1. Introduction
Analyzing the behavior of solutions near singularities is essential to understand geometric equations such as minimal surfaces, harmonic maps and mean curvature flows. Many open problems reduce to questions on the singularity formation and its asymptotics. In pioneering works of Leon Simon [Sim83, Sim85], the idea of using Łojasiewicz gradient inequality [Łoj63] from real algebraic geometry was introduced for the first time, and the uniqueness of blow-ups was shown for a class of elliptic and parabolic equations. This uniqueness shows that the solution converges to the unique tangent cone or tangent flow as it approaches a singular point or infinity.
A natural subsequent question is to investigate the rate of convergence and the next order asymptotics that describes the difference between the solution and the limit. This often serves as a crucial starting point for further analysis. For example, recent progresses on the classification of ancient solutions to geometric flows [SAS20][DH21][BCS22][CM22] and complete non-compact solutions to minimal surface [SS86] are based on higher order asymptotics and its improvement. For the singularity formation in parabolic problem, the higher order asymptotics at a singularity gives structural results on the singularity set in a neighborhood [SX22][Gan21]. As the uniqueness of blow-ups implies the second differentiability of arrival time [CM15, CM18], the higher order asymptotics and the convergence rate have a strong relation to further regularity of arrival time [KS06][Ses08].
The convergence rate and direction (i.e., the secant at the limit) are mostly understood when the solution converges at an exponential rate. This is because when the solution decays exponentially, the equation is well-approximated by its linearization. However, without the integrability of the limit, it is possible to have solutions that converge algebraically slowly. In [AS88], Adams-Simon discovered a sufficient condition (later called the Adams-Simon positivity condition or simply condition) on the limit so that they constructed a slowly converging solution to elliptic equations of the form (1.1). Carlotto-Chodosh-Rubinstein [CCR15] found explicit examples of critical points of normalized Yamabe functional with the condition and constructed normalized Yamabe flows converging slowly. As questioned in [CCR15, p.1533], it is of great interest to understand the general behavior of slowly converging solutions. More precisely, one can ask whether such a solution must satisfy the Adams-Simon positivity condition and whether the higher order asymptotics follows the ansatz used in the construction of [AS88] and [CCR15].
We answer this question by showing that the Adams-Simon ‘non-negativity’ condition is necessary for slowly converging solutions to exist. Moreover, when a positivity is satisfied, the convergence rate and the higher asymptotics agree with those of previously constructed examples. (See Theorem 2.3 and 2.5 for detailed statements.) We concern with the elliptic and parabolic equations of forms
| (1.1) |
|
|
|
and
| (1.2) |
|
|
|
Here, is the Euler-Lagrange operator of an analytic functional as in (2.2)-(2.3), is a nonzero constant, and we assume and satisfy the structure (2.6). The presence of nonzero constant accounts that the equation (1.1) appears in the study of the isolated singularity or the asymptotics near the infinity for geometric variational problems. In particular, minimal surfaces, harmonic maps [Sim83, Sim85, AS88] and manifolds [Che18] fall into this category. Equation (1.2) models a geometric flow which converges to a stationary solution when goes to or For instance, the (rescaled) mean curvature flow (see Appendix B) and the harmonic map heat flow [CM22] can be described by (1.2). Our results also apply to the normalized Yamabe flow. See Remark 2.9 (3) for more details.
The main results of the paper, Theorems 2.3-2.6, can be placed in line as a generalization of Thom’s gradient conjecture which we discuss in this paragraph. Let be a solution to the gradient flow with an analytic potential and suppose converges to the origin as . In [Tho89], Thom conjectured that the secant should converge to a direction as . Kurdyka-Mostowski-Parusiński [KKP00] settled this conjecture by showing a stronger result that the secant has a finite length in . However, the precise convergence rate is not yet completed revealed. If does not belong to , one obtains that has to be an eigenvector of the hessian for a positive eigenvalue , and the solution decays with higher order asymptotics for some . If belongs to the kernel of hessian, however, we merely know that the convergence takes place between algebraic rates for some . It is unknown if each solution has a specific convergence rate some .
Let us briefly summarize the main motive and results of the paper. In the slow decaying regime, the equation becomes well-approximated by a gradient flow on the kernel of , the linearization of at . The potential is given by the pull-back of through Lyapunov-Schmidt reduction (see Proposition 2.1). Near the origin, the flow dynamics shall be determined by the first non-constant homogeneous polynomial appearing in the expansion of potential . Let us denote this polynomial by . We reserve to denote the degree of this polynomial and we call it the order of integrability. When the gradient flow is considered, one readily finds a radial solution converging to the origin whenever there exists a critical point of with positive critical value. The slowly decaying solutions of [AS88] and [CCR15] are obtained as perturbations of such radial ansatzes.
In Theorem 2.3 and 2.5, we prove slowly decaying solution has a dichotomy on its convergence rate that either converges to a positive number or diverges to infinity. In the first case, we further show converges smoothly to a critical point of , say , with positive critical value. This shows the secant has a limit and confirms Thom’s gradient conjecture for the concerned case. Note that one may interpret the theorem as the higher order asymptotics , where is a constant of and . Next, the second alternative shows a solution possibly decays even at a slower rate than . As seen in the classical gradient flow of potential on (here, but a solution may decay at rate ), the second alternative can actually take place. We show this can only occur if admits critical point(s) of zero critical value and the secant accumulates on those critical point(s) as . It remains open whether the secant has a unique limit in the second case, and thereby, Thom’s conjecture holds for all slowly decaying solutions. We would like to point out that Theorem 2.5 is also novel for finite-dimensional gradient flows. For a finite-dimensional gradient flow given by , the theorem states that there exists an integer which depends only on , such that any slowly converging solution satisfies either for some , or .
In Theorem 2.4 and 2.6, we show the higher order asymptotic behavior of fast (exponentially) decaying solutions for elliptic and parabolic equations, respectively. This type of result has been expected among researchers and has been obtained for specific problems such as [SS86][CM22]. Nevertheless, we could not find proper literature covering the general forms (1.1) and (1.2), so we provide a proof in Section 4. The elliptic equation (1.1) can be viewed as a perturbation of the second order ODE . Suppose has an eigenvalue larger than , a solution might oscillate while decaying exponentially. This type of solutions has been constructed for minimal graphs over Simons cones in and . See [CP18, Remark 1.21] and [BDGG69] for more details. For this reason, the original form of Thom’s gradient conjecture is not true in the elliptic problem. Moreover, if is an eigenvalue of and , a resonance might occur and result in a solution that decays at a rate . Note [HS85] considered this possibility for minimal graphs over stable (but not strictly stable) minimal cones.
This paper is organized as follows. In Section 2, we introduce the notation, condition, spectral property of linearized operator , and main theorems. In Section 3, we set up the elliptic problem (1.1) as a first order ODE system on function spaces. In Section 4, Theorem 2.4 and Theorem 2.6 for exponentially decaying solutions to elliptic and parabolic equations are proved, respectively. In Section 5, we show in Proposition 5.1 that slowly converging solutions to (1.1) are governed by a finite-dimensional gradient flow with a small perturbation. The parabolic analogue, Proposition 6.1, is proved in Section 6. In Section 7, we complete the proofs for Theorems 2.3 and 2.5 by analyzing a finite-dimensional gradient flow with a small perturbation, using a version of the Łojasiewicz argument motivated by [KKP00]. Appendix A contains auxiliary tools we need in the paper. In Appendix B, we show that (rescaled) mean curvature flows can be written in the form (1.2).
2. Preliminary
Let us introduce our setting to study (1.1) and (1.2) and state the main results. Let be a closed -dimensional Riemannian manifold and be a smooth volume form on which is mutually absolutely continuous with respect to the volume form induced by the metric. Let be a smooth vector bundle equipped with a smooth inner product . For , we write . For , we denote by the fiber over . Let be a connection on which is compatible with the inner product.
We denote by the space of -sections of with respect to . Namely, a section belongs to provided
|
|
|
For , we denote by the collection of sections that satisfies
|
|
|
For , we define and equip with the product metric. We denote by the pull back bundle of through the projection . For and , is defined through . Its -norm is denoted by
|
|
|
Also,
| (2.1) |
|
|
|
Here, note that the norms of time derivatives are included in the definition of .
We often use as an abbreviation of .
Our assumption on is almost identical to the one in [Sim83]. The only difference is that the volume form, , is not necessarily the one induced from the Riemannian metric. Let be a functional defined for by
| (2.2) |
|
|
|
Here the integrand satisfies the following:
-
(1)
is a smooth function defined on an open set of that contains the zero section.
-
(2)
For each , is analytic on .
-
(3)
satisfies the Legendre-Hadamard ellipticity condition
|
|
|
for independent of , and .
is defined to be the negative Euler-Lagrange operator of . Namely, for any ,
| (2.3) |
|
|
|
We further assume is a critical point of . Namely, we assume . We denote by the linearization of at . It is clear that the difference between and is a quadratic term of the form
| (2.4) |
|
|
|
where are smooth with . The Legendre-Hadamard condition implies is elliptic. In particular, there exists a constant such that for any ,
| (2.5) |
|
|
|
Furthermore, is self-adjoint with respect to . This can be seen from
|
|
|
We suppose and in (1.1) and (1.2) are of the form
| (2.6) |
|
|
|
Here and are smooth with and ; .
Let be the eigenvalues of and be the corresponding eigensections which form a complete orthonormal basis of . We separate into four parts according to the eigenvalues.
| (2.7) |
|
|
|
Note that , and are (possibly empty) finite sets and is spanned by . Let be the cardinality of , the dimension of . This implies for some .
We denote by and the orthogonal projection of to and respectively. The following is a version of the implicit function theorem. See [Sim96, §3].
Proposition 2.1 (Lyapunov–Schmidt reduction).
Let be the open ball of radius in . There exist and a map
|
|
|
such that the following statements hold. First, for any and , is an analytic map from to . Second,
|
|
|
Lastly,
|
|
|
Here is given by
The function plays a crucial role in this paper and we will call it reduced functional. The reduced functional is real analytic with and . We may view as an analytic function defined on an open ball in through the identification
| (2.8) |
|
|
|
Let us review the integrable condition. The kernel is called integrable if for any , there exists a family such that in , and in . It is well-known [AS88, Lemma 1] that is integrable if and only if the reduced functional is a constant. Moreover, if integrable condition is satisfied, any decaying solution to (1.1) or (1.2) decays exponentially [AS88, CCR15]. For this reason, whenever non-exponentially decaying solution is considered, the integrable condition should necessarily fails. Namely, the reduced functional is not a constant function. In particular, there exists an integer such that
| (2.9) |
|
|
|
where are homogeneous polynomials with degree and . This integer is called the order of integrability [CCR15].
As explained in Introduction, the gradient flow of has a dominant role in the asymptotic behavior. Let be the restriction of on . Consider the critical points of :
| (2.10) |
|
|
|
If satisfies , then one checks that
|
|
|
becomes a radial solution to the flow . The higher order asymptotics in Theorem 2.3 and 2.5 will be modeled on such solutions.
Definition 2.2 (Adams-Simon conditions. c.f. (4.1) in [AS88]).
We say satisfies the Adams-Simon non-negativity condition for (1.2) if there exists such that .
We say satisfies the Adams-Simon non-negativity condition for (1.1) if there exists such that .
In both equations, the Adams-Simons positivity conditions are defined similarly by requiring that the critical values are positive.
Let us state main results concerning the asymptotic behavior and the convergence rate of decaying solutions. We begin with the elliptic equation (1.1).
Theorem 2.3 (slow decay in elliptic equation).
Let be a solution to (1.1) with that does not decay exponentially as . Then satisfies the Adams-Simon non-negativity condition for (1.1). Moreover, one of the following alternatives holds:
-
(1)
We have
|
|
|
Moreover,
|
|
|
where with .
-
(2)
We have
|
|
|
Moreover,
|
|
|
Theorem 2.4 (fast decay in elliptic equation).
Let be a solution to (1.1) with for some as . If is not the zero section, then one of the following alternatives holds:
-
(1)
There exists an eigenvalue of such that
|
|
|
or
|
|
|
Here
Moreover, for some eigensection with ,
|
|
|
-
(2)
We have
|
|
|
Moreover, for some eigensection with ,
|
|
|
-
(3)
We have
|
|
|
Moreover, there exist for and for such that
|
|
|
Here and .
Next, we consider decaying solutions to the parabolic equation (1.2).
Theorem 2.5 (slow decay in parabolic equation).
Let be a solution to (1.2) with that does not decay exponentially as . Then satisfies the Adams-Simon non-negativity condition for (1.2). Moreover, one of the following alternatives holds:
-
(1)
We have
|
|
|
Moreover,
|
|
|
where and is a critical point of with .
-
(2)
We have
|
|
|
Moreover,
|
|
|
Theorem 2.6 (fast decay in parabolic equation).
Let be a solution to (1.2) with for some as . Then there exists a negative eigenvalue of such that
|
|
|
Moreover, for some eigensection with ,
|
|
|
In fact, it is possible that no slowly converging solution exists even in the presence of non-integrable kernel. e.g., we refer [CCK21][CS20]. Theorem 2.3 and 2.5 provide a criterion for this non-existence result.
Corollary 2.7.
Let be a solution to (1.1) with as . Suppose the Adams-Simon non-negative condition for (1.1) fails. Then decays exponentially.
Corollary 2.8.
Let be a solution to (1.2) with as . Suppose the Adams-Simon non-negative condition for (1.2) fails. Then decays exponentially.
We finish this section with two lemmas related to the Lyapunov-Schmidt reduction map given in Proposition 2.1. We decompose as follows: let and define by
| (2.11) |
|
|
|
Since , can be written as a linear combination of as
| (2.12) |
|
|
|
Lemma 2.10.
Let be the constant given in Proposition 2.1 and be a section with . Let be given by (2.12). Then for any , there exists a positive constant such that
| (2.13) |
|
|
|
Proof.
In the proof we use to represent a positive constant that depends on , , and its value may vary from one line to another.
Let be the linearization of at . From Proposition 2.1 (Lyapunov-Schmidt reduction) and (2.11),
|
|
|
|
|
|
|
|
|
|
|
|
Because is an analytic map from to , the above is bounded by
|
|
|
From (2.12) and the analyticity of , there holds Together with (2.11), we have Hence the assertion holds.
Lemma 2.11 (boundedness of decomposition).
Let be the constant given in Proposition 2.1 and be a smooth section with . Then when ,
|
|
|
for some constant .
Proof.
In the proof we use to represent a positive constant that depends on , , , and its value may vary from one line to another. Let be the coefficients given by (2.12) for . Then
|
|
|
Through (2.8), we may view as a map from an open ball in to . We abuse the notation and write for . For ,
|
|
|
Let be the Banach space of bounded linear maps from to equipped with the operator norm. From Proposition 2.1, is an analytic map from to . In particular, the operator norm of is bounded in . Therefore, provided ,
|
|
|
We used the assumption in the second inequality. This ensures Then because of (2.11), holds.
∎
3. First order ODE system
In this section, we transform (1.1) into a first order ODE system. Let be a solution to (1.1). By setting , we may rewirte (1.1) as
| (3.1) |
|
|
|
From (2.4) and (2.6), the error term has the structure
| (3.2) |
|
|
|
where are smooth with . We aim to vectorize (3.1) and view it as a first order ODE.
Let us begin to set up some notion.
Definition 3.1.
Let be an operator from to given by
|
|
|
Definition 3.2.
For a section , define
|
|
|
If there is no confusion, we often omit the argument and write , to denote and , respectively. Moreover, and denote the restriction of and on , respectively.
The equation (3.1) can be rewritten as
| (3.3) |
|
|
|
We now identify the eigenvalues and eigensections of . For , let
|
|
|
Recall that is divided into in (2.7). The eigenvalues are real and different from if and only if . For , let
| (3.4) |
|
|
|
From , one can check that
| (3.5) |
|
|
|
For , are not real with the imaginary part . Set
|
|
|
Then
| (3.6) |
|
|
|
For , let
|
|
|
Then
| (3.7) |
|
|
|
Next, we introduce a bilinear form . Let be defined by
|
|
|
|
|
|
|
|
We denote . The positive-definiteness of will be justified in Lemma 3.3 below. We write for the adjoint operator of with respect to . Namely,
|
|
|
We define the collection of vectors
| (3.8) |
|
|
|
Lemma 3.3 (spectral decomposition of ).
-
(1)
The bilinear form is equivalent to the standard inner product. Namely,
| (3.9) |
|
|
|
-
(2)
The collection forms a complete -orthonormal basis.
-
(3)
For ,
| (3.10) |
|
|
|
For ,
| (3.11) |
|
|
|
For ,
| (3.12) |
|
|
|
Proof.
We start to prove (3.9). By expressing as ,
|
|
|
for some . In view of (2.5), this implies (3.9). It is straightforward to check that vectors in are -orthonormal. Suppose is -orthogonal to every vector in . Note that for all , and lie in the linear span of . This implies that
|
|
|
and that
|
|
|
Here for and for . Therefore, and is complete. Lastly, (3.10)-(3.12) follow from (3.5)-(3.7) and the -orthogonality of .
∎
The next lemma shows that behaves like .
Lemma 3.4 (equivalence of and angular derivative).
For each ,
| (3.13) |
|
|
|
Proof.
We use an induction argument. The assertion for follows directly from (3.9). Now we assume (3.13) holds for and prove it for . From the induction hypothesis,
| (3.14) |
|
|
|
In view of Definition 3.1, . Therefore,
| (3.15) |
|
|
|
To obtain the inequality in the other direction, we use
|
|
|
|
|
|
|
|
Combining Definition 3.1, (3.14) and the induction hypothesis, we obtain
| (3.16) |
|
|
|
The assertion then follows from (3.15) and (3.16).
∎
Definition 3.5.
Fix . For a section , we define
|
|
|
Here and are given in Definition 3.2. We often abbreviate them to , and denote by , the restriction of , on , respectively.
Corollary 3.6.
Fix . Then for all and ,
| (3.17) |
|
|
|
Moreover, suppose satisfies . Then
| (3.18) |
|
|
|
Proof.
From Definition 3.2 and Lemma 3.4 (equivalence of and angular derivative),
|
|
|
|
|
|
|
|
This gives (3.17). Similarly, from Lemma 3.4 and Definition 3.2,
|
|
|
|
From (3.2) and the assumption,
|
|
|
Then (3.18) follows from (3.17).
∎
Let us project the equation (3.3) onto vectors in . Let
| (3.19) |
|
|
|
Also, let
| (3.20) |
|
|
|
We can then rewrite (3.3) as follows. For ,
| (3.21) |
|
|
|
For ,
| (3.22) |
|
|
|
For ,
| (3.23) |
|
|
|
4. Fast decaying solutions
In this section, we consider solutions to (1.1) that decay exponentially and prove Theorem 2.4. The proof for Theorem 2.6 is simpler so we will omit it. We assume throughout this section that and are non-empty. The proof can be generalized easily to other cases. We begin with a unique continuation property at infinity. Though we closely follow the argument in [Str20], we include the proof for readers’ convenience.
Proposition 4.1.
Let be a solution to (1.1) that satisfies as for all . Then .
Proof.
By the elliptic regularity, Lemma A.3, for all and . Let be given in Definition 3.2. Take . Let be the projection of onto the eigenspace of whose eigenvalues are less than or equal to . Namely, in terms of the coefficients introduced in (3.19),
|
|
|
We claim that for any ,
| (4.1) |
|
|
|
Suppose (4.1) is true at the moment. From (3.2), there exists a uniform constant such that if , then there holds
|
|
|
Define . Suppose is large enough such that for . Then from (4.1),
|
|
|
From the exponential decay assumption we may choose large so that . Hence
| (4.2) |
|
|
|
It is clear that is non-increasing in . Therefore,
|
|
|
We conclude (and thus ) for . Next, we show upto . Suppose on the contrary . By the smoothness of , we may find a small such that for and
|
|
|
which implies that (4.2) holds for . This gives a contradiction and proves the statement.
It remains to show (4.1). Define non-negative functions by
|
|
|
|
|
|
|
|
From (3.21)-(3.23), equals,
|
|
|
|
|
|
|
|
|
From and ,
| (4.3) |
|
|
|
Here is given by
|
|
|
|
|
|
|
|
Similarly, we have
| (4.4) |
|
|
|
where is given by . Fix . By integrating (4.4) from to , we obtain
| (4.5) |
|
|
|
Take any . By integrating (4.3) from to , we obtain
|
|
|
|
|
|
|
|
By the decay assumption, goes to zero when goes to infinity. Hence
| (4.6) |
|
|
|
Note that , and . Then (4.1) follows from (4.5) and (4.6).
∎
Let be an exponentially decaying solution to (1.1). Namely, for some . We further assume is not identically zero. In view of Proposition 4.1, the set
|
|
|
has an infimum . From the elliptic regularity, Lemma A.3, for all and . Let , , be the coefficients defined in (3.19). Let , , be given by (3.20). The quadratic nature of (see (3.2)), in particular, implies
| (4.7) |
|
|
|
Here we used .
Lemma 4.3.
There holds .
Proof.
Suppose the assertion fails. This implies there exists such that there is no element of in the interval . We show this leads to a contradiction for the case . The argument for the case is similar. Define an non-negative function by
|
|
|
From (3.23),
|
|
|
|
|
|
|
|
Here is given by
|
|
|
From the Cauchy-Schwarz inequality and (4.7), . Using
|
|
|
we can integrate
|
|
|
from to to obtain
| (4.8) |
|
|
|
Let
|
|
|
From (3.21)-(3.23), equals
|
|
|
|
|
|
|
|
|
|
|
|
From ,
|
|
|
Here is given by
|
|
|
|
|
|
|
|
From the Cauchy-Schwarz inequality and (4.7), . Integrating
|
|
|
from to , we obtain
| (4.9) |
|
|
|
Combining (4.8) and (4.9), . From the elliptic regularity, Lemma A.3, we then have . This contradicts to the definition of .
∎
Lemma 4.4.
Suppose . Then there exists for and for such that the following holds. For
|
|
|
there exists such that .
Proof.
Define non-negative functions by
|
|
|
|
|
|
|
|
By shrinking the value of if necessary, we may assume for all . An argument similar to the one in the proof of Lemma 4.3 shows . From (3.21), we derive for
|
|
|
Integrating the above from to and using (4.7) yield
|
|
|
for some . A similar argument applying to gives
|
|
|
for some . Lastly, from
|
|
|
|
we derive for some .
∎
For or , an analogous result holds. We omit the proof because it is similar to and simpler than the one for Lemma 4.4.
Lemma 4.5.
Suppose for some and let be the multiplicity of . We may assume . Then there exists and such that
|
|
|
Similar result holds when for some
We are ready to prove Theorem 2.4.
Proof of Theorem 2.4.
Suppose or . From Lemma 4.5 and Corollary 3.6, there exists with and such that . Let . Then
|
|
|
We may upgrade the convergence to by taking derivatives to (3.3). As a result, case (1) in Theorem 2.4 holds.
Suppose . From Lemma 4.4 and Corollary 3.6, there exists for and such that
|
|
|
Then case (2) or case (3) in Theorem 2.4 holds depending on whether or not.
∎
5. Slowly decaying solutions to elliptic equation
In this section, we show that if a solution to (1.1) decays slowly, then the neutral mode, the projection of onto the -eigenspace of , dominates the solution. Moreover, the neutral mode evolves by a gradient flow up to a small error. That is the content of Proposition 5.1.
Let be a solution to (1.1) with as . From the elliptic regularity, Lemma A.3, for all . We further assume that does not decay exponentially. Namely, for any ,
| (5.1) |
|
|
|
Recall that we rewrote (1.1) as an ODE system (3.21)-(3.23). For brevity, we assume throughout this section that . With notational changes, the proof can be readily extended to cover the case where . Since the ODE system consists of (3.21) and (3.23). It is convenient to relabel the coefficients in (3.23). For , we set
|
|
|
|
|
|
and define , a relabelling of for , by
| (5.2) |
|
|
|
Recall that . If , we set for and . If , we set and . This arrangement ensures that
| (5.3) |
|
|
|
Let
| (5.4) |
|
|
|
and
| (5.5) |
|
|
|
We rewrite (3.23) as (5.6) and (5.7) below. For ,
| (5.6) |
|
|
|
and for ,
| (5.7) |
|
|
|
We denote and , and use and to denote their Euclidean norms respectively. Recall the reduced functional is introduced in Proposition 2.1. The goal of this section is to prove the following proposition.
Proposition 5.1.
For a slowly decaying solution to (1.1) as described above, there holds
| (5.8) |
|
|
|
Moreover, for all there exists a positive constant such that
| (5.9) |
|
|
|
The remainder of this section is dedicated to proving Proposition 5.1, which involves two main parts. In Subsection 5.1, we demonstrate (as shown in Corollary 5.3) that any norm of can be bounded by . In Subsection 5.2, we obtain an enhanced decay rate in Lemma 5.11 through the decomposition (2.11). Proposition 5.1 is then a simple consequence of Lemma 5.11.
5.1. Bounding
To estimate norms of , we need higher-derivative versions of the ODE system. Fix . Recall that and are given in Definition 3.5. Let
| (5.10) |
|
|
|
Also, let
| (5.11) |
|
|
|
Then for ,
| (5.12) |
|
|
|
for ,
| (5.13) |
|
|
|
and for ,
| (5.14) |
|
|
|
In the next lemma, we use the Merle-Zaag ODE lemma [MZ98] (see Lemma A.1) to show that dominates the other coefficients, thereby obtaining a stronger version of (5.8).
Lemma 5.2 (dominance of neutral mode).
For any ,
|
|
|
|
|
|
|
|
Proof.
Fix . We give the proof for the case . The argument for is similar. Define three non-negative functions , and by
|
|
|
|
and
|
|
|
From (5.12), (5.13) and (5.14), we compute
|
|
|
|
|
|
|
|
Denote the terms on the second line by and let be the minimum among and . We have
|
|
|
Similarly, define and by
|
|
|
and
|
|
|
It holds that
| (5.15) |
|
|
|
We now compare and .
|
|
|
|
|
|
|
|
|
From the Cauchy-Schwarz inequality and (5.11),
|
|
|
|
|
|
|
|
|
From (3.18) in Corollary 3.6, We can then apply the ODE lemma, Lemma A.1. In view of (A.4), the slow decay assumption (5.1) rules out the possibility that dominates. Hence
| (5.16) |
|
|
|
It remains to show that
| (5.17) |
|
|
|
For ,
|
|
|
For ,
|
|
|
These imply
|
|
|
|
|
|
|
|
|
|
|
|
We used (5.16) in the last inequality. Therefore (5.17) holds and the proof is finished.
∎
In view of Corollary 3.6 and Lemma 5.2, for any ,
|
|
|
Applying the Sobolev embedding on , we can bound any norm of .
Corollary 5.3 (control on higher derivatives).
For any , there exists a positive constant such that
Lemma 5.4.
There hold the following statements:
-
(1)
for , where .
-
(2)
For any there exists , such that on , non-increasing and non-decreasing.
Proof.
The proof directly follows from and .
∎
5.2. Enhanced decay rate
Recall that the decomposition of in (2.11)
| (5.18) |
|
|
|
Let us introduce an auxiliary quantity
| (5.19) |
|
|
|
Lemma 5.5.
There exists a positive constant such that
| (5.20) |
|
|
|
and
| (5.21) |
|
|
|
Proof.
In the proof we use to represent a positive constant that depends on , , in (2.6), and its value may vary from one line to another.
Note that the coefficients of are given by .
|
|
|
Fix . From Lemma 2.10,
|
|
|
Clearly and . From Corollary 5.3, . Therefore,
|
|
|
Because vanishes at the origin of degree , near the origin. Together with ,
|
|
|
Hence (5.20) holds. To show (5.21), recall that , where are smooth with . From Corollary 5.3, . Hence
|
|
|
Together with (5.20), (5.21) follows.
∎
Lemma 5.6.
There exists a positive constant such that
| (5.22) |
|
|
|
and
| (5.23) |
|
|
|
Proof.
In view of (5.7), it suffices to show
|
|
|
From (5.5) and , we actually have . Recall that
|
|
|
where . Because ,
|
|
|
Then the assertion follows from (5.20) and (5.21).
∎
Corollary 5.7.
Suppose for some . Then there exists a positive constant such that
|
|
|
Proof.
Because vanishes at the origin of degree , . Hence the bound for follows from (5.22). The the bound for can be obtained by integrating (5.23).
∎
We vectorize and perform the projection. Set
and
|
|
|
Note that because is orthogonal to , those coefficients completely characterize . The projections of higher order derivatives of , namely , are defined similarly. In the lemma below, we show that , and the higher order coefficients are bounded by .
Lemma 5.8 (control on higher derivatives of ).
For any , there exists a positive constant such that
|
|
|
Proof.
Let .
|
|
|
The assertion then follows from Lemma 2.11 (boundedness of decomposition) and Corollary 5.3 (control on higher derivatives).
∎
Let us define
| (5.24) |
|
|
|
|
Set and be the coefficients of . Namely,
| (5.25) |
|
|
|
Then we have, for ,
| (5.26) |
|
|
|
for ,
| (5.27) |
|
|
|
Lemma 5.9.
There exists a positive constant such that
|
|
|
Proof.
In the proof we use to represent a positive constant that depends on , , , in (2.6), and its value may vary from one line to another.
We compute
|
|
|
|
|
|
|
|
|
|
|
|
where
|
I |
|
|
From (5.20) and (5.21), we have From a direct computation,
|
|
|
Since , the above is bounded by
|
|
|
Together with Corollary 5.3, we get
|
|
|
∎
Corollary 5.10.
Suppose for some . Then there exists a positive constant such that
|
|
|
Proof.
In the proof we use to represent a positive constant that depends on , , , in (2.6), , , and its value may vary from one line to another.
We shall prove
| (5.28) |
|
|
|
by directly solving evolution equations. The other terms can be treated similarly.
For , from (5.27),
|
|
|
For any sequence with , we compute
|
|
|
|
|
|
|
|
From (1) in Lemma 5.4, . From (5.24), (5.25), Lemma 5.9 and the assumption,
|
|
|
By the Cauchy-Schwarz inequality,
|
|
|
Let be the constant given in (2) of Lemma 5.4. Then
|
|
|
|
|
|
|
|
In short, we deduce for all sequence with . This implies (5.28).
Lemma 5.11 (improvement in decay).
For , suppose the following holds.
| (5.29) |
|
|
|
Then for any , there exists a positive constant such that
| (5.30) |
|
|
|
Moreover,
| (5.31) |
|
|
|
Proof.
Fix . In the proof we use to represent a positive constant that depends on , , , in (2.6), , , , and its value may vary from one line to another. Using the assumption (5.29), Lemma 5.2 (dominance of neutral mode) and Lemma A.2 (interpolation), we have for all , Similarly, applying Lemma A.2 with Lemma 5.8,
| (5.32) |
|
|
|
By the Sobolev embedding, . From (5.18) and (5.32), we also have . In view of the definition of in (5.19),
|
|
|
From Corollary 5.10,
| (5.33) |
|
|
|
From Corollary 5.7, (5.31) and
| (5.34) |
|
|
|
Note that (5.31) implies
| (5.35) |
|
|
|
Combining (5.33), (5.35) and (5.34) yields (5.30).
∎
Proof of Proposition 5.1.
The bound (5.8) is a weaker form of Lemma 5.2 (dominance of neutral mode). From Lemma 5.2 (dominance of neutral mode) and Lemma 5.8 (control on higher derivatives of ), the assumption (5.29) in Lemma 5.11 (improvement in decay) holds for . By iterating Lemma 5.11, (5.29) holds for . Then the (5.9) follows from (5.31).
∎
6. Slowly decaying solutions to parabolic equation
In this section, we consider the slowly decaying solutions to the parabolic equation (1.2). The main goal is to prove Proposition 6.1, which is analogous to Proposition 5.1.
Let be a solution to (1.2) with as . From Lemma A.4 (parabolic regularity), for all . We further assume that does not decay exponentially. Namely, for any ,
| (6.1) |
|
|
|
We project onto the eigensections . Set
| (6.2) |
|
|
|
Recall that . The neutral mode will play a special role and we denote it by
| (6.3) |
|
|
|
Proposition 6.1.
For a slowly decaying solution to (1.2) as described above, there holds
| (6.4) |
|
|
|
Moreover, for all , there exists a positive constant such that
| (6.5) |
|
|
|
The proof of Proposition 6.1 follows a similar structure to the one of Proposition 5.1. The first part is to show (in Corollary 6.4) that dominates any norm of .
Let . Then (1.2) becomes
| (6.6) |
|
|
|
From (2.4) and (2.6), the error term has the structure
| (6.7) |
|
|
|
where are smooth with . With , the quadratic nature of allows us to bound higher derivatives of in terms of higher derivatives of .
Lemma 6.2.
For any ,
|
|
|
Proof.
We present the proof for
|
|
|
The other terms can be treated similarly. For simplicity, we write for . For , we write for the partial derivative of of order . Fix with . Then the terms in the expansion of are of the form
|
|
|
where , and . It suffices to show the pointwise bound
|
|
|
We first discuss the case . The above becomes . From
|
|
|
the assertion holds.
Next, we consider the case . It suffices to show as the other terms can be bounded by . Suppose this fails. In other words, for all . Then
|
|
|
In view of and , this is a contradiction.
∎
Now we project (6.6) onto the eigensections . Let
|
|
|
Then (6.6) becomes
| (6.8) |
|
|
|
For the neutral mode, we denote
|
|
|
Lemma 6.3.
For any ,
|
|
|
Proof.
Fix . Define three non-negative functions , and by
|
|
|
|
We note that these coefficients are grouped together according to the sign of the eigenvalues. From (6.8), we compute
|
|
|
|
|
|
|
|
here is defined by the last equality. Let . We have
|
|
|
Similarly, define and by
|
|
|
and
|
|
|
It holds that
|
|
|
We now compare and .
|
|
|
From the Cauchy-Schwarz inequality,
|
|
|
|
From Lemma 6.2,
|
|
|
Therefore, We can then apply the ODE lemma, Lemma A.1. The slow decay assumption (6.1) rules out the possibility that dominates. Hence
| (6.9) |
|
|
|
It remains to show that
| (6.10) |
|
|
|
For ,
|
|
|
For ,
|
|
|
These imply
|
|
|
|
|
|
|
|
|
|
|
|
We used (6.9) in the last inequality. Therefore (6.10) holds and the proof is finished.
∎
Corollary 6.4.
For any , there exists a positive constant such that
The rest of the arguments are simpler than ones in the elliptic case. We omit the details and only provide the main steps. Their elliptic counterparts can be found in Subsection 5.2. Let
|
|
|
Lemma 6.5 (c.f. Lemma 5.5).
There exists a positive constant such that
|
|
|
Lemma 6.6 (c.f. Lemmas 5.6 and 5.9).
There exists a positive constant such that
|
|
|
Corollary 6.7 (c.f. Corollaries 5.7 and 5.10).
Suppose for some . Then there exists a positive constant such that and .
Lemma 6.8 (c.f. Lemma 5.11).
For , suppose the following holds.
|
|
|
Then for any , there exists a positive constant such that
|
|
|
Moreover,
|
|
|
With Lemma 6.8, Proposition 6.1 holds through a simple iteration argument.
7. Gradient flow
In this section, we study the gradient flow on Euclidean space with a perturbative vector field: let be a curve on which satisfies
| (7.1) |
|
|
|
where is an analytic potential function and is a smooth vector field. Let satisfy the assumptions:
| (7.2) |
|
|
|
For some positive integer , has an expansion at zero
|
|
|
where is homogeneous polynomial of degree and . The restriction of on is denoted by . We write for the standard connection on . Consider the critical points and critical values of as
|
|
|
We further assume the perturbative vector field has a bound
| (7.3) |
|
|
|
for some uniform constants and . The main theorem of this section concerns the secant direction when the solution converges to the origin.
Theorem 7.1.
Suppose . Then converges to a non-negative critical value . Moreover, one of the following alternatives holds:
-
(1)
Suppose . Then the secant converges to a critical point with . Moreover,
-
(2)
Suppose . Then
. Here is some connected component of Moreover,
.
We are ready to prove Theorems 2.3 and 2.5.
Proof of Theorem 2.3.
From Proposition 5.1, Theorem 7.1 applies with the potential function being . Suppose case (1) in Theorem 7.1 occurs. This ensures converges to , a critical point of with . Moreover,
|
|
|
This implies
|
|
|
Let be the section corresponding to through (2.8). Clearly converges to in . From Corollary 5.3, for any , is uniformly bounded in . Therefore converges to in . We then obtain case (1) in Theorem 2.3. The other possibility, case (2) in Theorem 7.1, leads to case (2) in Theorem 2.4.
∎
Proof of Theorem 2.5.
Theorem 2.5 can be obtained through replacing Proposition 5.1 in the above proof by Proposition 6.1.
∎
The rest of the section is devoted to proving Theorem 7.1. Let us write the problem in terms of the polar coordinates where
and . Because is homogeneous of degree , . We compute the gradient .
Let and be the radial and tangential parts of . Then equation (7.1) can be decomposed into the radial and tangential parts:
| (7.4) |
|
|
|
Here
|
|
|
|
|
|
|
|
From (7.3), there exists a constant such that
| (7.5) |
|
|
|
Lemma 7.3.
Proof.
It is convenient to work with . The equation (7.4) becomes
| (7.6) |
|
|
|
The assumption becomes .
Suppose From (7.5), there exists such that and for all . This implies for all ,
|
|
|
From the first equation in (7.6), we infer that grows at most linearly. As a result,
|
|
|
which is a contradiction.
∎
By Lemma 7.3, there exists a sequence goes to infinity such that goes to zero. Since is compact, there exists a subsequence (still denoted by ) such that converges to a critical point Let .
Lemma 7.4.
We have and .
Proof.
Suppose . From (7.5) and the first equation of (7.4), is strictly increasing for large. This contradicts the assumption .
To show the second assertion, let us assume is not a constant as otherwise the assertion is trivial. By the Łojasiewicz gradient inequality [Łoj63], for each critical point , there exists a neighborhood such that for all , . Therefore, is a finite set. Because is not a constant, has at least two distinct elements. Fix small such that the critical values are separated at least by . That is,
|
|
|
For every , let
|
|
|
It is clear that . Now assume . There exist and such that Since is a continuous and differentiable function of , there exist such that
|
|
|
From (7.5), for large enough, . Then we compute from (7.6)
| (7.7) |
|
|
|
|
|
|
|
|
This is a contradiction.
∎
Lemma 7.5.
Let be the component of containing . Then
|
|
|
Proof.
Let denote the (open) -neighborhood of . If the assertion is false, there is a small such that for any , there exists such that . By taking small enough, we may assume . Set . Because is a limit point of , there exist a sequence of strictly increasing times such that for , , and for all positive integers . Since is a closed set with , there exists such that for all . From (7.5), for some large . This implies for any with ,
|
|
|
|
|
|
|
|
Therefore,
|
|
|
In particular, does not converge. This contradicts to Lemma 7.4.
Proposition 7.6.
If , then . Moreover, the length of curve on is finite.
Proof.
Recall Consider the control function
|
|
|
Our goal is to show the following. There exist , , and a neighborhood of such that , for , and moreover,
| (7.8) |
|
|
|
Let us first show this implies the proposition. First, since decreases monotonically to zero, for . Suppose for with . By integrating (7.8),
|
|
|
This shows that if , is sufficiently small and is sufficiently close to (we may find such by Lemma 7.4 and ), then remains inside of for all and the length of the curve for is finite. In particular, converges as and this proves the proposition.
It remains to prove and (7.8). Let be the constant in (7.5). Fix large so that for ,
| (7.9) |
|
|
|
| (7.10) |
|
|
|
This is possible because of Lemma 7.4 and . Observe from (7.4),
| (7.11) |
|
|
|
For ,
|
|
|
|
|
|
|
|
|
|
|
|
This shows .
Next, applying the Łojasiewicz gradient inequality [Łoj63] to by viewing it as an analytic function on , we obtain a neighborhood of in , and such that if then
| (7.12) |
|
|
|
Now we suppose for some and show that (7.8) holds. We divide the discussion into two cases depending on the term we utilize from the right hand side of (7.11).
Case 1. Suppose . This implies
| (7.13) |
|
|
|
and
| (7.14) |
|
|
|
From (7.12) and (7.13),
| (7.15) |
|
|
|
Combining (7.14) and (7.15), we derive
|
|
|
|
|
|
|
|
As a result, for and some ,
|
|
|
Case 2. Suppose . Then This implies
|
|
|
Here is defined by the last equality. From (7.10),
| (7.16) |
|
|
|
Hence
|
|
|
Moreover, from (7.16) and (7.12),
|
|
|
for some and . The inequality (7.8) is obtained and this finishes the proof.
∎
Proof of Theorem 7.1.
By Lemma 7.4, there exists a critical point such that
|
|
|
As previously, we denote . If , then the convergence of is an immediate consequence of Proposition 7.6. Moreover, by solving the first line in (7.4) with known asymptotics, we obtain
|
|
|
Suppose . Let be the connected component of which contains . Since there are finitely many critical values (see the proof of Lemma 7.4) and is connected, . In particular, is a connected component of . The convergence of follows by Lemma 7.5 and the limit of follows by a similar argument.
Appendix B Mean curvature flow
In this section, we show that (rescaled) mean curvature flows close to a stationary solution can be described by (1.2).
We begin with the mean curvature flow (MCF). Let be an embedded closed -dimensional submanifold in an -dimensional analytic ambient space . Let be the volume form induced from the metric on . Let be the normal bundle of equipped with the inner product and the connection induced from . For , let and let be the -tubular neighborhood of in . For small enough, and can be identified through the exponential map:
| (B.1) |
|
|
|
Consider a local coordinate chart of as , , , where are coordinates of and are fiber coordinates. We write and for the metric on and the inner product on , respectively. Under the identification (B.1), the ambient metric is of the form
|
|
|
When , , and . In particular, the induced volume form of is given by
|
|
|
Let be a section of which is contained in . From (B.1), can be identified as a submanifold in and we denote this submanifold by . The induced metric on is
|
|
|
Let be the inverse metric of . It is straightforward to check that
|
|
|
form a basis for the normal space of graph. Let and be its inverse metric.
Now we compare the mean curvature vector and the Euler-Lagrange operator of the area functional. The (local) area functional is given by
|
|
|
Denote by the mean curvature vector of . Let be a smooth section of . From the first variational formula,
|
|
|
Comparing the above with
|
|
|
we have
| (B.2) |
|
|
|
In particular, if is a minimal submanifold.
Next, we compute the non-parametric form of the MCF. In this case, the speed is determined by for all . From and , we have . Hence the non-parametric form of the MCF is given by
| (B.3) |
|
|
|
Using (B.2), (B.3) is equivalent to
| (B.4) |
|
|
|
In view of (2.6), (B.4) is of the form (1.2).
Next, we consider the rescaled MCF close to a shrinker. Let be an -dimensional embedded closed submanifold in . We continue to use notation introduced above. For a section of the normal bundle , we write for the position vector. In particular, stands for the position vector of . Let be the volume form with a Gaussian weight. Namely,
|
|
|
For a general section , the (local) Gaussian area functional is
|
|
|
Let be a smooth section of . From the first variational formula,
|
|
|
where stands for the normal part of . A similar argument as above yields
|
|
|
Moreover, the non-parametric form the the rescaled MCF is given by
|
|
|
It is then clear that the above is of the form (1.2).