.
Proposition 3.9 (R. Hermann, [3], Theorem 1).
Totally geodesic Riemannian submersions between closed manifolds are Riemannian fiber bundles.
Proof.
∎
Thus, we can prove another rigidity result under strong condition .
Theorem 3.10.
Let and be closed connected manifolds. Assume and satisfies the strong condition for some . Then for every smooth map with , one of the following holds:
-
(i)
, and is a Riemannian fiber bundle, where and ;
-
(ii)
, and is a totally geodesic isometric immersion with ;
-
(iii)
the harmonic map heat flow exists for all time and converges to a constant map smoothly.
Proof.
By Proposition 3.5 and Lemma 3.8, the harmonic map heat flow exists for all time and subsequently converges to some , which is a Riemannian subimmersion or a constant map. When is constant, case (iii) holds by Proposition 2.3. Therefore, assume is a Riemannian subimmersion from now on.
We claim that . If so, then by Proposition 2.4. By direct computation, we have . On the other hand, since , , and
4. Contracting Conditions
In this section, we study which conditions are satisfied under different curvature conditions, focusing on the curvature pinching conditions.
Let us consider the case of as an example, which has been investigated in [5]. Let be a symmetric matrix with non-negative entries satisfying . Then for each of its principal -submatrices
and any vector with non-negative entries and , we have
Furthermore, when , the equality holds if and only if or . Therefore, for every Riemannian manifold of positive sectional curvature, if its partial Ricci curvature is bounded above by a constant , then satisfies the strong condition .
It can be seen that is quite a special case. When we try to extend the above method to the case of , the first problem is that the submatrix is no longer always a multiple of . We address this problem by considering curvature pinching conditions.
First, by considering each direction of vector u, let us investigate for any given matrix , satisfies which conditions .
Definition 4.1.
Let be a symmetric matrix with non-negative entries such that .
-
(a)
For any principal -submatrix of and any vector in the standard simplex , we define
-
(b)
and we define
-
(c)
Now, we define , where runs over all principal -submatrix of .
According to our definitions, we can reformulate conditions through .
Proposition 4.2.
Let be a symmetric matrix with non-negative entries such that for all . Then satisfies the condition if and only if .
Proof.
Notice that satisfies the condition if and only if for every , and principal -submatrix of ,
which is equivalent to say for every principal -submatrix of . Thus, these two conditions are equivalent. ∎
Since is a quadric polynomial of of non-positive leading coefficient, we have:
Lemma 4.3.
Let be a symmetric matrix with non-negative entries such that , . Then for every , we have
-
(i)
. (Here, is defined to be if )
-
(ii)
Moreover, if for , then
Proof.
-
(i)
When , we get directly. When ,
so (i) holds.
-
(ii)
In this case, every entry in is positive, and hence . When , using the equality in the proof of (i), we deduce (ii); when ,
for all , so (ii) still holds.
∎
For any manifold of positive sectional curvature, if is a matrix such that , and a principal -submatrix of , always satisfies the extra condition in Lemma 4.3-(ii) always holds on . Therefore, we can reformulate strong conditions through and in this case.
Proposition 4.4.
Let be a symmetric matrix such that , for all for all , . Then the following statements hold:
-
(i)
when , satisfies the strong condition if and only if satisfies the condition .
-
(ii)
when , satisfies the strong condition if and only if for every and only can be attained by .
Proof.
- (i)
-
(ii)
The proof is essentially similar to (i). In this case, since only can be attained by , for every , , and we have
Therefore, these two statements are equivalent.
∎
By the formula of , we also can show the minimizer exists when for , which allows us to apply the variational method.
Proposition 4.5.
Let be a symmetric matrix such that , for all for all , where . Then
Proof.
As in the proof of Lemma 4.3, since for every with , is a continuous function on compact set, so the minimizer exists. ∎
Based on the above discussion, we build on a translation between the conditions and minimizing problems for . We are going to apply this technique to investigate conditions in the remaining of this section.
According to the curvature conditions of manifolds, let us make some terminologies for matrices. Corresponding to Einstein manifolds, for a symmetric matrix with non-negative entries such that , if , we say is Einstein (of constant ) . Similarly, for positive curvature pinching condition, for a symmetric matrix and a constant , if , for all , and for any pairs of distinct elements, we say is of pinching .
Notice that is Einstein if and only if is an eigenvector of . Hence, we can prove every matrix satisfying conditions is Einstein:
Lemma 4.6.
Let be a symmetric matrix with non-negative entries such that .
-
(i)
If is a minimizer of , then lies in the normal cone of .
-
(ii)
Therefore, if attains minimum of , then is Einstein.
Proof.
When , the above statements hold trivially, so we assume . In this case, since , , and . So, we can differentiate . By definition of normal cone, we have
By (i), we have , i.e. is Einstein. ∎
Therefore, when researching condition , we may assume is Einstein.
Corollary 4.7.
Let be a symmetric matrix with non-negative entries such that . Then
-
(i)
if and only if is Einstein and the second eigenvalue of is non-positive.
-
(ii)
and this minimum is only attained by if and only if is Einstein and the second eigenvalue of is negative.
Proof.
By Lemma 4.6, we may assume is Einstein.
-
(i)
if and only if
for every if and only if for every if and only if the second eigenvalue of is non-positive.
-
(ii)
Similar to (i), and this minimum is only attained by if and only if
for every if and only if the second eigenvalue of is negative.
∎
We need the following lemma to apply Corollary 4.7 to general conditions .
Lemma 4.8.
Let be a symmetric matrix with non-negative entries such that for all . If there is an integer such that every principal -submatrix of is Einstein, then is either 0 or of pinching .
Proof.
For every two pairs and of distinct elements, we want to show . Since is symmetric, when , we have . If , then by considering , it suffices to show for any distinct three elements .
For any four distinct elements , extend them to distinct elements . By assumption, we have
and thus
Now, for any distinct three elements , extend them to distinct elements . Let . Since every principal -submatrix is Einstein, we have
Hence, , and . ∎
Now, we can describe matrices satisfying condition completely.
Proposition 4.9.
Matrices satisfy conditions are as follows:
-
(i)
Matrices satisfying the condition are such that and .
-
(ii)
Matrices satisfying the strong condition are such that for all and .
-
(iii)
For , matrices satisfying the condition are , where .
-
(iv)
Matrices satisfying the (strong) condition are Einstein matrices of constants at most and (negative) non-positive second eigenvalue. In particular, Einstein matrices of constants at most and pinching (less than) at most satisfy the (strong) condition .
Proof.
Let be a symmetric matrix with non-negative entries such that for all . We discuss each condition respectively.
-
(i)
For every principal submatrix of , has the second eigenvalue , so by Corollary 4.7, and (i) is proved.
-
(ii)
As in (i), satisfies the strong condition if and only if for every principal -submatrix and , is only attained by if and only if for every .
- (iii)
-
(iv)
The first statement follows from Corollary 4.7 directly. For the second statement, we may assume and decompose into
where the latter term is a multiple of transition matrix and hence of second eigenvalue at most . Thus, the second eigenvalue of is at most , and the second statement is proved by the first one.
∎
Apply Proposition 4.9 to manifolds. Then we get the following corollary.
Corollary 4.10.
Let be a Riemannian manifold with . Then the following statements hold:
-
(i)
always satisfies the condition .
-
(ii)
for every , always satisfies the strong condition , and satisfies the strong condition if and only if the maximum of is only attained by vectors such that
-
(iii)
for , satisfies the condition if and only if . In this case, when , satisfies all strong conditions .
-
(iv)
if is an Einstein manifold of positive sectional curvature with , then satisfies the strong condition .
-
(iv)’
if is an Einstein manifold of positive sectional curvature with , then satisfies the condition .
Proof.
-
(i)
Denote by . For every orthonormal vectors ,
is a matrix satisfying the condition and . Hence, satisfies the condition .
-
(ii)
For the first statement, let such that . Then is our goal.
Assume satisfies the strong condition and attain maximum of . Let be the matrix satisfying strong condition and . Since
we have . Furthermore, since satisfies the strong condition , for , and hence for all .
Conversely, assume the maximum of is only attained by such vectors. Let be orthonormal vectors. By rearrangement, we may assume
where . Then, by assumption, if and one of is smaller than . Hence, in order to find suitable , we only have to adjust the lower right -principal submatrix.
Let be a small constant such that
Then
is a matrix satisfying the strong condition and .
-
(iii)
Assume satisfies the condition . Then for every orthonormal vectors , , so . Conversely, if , is our goal.
-
(iv)&(iv)’
is our goal by Proposition 4.9.
∎
Corollary 4.10 demonstrates the constraints of condition . That is why we should consider parameter for weaker curvature conditions.
The main difficulty in investigating general condition is that the restrictions to principal submatrices don’t preserve algebraic properties of matrices. There are two situations where we can avoid it. The first one is when we consider condition for Einstein manifolds, where we don’t restrict the matrices. The second one is solely considering the curvature pinching condition, which is evidently preserved under restriction.
Proposition 4.11.
Let be an Einstein manifold with and positive sectional curvature of pinching . Then satisfies the strong condition .
Proof.
For every orthonormal basis of , we are going to show satisfies the strong condition . Notice that it suffices to show .
First, we estimate the second eigenvalue of . Let . Then we have
by curvature pinching condition. Decompose into
Then is an Einstein matrix of entries at most , and its second eigenvalue (that is, the largest eigenvalue on ) is at most . Hence, the second eigenvalue of is at most . Now, we solve maximum of subject to . When is fixed, the maximum of occurs when , so we assume . In this case, , and hence the second eigenvalue is smaller than .
Since tends to as tends to corners of , we have
Therefore, satisfies the strong condition , and so is . ∎
The parameter for general curvature pinching conditions can be characterized by some extreme cases. Given a positive integer , we let and for every , define by
( for every ) | ||||
(for other ) |
Then we have the following proposition:
Proposition 4.12.
For any -matrix of pinching ,
Moreover, for every , .
Proof.
For convenience, we may assume . Then . For every vector with , we have
(where ) | ||||
(since ) |
so
Now, suppose . Then, the minimum of some is attained by a vector . In this case, . Since each is non-negative, they are all zero, and . However, this is impossible since one can show is an isolated zero of by its first variation. Hence, . ∎
Corollary 4.13.
For every Riemannian manifold of positive sectional curvature, curvature pinching , and , satisfies the condition .
Proof.
This corollary is directly proved by Proposition 4.12. ∎
Corollary 4.14.
Let be a Riemannian manifold of positive sectional curvature with quarter pinching. Then satisfies the condition for every . Moreover, when , satisfies the strong condition .
Proof.
For every orthonormal vectors , consider . By symmetry, we have . Let . Notice that is attained by some , and
If , then . Now, suppose . Then , and
Therefore,
with discriminant and solution , so . Evaluate at again. We get
Thus, , and , so it is proved by Proposition 4.12. ∎
5. Rigidity of Contracting Maps
In this section, we build upon Theorems 3.7 and Theorem 3.10 to establish the rigidity of contracting maps. The contracting conditions include upper bound for energy density, 2-nonnegativity of and general -nonnegativity of . By rescaling the metric, these results can be extended to the setting where .
The assumption about an upper bound on the energy density corresponds to condition , valid for the space-form case.
Corollary 5.1.
Let and be closed connected manifolds. Assume their curvatures satisfy:
-
(a)
-
(b)
for some positive constant . Then for every smooth map with , one of the following holds:
-
(i)
is a Riemannian fiber bundle, and ;
-
(ii)
is a totally geodesic isometric immersion, and ;
-
(iii)
is homotopically trivial.
Since spheres are simply-connected, we have:
Corollary 5.2.
For any smooth map with , is either an isometry embedding into a big-circle or homotopically trivial.
Theorem 1.3 also follows from this theory directly:
Proof of Theorem 1.3.
For Einstein manifolds of positive sectional curvature, we have the following result:
Corollary 5.3.
Let and be closed, connected manifolds. Suppose their curvatures satisfy the following conditions:
-
(a)
,
-
(b)
is an Einstein manifold of positive sectional curvature,
-
(c)
for some .
For a smooth map , if , then is homotopically trivial.
The general contracting properties and curvature pinching conditions also can be studied in this way:
6. Appendix
Proposition 6.1.
Let be a compact manifold with time-dependent metric . Assume is a time-dependent self-adjoint -tensor on such that is -nonnegative on and that there are
-
(i)
a locally Lipschitz time-dependent bundle map such that if for singular value decomposition with singular values
-
(ii)
a vector field such that
Then is -nonnegative on .
Proof.
First, observe that the set is closed. If we can demonstrate its openness, we can conclude its connectivity. Therefore, our objective is to establish the existence of a small such that for all . In this context, we can choose a space-time-uniform Lipschitz constant for (with respect to each metric )
For every and , consider the -tensor . Our strategy is to show that there exists a such that for every , is -nonnegative for all . Let be an undetermined positive constant. Suppose not. Then there is an and a least such that is not -positive at some . By continuity, is -nonnegative, and
where are the eigenvectors with smallest eigenvalues. By assumption, we have , and hence
On the other hand, consider SVD frame and the function
Then we have for all , and hence at
So, in this case, , and thus for every , is non-negative on . By taking , we deduce that is non-negative on . ∎
References
- [1] J. Eells, Jr. and J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math., 86 (1964), pp. 109–160, https://doi.org/10.2307/2373037, https://doi.org/10.2307/2373037.
- [2] M. Grayson and R. S. Hamilton, The formation of singularities in the harmonic map heat flow, Comm. Anal. Geom., 4 (1996), pp. 525–546, https://doi.org/10.4310/CAG.1996.v4.n4.a1, https://doi.org/10.4310/CAG.1996.v4.n4.a1.
- [3] R. Hermann, A sufficient condition that a mapping of Riemannian manifolds be a fibre bundle, Proc. Amer. Math. Soc., 11 (1960), pp. 236–242, https://doi.org/10.2307/2032963, https://doi.org/10.2307/2032963.
- [4] T. Huang, Regularity and uniqueness of some geometric heat flows and it’s applications, ProQuest LLC, Ann Arbor, MI, 2013, http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqm&rft_dat=xri:pqdiss:3583738. Thesis (Ph.D.)–University of Kentucky.
- [5] M.-C. Lee, L.-F. Tam, and J. Wan, Rigidity of area non-increasing maps, 2024, https://arxiv.org/abs/2312.10940, https://arxiv.org/abs/2312.10940.
- [6] M.-C. Lee and J. Wan, Rigidity of contracting map using harmonic map heat flow, 2023, https://arxiv.org/abs/2306.12258, https://arxiv.org/abs/2306.12258.
- [7] L. Lin, Uniformity of harmonic map heat flow at infinite time, Anal. PDE, 6 (2013), pp. 1899–1921, https://doi.org/10.2140/apde.2013.6.1899, https://doi.org/10.2140/apde.2013.6.1899.
- [8] Z. M. Shen, On complete manifolds of nonnegative th-Ricci curvature, Trans. Amer. Math. Soc., 338 (1993), pp. 289–310, https://doi.org/10.2307/2154457, https://doi.org/10.2307/2154457.
- [9] P. M. Topping, Rigidity in the harmonic map heat flow, J. Differential Geom., 45 (1997), pp. 593–610, http://projecteuclid.org/euclid.jdg/1214459844.
- [10] C.-J. Tsai, M.-P. Tsui, and M.-T. Wang, A new monotone quantity in mean curvature flow implying sharp homotopic criteria, 2023, https://arxiv.org/abs/2301.09222, https://arxiv.org/abs/2301.09222.
- [11] M.-P. Tsui and M.-T. Wang, Mean curvature flows and isotopy of maps between spheres, Comm. Pure Appl. Math., 57 (2004), pp. 1110–1126, https://doi.org/10.1002/cpa.20022, https://doi.org/10.1002/cpa.20022.
- [12] M.-T. Wang, Long-time existence and convergence of graphic mean curvature flow in arbitrary codimension, Invent. Math., 148 (2002), pp. 525–543, https://doi.org/10.1007/s002220100201, https://doi.org/10.1007/s002220100201.
- [13] H. Wu, Manifolds of partially positive curvature, Indiana Univ. Math. J., 36 (1987), pp. 525–548, https://doi.org/10.1512/iumj.1987.36.36029, https://doi.org/10.1512/iumj.1987.36.36029.