Asymptotic convergence of evolving hypersurfaces
Abstract.
If is a smooth immersed closed hypersurface, we consider the functional
where is a local unit normal vector along , is the Levi–Civita connection of the Riemannian manifold , with the pull–back metric induced by the immersion and the associated volume measure. We prove that if then the unique globally defined smooth solution to the –gradient flow of , for every initial hypersurface, smoothly converges asymptotically to a critical point of , up to diffeomorphisms. The proof is based on the application of a Łojasiewicz–Simon gradient inequality for the functional .
Key words and phrases:
Geometric flows, Łojasiewicz–Simon gradient inequality, smooth convergence2010 Mathematics Subject Classification:
53E40, 35R01, 46N201. Introduction
We consider a closed connected differentiable manifold of dimension and a smooth immersion of in the Euclidean space . We shall usually omit the superscript denoting the dimension of . For such an immersion, we always assume that is endowed with the metric tensor , induced by the immersion that we also sometimes simply denote as . The Levi–Civita connection of the Riemannian manifold is denoted by and the associated volume measure by . Then, for with , we consider the functional
on the smooth immersions , where is a (locally defined) unit normal vector field along . Let us specify that in the above definition, if and is the standard basis in , we mean . Notice that is independent of the local choice of the unit normal and it is well-defined without further hypotheses on . However, by the discussion below in Remark 1.2, we shall always assume without loss of generality that is orientable and that a global choice of unit normal field is understood.
We observe that in case , we recognize the well-known elastic energy of closed curves , i.e., , where is the curvature of . If instead , , then one has , where is the second fundamental form (see (1.1)), and then yields the sum of the area and of (an equivalent form of) the Willmore energy of an immersed surface . Let us also notice that if instead , then for any given immersion , where is some polynomial as in (1.3) below.
By the formula for the first variation of (see Theorem 2.1), one can prove that the associated –gradient flow is defined by an evolution equation
for a smooth map (where describes the moving hypersurface and is its unit normal vector field), which turns out to be a (quasilinear and degenerate) parabolic system of PDEs.
If , the study carried out in [17] shows that for every initial smooth immersed hypersurface , there exists a unique smooth solution with initial datum , defined for all positive times; moreover, sub–converges to a critical point of the functional , that is, such that (see Theorem 4.3). By sub–convergence we mean that for some sequence of times , the sequence smoothly converges to , up to diffeomorphisms and translations in . More precisely, there exist a sequence of smooth diffeomorphisms and a sequence of points such that the sequence of immersions converge to in , for any . From such a sub–convergence result it is anyway not possible to immediately deduce that the flow fully converges, i.e., that there exists the full limit of as in for any (up to diffeomorphisms). Actually, the sub–convergence of the flow does not even guarantee that the limits of the flow along different diverging sequences of times coincide. Moreover, as the evolution equations involved here are of order greater or equal than four with respect to the parametrization, it is not even possible to conclude that the flow stays in a compact set of for all times by means of comparison arguments, as maximum principles are not applicable.
In this work we address this issue, that is, we prove that the gradient flow of does actually converge, for any initial hypersurface. Our main result is the following theorem.
Theorem 1.1.
Let be a smooth immersion of a closed hypersurface and let . Then the unique smooth solution to the evolution problem
converges in to a smooth critical point of as , for every up to diffeomorphisms of ; more precisely, there exists a one-parameter family of diffeomorphisms such that the flow converges in to a smooth critical point of as , for every .
In particular, there exists a compact set such that for any .
We remark that the assumption is sharp, in fact if then one gets flows that may develop singularities in finite time.
A relevant motivation for the study of the gradient flow of the functionals goes back to Ennio De Giorgi. In one of his last papers, he conjectured that any compact –dimensional hypersurface in , evolving by the gradient flow of certain functionals depending on sufficiently high derivatives of the curvature does not develop singularities during the flow (see [10] and [11, Section 5, Conjecture 2] for an English translation, see also [17, Section 9]). This result was central in his program to approximate singular geometric flows, as the mean curvature flow, with sequences of smooth ones (see [17, Sec. 9] and [2] for a result in this direction). The functionals are strictly related to the ones proposed by De Giorgi since, roughly speaking, the derivative of the normal field yields the curvature of (see (1.1)). Though not exactly the same, the energies can then play the same role in the approximation process he suggested and the analysis of the asymptotic behavior of their gradient flow is another step in understanding such process.
The main tool in the proof of Theorem 1.1 is a Łojasiewicz–Simon gradient inequality for the functional (see Corollary 4.2). Such an estimate bounds a less–than– power of the difference in “energy” (the value of the functional) between a critical point and a point sufficiently close to it in terms of a suitable norm of the first variation of the functional. The use of this kind of inequalities in the study of the convergence of parabolic equations of gradient–type goes back to Łojasiewicz [15, 16] and to the seminal paper of Simon [23]. More recently, useful sufficient hypotheses implying a Łojasiewicz–Simon gradient inequality have been derived in [4] (see also [13] for several recent generalizations). Building on the abstract tools developed in [4], a first recent application of the inequality to get convergence of an extrinsic geometric flow is contained in [5], where the authors investigate the Willmore flow of surfaces in neighborhoods of critical points. In the last years and in the context of higher order geometric gradient flows, the Łojasiewicz–Simon inequality appeared as a tool for “promoting” the sub–convergence of a flow to its full convergence. As applications of this method we mention [9], in which it is proved the full convergence of the elastic flow of open clamped curves, and [20], in which the sub–convergence of the –elastic flow of closed curves on Riemannian manifolds is shown to imply the full convergence. The analysis in [20] led to a further simplification and deeper understanding of the method, which is exposed in [19]. In this work we essentially generalize the strategy employed in [19] to the gradient flows of the functionals . Moreover we tried to keep most of the arguments as general as possible, in order that the method could be possibly applied also to other geometric gradient flows, also in the context of extrinsic geometric flows in higher codimension possibly in Riemannian manifolds.
In the broad framework of geometric flows, Łojasiewicz–Simon gradient inequalities found many other notable applications. For example, the study of singularities of mean curvature flow can be reconducted to the study of the smooth convergence of a suitable extrinsic geometric flow. Smooth convergence of such flow has been proved exploiting Łojasiewicz–Simon inequalities in some relevant particular cases in [22, 8, 6]. Let us also mention [7], where a classification of ancient solutions to a family of geometric flows in Riemannian manifolds is derived. Łojasiewicz–Simon inequalities have been employed also in the context of intrinsic geometric flows. We refer, for instance, to the study of the rate of convergence of Yamabe flows in [3], or to the deep investigation on the Yang–Mills flow contained in [12] (see also references therein).
Notation and geometry of submanifolds
Let be closed, connected, and orientable. Let be a smooth immersion of and let be global unit normal field on along .
Remark 1.2.
In case is not orientable, given an initial immersion , we can consider the canonical two–fold cover , where is orientable and the initial immersion . By uniqueness of the flow starting at (Theorem 4.3), it follows that the flow starting at is just . Therefore, if we prove that smoothly converges, then the same holds for the flow . Hence, also in this case Theorem 1.1 holds.
As the metric is obtained pulling it back with , in local coordinates on , we have
and the canonical volume measure induced by the metric is given in local coordinates by where
is the standard Lebesgue measure on .
The induced covariant derivative on of a tangent vector field is given by
(in the whole paper we will adopt the Einstein convention of summation over repeated indices) where the Christoffel symbols are expressed by the formula
We will write for the coordinates derivatives, opposite to the covariant ones . With we will mean the –th iterated covariant derivative of a tensor . If is a smooth function on a smooth immersed hypersurface, the symbol denotes its gradient and its Hessian, whose trace is the Laplacian .
The second fundamental form of the immersion is the bilinear symmetric form acting on any pair of tangent vector fields to the hypersurface as
given a (global, since we assumed orientable) choice of the unit normal vector (we will usually identify with and in this formula the field is extended locally around in ). Hence is defined up to a sign, that is, up to the choice of , while is independent of the choice of . In local coordinates, the components of are given by
We recall that the following Gauss–Weingarten relations hold
(1.1) |
The mean curvature of is the trace of , that is
By means of the Gauss equation, the Riemann tensor can be expressed via the second fundamental form, in local coordinates, as follows
Hence, the formulae for the interchange of covariant derivatives become
(1.2) |
where we recall that by we mean the -th component of the field .
Abusing a little the notation, if is a finite family of tensors, we denote by
a generic contraction of some indices of the tensors using the coefficients or . We will also denote
(1.3) |
for some constants . Notice that in every additive term of each tensor appears exactly once (there are no repetitions).
We will use instead the symbol for “polynomials” of the form
with for any and with
Hence, repetitions are allowed in and in every additive term there must be present every argument of .
We notice that, by the above relations, the Riemann tensor of the hypersurface can be written as , exploiting the above notation.
2. Preliminary computations
Let us recall the first variation formula for the functional .
Theorem 2.1 ([17, Theorem 3.7]).
Let be a smooth family of immersions smoothly depending on and . Then, for every , there holds
with
where all the quantities are relative to the hypersurface .
The next lemma states the evolution formulae for the geometric quantities that we need in the computation of the second variation of the functional .
Lemma 2.2.
Let be a smooth family of immersions smoothly depending on and . Let and assume that is a normal vector field along . Then, we have
(2.1) |
(2.2) |
(2.3) |
for any smooth function and with , where and
(2.4) |
Proof.
The first four formulae are computed explicitly at page 150 of [17].
By means of the Gauss–Weingarten relations (1.1), setting , hence , we compute
that is, , hence it follows
(2.5) |
We now deal with equation (2.3) arguing by induction on . Using the previous evolution formulae, for we compute
and the claim follows. Now for , by induction we get
Finally, in order to show equation (2.4), we need to differentiate a generic term of the form
with .
For any component of we can apply [17, Proposition 3.6] in order to get
where denotes the –th component in of the gradient . Also, by [17, Lemma 3.5] and formula (2.1), we have
Therefore, using these formulae and the ones above for the derivative of the metric and its inverse , formula (2.4) follows. ∎
We can now compute the second variation of .
Theorem 2.3.
Let be a smooth family of immersions smoothly depending on . Denote and assume that is a critical point for , i.e., . Let and assume that is normal along . Then
where is linear in and depends on its covariant derivatives of order at most.
Proof.
It follows that, by polarization, we can define the bilinear form
(2.6) |
for any pair of smooth functions and is as in Theorem 2.3.
3. Analysis of the second variation
Suppose that is a smooth critical point of , i.e., . The formula for the second variation given above shows that is well-defined for and . This means that
for any and it is well-defined the map
We are going to exploit the theory of Fredholm operators between Banach spaces. For definitions and results on the subject we refer the reader to [14, Section 19.1]. We recall that if is a Fredholm operator between Banach spaces, its index is defined to be the integer number
where denotes the dimension of a finite dimensional vector space.
Proposition 3.1.
Let be a smooth critical point of , i.e., . Then the second variation functional
is a Fredholm operator of index zero.
In order to prove Proposition 3.1 we need the following commutation rule.
Lemma 3.2.
Let be a smooth immersion and let be a tensor defined on . Assume is endowed with the pull-back metric induced by . Then
for any with .
Proof.
As we need to prove a pointwise identity, we can take a local coordinate frame which is orthonormal at a given point (that is, ) and at . In this way we can compute
at the point . On the other hand, using that for any tensor we have the commutation rule
for any and , we obtain
where we used that , by Gauss equations. Hence, the thesis is proved for . Letting now , by induction we obtain
and the thesis follows. ∎
We are now ready to prove Proposition 3.1. A relevant property about Fredholm operators that we are going to use is the following. If is a Fredholm operator between Banach spaces and is a compact operator, then is Fredholm and (see [14, Corollary 19.1.8]).
Proof of Proposition 3.1.
For the functional is given by
where is
and is as in Theorem 2.3, hence is a compact operator. Therefore
is Fredholm of index zero if and only if the same holds for .
We then claim that the operator
is invertible for sufficiently large, thus it is Fredholm of index zero. As the inclusion is compact, this eventually implies that is Fredholm of index zero.
The injectivity of the above operator immediately follows, suppose indeed that we have , if , multiplying by and integrating, we get
then . If instead , multiplying by and integrating we get
then as well.
About the surjectivity, given we aim at finding such that . We shall minimize the functional
defined by
We can prove that is coercive on , up to choosing sufficiently large (depending on and the geometry of ).
We first consider the case . Integrating by parts in the integral , that is, using the divergence theorem and applying the commutation rule of Lemma 3.2 we get
Moreover, by definition of , we can apply the divergence theorem on the integral in the above expression so that in the polynomial there appear derivatives of of order at most.
We recall that for any covariant tensor there holds the general inequality (see [1, Chapter 3, Section 7.6])
(3.1) |
for any and . Therefore we can estimate
where and then
Therefore, taking and sufficiently large, we estimate
that by inequality (3.1) implies that is coercive on . Analogously, one can prove the coercivity of also in the case .
It follows that there exists a function solving
if , or
if . In any case, is a weak solution to an elliptic equation with constant coefficients and datum (in the sense of [1, Point (d), Page 85]). Therefore, the standard regularity theory for distributional solutions applies (see [1, Theorem, Page 85]), hence belongs to . Integrating by parts, we then get that solves , as required. ∎
4. Convergence
Suppose that is a smooth critical point of , that is, . Then for suitably small, it is well-defined the functional given by
The advantage of the above definition is that the functional is now defined on an open set of a Banach space and we can then look at first and second variation functionals in the classical sense of functional analysis. More precisely, by Theorem 2.1 we have
where (resp. ) is a unit normal vector along (resp. ) and is the volume measure induced by . In this way we see that
Analogously, by Theorem 2.3 and formula (2.6) the second variation of evaluated at is given by
for as in Theorem 2.3, so that
and it is a Fredholm operator of index zero by Proposition 3.1.
In this setting we can apply the following abstract result stating sufficient conditions implying a Łojasiewicz–Simon gradient inequality.
Proposition 4.1 ([20, Corollary 2.6]).
Let be an analytic map, where is a Banach space. Suppose that is a critical point for , i.e., . Assume that there exists a Banach space such that , the first variation is –valued and analytic and the second variation evaluated at is –valued and Fredholm of index zero.
Then there exist constants and such that
for every .
The above functional analytic result is a corollary of the useful theory developed in [4] and it has been also proved in [21] independently.
Applying Proposition 4.1 to the functional we obtain the following corollary.
Corollary 4.2.
Let be a smooth critical point of , i.e., . Let such that is well-defined.
Then, there exist constants and such that
for every .
Proof.
We want to apply Proposition 4.1 with and . By Proposition 3.1 and the discussion at the beginning of the section, we just need to check that and are analytic as maps between Banach spaces.
We can rewrite
where is a unit normal along and is the volume measure induced by . If is any immersion, we have that a unit normal along is , where denotes the Euclidean Hodge star operator. As is an immersion, we see that is analytic. It follows that is analytic as well. As the metric tensor induced by an immersion has components , we get that the metric tensor of depends analytically on and then it is analytic the dependence of and of Christoffel symbols (and thus of the connection) on . Then the integrand in the definition of is just a sum of compositions and multiplications of functions which are analytic in . Finally, integration is linear on , then is analytic for .
By the very same arguments, one can check that also is analytic. Hence, all the hypotheses of Proposition 4.1 are satisfied and the thesis follows. ∎
The starting point for proving the smooth convergence of the gradient flow of is the following sub–convergence theorem.
Theorem 4.3 ([17, Theorem 7.8, Theorem 8.2]).
Let be a smooth immersion and let . Then there exists a unique smooth solution to the evolution equation
where denotes a unit normal vector field along . Moreover, the solution satisfies the estimates
(4.1) |
for any , where and are the second fundamental form and the metric of respectively and there exists a smooth critical point of , a sequence of times and a sequence of points such that
for any , where is a sequence of diffeomorphisms of .
We need a preliminary lemma.
Lemma 4.4.
Let be as in Theorem 4.3. Then, for any there is such that for any there exists such that the immersion coincides with up to diffeomorphism, where is a unit normal vector along , for some “height” functions smoothly depending on . Moreover,
for any .
Proof.
Fixed and , by Theorem 4.3 there is such that for any we have
(4.2) |
for every , for some .
Let us assume that is an embedding. The general statement analogously follows by recalling that immersions are local embeddings. So for large enough, is an embedding as well for every . Moreover there exists open set containing such that it is well-defined the projection map as
where is the distance function from . The vector is orthogonal to at , is smooth on and for sufficiently large we have that for every (for a proof of these facts see [18, Proposition 4.2]).
Hence, for the “height” function is uniquely determined by the identity
that is,
(4.3) |
Then, the map is smooth on and as , by inequality (4.2) and the fact that .
Hence, for the chosen , taking a suitable we have the estimate in the statement of the lemma.
∎
We are now ready for proving our main result. The proof of Theorem 1.1 is essentially a generalization of the strategy employed in [19] to show the smooth convergence of the elastic flow of closed curves in .
Proof of Theorem 1.1..
Let be as in Theorem 4.3. Fixed and chosen smaller than the constant given by Corollary 4.2, relative to the critical point , by Theorem 4.3 and Lemma 4.4, there exists such that for every we have
(4.4) |
for every with some , moreover, coincides with , up to diffeomorphism, for the functions given by Lemma 4.4 (we recall that depends on ), satisfying
(4.5) |
for every .
We claim that it is possible to choose small enough such that for any fixed , the hypersurfaces coincide with (up to diffeomorphism) for some smooth functions with for any .
We define
where is as in Corollary 4.2 applied to the critical point and, without loss of generality, we can clearly assume that for any . As , by Corollary 4.2 we have
where , , are the unit normal, metric tensor and volume measure on and we estimated , for any such that
Differentiating and using the above inequality, we obtain
for any such that . For such times, possibly choosing a smaller , we can assume that . Letting we thus get
and the above estimate becomes
for any such that . Integrating the above differential inequality and estimating , we obtain
then, since possibly choosing a larger we can assume that , we see that
(4.6) |
for any such that on . Finally, since , we get
(4.7) |
for any such that .
Since , estimate (4.5) implies that the hypersurfaces are represented as graph on by means of functions with uniformly equibounded gradients (such bound clearly depends on and goes to zero with it). Also, the inequalities (4.1) clearly hold also for the second fundamental form of the hypersurfaces and , since they coincide with up to diffeomorphism (and translation). These facts imply uniform estimates on the “height” functions in ; namely, for any we have
(4.8) |
for any (a tedious but straightforward way to see this is to differentiate formula (4.3) and use Gauss–Weingarten relations (1.1), taking into account that the closeness in implies that the metric tensor and the Christoffel symbols of the covariant derivative of are mutually “comparable” with the ones relative to ). Hence, if and is small enough, combining estimates (4.7) and (4.8), the interpolation inequalities (3.1) imply that
for any . By a maximality argument, it clearly follows that we can take , for every . Hence, the estimate (4.6), which then holds for any , implies that the flow satisfies the Cauchy criterion for convergence in , hence converges in , as . Interpolating as before by means of inequalities (4.8), the same holds for in , for any and, by Sobolev embeddings, we thus deduce that there exists the limit in for any . Therefore, the same conclusion holds for the original flow , up to diffeomorphism. ∎
References
- [1] T. Aubin, Some nonlinear problems in Riemannian geometry, Springer Monographs in Mathematics, Springer–Verlag, Berlin, 1998.
- [2] G. Bellettini, C. Mantegazza, and M. Novaga, Singular perturbations of mean curvature flow, J. Differential Geom. 75 (2007), no. 3, 403–431.
- [3] A. Carlotto, O. Chodosh, and Y. Rubinstein, Slowly converging Yamabe flows, Geom. Topol. 19 (2015), no. 3, 1523–1568.
- [4] R. Chill, On the Łojasiewicz–Simon gradient inequality, J. Funct. Anal. 201 (2003), no. 2, 572–601.
- [5] R. Chill, E. Fašangová, and R. Schätzle, Willmore blowups are never compact, Duke Math. J. 147 (2009), no. 2, 345–376.
- [6] O. Chodosh and F. Schulze, Uniqueness of asymptotically conical tangent flows, ArXiv Preprint Server – arXiv:1901.06369. To appear: Duke Math. J., 2020.
- [7] K. Choi and C. Mantoulidis, Ancient gradient flows of elliptic functionals and Morse index, ArXiv Preprint Server – arXiv:1902.07697. To appear: Amer. J. Math., 2020.
- [8] T. H. Colding and W. P. Minicozzi II, Uniqueness of blowups and Łojasiewicz inequalities, Ann. of Math. 182 (2015), no. 1, 221–285.
- [9] A. Dall’Acqua, P. Pozzi, and A. Spener, The Łojasiewicz–Simon gradient inequality for open elastic curves, J. Differential Equations 261 (2016), no. 3, 2168–2209.
- [10] E. De Giorgi, Congetture riguardanti alcuni problemi di evoluzione. A celebration of John F. Nash, Jr., Duke Math. J. 81 (1996), no. 2, 255–268.
- [11] E. De Giorgi, Congetture riguardanti alcuni problemi di evoluzione. A celebration of John F. Nash, Jr. (English translation), CvGmt Preprint Server – Scuola Normale Superiore di Pisa, http://cvgmt.sns.it, 1996.
- [12] P. M. N. Feehan, Global existence and convergence of solutions to gradient systems and applications to Yang–Mills gradient flow, ArXiv Preprint Server – arXiv:1409.1525, 2016.
- [13] P. M. N. Feehan and M. Maridakis, Łojasiewicz–Simon gradient inequalities for analytic and Morse–Bott functions on Banach spaces, J. Reine Angew. Math. 765 (2020), 35–67.
- [14] L. Hörmander, The analysis of linear partial differential operators. III, Classics in Mathematics, Springer, Berlin, 2007, –differential operators, Reprint of the 1994 edition.
- [15] S. Łojasiewicz, Une propriété topologique des sous–ensembles analytiques réels, Les Équations aux Dérivées Partielles (Paris, 1962), Éditions du Centre National de la Recherche Scientifique, Paris, 1963, pp. 87–89.
- [16] by same author, Sur les trajectoires du gradient d’une fonction analytique, Seminari di Geometria (1982/83), Università degli Studi di Bologna (1984), 115–117.
- [17] C. Mantegazza, Smooth geometric evolutions of hypersurfaces, Geom. Funct. Anal. 12 (2002), no. 1, 138–182.
- [18] C. Mantegazza and A. C. Mennucci, Hamilton–Jacobi equations and distance functions on Riemannian manifolds, Appl. Math. Opt. 47 (2003), no. 1, 1–25.
- [19] C. Mantegazza and M. Pozzetta, The Łojasiewicz–Simon inequality for the elastic flow, Calc. Var. 60 (2021), no. 56.
- [20] M. Pozzetta, Convergence of elastic flows of curves into manifolds, ArXiv Preprint Server – arXiv:2007.00582. To appear: Nonlinear Anal., 2021.
- [21] F. Rupp, On the Lojasiewicz–Simon gradient inequality on submanifolds, J. Funct. Anal. 279 (2020), no. 8, 1–32.
- [22] F. Schulze, Uniqueness of compact tangent flows in Mean Curvature Flow, J. Reine Angew. Math. 690 (2014), 163–172.
- [23] L. Simon, Asymptotics for a class of nonlinear evolution equations, with applications to geometric problems, Ann. of Math. (2) 118 (1983), no. 3, 525–571.