Abstract.
Motivated by the theory of harmonic maps on Riemannian surfaces, conformal-harmonic maps between two Riemannian manifolds and were introduced in search of a natural notion of βharmonicityβ for maps defined on a general even dimensional Riemannian manifold . They are critical points of a conformally invariant energy functional and reassemble the GJMS operators when the target is the set of real or complex numbers. On a four dimensional manifold, conformal-harmonic maps are the conformally invariant counterparts of the intrinsic bi-harmonic maps and a mapping version of the conformally invariant Paneitz operator for functions.
In this paper, we consider conformal-harmonic maps from locally conformally flat 4-manifolds into spheres. We prove a quantitative uniqueness result for such conformal-harmonic maps as an immediate consequence of convexity for the conformally-invariant energy functional. To this end, we are led to prove a version of second order Hardy inequality on manifolds, which may be of independent interest.
1. Introduction
The most prominent and classic problem in calculus of variations for mappings between two Riemannian manifolds and is the study of harmonic maps, which are critical points of the Dirichlet energy
|
|
|
where is the volume measure on defined by the metric and is the Hilbert-Schmidt norm square of . The conformal invariance of on a Riemannian surface (with respect to the metric of ) and its connection to the theory of minimal surfaces make harmonic maps on two dimensional domains the most widely studied topic in the field of geometric analysis ever since the pioneering work of J. Eells and J. Sampson [13], see also [19]. Motivated by the theory of harmonic maps on Riemannian surfaces, the Paneitz operator [39] and GJMS operators [17], in a recent work [2] V. BΓ©rard has shown the existence of an intrinsically defined energy functional for smooth maps between two Riemannian manifolds (of even dimension where ) and , denoted by , which is conformally invariant with respect to and coincides with the above Dirichlet energy when . Following the terminology of A. Gastel and A. Nerf in [15] who considered an extrinsic analogue (i.e., a variant dependent of the embedding ) of , we will call an intrinsic Paneitz energy when and intrinsic Paneitz poly-energy when . The critical points of are called conformal-harmonic maps or C-harmonic maps, which generalize the harmonic maps on surfaces (i.e., ) and satisfy a system of nonlinear PDEs with leading term . When or , the induced operators for the critical points reassemble the GJMS operators [17], see also Chang-Yang [6]. In particular, when the intrinsic Paneitz energy reads as
(1.1) |
|
|
|
where is the tension field, we denote the scalar curvature and Ricci curvature of by and , respectively. We remark that is conformally invariant on a four dimensional manifold (see Appendix A) and critical points of are the conformal-invariant counterparts of the intrinsic bi-harmonic maps, i.e., the critical points of the intrinsic bi-energy
|
|
|
Conformal-harmonic maps are also generalizations of the Paneitz operator in the context of maps, and intrinsically they satisfy the following system of -th order PDEs (c.f. [22, 27, 37]):
|
|
|
where is the induced Laplace operator on the pullback vector bundle over or equivalently and extrinsically (c.f. [32, equation (1.2)], [31, equation (9)])
|
|
|
|
|
|
|
|
(1.2) |
|
|
|
|
|
|
|
|
|
|
|
|
see Appendix B for a detailed derivation of this equation. Here is the curvature tensor of , and are the Levi-Civita connection and LaplaceβBeltrami operator on respectively, is the tension field, is the orthogonal projection to the tangent plane , is the derivative with respect to the standard coordinates of and denotes the second fundamental form of . Note that this conformal-harmonic map equation (1) differs from the intrinsic biharmonic map
equation by lower order terms which make (1) conformally invariant.
Higher order geometric variational problems, including the study of (both extrinsic and intrinsic) biharmonic maps and polyharmonic maps, have attracted much attention in the last two decades, see e.g. [11, 5, 43, 45, 29, 40, 42, 16, 30, 15, 35, 1, 18, 10] for the extrinsic case and [45, 25, 38, 16, 32] for the intrinsic case. The corresponding heat flows have been studied extensively as well, as a tool to prove the existence of biharmonic maps and polyharmonic maps in a given homotopy class, see e.g. [26, 27, 14, 46, 20, 21, 32]. It should be noted that the extrinsic and intrinsic cases come in two different flavors: the intrinsic variants are considered more geometrically natural because they do not depend on the embedding of the target manifold , although they are less natural from the variational point of view due to the lack of coercivity for the intrinsic energies (and thus they are considerably more difficult analytically and less studied); on the other hand, the extrinsic variants are more natural from the analytical point of view but in turn they do depend on the embedding of . Among them, the conformally invariant problems (both extrinsic and intrinsic) are considered the most geometric, see e.g., [11, 15]. Most of the results in the current literature concern the regularity and existence of (the weak solutions of) the critical points of the higher order geometric variational problems because they are associated to systems of higher order PDEs with critical growth nonlinearities. However, the uniqueness problem of these critical points have been left largely open. In [32] P. Laurain and the first author proved a version of bienergy convexity (and thus the uniqueness) for weakly intrinsic biharmonic maps in with small bienergy and prescribed boundary data, where is the unit 4-ball and is the standard unit sphere. We shall remark that such energy convexity plays an essential role in the discrete replacement min-max construction of geometric objects of interest (such as minimal spheres and free boundary minimal disks in manifolds), see e.g. Colding-Minicozzi [8], Lamm-Lin [28], Lin-Sun-Zhou [34] and Laurain-Petrides [33].
We shall remark that the intrinsic Paneitz energy functional defined in (1.1) on was already used by T. Lamm in [27] as a tool to prove that every weakly intrinsic biharmonic maps from into a non-positively curved target manifold with finite bienergy has to be constant. More recently, O. Biquard and F. Madani in [3] used the corresponding heat flow for to prove an existence result for conformal-harmonic maps from a certain class of 4-manifolds into a non-positively curved manifold . Here the conformal-harmonic map heat flow (or, the negative -gradient flow of ) is defined as follow (when ):
(1.3) |
|
|
|
and (when )
(1.4) |
|
|
|
where . More precisely, in [3] O. Biquard and F. Madani proved
Theorem 1.1.
Let and be compact Riemannian manifolds of four dimensions and dimensions respectively. Assume that has non-positive curvature, the Yamabe invariant
|
|
|
and the conformally invariant total -curvature
|
|
|
satisfies . Then the conformal-harmonic map heat flow 1.3 exists and is smooth for all time and converges subsequently to a smooth conformal-harmonic map as . Consequently, there exists a conformal-harmonic map in each homotopy class in .
In this paper, we consider the intrinsic Paneitz energy for maps from locally conformally flat 4-manifolds with smooth boundaries into spheres and study the (quantitative) uniqueness of the critical points. The first main result of this paper is a version of energy convexity for , more precisely, we prove
Theorem 1.2.
Let be a four dimensional compact locally conformally flat Riemannian manifold with a smooth boundary . There exist depending only on such that for any with , on ,
(1.5) |
|
|
|
and is a weakly conformal-harmonic map, we have
(1.6) |
|
|
|
Remark 1.3.
Our proof relies locally on the result of Laurain and the first author [32], which was established on the unit ball in Euclidean space. Our results could also extend to general source and target manifolds if the Laurain-Lin result is generalized to such settings.
In order to prove this theorem, we are led to show a second order Hardy inequality on smooth manifolds with smooth boundaries. Inspired by the recent work of DβAmbrosio-Dipierro [9] (which followed the techniques introduced by Mitidieri in [36]), we are able to prove the following version of second order Hardy inequality on certain smooth manifolds, for details see Section 3.
Theorem 1.4.
Let be an dimensional compact Riemannian manifold with a smooth boundary and be a nonnegative function such that on in the weak sense and a.e. on for some , then there exits such that
|
|
|
for any .
In fact, based on Theorem 1.4 we are able to prove a general version of second order Hardy inequality that is valid on any smooth manifold.
Theorem 1.5.
Let be an dimensional compact Riemannian manifold with a smooth boundary and be the distance function to the boundary. Then there exits such that
|
|
|
for any .
Proof.
Let be the eigenfunction associated to the first eigenvalue of on with . Then we have and on by the Hopf boundary point Lemma (see e.g. [7, Theorem E.4]), noting that and the tangential derivative of vanishes on the boundary. Now as in the proof of [9, Theorem 6.2], for fixed , define
|
|
|
and
|
|
|
Then there exist constants depending on and a smooth cut-off function such that
|
|
|
on , on , and on . For any , denote
|
|
|
Then by Theorem 1.4 (replacing by , by and by ) we get
|
|
|
|
|
|
|
|
(1.7) |
|
|
|
|
Now let , then by [9, Theorem 6.3] (with ) we have
|
|
|
|
|
|
|
|
(1.8) |
|
|
|
|
Combining (1) and (1) completes the proof.
β
As an immediate corollary of Theorem 1.2 we get the following uniqueness result for weakly conformal-harmonic maps from a locally conformally flat -manifold into spheres.
Corollary 1.6.
Let be a four dimensional compact locally conformally flat Riemannian manifold with a smooth boundary . Then there exists depending only on such that for any weakly conformal-harmonic maps with , on and
|
|
|
we have on .
Acknowledgement
We would like to thank the referees for their valuable comments, which have significantly improved the presentation of the results in this paper. The first author acknowledges partial support from a COR research funding at UC Santa Cruz. The second author would like to thank the support by the National Natural Science Foundation of China (Grant No. 12201440) and the Fundamental Research Funds for the Central Universities (Grant No. YJ2021136).
2. Preliminary
In this section, we will fix some notations and recall a technical theorem (-regularity) that will be used later. Throughout this section, denotes a smooth -dimensional compact Riemannian manifold with a smooth boundary and is an -dimensional smooth closed Riemannian manifold which can be embedded into . As mentioned in the introduction, a weakly conformal-harmonic map from into is a map in that is a critical point of the conformally invariant energy defined in (1.1) and takes values almost everywhere in .
Note that the dimension is critical for the analysis of weakly conformal-harmonic maps (e.g. a map falls in for any but barely fails to be continuous in dimension ). Now let be the nearest point projection map, which is well defined and smooth for small enough. Here . For , let
|
|
|
be the orthogonal projection onto the tangent plane , and
|
|
|
where is the derivative with respect to the standard coordinates of . In the following, we will write (resp. ) instead of (resp. ) and we will identify these linear transformations with their matrix representations in . We also note that these projections are in as soon as is in . Finally, note that the second fundamental form of is defined by
|
|
|
We know that is an intrinsic bi-harmonic map if it satisfies the fourth order PDE (see [44] and [31] for details)
|
|
|
|
|
|
|
|
(2.1) |
|
|
|
|
|
|
|
|
Here . Note that
(2.2) |
|
|
|
and the following two terms in (2) are equivalent to:
(2.3) |
|
|
|
and
(2.4) |
|
|
|
When , we have (note that ).
(2.5) |
|
|
|
and therefore
(2.6) |
|
|
|
In particular, when , the intrinsic bi-harmonic map equation can be rewritten as (see e.g. Lamm-Rivière [29])
(2.7) |
|
|
|
where
(2.8) |
|
|
|
Remark 2.1.
In local coordinates, the terms in (2) read as
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where and is a local orthonormal frame on .
To end this section, let us recall the following version of -regularity for approximate intrinsic and extrinsic bi-harmonic maps into spheres, which will be useful later. Throughout the rest of this paper, and will denote the connection and Laplacian with respect to the Euclidean metric.
Theorem 2.2.
([32, Theorem A.4], c.f. [29, 31])
Let be the unit 4-ball. There exist , , and independent of such that if is a solution of
(2.9) |
|
|
|
where , , and , which satisfy
(2.10) |
|
|
|
almost everywhere (where is a constant depending only on ) and
(2.11) |
|
|
|
then we have and
|
|
|
for all and . Moreover, and for we have
(2.12) |
|
|
|
for some constant . In particular, by rescaling we have for and :
(2.13) |
|
|
|
4. Proof of the main result
In this section, we prove Theorem 1.2. Let us first fix some notations. Since is compact and locally conformally flat with a smooth boundary , we can choose a smooth atlas
|
|
|
for
such that and is conformally flat for each . We assume that
(4.1) |
|
|
|
where is the Euclidean metric. Here denotes the set
|
|
|
Moreover,
|
|
|
and every point in is covered by at most times, see e.g. [41, Lemma 3.3].
For , let
|
|
|
When choosing the smooth atlas, we additionally require that there exists depending only on and such that for any and any point we have
(4.2) |
|
|
|
where . Moreover,
(4.3) |
|
|
|
Let and
(4.4) |
|
|
|
be a disjoint partition of . Now for any we define
(4.5) |
|
|
|
Remark 4.1.
Note that by Proposition A.1 we know that if is a conformal-harmonic map on , then defined in (4.5) is an intrinsic biharmonic map on .
In what follows, we will denote the norm with respect to the Euclidean metric in . As before, and will denote the connection and Laplacian with respect to the Euclidean metric , but note that on each the information of is embedded in these two operators and we do not differentiate the notations.
Lemma 4.2.
There exists depending only on such that if are as in Theorem 1.2, then we have
(4.6) |
|
|
|
where is the tension field of with respect to the flat connection.
Proof.
Let be defined as in (4.5), then use the conformal change of Laplacian (c.f. (A.6)), for we get
|
|
|
|
(4.7) |
|
|
|
|
and
(4.8) |
|
|
|
where are positive constants. Moreover, using (4), (4.8), (1.5) and
|
|
|
we have
(4.9) |
|
|
|
and
(4.10) |
|
|
|
where may depend on all and . Now using the decomposition , we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we have used (1.5), (4.9), (4.10),
(4.11) |
|
|
|
and
(4.12) |
|
|
|
where is a universal constant.
For the last term above, we claim that
(4.13) |
|
|
|
where and . If the claim (4.13) is true, then the second Hardy inequality (Theorem 1.5) imply that (since and on )
|
|
|
|
|
|
|
|
|
|
|
|
(4.14) |
|
|
|
|
where we used
for each , which is similar to (4). Therefore, we obtain (4.6) by choosing sufficiently small. To complete the proof of this lemma, it remains to show that the claim (4.13) holds.
Using (4), Remark 4.1 and Theorem 2.2, and choosing small enough we have
(4.15) |
|
|
|
for any and , where depends on the above.
In particular, define
|
|
|
then we have
(4.16) |
|
|
|
(4.17) |
|
|
|
and therefore
(4.18) |
|
|
|
for any and , where the in (4.17) and (4.18) may depend on and
|
|
|
Now by the choice of the smooth atlas, in particular, (4.2), with a similar argument as above, we have
(4.19) |
|
|
|
for any and . Here may depend on all and Then with a argument similar to (4) we get (4.13). This completes the proof of the lemma.
β
Lemma 4.3.
There exists depending only on such that if are as in Theorem 1.2, then we have
|
|
|
|
(4.20) |
|
|
|
|
Proof.
By Remark 4.1, we know that is an intrinsic biharmonic map on each with respect to the flat connection , namely, we have
|
|
|
|
|
|
|
|
Since , on , we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
For term I, we can use 1.5, a similar argument as in the proof of (4.13) and the second order Hardy inequality (Theorem 1.5) to get
|
|
|
|
|
|
|
|
|
|
|
|
(4.21) |
|
|
|
|
For term II, we note that the integrand reads as
|
|
|
Now since the target manifold is , we know that is the unit normal vector at the point and for any vector so that
|
|
|
Therefore, in this case the integrand in term II becomes
(4.22) |
|
|
|
For term III, we have (using again )
(4.23) |
|
|
|
For term IV, we use again and also on to get
|
|
|
|
|
|
|
|
|
|
|
|
Then by (4.13) and (4) we have
(4.24) |
|
|
|
Now (4.3) follows directly by combining the estimates above.
β
Lemma 4.4.
There exists depending only on such that if are as in Theorem 1.2, then we have
|
|
|
|
(4.25) |
|
|
|
|
Proof.
It follows from Lemma 4.3 and (4.3) that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore, using and , we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we have used so that and . Thus, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we used (4.12), (4.18), (4.19), Hardy inequality (Theorem 1.5), (4.10) and (4.11).
β
Proof of Theorem 1.2.
From (4), we have
(4.26) |
|
|
|
Combining (4.6) and (4.4) we get
|
|
|
|
(4.27) |
|
|
|
|
Now choosing sufficiently small we get (using (4.26) and (4))
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
This completes the proof of Theorem 1.2.
β
Appendix A Conformal invariance of
In this appendix we give a quick check of the conformal invariance of in four dimensions which does not seem to be in the literature. Let be a smooth Riemannian 4-manifold with or without boundary and be another smooth Riemannian manifold. Given , we denote by the pull-back on of the tangent bundle of . Then the tension field is defined by
(A.1) |
|
|
|
where is an orthonormal frame of and is the metric connection on . Fix local coordinates and of and respectively, we have
(A.2) |
|
|
|
where and are the Christoffel symbols on and respectively. Define
|
|
|
Then
(A.3) |
|
|
|
where is the connection on and .
Now under the conformal change (with if ), we have
(A.4) |
|
|
|
and
(A.5) |
|
|
|
where we have used
(A.6) |
|
|
|
Therefore, combing (A.4) and (A.5) together, we get
(A.7) |
|
|
|
and
(A.8) |
|
|
|
Recall the following formulas for the conformal change of Ricci tensor and scalar curvature:
(A.9) |
|
|
|
and
(A.10) |
|
|
|
where denotes the Hessian of . Consequently, we obtain
(A.11) |
|
|
|
|
|
|
|
|
and
|
|
|
|
|
|
|
|
|
|
|
|
(A.12) |
|
|
|
|
|
|
|
|
where Einstein summation convention was used and we used different subscripts to distinguish different metrics on the bundle . Combining these we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Hence, the energy functional is conformally invariant in four dimensions.
Now using the notations in (4.1), on each we have
|
|
|
where is the standard Euclidean metric. By the above conformal change of , for any we get
|
|
|
|
|
|
|
|
Thus, for any variation of such that where , the first variation formula for is (cf. Appendix B)
(A.13) |
|
|
|
|
|
|
|
|
|
|
|
|
(A.14) |
|
|
|
|
(A.15) |
|
|
|
|
In particular, for any fixed and any we have
|
|
|
|
(A.16) |
|
|
|
|
|
|
|
|
noting that since , the terms in (A.14) and (A.15) cancel out to zero using integration by parts over . Therefore we have
Proposition A.1.
If is a conformal-harmonic map on a locally conformally
flat -manifold , then is locally an intrinsic biharmonic map on (with respect to the flat connection) up to a conformal change of .
Appendix B Conformal-harmonic map equation
For the sake of completeness, in this appendix we include a derivation of the conformal-harmonic map equation (1), i.e., the Euler-Lagrangian equation for . Consider a variation of such that where . Then
|
|
|
|
|
|
|
|
(B.1) |
|
|
|
|
We compute the three terms in (B) as follow. Note that the first term I is the first variation of the intrinsic bienergy and therefore yields exactly the intrinsic biharmonic map equation.
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(B.2) |
|
|
|
|
|
|
|
|
Now letβs look at terms II and III which give the additional lower order terms to the intrinsic biharmonic map equation that make the conformal-harmonic map equation (1) conformally invariant.
|
II |
|
|
|
|
|
|
|
|
|
|
(B.3) |
|
|
|
|
|
III |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(B.4) |
|
|
|
|
where we used the subscripts to denote for short.
Combining these together, we get
|
|
|
|
|
|
|
|
(B.5) |
|
|
|
|
|
|
|
|
Thus, the critical point of satisfies the fourth order PDE:
|
|
|
|
(B.6) |
|
|
|
|
|
|
|
|
Note that
|
|
|
|
|
|
|
|
(B.7) |
|
|
|
|
|
|
|
|
and
|
|
|
|
|
|
|
|
Therefore, (B) can be rewritten as
|
|
|
|
|
|
|
|
(B.8) |
|
|
|
|
|
|
|
|
or equivalently (c.f. [32, equation (1.2)], [31, equation (9)])
|
|
|
|
|
|
|
|
(B.9) |
|
|
|
|
|
|
|
|
|
|
|
|
where
|
|
|
and
|
|
|