Entire solutions of two-convex Lagrangian
mean curvature flows
Abstract.
Given an entire function on , we consider the graph of as a Lagrangian submanifold of , and deform it by the mean curvature flow in . This leads to the special Lagrangian evolution equation, a fully nonlinear Hessian type PDE. We prove long-time existence and convergence results under a 2-positivity assumption of . Such results were previously known only under the stronger assumption of positivity of .
1. Introduction
Fundamental to a fully nonlinear Hessian type partial differential equation is the convexity assumption, which guarantees ellipticity or parabolicity and facilitates essential regularity estimates. Among these equations, the special Lagrangian equation is rather unique in that the variational structure always implies ellipticity or parabolicity without the need of any convexity assumption. Still most previous results were obtained under the convexity condition or its equivalence. In [TTW22l], a substantial improvement was achieved by removing the convexity assumption and replacing it by the two-convexity assumption. The main purpose of this paper is to extend results in [TTW22l] from the compact setting to the non-compact global setting. Essential difficulties of this extension were dealt with by several new ingredients in this article. These include two new evolution equations (Proposition 4.1 and Proposition 4.2) which allow us to localize estimates in [TTW22l] in the global setting. There is also a new regularization procedure that works without the convexity assumption. In particular, new non-convex self-expanders of Lagrangian mean curvature flows were discovered.
Given an entire function on , we consider the graph of as a Lagrangian submanifold of and deform it by the mean curvature flow, which corresponds to the following initial value problem for the potential function :
(1.1) |
As remarked above, this is a fully nonlinear Hessian type parabolic equation. We assume that is 2-convex, which is a 2-positivity assumption in terms of (see Definition 2.1 for the definition in terms of the eigenvalues of ). This is a natural quantity as it corresponds to the Hessian of as measured with respect to the induced metric on the corresponds Lagrangian submanifold.
The main result of this paper is the following long-time existence result.
Theorem 1.1 (Theorem 5.6).
Let be a -convex function with for some . Then, (1.1) admits a unique solution in the space such that
-
•
for any , is -convex;
-
•
there exists for any such that for any .
Depending on the asymptotic behavior of the initial data, we prove the following convergence theorems.
Theorem 1.2 (Theorem 6.1).
In other words, converges locally smoothly to as .
Theorem 1.3 (Theorem 6.2).
The paper is organized as follows. In Section 2, we consider the geometry of a Lagrangian submanifold in terms of its potential function. In Section 3, we review some known results about Lagrangian mean curvature flows that are needed in the article. In Section 4, we derive two important evolution equations that play critical roles in the proof of long-time existence. The long-time existence results are established in Section 5 and the convergence results in Section 6.
2. The Lagrangian Geometry in Potential
Endow the standard metric, symplectic form, and complex structure. Given a function , the graph of its gradient is a Lagrangian submanifold in . Denote it by .
At any , one may find a orthonormal basis to diagonalize the Hessian of , . Specifically, with respect to an orthonormal basis for . It follows that the tangent space of has the orthonormal basis
(2.1) |
for . In terms of the parametrization (the so-called non-parametric form in the minimal graph theory), the induced metric has metric coefficients
(2.2) |
We shall study two parallel tensors on . The first tensor is the volume form of . It is an -form on , and denote it by . The restriction of on is equivalent to the scalar-valued function , where is the Hodge star of the induced metric on . By using the frame (2.1),
(2.3) |
It is clear that takes value in .
The second one is a -tensor:
(2.4) |
With respect to the frame (2.1), the restriction of on is
In particular, it is positive definite if and only if is convex. The main interest of this paper is the -positivity case. Namely,
(2.5) |
for any . Note that does not correspond to a connected region in the -plane.
Definition 2.1.
A -function is said to be -convex if the eigenvalues of its Hessian satisfy everywhere
for any . It is said to be strictly -convex if both inequalities are strict.
As in [CNS85], we introduce a symmetric endomorphism on to study the -convexity of . It is denoted by , and is given by
with respect to an orthonormal frame of . The -convexity (2.5) condition corresponds to the positivity of , and thus it is useful to study the scalar valued function . In terms of (2.1),
(2.6) |
It is not hard to see that if is strictly -convex, takes value within .
When and , one can deduce that
(2.7) | ||||
(2.8) | ||||
(2.9) |
for any (under the strict -convexity assumption).
3. The Lagrangian Mean Curvature Flow
Given any , it is known that the Lagrangian mean curvature flow (up to a tangential diffeomorphism) equation on reduces to (1.1) for the potential function .
The uniqueness of the solution to (1.1) was established by Chen and Pang in [CP09].
Remark 3.1.
Consider the function
on the space of symmetric matrices. A direct computation shows that its derivative is
where .
We recall some preliminary results about short-time existence and finite time singularity established in [CCH12, CCY13].
Proposition 3.2 ([CCY13]*Proposition 2.1).
Suppose that is a smooth function with for any . Then, (1.1) admits a smooth solution for some . Moreover, for any and .
Lemma 3.3 ([CCH12]*Lemma 4.2).
Let be a smooth solution to (1.1) on for some . Suppose that
-
•
for any and ;
-
•
and are uniformly bounded on . Namely, there exist such that and on .
Then, there exists for any such that on .
Corollary 3.4.
Proof.
Suppose that the above limit is finite. Due to Lemma 3.3, all the higher order () derivatives of are uniformly bounded on . In particular, Proposition 3.2 applies to the family of initial data , and there is a such that the solution starting from exists on the time interval . It follows that the solution from can be extended over time , which contradicts to . ∎
We will also need the Liouville theorem and a priori estimate established by Nguyen and Yuan in [NY11]. Denote by . For any , denote by .
Proposition 3.5 ([NY11]*Proposition 2.1).
Theorem 3.6 ([NY11]*Theorem 1.1).
There exists a constant with the following significance. Let be a smooth solution to (1.1) in . Suppose that for any everywhere in . Then,
(3.1) |
4. Two Evolution Equations
We derive the evolution equations for and , defined in (2.3) and (2.6), respectively. They will be considered in the parametric form: the parametrization given by . That is to say, suppose that is a solution to (1.1), then is composing with a time-dependent diffeomorphism.
Proposition 4.1.
Let be a solution to (1.1). Suppose that is -convex for all . Then. satisfies
(4.1) |
in the parametric form of the mean curvature flow.
Proof.
With respect to the orthonormal frame (2.1),
By the Cauchy–Schwarz inequality,
According to [TTW22l]*Proposition 2.2,
Since , it finishes the proof of this proposition. ∎
Proposition 4.2.
Let be a solution to (1.1). Suppose that is strictly -convex for all . Then. satisfies
(4.2) |
in the parametric form of the mean curvature flow.
5. Long-Time Existence
5.1. A Maximum Principle
We first establish a maximum principle, whose proof is based on the argument in [EH91]*Theorem 2.1.
Lemma 5.1.
Let be a solution to the mean curvature flow111in the parametric form, . Suppose that is a positive smooth function satisfying
(5.1) |
for some constant , and for every . Then,
(5.2) |
for every .
Proof.
Let . It follows from that
(5.3) |
Fix , and consider the function . It follows from (5.3) that
(5.4) | ||||
(5.5) |
Choosing gives
(5.6) |
By replacing with , the computation remains valid. Due to the maximum principle,
(5.7) |
By letting , it implies that . ∎
5.2. Smooth Initial Condition
Proposition 5.2.
Let be a smooth function with for any . Suppose that
(5.8) |
for some . Then, (1.1) admits a unique smooth solution which has the following properties.
-
(i)
The solution is strictly -convex for all . Specifically, (5.8) is preserved along the flow.
-
(ii)
For any , there exists a constant such that for any .
Proof.
Step 1. Preservation of (5.8). Denote by the maximal existence time. Since
is a closed subset of , it remains to show that if , for some .
It suffices to do it for . Note that (2.7), (2.8) and (2.9) hold true everywhere at . By Proposition 3.2 and Remark 3.1, there exists such that remains -convex for . According to Proposition 4.2,
for . By applying Lemma 5.1 for , we find that for .
Step 2. Long Time Existence. The next step is to show the maximal existence time is infinity. Suppose not. By Corollary 3.4,
is unbounded as . It follows that there exists a sequence such that
-
•
as ;
-
•
as ;
-
•
for all .
Denote by . Let
for and . Then one has , , and on . It is straightforward to see that solves (1.1) on . Note that the eigenvalues of is the same as those of .
By using Lemma 3.3 and the Arzelà–Ascoli theorem, admits a subsequence which converges to in . Hence, is an ancient solution to (1.1) with , and the eigenvalues of satisfies (2.7), (2.8) and (2.9). Due to Proposition 3.5, is stationary. Since and ’s are bounded, [TW02]*Theorem A asserts that is an affine -plane. It contradicts to . Thus, the maximal existence time cannot be finite.
Step 3. Estimates. The argument for assertion (ii) follows from the a priori estimate [NY11]*Theorem 1.1 and the scaling argument; see [CCY13]*p.173. The case is included here for completeness. Fix . Let
Then, is a solution to (1.1) on . Since the second order derivative remains unchanged in the rescaling, it follows from Theorem 3.6 that . Due to (2.7), is uniformly bounded. Therefore,
for any with and . The estimate holds true for any , and it finishes the proof for . The estimate for follows a scaling argument; see the last part in the proof of [CCH12]*Lemma 5.2. ∎
5.3. Initial Condition
With Proposition 5.2, we can now prove the main theorem of this paper. In [CCH12]*section 5 and 6 and [CCY13]*section 3, the long-time existence was proved by constructing smooth approximations to the initial condition. Unlike their situation, the convolution between the standard mollifier and a -convex function needs not to be -convex. It requires some extra work to handle this issue.
The following lemma says that for a strict -convex region, one can find a positive cone such that the corresponding translation leaves the region invariant. In fact, the slope of the cone is determined by the corner points of the boundary of the region. See Figure 2 for the -convex region, and Figure 2 for a strictly -convex region.
Lemma 5.3.
For any and , let . There exists a depending on such that
where . The above sum of two sets is the Minkowski sum.


Proof.
Since and , and have exactly two intersection point; one in the second quadrant, and the other in the fourth quadrant. Denote the intersection point in the second quadrant by . It must belong to the region . It is not hard to see that is a one-to-one correspondence. The inverse map is given by and .
Let
Pick any . It is clear that if . It remains to verify that if and . Note that the condition implies that . Assume that ,
The argument for is similar. If and , it is obvious. ∎
Lemma 5.3 allows to construct some functions such that is still -convex, and becomes convex outside .
Lemma 5.4.
Given any , there exists a sequence of smooth functions with the following significance.
-
(i)
when .
-
(ii)
everywhere.
-
(iii)
At any , the ratio between any two eigenvalues of belongs to .
-
(iv)
For any , converges to uniformly over as , so do their derivatives.
Proof.
Let be the distance to the origin. For , its Hessian has two eigenvalues, and . The first one is simple, and the second one has multiplicity . Denote by . Let
(5.9) |
and then
(5.10) |
Thus, one only needs to specify , , and to construct . By requiring , the function is uniquely determined.
Fix a constant . For any , choose a smooth function for with
-
•
if or ;
-
•
if ;
-
•
if ;
-
•
if .
The base point is set to be , and the base value is set to be . One can see Figure 3 for an illustration of some ’s (by using piecewise constant ).
It follows that for . When ,
When ,
With a similar argument, when , where
It is not hard to see that the corresponding satisfies the assertions of this lemma. ∎

Note that Lemma 5.4 (ii) and (iii) implies that belongs to defined in Lemma 5.3, for any unit vectors .
Theorem 5.5.
Proof.
Let
By (2.8) and (2.9), the eigenvalues of satisfy and for any . Let be given by Lemma 5.3. Apply Lemma 5.4 for this to get a sequence of functions, .
It follows from Lemma 5.3, Lemma 5.4 and (2.7) that the eigenvalues of satisfy
(5.11) |
A complete discussion on the perturbation theory of eigenvalues of matrices can be found in [Kato76]*ch.1 and 2. By (5.11), there exist depending222The dependence on the dimension is always omitted in this paper. on such that
(5.12) |
for . Let be a symmetric matrix whose eigenvalues satisfy (5.11). It is not hard to see that there exists such that . Note that is a convex set, and there exists such that
(5.13) |
for any symmetric matrix333The functions and are defined for a symmetric matrix by using (2.3) and (2.6), respectively. with .
Let be the standard mollifiers; see for instance [Evans98]*Appendix C.4. They have the following properties: , , for all , and depends only on . For any and , let
(5.14) |
We claim that there exists a sequence of positive numbers with as such that obeys
(5.15) |
Note that
When , it follows from Lemma 5.4 (i) and (5.13) that for , and for any . When , it follows from (5.11), (5.12) and the uniform continuity of on that (5.15) holds true for on , provided that is sufficiently small. This finishes the proof of the claim.
Now, set to be . With (5.15) and
for any , one can apply Proposition 5.2 to the initial conditions . Denote the solution by , and it obeys (5.15) for all .
It follows from (5.15) and (2.7) that is uniformly bounded on . Since solves (1.1), is also uniformly bounded over . Fix and . Let . It follows from (5.14) and Lemma 5.4 (iv) that is uniformly bounded over . With the uniform boundedness of the time derivative, is uniformly bounded over . Together with the uniformly boundedness of , is uniformly bounded over . Hence, there exists a subsequence of which converges uniformly on any compact subset of . By Lemma 5.4 (iv) and as , .
This together with assertion (ii) of Proposition 5.2 implies that (a subsequence of) converges smoothly on any compact subset of . The limit is a solution to (1.1) with initial condition . The decay estimate on the derivatives of follows from that of . Since the constant depends on the , it is not hard to see from the last item of (5.11) that depends only on .
By construction, satisfies (5.15) for all . To see that still satisfies (5.8), we will need the local estimate (5.7) in Lemma 5.1. Fix an . For any , let be its compact neighborhood . Since is uniformly bounded over , there exists an such that
on . By Proposition 4.1 and Lemma 5.1, (5.7) for gives that
Due to the uniform continuity of over , Lemma 5.4 (iv), and ,
It follows that of over is no greater than . Since it is true for any , of is always greater than or equal to . The argument for is similar. ∎
Since a -convex initial can be approximated by strictly -convex functions, we can extend Theorem 5.5 to -convex initial condition.
Theorem 5.6.
Let be a -convex function with for some . Then, (1.1) admits a unique solution in the space such that
-
•
for any , is -convex and ;
-
•
there exists for any such that for any .
Proof.
Consider for . Since , it is not hard to see that there exist and such that and for . Denote by the solution to (1.1) with initial condition given by Theorem 5.5. Note that the bound of the derivatives given by Theorem 5.5 only depends on . By the same argument as that in the proof of Theorem 5.5, there exists a subsequence of which converges in and in . The properties asserted by this theorem follows from and the smooth convergence of in . ∎
6. Some Convergence Results
6.1. Potentials of Bounded Gradient
Theorem 6.1.
Proof.
The key step is to show that the bounded gradient condition is preserved along the flow. Denote by , by , etc. By (1.1) and Remark 3.1,
(6.1) |
where is the inverse of . It follows that
and hence
(6.2) |
Since along the flow, is uniformly bounded over . According to Theorem 5.6, is uniformly bounded. With (6.1), there exits for any a constant such that on . Thus, there is a such that on . By the maximum principle in [Fr64]*Theorem 9 on p.43, on .
With for any , converges in to some as . Since the convergence is , the image of is still a minimal Lagrangian submanifold. Thus, for some satisfying for some and . According to the Bernstein theorem [TW02]*Theorem A, must be an affine subspace. Since , the affine subspace must be . ∎
6.2. Convergence to Self-Expanders
A solution to the mean curvature flow is called a self-expander if the submanifold is the dilation of with the factor . It is a natural model for immortal solutions to the mean curvature flow, and can also be used to study the mean curvature flow of conical singularities. In our setting , one finds that being a self-expander means that
(6.3) |
for any . Therefore, for the entire Lagrangian mean curvature flow (1.1), is a self-expander if and only if obeys
(6.4) |
In [CCH12]*Theorem 1.2, it is shown that if the initial condition is strictly distance-decreasing, and is asymptotic to a cone at spatially infinity, then the rescaled flow converges to a self-expander. In the following proposition, we prove that the result holds true in the -convex case.
Theorem 6.2.
Proof.
It is straightforward to see that for any ,
is a solution to (1.1) with initial condition . Since the spatial Hessian remains unchanged in this rescaling, it follows from Theorem 5.6 that
for any , and on . With Remark 3.1, there exists for any and a constant such that
It is clear that and are uniformly bounded for . These estimates together with the Arzelà–Ascoli theorem implies that converges to some in as . Moreover, satisfies the equation (1.1), is -convex for all , and for all . It follows from the construction that satisfies for any , and thus is a self-expander.
Since , converges to a function as . Again by , one finds that
It finishes the proof of this proposition. ∎
On the other hand, it is shown in [CCH12]*Lemma 7.1 that for any which is strictly distance-decreasing and homogeneous of degree , there exists a self-expander to (1.1) which is strictly distance-decreasing for all time. A function is said to be homogeneous of degree if
for all . If is over , we can invoke Theorem 5.5. It is more interesting to assume that is only on , and it requires some more work to construct the solution.
Proposition 6.3.
Suppose that is homogeneous of degree , is on , and satisfy (5.8) on . Then, (1.1) admits a unique self-expanding solution in the space with initial condition such that and (5.8) is preserved along the flow.
Moreover, there exists for any such that for any .
It follows from the assumption that is uniformly Lipschitz. The following lemma is analogous to Lemma 5.4, and will be used to handle near the origin.
Lemma 6.4.
Given any , there exists a sequence of smooth functions with the following significance.
-
(i)
when .
-
(ii)
everywhere.
-
(iii)
At any , the ratio between any two eigenvalues of belongs to .
-
(iv)
For any , converges to uniformly over as , so do their derivatives.
Proof.
For any , choose a smooth function for with
-
•
if ;
-
•
if ;
-
•
for all .
Set the to be , and set to be . It follows that for . When ,
When ,
The function is determined by setting . It is not hard to see that the corresponding satisfies the assertions of this lemma. ∎
Proof of Proposition 6.3.
To start, note that there exists a such that on where . With this observation, the argument of Theorem 5.5 works for the initial condition . Denote the solution by . By the same argument as that in Theorem 5.5, (subsequentially) converges on compact subsets of , and converges smoothly on compact subsets of . The limit, , is a solution to (1.1) with initial condition .
Since is homogeneous of degree , (for any ) is also a solution to (1.1) with initial condition . It follows from the uniqueness theorem in [CP09] that for any . Hence, the solution is a self-expander. ∎