May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Communications in Contemporary Mathematics Vol. 16, No. 3 (2014) 1350027 (41 pages) c World Scientific Publishing Company  DOI: 10.1142/S0219199713500272

On the horizontal mean curvature flow for axisymmetric surfaces in the Heisenberg group

Fausto Ferrari Dipartimento di Matematica dell’Universit` a di Bologna Piazza di Porta S. Donato 5 40126, Bologna, Italy [email protected] Qing Liu∗ and Juan J. Manfredi† Department of Mathematics, University of Pittsburgh Pittsburgh, PA 15260, USA ∗[email protected][email protected] Received 19 November 2012 Revised 29 April 2013 Accepted 15 May 2013 Published 7 August 2013 We study the horizontal mean curvature flow in the Heisenberg group by using the levelset method. We prove the uniqueness, existence and stability of axisymmetric viscosity solutions of the level-set equation. An explicit solution is given for the motion starting from a subelliptic sphere. We also give several properties of the level-set method and the mean curvature flow in the Heisenberg group. Keywords: Mean curvature flow equation; Heisenberg group; viscosity solutions; level-set method. Mathematics Subject Classification 2010: 35K93, 35R03, 35D40

1. Introduction We are interested in a family of compact hypersurfaces {Γt }t≥0 parametrized by time t ≥ 0 in the Heisenberg group H. The motion of the hypersurfaces is governed by the following law: VH = κH ,

(1.1)

where VH denotes its horizontal normal velocity and κH stands for the horizontal mean curvature in the Heisenberg group. Equation (1.1) is thus called horizontal mean curvature flow. The objective of this work is to investigate the evolution of the surface Γt for t > 0 for a general class of initial surface Γ0 . 1350027-1

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

We implement a version of the level-set method adapted to the Heisenberg group. Let us assume, for the moment, that Γt is smooth for any t ≥ 0. If there exists u ∈ C 2 (H × [0, ∞)) such that Γt = {p ∈ H : u(p, t) = 0} for t ≥ 0, then one may represent the horizontal normal velocity VH as ut VH = |∇H u| and the horizontal mean curvature κH as      ∇H u ∇H u ⊗ ∇H u 1 2 ∗ tr I − u) κH = divH (∇ . = H |∇H u| |∇H u| |∇H u|2 2 u)∗ respectively denote the derivative in t, the horizontal Here ut , ∇H u and (∇H gradient and the (symmetrized) horizontal Hessian of u. The notation divH denotes the horizontal divergence operator. The horizontal gradient of u is given by ∇H u = (X1 u, X2 u), where

X1 =

∂ p2 ∂ − , ∂p1 2 ∂p3

X2 =

∂ p1 ∂ + . ∂p2 2 ∂p3

In order to understand the law of motion by curvature (1.1), it therefore suffices to solve      2 ∗  ut − tr I − ∇H u ⊗ ∇H u (∇H u) = 0 in H × (0, ∞), (1.2) |∇H u|2 (MCF) =   u(p, 0) = u0 (p) in H, (1.3) with a given function u0 ∈ C(H) satisfying Γ0 = {p ∈ H : u0 (p) = 0}. We refer the reader to [8, 13, 17] for a detailed derivation of (MCF) in the Euclidean spaces and to [7, 6] for the analogue in the Heisenberg group. In this work we aim to establish the uniqueness, existence and stability of the solutions of (MCF) that are spatially axisymmetric about the third coordinate axis. Namely, we are interested in the solutions u satisfying u(p1 , p2 , p3 , t) = u(p1 , p2 , p3 , t) when (p1 )2 + (p2 )2 = p21 + p22 .

(1.4)

The symmetric structure of the functions is useful to obtain positive results. We thus consider our contribution as a first step in order to prove more general results. See [1, 31] for the results on motion by mean curvature for axisymmetric surfaces in the Euclidean spaces. The symmetry with respect to the third axis in the Heisenberg group is not accidental. Indeed it is well known that this coordinate plays a key role in the 1350027-2

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Heisenberg group in several cases. In particular, we recall, for example, that {(0, 0, p3 ) ∈ H : p3 ∈ R} is the center of the Heisenberg group and moreover the points along the p3 -axis correspond to conjugate points for the exponential map [27]. We warn the reader that, in general, our results do not apply to functions with different axes of symmetry. Hereafter the property (1.4) is sometimes referred to as “spatial symmetry about the vertical axis” or simply as “axisymmetric”. Since the general regularity of u is not known a priori, we discuss the problem in the framework of viscosity solutions [10]. As it is easily observed from the equation, a key difficulty lies at the characteristic set of the level set Γt , i.e. at the points where ∇H u = 0. 1.1. Uniqueness Even in the Euclidean case [8, 13, 30, 17], the proof of the comparison principle and the uniqueness of solutions for this type of degenerate equations need special techniques to deal with the characteristic set. The comparison principle we expect is as follows: for any upper semicontinuous subsolution u and lower semicontinuous supersolution v defined on H × [0, ∞) satisfying u(p, 0) ≤ v(p, 0) for all p ∈ H, we have u(p, t) ≤ v(p, t) for any t ≥ 0. Capogna and Citti [6] extended the results of [13] and proved a comparison principle by excluding the characteristic points. Their comparison principle further required that (i) either u or v be uniformly continuous and (ii) the initial surfaces are completely separated in the horizontal directions, i.e. u(p, 0) ≤ v(q, 0) for all p = (p1 , p2 , p3 ), q = (q1 , q2 , q3 ) ∈ H such that pi = qi for i = 1, 2. The general comparison principle, as stated above, remains an open question. In this paper, we give a comparison principle without assuming (i) and (ii) above but requiring that either u or v be axisymmetric. We also restrict ourselves to the case of compact surfaces for simplicity. The comparison theorem we present is as follows. Theorem 1.1 (Comparison theorem). Let u and v be respectively an upper semicontinuous subsolution and a lower semicontinuous supersolution of    ∇H u ⊗ ∇H u 2 ∗ u) (∇ =0 ut − tr I − H |∇H u|2 in H × (0, T ) for any T > 0. Assume that there is a compact set K ⊂ H and a, b ∈ R with a ≤ b such that u(p, t) = a and v(p, t) = b for all p ∈ H\K and t ∈ [0, T ]. Assume that either u or v is spatially axisymmetric about the vertical axis. If u(p, 0) ≤ v(p, 0) for all p ∈ H, then u(p, t) ≤ v(p, t) for all (p, t) ∈ H×[0, T ). The uniqueness of the axisymmetric solutions follows immediately from the theorem above. It is worth remarking that when showing comparison principles 1350027-3

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

involving viscosity solutions, one usually needs to double the variables and maximize φ(p, q) + |t − s|2 , ε where ε > 0, p, q ∈ H, t, s ∈ [0, ∞) and φ is a smooth penalty function on H × H, and to argue by contradiction. The typical choice of φ in the Euclidean spaces, as discussed in [10, 17], is a quadratic function φ(x, y) = |x − y|2 usually or a quartic function φ(x, y) = |x − y|4 for mean curvature flow equation (for x, y ∈ Rn ). The advantages of these choices are: u(p, t) − v(q, s) −

(a) The derivatives of φ with respect to x and y are opposite, i.e. ∇x φ = −∇y φ. We would plug these derivatives in the viscosity inequalities, since they serve as semi-differentials for the compared functions. This construction enables us to derive a contradiction. (b) When discussing (mild) singular equations such as curvature flow equations, it will be convenient to have the second derivatives be 0 whenever the first derivatives are 0, as in the case of |x − y|4 . The analogue of the choice |x − y|4 is not immediate in the Heisenberg group. Since the group multiplication is not commutative, the two natural choices f (p, q) = |q −1 · p|4 and g(p, q) = |p · q −1 |4 are different. It seems that we have more options but it turns out that neither of them satisfies both conditions above. By direct calculation, we may find that g fulfills the requirement (a) above but its derivatives do not satisfy (b). The function f is good for our requirement (b) but unfortunately fails to have the property (a). Hence, the main difficulty of the uniqueness argument in the Heisenberg group consists in a wise choice of the penalty function φ. Our approach combines both choices f and g. On the one hand, we use f to derive a relaxed definition of solution to (1.2), facilitating us to overcome the singularity; see Definition 3.2. On the other hand, under the help of axial symmetry, we obtain the property (b) when employing g type of penalty functions in the proof of the comparison principle. The symmetry plays an important role since it largely simplifies the structure of characteristic points; see [16] for some geometric details. Roughly speaking, when a smooth function u : H×R → R is spatially symmetric about the vertical axis, i.e. u = u(r, p3 , t), where r = (p21 + p22 )1/2 , we get X1 u =

p2 ∂ p1 ∂ u− u, r ∂r 2 ∂p3

X2 u =

p1 ∂ p2 ∂ u+ u. r ∂r 2 ∂p3

Then ∇H u(p, t) = 0 implies that either ∂u/∂r = ∂u/∂p3 = 0 or p21 + p22 = 0. This observation enables us to obtain property (b) for the function g. Our definition of viscosity solutions is actually an extension of that introduced in [8, 17] to the Heisenberg group. In Sec. 3, we discuss the equivalence of this definition and the others. 1350027-4

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

1.2. Existence Generally speaking, there are at least three possible approaches to get the existence of solutions of (MCF). One may follow [13] to use the uniformly parabolic theory by considering a regularized equation      ut − tr I − ∇H u ⊗ ∇H u (∇2 u)∗ = 0 in H × (0, ∞), H |∇H u|2 + ε2  u(p, 0) = u (p) in H 0

and take the limit of its solution as ε → 0; see [6] for results in the Carnot groups with this method. Another possible option is to employ Perron’s method by considering the supremum of all subsolutions or the infimum of all supersolutions, as shown in [8, 17] for the Euclidean case. We refer to [20, 10] for a general introduction of this method in the framework of viscosity solutions. A third method for existence is based on the representation theorem involving optimal control or game theory, which recently generated a spur of activity. Consult the works [9, 21, 22, 25, 26, 28, 29, 32] for the development of this approach to various equations in the Euclidean spaces. For the mean curvature flow in the sub-Riemannian geometry, a stochastic control-based formulation analogous to [32] is addressed in [12], where the authors found a solution via a suitable optimal stochastic control problem. In this work, we adapt the deterministic game-theoretic approach of Kohn and Serfaty [21] to the Heisenberg group. For any given axisymmetric continuous function u0 , we set up a family of games, whose value functions uε converge to the solution u to the mean curvature flow equation. We not only get the existence of solutions but also obtain a game interpretation of the equation in the Heisenberg group. The proof is based on the dynamic programming principle, which can be regarded as a (nonlinear) semigroup. Our convergence theorem relies on the comparison principle given in Theorem 1.1. More precisely, taking the half relaxed limits, defined on H × [0, ∞), u(p, t) := lim sup∗ uε (p, t) ε→0

= lim sup{uε (q, s) : s ≥ 0, 0 < ε < δ, |p − q| + |t − s| < δ} δ→0

(1.5)

and u(p, t) := lim inf∗ uε (p, t) ε→0

= lim inf{uε (q, s) : s ≥ 0, 0 < ε < δ, |p − q| + |t − s| < δ}, δ→0

(1.6)

we show that u and u are respectively a subsolution and a supersolution of (1.2) using the dynamic programming principle. We also show that u(p, 0) ≤ u0 (p) ≤ u(p, 0) and that uε , u and u are spatially axisymmetric about the vertical axis. Our game approximation then follows immediately from the comparison theorem. See Sec. 5 for more details on the game setting and the existence theorem. 1350027-5

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

We discuss asymptotic mean value properties related to random tug-of-war games for p-harmonic functions on the Heisenberg group in [15]. 1.3. Stability and uniqueness of the evolution We present a stability theorem (Theorem 6.1), which is used to show that Eq. (1.2) is invariant under a change of dependent variable. Once the comparison principle is known, the stability result is a straightforward adaptation of that in the Euclidean spaces, as given in [13, 8, 4]. We prove that even in the Heisenberg group for any continuous function θ : R → R, the composition θ ◦ u is a solution provided that u is a solution. Note that this is clear if θ is smooth and strictly monotone, since the mean curvature flow equation is geometric and orientation-free; see [17] for more explanation. This stability result is applied to weaken the regularity of θ. It follows from the invariance property that any axisymmetric evolution Γt does not depend on the particular choice of u0 but depends on Γ0 only, which is important for the level-set method. 1.4. Evolution of spheres Our uniqueness and existence results enable us to discuss motion by mean curvature with a variety of initial hypersurfaces including spheres, tori and other compact surfaces. We are particularly interested in the motion of a subelliptic sphere. It turns out that if u0 is a defining function of the sphere centered at 0 with radius r, say u0 (p) = min{(p21 + p22 )2 + 16p23 − r4 , M } with p = (p1 , p2 , p3 ) ∈ H and M > 0 large, then the unique solution of (MCF) is u(p, t) = min{(p21 + p22 )2 + 12t(p21 + p22 ) + 16p23 + 12t2 − r4 , M } for any t ≥ 0. We need to truncate the initial function and the solution by a constant M because all of our well-posedness results are for solutions that are constant outside a√compact set. It is obvious that the zero level set Γt of u vanishes after time t = r2/ 12, which, by Theorem 1.1, indicates that all compact surfaces under the motion by horizontal mean curvature disappear in finite time. To understand the asymptotic profile at the extinction time, we normalize the √ evolution Γt initialized from the sphere and find that the normalized surface Γt / r4 − 12t2 converges to an ellipsoid given by the following equation: √ 2 2 12r (p1 + p22 ) + 16p23 = 1. The asymptotic profile above depends on r, the size of the initial surface, which is quite different from the Euclidean case. The paper is organized in the following way. We present an introduction in Sec. 2.1 about the Heisenberg group including calculations of some particular functions we will use later. In Sec. 3, we discuss various kinds of definitions of solution 1350027-6

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

to (1.2). We propose a new definition and show its equivalence with the others. An explicit solution related to the evolution of a subelliptic sphere is given at the end of this section. The comparison principle, Theorem 1.1, is proved in Sec. 4. We establish the games and show the existence theorem in Sec. 5. Section 6 includes a stability result, and Sec. 7 is intended to show further properties of the evolution including the uniqueness and finite extinction with the interesting asymptotic profile. In order to improve the readability of this paper, many technical proofs are collected in Appendix A. Moreover in Appendix B we present the relevant results on viscosity solutions and semijets in the Heisenberg group. 2. Tools from Calculus in H Good references for this section are the course notes [24] and the monograph [7]. 2.1. Preliminaries Recall that the Heisenberg group H is R3 endowed with the noncommutative group multiplication   1 (p1 , p2 , p3 ) · (q1 , q2 , q3 ) = p1 + q1 , p2 + q2 , p3 + q3 + (p1 q2 − q1 p2 ) 2 for all p = (p1 , p2 , p3 ) and q = (q1 , q2 · q3 ) in H. Note that the group inverse of p = (p1 , p2 , p3 ) is p−1 = (−p1 , −p2 , −p3 ). The Haar measure in H is the usual anyi gauge is given by Lebesgue measure in R3 . The Kor´ |p| = ((p21 + p22 )2 + 16p23 )1/4 and the left-invariant Kor´ anyi or gauge metric is d(p, q) = |q −1 · p|. The Kor´ anyi ball of radius r > 0 centered at p is Br (p) := {q ∈ H : d(p, q) < r}. The Lie Algebra of H is generated by the left-invariant vector fields X1 =

∂ p2 ∂ − , ∂p1 2 ∂p3

X2 =

∂ p1 ∂ + , ∂p2 2 ∂p3

X3 =

∂ . ∂p3

One may easily verify the commuting relation X3 = [X1 , X2 ] = X1 X2 − X2 X1 . 1350027-7

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

For any smooth real-valued function u defined in an open subset of H, the horizontal gradient of u is ∇H u = (X1 u, X2 u), while the complete gradient of u is ∇u = (X1 u, X2 u, X3 u). For further details about the relation between sub-Riemannian metrics in Carnot group and Riemannian metrics; see, for example, [2]. 2 u)∗ is the 2 × 2 symmetry The symmetrized second horizontal Hessian (∇H matrix given by   (X1 X2 u + X2 X1 u) 2 u X 1   2   2 u)∗ :=  (∇H .   (X X u + X X u) 1 2 2 1 X22 u 2 We will also consider the symmetrized complete Hessian (∇2 u)∗ defined as the 3 × 3 symmetric matrix   (X1 X2 u + X2 X1 u) (X1 X3 u + X3 X1 u) 2 X1 u   2 2        (X1 X2 u + X2 X1 u) X u + X X u)  (X 2 ∗ 2 3 3 2 (∇ u) :=  . X22 u   2 2      (X X u + X X u) (X X u + X X u)  1 3 3 1 2 3 3 2 2 X3 u 2 2 2.2. Derivatives of auxiliary functions Here we include several basic calculations for some test functions related to the Kor´ anyi distance, which will be used in the proof of comparison theorem for generalized horizontal mean curvature flow. We are interested in the first and second horizontal derivatives of f (p, q) := d(p, q)4  2 1 1 = ((p1 − q1 )2 + (p2 − q2 )2 )2 + 16 p3 − q3 − q1 p2 + q2 p1 . 2 2 We use the super index p to denote derivatives with respect to the p variable and follow the same convention for derivatives with respect to q. Let us record the results of our calculation: X1p f (p, q) = 4((p1 − q1 )2 + (p2 − q2 )2 )(p1 − q1 )   1 − 16(p2 − q2 ) p3 − q3 + (q2 p1 − q1 p2 ) , 2 1350027-8

(2.1)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

X2p f (p, q) = 4((p1 − q1 )2 + (p2 − q2 )2 )(p2 − q2 )   1 + 16(p1 − q1 ) p3 − q3 + (q2 p1 − q1 p2 ) , 2 X1q f (p, q) = −4((p1 − q1 )2 + (p2 − q2 )2 )(p1 − q1 )   1 − 16(p2 − q2 ) p3 − q3 + (q2 p1 − q1 p2 ) , 2 X2q f (p, q) = −4((p1 − q1 )2 + (p2 − q2 )2 )(p2 − q2 )   1 + 16(p1 − q1 ) p3 − q3 + (q2 p1 − q1 p2 ) . 2

(2.2)

(2.3)

(2.4)

p q It is clear that in general ∇H f (p, q) = −∇H f (p, q), which is not the case in the Euclidean case. But the following Euclidean property still holds here. p q (|q −1 · p|4 ) = 0 or ∇H (|q −1 · p|4 ) = 0, then the Proposition 2.1. If either ∇H horizontal components of p and q are equal, i.e. p1 = q1 and p2 = q2 .

Proof. Set A := 4((p1 − q1 )2 + (p2 − q2 )2 ),   1 B := 16 p3 − q3 + (q2 p1 − q1 p2 ) . 2 p When ∇H (|q −1 · p|4 ) = 0, the calculations (2.1) and (2.2) read  A(p1 − q1 ) − B(p2 − q2 ) = 0,

B(p1 − q1 ) + A(p2 − q2 ) = 0

(2.5)

A −B 2 2 2 2 with det(B A) = A + B ≥ 0. Since A + B = 0 implies that pi = qi for i = 1, 2, 2 2 the desired result is trivial if A + B = 0. If the determinant is not zero, then we also obtain q1 = p1 and q2 = p2 by solving the linear system (2.5). The same q (|q −1 · p|4 ) = 0. argument applies to the case when ∇H

We next calculate the second horizontal derivatives. X12,p f (p, q) = X12,q f (p, q) = 12(p1 − q1 )2 + 12(p2 − q2 )2 , X22,p f (p, q) = X22,q f (p, q) = 12(p2 − q2 )2 + 12(p1 − q1 )2 ,   1 X2p X1p f (p, q) = X1q X2q f (p, q) = −16 p3 − q3 + (q2 p1 − q1 p2 ) = −B, 2   1 X1p X2p f (p, q) = X2q X1q f (p, q) = 16 p3 − q3 + (q2 p1 − q1 p2 ) = B. 2 It is clear that 1 p p 1 (X X f + X2p X1p f ) = (X1q X2q f + X2q X1q f ) = 0. 2 1 2 2 1350027-9

(2.6) (2.7) (2.8) (2.9)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

For later use, let us investigate the derivatives of another function. Take g(p, q) := |p · q −1 |4

2  1 1 = ((p1 − q1 )2 + (p2 − q2 )2 )2 + 16 p3 − q3 − p1 q2 + p2 q1 . 2 2

(2.10)

Then X1p g(p, q) = 4((p1 − q1 )2 + (p2 − q2 )2 )(p1 − q1 )   1 1 − 16(p2 + q2 ) p3 − q3 − p1 q2 + p2 q1 , 2 2 X2p g(p, q) = 4((p1 − q1 )2 + (p2 − q2 )2 )(p2 − q2 )   1 1 + 16(p1 + q1 ) p3 − q3 − p1 q2 + p2 q1 , 2 2 X1q g(p, q) = −4((p1 − q1 )2 + (p2 − q2 )2 )(p1 − q1 )   1 1 + 16(p2 + q2 ) p3 − q3 − p1 q2 + p2 q1 , 2 2 X2q g(p, q) = −4((p1 − q1 )2 + (p2 − q2 )2 )(p2 − q2 )   1 1 − 16(p1 + q1 ) p3 − q3 − p1 q2 + p2 q1 . 2 2

(2.11)

(2.12)

(2.13)

(2.14)

p q Remark 2.1. In this case, we do have ∇H g(p, q) = −∇H g(p, q). But the property in Proposition 2.1 does not hold in general.

The second derivatives are given below: X12,p g(p, q) = X12,q g(p, q) = 12(p1 − q1 )2 + 4(p2 − q2 )2 + 8(p2 + q2 )2 ,

(2.15)

X22,p g(p, q) = X22,q g(p, q) = 4(p1 − q1 )2 + 12(p2 − q2 )2 + 8(p1 + q1 )2 ,

(2.16)

X1p X2p g(p, q) = X2q X1q g(p, q) = 8(p1 − q1 )(p2 − q2 ) − 8(p1 + q1 )(p2 + q2 )   1 1 + 16 p3 − q3 − p1 q2 + p2 q1 , 2 2 X2p X1p g(p, q) = X1q X2q g(p, q) = 8(p1 − q1 )(p2 − q2 ) − 8(p1 + q1 )(p2 + q2 )   1 1 − 16 p3 − q3 − p1 q2 + p2 q1 , 2 2

(2.17)

(2.18)

1 p p 1 (X1 X2 g + X2p X1p g) = (X1q X2q g + X2q X1q g) 2 2 = 8(p1 − q1 )(p2 − q2 ) − 8(p1 + q1 )(p2 + q2 ). 1350027-10

(2.19)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

2.3. Extrema in the Heisenberg group As |p|2 ≈ p21 + p22 + |p3 | in Heisenberg group, the Taylor formula reads 1 2 p) + (∇H u)∗ (ˆ p)h, h + o(|ˆ p−1 · p|2 ), u(p) = u(ˆ p) + ˆ p−1 · p, ∇u(ˆ 2

(2.20)

where h = (p1 − pˆ1 , p2 − pˆ2 ) is the horizontal projection of pˆ−1 · p. The following proposition follows easily from the Euclidean analog. Proposition 2.2 (Maxima on Heisenberg group). Suppose O is an open subp) for all p ∈ O, then ∇u(ˆ p) = 0 set of H. Let u ∈ C 2 (O) and pˆ ∈ O. If u(p) ≤ u(ˆ 2 u)∗ (ˆ p) ≤ 0. and (∇H Analogously, for minima we have that if u(p) ≥ u(ˆ p) for all p ∈ O, then 2 u)∗ (ˆ p) ≥ 0. ∇u(ˆ p) = 0 and (∇H 3. Definitions of Solutions 3.1. General definitions For a vector η ∈ R2 and a 2 × 2 symmetric matrix Y ∈ S2 we define    η⊗η F (η, Y ) = −tr I− Y . |η|2 In any open subset O ⊂ H × (0, ∞) the mean curvature flow equation    ∇H u ⊗ ∇H u 2 ∗ u) (∇ = 0 in O ut − tr I − H |∇H u|2

(3.1)

(3.2)

can be written as 2 ut + F (∇H u, (∇H u)∗ ) = 0 in O.

We next define the semicontinuous envelopes in the following way: for any function h defined on a set O of a metric space M with values in R ∪ {±∞}, we take h (x) = lim sup{h(y) : y ∈ O ∩ Br (x)}

(3.3)

h (x) = lim inf{h(y) : y ∈ O ∩ Br (x)}

(3.4)

r→0

and r→0

for any x ∈ O, where Br (x) denotes the ball with radius r > 0 centered at x. It is easily seen that F  (0, 0) = F (0, 0) = 0, F  (η, X) = F (η, X) = F (η, X) for all (η, X) ∈ R2 \{0} × S2 . One type of definition [8, 10] of viscosity solutions of (3.2) is as follows. Definition 3.1. An upper (respectively, lower) semicontinuous function u : O → R ∪ {±∞}, where O ⊂ H × (0, ∞), is a subsolution (respectively, supersolution) 1350027-11

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

of (3.2) if (i) u < ∞ (respectively, u > −∞) in O; (ii) for any smooth function φ and (ˆ p, tˆ) × O such that max(u − φ) = (u − φ)(ˆ p, tˆ), O

  respectively, min(u − φ) = (u − φ)(ˆ p, tˆ) , O

the function φ satisfies 2 φt + F (∇H φ, (∇H φ)∗ ) ≤ 0 at (ˆ p, tˆ), 2 φ)∗ ) ≥ 0 at (ˆ p, tˆ)). (respectively, φt + F  (∇H φ, (∇H

A function u is called a solution of (3.2) if it is both a subsolution and a supersolution. We now propose another definition for the horizontal mean curvature flow equation. This definition appears in [3, 11, 17] in the Euclidean case. Definition 3.2. An upper (respectively, lower) semicontinuous function u : O → R ∪ {±∞}, where O ⊂ H × (0, ∞), is a subsolution (respectively, supersolution) of (3.2) if (i) u < ∞ (respectively, u > −∞) in O; (ii) for any smooth function φ and (ˆ p, tˆ) × O such that 

p, tˆ), max(u − φ) = (u − φ)(ˆ O

 respectively, min(u − φ) = (u − φ)(ˆ p, tˆ) , O

the function φ satisfies 2 φ)∗ ) ≤ 0 at (ˆ p, tˆ), φt + F (∇H φ, (∇H 2 (respectively, φt + F (∇H φ, (∇H φ)∗ ) ≥ 0 at (ˆ p, tˆ)),

when ∇H φ(ˆ p, tˆ) = 0 and φt (ˆ p, tˆ) ≤ 0, (respectively, φt (ˆ p, tˆ) ≥ 0), 2 when ∇H φ(ˆ p, tˆ) = 0 and (∇H φ)∗ (ˆ p, tˆ) = 0.

A function u is called a solution of (3.2) if it is both a subsolution and a supersolution. Remark 3.1. One may replace the maximum (respectively, minimum) in condition (ii) of the above definitions with a strict maximum by adding a positive (respectively, negative) smooth gauge to φ. 1350027-12

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Roughly speaking, in Definition 3.2 we restrict the test function space to the following A0 = {φ ∈ C ∞ (H) : ∇H φ(p) = 0 implies (∇2H )∗ φ(p) = 0}. It brings us convenience in the proof of comparison principle. Another type of definitions, as below, involves subelliptic semijets, whose definition is presented in Appendix B. Definition 3.3. An upper (respectively, lower) semicontinuous function u : O → R ∪ {±∞}, where O ⊂ H × (0, ∞), is a subsolution (respectively, supersolution) of (3.2) if (1) u < ∞ (respectively, u > −∞) in O; 2,+ 2,− p, tˆ) (respectively, (τ, η, X ) ∈ J H u(ˆ p, tˆ)) with (ˆ p, tˆ) ∈ (2) for any (τ, η, X ) ∈ J H u(ˆ O, we have 2 φ)∗ ) ≤ 0 φt + F (∇H φ, (∇H

at (ˆ p, tˆ),

2 φ)∗ ) ≥ 0 (respectively, φt + F  (∇H φ, (∇H

at (ˆ p, tˆ)).

A function u is called a solution of (3.2) if it is both a subsolution and a supersolution. It is not hard to see that Definition 3.3 is equivalent to Definition 3.1. The next result, which is actually a variant of [17, Proposition 2.2.8] for the Heisenberg group, indicates the equivalence between Definitions 3.2 and 3.1. Proposition 3.1 (Equivalence of definitions). An upper (respectively, lower) semicontinuous function u : O → R is a subsolution (respectively, supersolution) of (3.2) defined as in Definition 3.2 (in O) if and only if it is a subsolution (respectively, supersolution) in O in the sense of Definition 3.1. Proof. It is obvious that Definition 3.2 is a relaxation of Definition 3.1. We prove the reverse implication only for subsolutions. The statement for supersolutions can be proved similarly. Suppose there are a smooth function φ and (ˆ p, tˆ) ∈ O such that p, tˆ). max(u − φ) = (u − φ)(ˆ O

By usual modification in the definition of viscosity solutions, we may assume it is a strict maximum. We construct 1 Ψε (p, q, t) := u(p, t) − |q −1 · p|4 − φ(q, t). ε It is clear that  u(p, t) − φ(p, t) if p = q, ∗ ∗ Ψ (p, q, t) := lim sup Ψε (p, q, t) = ε→0 −∞ if p = q 1350027-13

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

attains a strict maximum at (ˆ p, pˆ, tˆ). By the convergence of maximizers ([17, Lemma ε ε ε 2.2.5] or [10]), we may take p , q , t converging to pˆ, pˆ, tˆ respectively as ε → 0 such that Ψε attains a maximum at (pε , q ε , tε ). It follows that q → − 1ε |q −1 · pε |4 − φ(q, t) has a maximum at q ε , which, by Proposition 2.2, implies that 1 q − ∇H f (pε , q ε ) = ∇H φ(q ε , tε ), ε

(3.5)

1 2,q ∗ ε ε 2 f ) (p , q ) ≤ (∇H φ)∗ (q ε , tε ), − (∇H ε

(3.6)

where f (p, q) = |q −1 · p|4 . We next discuss the following two cases. Case A. ∇H φ(q ε , tε ) = 0 for a subsequence of ε → 0. (We still use ε to denote the subsequence.) Since the maximality of Ψ at (pε , q ε , tε ) implies that 1 (p, t) → u(p, t) − f (pε , q ε ) − φ(p · (pε )−1 · q ε , t) ε attains a maximum at (pε , tε ) ∈ O. Denote φε (p, t) = φ(p · (pε )−1 · q ε , t). We apply Definition 3.2 to get 2 ε ∗ φt + F (∇H φε , (∇H φ ) ) ≤ 0 at (pε , tε ).

(3.7)

Since the derivative of the right multiplication tends to 0 as ε → 0 and its second derivatives are 0, we have 2 ε ∗ ε ε 2 ∇H φε (pε , tε ) → ∇H φ(ˆ p, tˆ) and (∇H φ ) (p , t ) → (∇H φ)∗ (ˆ p, tˆ) as ε → 0.

It follows immediately that 2 φt + F (∇H φ, (∇H φ)∗ ) ≤ 0 at (ˆ p, tˆ).

Case B. ∇H φ(q ε , tε ) = 0 for all sufficiently small ε > 0. q f (pε , q ε ) = 0, which by Proposition 2.1 yields that It follows from (3.5) that ∇H pεi = qiε

for i = 1, 2.

(3.8)

In terms of (2.1)–(2.2) and (2.6)–(2.9), we have p 2,p ∇H f (pε , q ε ) = 0 and ∇H f (pε , q ε ) = 0.

(3.9)

Since (pε , tε ) is a maximizer of 1 (p, t) → u(p, t) − f (p, q ε ) − φ(q ε , t), ε applying Definition 3.2 and sending the limit, we obtain φt (ˆ p, tˆ) ≤ 0,

(3.10)

On the other hand, by passing to the limit in (3.5) and (3.6), we have ∇H φ(ˆ p, tˆ) = 0 1350027-14

(3.11)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

and 2 ∇H φ(ˆ p, tˆ) ≥ 0.

(3.12)

By (3.11), (3.10) is equivalent to p, tˆ) + F (∇H φ(ˆ p, tˆ), 0) ≤ 0, φt (ˆ which, thanks to (3.12) and the ellipticity of F , implies that 2 φt (ˆ p, tˆ) + F (∇H φ(ˆ p, tˆ), (∇H φ)∗ (ˆ p, tˆ)) ≤ 0.

3.2. An explicit solution We provide an example of solutions of (3.2) when the initial value is the fourth power of a smooth gauge of the Heisenberg group. We can actually express a solution explicitly. Proposition 3.2. For any p = (p1 , p2 , p3 ) ∈ H, let G(p) = |p|4 = (p21 + p22 )2 + 16p23 .

(3.13)

w(p, t) = (p21 + p22 )2 + 12t(p21 + p22 ) + 16p23 + 12t2

(3.14)

Then

is a solution of (1.2) and w(p, 0) = G(p). Proof. Since w is smooth, the proof is based on a straightforward calculation of the first derivatives of w wt = 12(p21 + p22 ) + 24t, X1 w = Kp1 − 16p2 p3 ,

(3.15)

X2 w = Kp2 + 16p1 p3 ,

where K := 4(p21 + p22 ) + 24t and the second derivatives X12 w = X22 w = 12p21 + 12p22 + 24t, X1 X2 w = 16p3 , X2 X1 w = −16p3,   12p21 + 12p22 + 24t 0 2 ∗ (∇H w) = . 0 12p21 + 12p22 + 24t

(3.16)

2 Noting that (∇H w)∗ is constant multiple of the identity, we easily conclude from our calculation that 2 2 w)∗ ) = F (∇H w, (∇H w)∗ ) F  (∇H w, (∇H    ∇H w ⊗ ∇H w 2 ∗ = tr I − w) (∇ = 12(p21 + p22 ) + 24t = wt , H |∇H w|2

which means that w satisfies (1.2) by Definition 3.1. 1350027-15

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

Remark 3.2. There is another way to understand that w is a solution of (1.2) by adopting Definition 3.2 when ∇H w = 0 at (p, t) ∈ H × (0, ∞). If ∇H w(p, t) = 0, we have p1 = p2 = 0 by solving a linear system      0 p1 K −16p3 = 0 16p3 K p2 with

 det

K

−16p3

16p3

K

In addition,

 = K 2 + 16p23 > 0. 

2 (∇H w)∗

=



24t

0

0

24t

.

Note that, by Proposition 2.2, it is not possible to take a smooth function φ touching w from above at (p, t) with ∇H φ(p, t) = 0

2 and (∇H φ)∗ (p, t) = 0.

(3.17)

Therefore w is a subsolution of (1.2) at (p, t) by Definition 3.2. On the other hand, whenever a test function φ touches w from below at (p, t) with (3.17), we get φt (p, t) = wt (p, t) = 24t > 0, which implies that w is also a supersolution due to Definition 3.2. Remark 3.3. A basic transformation keeps the solution (3.14) being a solution. To be more precise, for any fixed c ∈ R, L > 0 and pˆ ∈ H, we define w(p, ˆ t) = −1 ˆ is a solution of Lw(ˆ p · p, t) + c for all (p, t) ∈ H × [0, ∞). Then we claim that w (1.2). Indeed, our calculation above extends to ˆ = X22 w ˆ = 12L(p1 − pˆ1 )2 + 12L(p2 − pˆ2 )2 + 24Lt, X12 w   1 1 ˆ = −X2 X1 w ˆ = 16L p3 − pˆ3 − pˆ1 p2 + p1 pˆ2 . X1 X2 w 2 2 The conclusion follows immediately as in the proof of Proposition 3.2. A primary geometric observation for the explicit solution u in (3.14) is as follows. For any fixed µ > 0, the µ-level set, Γµt = {p ∈ H : w(p, t) = µ}, describes the position of surface at time t ≥ 0. It is obvious that even if Γµ0 = ∅, Γt will vanish when t is sufficiently large, which agrees with the usual extinction for mean curvature flows in the Euclidean spaces [15]. We will revisit this property in Sec. 7. A natural question now is whether the explicit solution we found is the only solution of (MCF) with the initial data (3.13). This is related to the open question on the uniqueness of solutions of (MCF). In the following sections we will give an 1350027-16

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

affirmative answer for the case when the initial data are cylindrically symmetric about the vertical axis.

4. Comparison Principle 4.1. Axisymmetric solutions Before presenting the proof of Theorem 1.1, let us investigate the properties for the solutions of (MCF) that are axisymmetric with respect to the vertical axis; in other words, we consider solutions of the form u = u(r, z, t) where r = (x2 + y 2 )1/2 . Lemma 4.1 (Tests for axisymmetric solutions). Let u be a subsolution (respectively, supersolution) of (3.2). Suppose that there exist (ˆ p, tˆ) ⊂ H × (0, ∞) 2 and φ ∈ C (O) such that   p, tˆ) respectively, min(u − φ) = (u − φ)(ˆ p, tˆ) . max(u − φ) = (u − φ)(ˆ O

O

If pˆ = (ˆ p1 , pˆ2 , pˆ3 ) satisfies pˆ21 + pˆ22 = 0 and u is axisymmetric about the vertical axis, then there exists k ∈ R such that ∂ φ(ˆ p, tˆ) = pˆ1 k ∂p1

and

∂ φ(ˆ p, tˆ) = pˆ2 k. ∂p2

(4.1)

 ∂ φ( pˆ21 + pˆ22 , pˆ3 , tˆ) provided that φ = φ(r, p3 , t), Remark 4.1. It is clear that k = ∂r i.e. φ is also axisymmetric about the vertical axis.  Proof of Lemma 4.1. Denote rˆ = pˆ21 + pˆ22 . We only prove the situation when p1 , pˆ2 , pˆ3 , tˆ) for all u is a subsolution. By the symmetry of u, u(p1 , p2 , pˆ3 , tˆ) = u(ˆ 2 2 2 p1 + p2 = rˆ . By assumption, we have (u − φ)(p1 , p2 , pˆ3 , tˆ) ≤ (u − φ)(ˆ p1 , pˆ2 , pˆ3 , tˆ) for all (p1 , p2 , p3 , t) ∈ O, which implies that φ(p1 , p2 , pˆ3 , tˆ) ≥ φ(ˆ p1 , pˆ2 , pˆ3 , tˆ) for all (p1 , p2 ) close to (ˆ p1 , pˆ2 ) with p21 + p22 = r2 . Applying the method of Lagrange multipliers, we get k ∈ R such that   ∂ k φ(p1 , p2 , pˆ3 , tˆ) − (p21 + p22 − rˆ2 ) = 0, ∂p1 2   ∂ k 2 2 2 φ(p1 , p2 , pˆ3 , tˆ) − (p1 + p2 − rˆ ) = 0 ∂p2 2 at (ˆ p1 , pˆ2 ). We conclude (4.1) by straightforward calculations. 1350027-17

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

4.2. Proof of the comparison theorem We are now in a position to prove Theorem 1.1. Proof of Theorem 1.1. Let us assume u is axisymmetric about the vertical axis. The same argument applies to the case when v is axisymmetric. Suppose by contradiction that there exists (¯ p, t¯) ∈ H × (0, T ) such that (u − v)(¯ p, t¯) > 0. Then we have u(¯ p, t¯) − v(¯ p, t¯) −

σ > 0, T − t¯

when σ > 0 is sufficiently small. Hence, for such a σ, there exists (ˆ p, tˆ) ∈ H × (0, T ) satisfying   σ σ ˆ ˆ u(ˆ p, t ) − v(ˆ p, t ) − = max u(p, t) − v(p, t) − = µ > 0. (4.2) T −t T − tˆ H×[0,T ) We fix this σ, double the variables and set up an auxiliary function 1 1 σ , Φε (p, t, q, s) = u(p, t) − v(q, s) − g(p, q) − (t − s)2 − ε 2ε T −t where g(p, q) = |p · q −1 |4 . Let (pε , tε , q ε , sε ) ∈ (H × [0, T ))2 be a maximizer of Φε ; then it is clear that Φε (pε , tε , q ε , sε ) =

sup (H×[0,T ))2

Φε ≥ Φε (ˆ p, tˆ, pˆ, tˆ),

which implies that 1 1 g(pε , q ε ) + (tε − sε )2 ≤ u(pε , tε ) − v(q ε , sε ) − u(ˆ p, tˆ) ε 2ε σ σ + v(ˆ p, tˆ) + . − ˆ T − tε T −t

(4.3)

By the boundedness of u from above and of v from below, we have |pε · (q ε )−1 | → 0 and |tε − sε | → 0 as ε → 0. Since u = a and v = b with a ≤ b outside K × [0, ∞), we may take a subsequence of ε, still indexed by ε, such that pε , q ε → p ∈ H and tε , sε → t ∈ [0, T ) as ε → 0. Sending the limit in (4.3) and applying (4.2), we get   1 1 g(pε , q ε ) + (tε − sε )2 ≤ 0. lim sup ε 2ε ε→0 In other words, we have 1 g(pε , q ε ) → 0 and ε

1 ε (t − sε )2 → 0 as ε → 0. 2ε 1350027-18

(4.4)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

We next claim that t = 0. Indeed, if t = 0, then, since u(p, 0) ≤ v(p, 0) for all p ∈ H, we are led to σ Φε (pε , tε , q ε , sε ) → u(p, 0) − v(p, 0) − < 0, T which contradicts the fact that Φε (pε , tε , q ε , sε ) ≥ µ. We next apply the Crandall– Ishii lemma and get   σ 1 ε 2,+ ε 1 p ε ε ε + (t − s ), ∇ g(p , q ), X ∈ J H u(pε , tε ), (T − tε )2 ε ε   1 ε 1 2,− (t − sε ), − ∇q g(pε , q ε ), Y ε ∈ J H v(q ε , sε ), ε ε 2,+

2,−

where J H and J H X ε , Y ε ∈ S2 satisfy

denote the closure of the semijets in Heisenberg group and

X ε ξ, ξ − Y ε ξ, ξ ≤

C ε C |p · (q ε )−1 |4 |ξ|2 = g(pε , q ε )|ξ|2 ε ε

for some C > 0 and all ξ ∈ R2 ; see Theorem B.2. See Appendix B and [5, 24] for more details on the semijets and the Crandall–Ishii lemma on the Heisenberg group. It follows from (4.4) that lim sup( X ε ξ, ξ − Y ε ξ, ξ) ≤ 0 ε→0

(4.5)

uniformly for all bounded ξ ∈ R2 . Moreover, as is derived from Remark 2.1, the following gradient relation holds: 1 p 1 q ∇ g(pε , q ε ) = − ∇H g(pε , q ε ). ε H ε p Let η ε denote 1ε ∇H g(pε , q ε ).

Case A. If η ε = 0 for all ε > 0 small, then, by Definition 3.3, σ 1 + (tε − sε ) + F (η ε , X ε ) ≤ 0 ε 2 (T − t ) ε

(4.6)

1 ε (t − sε ) + F (η ε , Y ε ) ≥ 0. ε

(4.7)

and

Taking the difference of (4.6) and (4.7) yields   σ ηε ⊗ ηε ≤ tr I − (X ε − Y ε ). (T − tε )2 |η ε |2 Passing to the limit as ε → 0 with an application of (4.5), we end up with σ ≤ 0, (T − t)2 which is clearly a contradiction. 1350027-19

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi p Case B. If η εj = ε1j ∇H g(pεj , q εj ) = we obtain, by computation, that

q 1 εj εj εj ∇H g(p , q )

= 0 for a subsequence εj → 0,

ε

∂ pj ∂ g(pεj , q εj ) + 1 g(pεj , q εj ) = 0 ∂p2 2 ∂p3 (4.8)

ε

∂ q j ∂ g(pεj , q εj ) + 1 g(pεj , q εj ) = 0. ∂q2 2 ∂q3 (4.9)

pj ∂ ∂ g(pεj , q εj ) − 2 g(pεj , q εj ) = 0, ∂p1 2 ∂p3

ε

and q j ∂ ∂ g(pεj , q εj ) − 2 g(pεj , q εj ) = 0, ∂q1 2 ∂q3 ε

ε

ε

We first claim that p1j = p2j = 0 on this occasion. Suppose by contradiction ε ε that (p1j )2 + (p2j )2 = 0. In terms of Lemma 4.1, there is k ∈ R such that (4.8) reduces to ε

p1j k −

ε

ε

p2j ∂ pj ∂ ε g(pεj , q εj ) = 0 and p2j k + 1 g(pεj , q εj ) = 0, 2 ∂p3 2 ∂p3

which yields that k = 0 and ∂ 1 ε ε 1 ε ε ε ε g(pεj , q εj ) = p3j − q3j − p1j q2j + p2j q1j = 0. ∂p3 2 2 It follows from (4.8), (2.11) and (2.12) that pεj = q εj , which contradicts the assumption that g(pεj , q εj ) = 0. This completes the proof of our claim. ε ε As p1j = p2j = 0, we apply (4.8), (2.11) and (2.12) again and get ε

ε

ε

ε

ε

ε

ε

ε

ε

ε

ε

ε

4((q1j )2 + (q2j )2 )(−q1j ) − 16q2j (p3j − q3j ) = 0, 4((q1j )2 + (q2j )2 )(−q2j ) + 16q1j (p3j − q3j ) = 0. ε

ε

We are then led to q1j = q2j = 0. Now simplifying the second derivatives of g in 2,p ∗ εj g) (p , q εj ) = 0. An analog of calculation by using (2.15)–(2.19), we obtain (∇H 2,q ∗ εj εj yields that (∇H g) (p , q ) = 0. The proof is complete since Definition 3.2 can be adopted to get 1 σ + (tεj − sεj ) ≤ 0 ε 2 j (T − t ) εj

(4.10)

1 εj (t − sεj ) ≥ 0 εj

(4.11)

and

and deduce a contradiction. 5. Existence Theorem by Games The game setting is as follows. A marker, representing the game state, is initialized at a state p ∈ H from time 0. The maturity time given is denoted by t. Let the 1350027-20

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

step size for space be ε > 0. Time ε2 is consumed for every step. Then the total number of game steps N is given by the integer floor t/ε2 . The game states for all steps are denoted in order by ζ 0 , ζ 1 , . . . , ζ N with ζ 0 = p. Two players, Player I and Player II participate the game. Player I intends to minimize at the final state an objective function, which in our case is u0 : H → R, while Player II is to maximize it. At the (k + 1)th round (k < N ), (1) Player I chooses in H a unit horizontal vector v k , i.e. v k = (v1k , v2k , 0) satisfying |v k |2 = (v1k )2 +(v2k )2 = 1. We denote by Sh1 the set of all unit horizontal vectors. (2) Player II has the right to reverse Player I’s choice, which bk = ±1. √ determines k k k k (3) The marker is moved from the present state ζ to ζ · ( 2εb v ). Then the state equation is written inductively as  √ ζ k+1 = ζ k · ( 2εbk v k ), k = 0, 1, . . . , N − 1, ζ 0 = p.

(5.1)

For any k ≥ 0, let Hk := {ζ0 , v1 , b1 , ζ1 , v2 , b2 , . . . , vk , bk , ζk } ⊂ H × (Sh1 × {±1} × H)k denote the history of the game after the kth round. For any h ∈ Hk , let us call a mapping αk : Hk → Sh1 with αk (h) = vk+1 a decision of Player I at (k + 1)th round. On the other hand, we call a mapping βk : Hk ×Sh1 → {±1} with βk (h, vk+1 ) = bk+1 a decision of Player II at (k + 1)th round. The collection {αk } (respectively, {βk }) is called a strategy of Player I (respectively, of Player II). The value function is defined to be max · · · min max u0 (ζ N ). uε (p, t) := min 1 1 v

b

vN

bN

(5.2)

By the dynamic programming:

√ uε (p, t) = min1 max uε (p · ( 2εbv), t − ε2 ) v∈Sh b=±1

(5.3)

with uε (p, 0) = u0 (p). Our main result of this section is given below. Theorem 5.1 (Existence theorem by games). Assume that u0 is uniformly continuous function in H and is constant C ∈ R outside a compact set. Assume also that u0 is spatially axisymmetric about the vertical axis. Let uε be the value function defined as in (5.2). Then uε converges, as ε → 0, to the unique axisymmetric viscosity solution of (MCF) uniformly on compact subsets of H × [0, ∞). Moreover, u = C in (H\K) × (0, ∞) for some compact set K ⊂ H. Before presenting the proof of Theorem 5.1, we first give bounds for the game trajectories under some particular strategies. Lemma 5.2 (Lower bound of the game trajectories). For any p ∈ H and t ≥ 0 with N = [t/ε2 ], let ζk be defined as in (5.1) for all k = 0, 1, . . . , N . Then the 1350027-21

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

following statements hold. (i) There exists a strategy of Player I such that (|ζ1N |2 + |ζ2N |2 )2 + 16|ζ3N |2 ≥ (|p1 |2 + |p2 |2 )2 + 16|p3 |2

(5.4)

under this strategy regardless of Player II ’s choices. (ii) There exists a strategy of Player II such that (5.4) holds under this strategy regardless of Player I ’s choices. Proof. (i) By direct calculation, we have  2 √ √ 1√ 2 2 2 2εb(p1 v2 − p2 v1 ) ((p1 + 2εbv1 ) + (p2 + 2εbv2 ) ) + 16 p3 + 2 = (p21 + p22 + 2ε2 )2 + 8ε2 (p21 + p22 ) + 16p23 √ + 4 2εb((p21 + p22 + 2ε2 )(p1 v1 + p2 v2 ) + 4(p1 p3 v2 − p2 p3 v1 )).

(5.5)

It is clear that Player I may take v = (v1 , v2 , 0) ∈ Sh1 satisfying v1 =

1 2 ((p + p22 + 2ε2 )p2 + 4p1 p3 ), ρ 1

1 v2 = − ((p21 + p22 + 2ε2 )p1 + 4p2 p3 ) ρ with ρ = (p21 + p22 )1/2 ((p21 + p22 + 2ε2 )2 + 16p23 )1/2 so that, no matter which b is picked, we have b((p21 + p22 + 2ε2 )(p1 v1 + p2 v2 ) + 4(p1 p3 v2 − p2 p3 v1 )) = 0 and, furthermore by (5.5),  2 √ √ 1√ 2 2 2 2εb(p1 v2 − p2 v1 ) ((p1 + 2εbv1 ) + (p2 + 2εbv2 ) ) + 16 p3 + 2 = (p21 + p22 + 2ε2 )2 + 8ε2 (p21 + p22 ) + 16p23 ≥ (p21 + p22 )2 + 16p23 .

(5.6)

We can iterate (5.6) to get (|ζ1k |2 + |ζ2k |2 )2 + 16|ζ3k |2 ≥ (|ζ1k−1 |2 + |ζ2k−1 |2 )2 + 16|ζ3k−1 |2 for all k = 1, 2, . . . , N and (5.4) follows easily. (ii) The proof of (ii) is similar and even easier. Note that Player II may take a proper b = ±1 so that b((p21 +22 +2ε2 )(p1 v1 + p2 v2 ) + 4(p1 p3 v2 − p2 p3 v1 )) ≥ 0 and therefore (5.6) holds immediately. We then complete the proof by iteration again. 1350027-22

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Lemma 5.3 (Upper bound of the game trajectories). For any p ∈ H and t ≥ 0 with N = [t/ε2 ], let ζ k be defined as in (5.1) for all k = 0, 1, . . . , N . Then the following statements hold. (i) There exists a strategy of Player I such that (|ζ1N |2 + |ζ2N |2 )2 + 16|ζ3N |2 ≤ (|p1 |2 + |p2 |2 + 6N ε2 )2 + 16|p3 |2

(5.7)

under this strategy regardless of Player II ’s choices. (ii) There exists a strategy of Player II such that (5.7) holds under this strategy regardless of Player I ’s choices. Remark 5.1. With the notation of the gauge G in (3.13), the inequality (5.7) can be simplified into G(ζ N ) ≤ (|p1 |2 + |p2 |2 + 6t)2 + 16|p3 |2 , which is intuitively natural, since the explicit solution given in (3.14) satisfies w(p, t) ≤ (|p1 |2 + |p2 |2 + 6t)2 + 16|p3 |2 . Proof of Lemma 5.3. By iteration, it suffices to show that there exist strategies of Player I or Player II such that √ √ ((p1 + 2εbv1 )2 + (p2 + 2εbv2 )2 + jε2 )2  2 1√ + 16 p3 + 2εb(p1 v2 − p2 v1 ) ≤ (p21 + p22 + (j + 6)ε2 )2 + 16p23 . (5.8) 2 Indeed, the left-hand side is calculated to be (p21 + p22 + (j + 2)ε2 )2 + 8ε2 (p21 + p22 ) + 16p23 √ + 4 2εb((p21 + p22 + 2ε2 )(p1 v1 + p2 v2 ) + 4(p1 p3 v2 − p2 p3 v1 )). As in the proof of Lemma 5.3, either Player I or Player II may let b((p21 + p22 + 2ε2 )(p1 v1 + p2 v2 ) + 4(p1 p3 v2 − p2 p3 v1 )) ≤ 0 with no regard for their opponents strategies. Hence, by a strategy of either Player I or Player II, we have  2 √ √ 1√ 2 2 2 2 2εb(p1 v2 − p2 v1 ) ((p1 + 2εbv1 ) + (p2 + 2εbv2 ) + jε ) + 16 p3 + 2 ≤ (p21 + p22 + (j + 2)ε2 )2 + 8ε2 (p21 + p22 ) + 16p23 ≤ (p21 + p22 + (j + 6)ε2 )2 + 16p23 , which proves (5.8). 1350027-23

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

Remark 5.2. For any pˆ ∈ H, c ∈ R and L > 0, let ˆ G(p) = c + LG(ˆ p−1 · p).

(5.9)

Our proof above can be directly generalized to show that ˆ N ) ≤ c + L(|p1 − pˆ1 |2 + |p2 − pˆ2 |2 + 6N ε2 )2 G(ζ  2   1 + 16L p3 − pˆ3 + (p1 pˆ2 − p2 pˆ1 ) 2 with either a strategy of Player I or a strategy of Player II. We now return to the proof of Theorem 5.1, which actually rests on showing that u and u, as defined in (1.5) and (1.6), are respectively a subsolution and a supersolution of (MCF). (Note that our definitions are valid since the game value uε are bounded uniformly for all ε > 0 by its definition.) Moreover, we show that u(p, 0) ≤ u(p, 0) and u and u are constant outside a compact set. Then it follows immediately from the comparison principle (Theorem 1.1) that u ≤ u and therefore uε → u locally uniformly as ε → 0. Proposition 5.1 (Constant value outside a compact set). Assume that u0 is uniformly continuous function in H and is a constant C ∈ R outside a compact set. Let uε be the value function defined by (5.2). Then for any T > 0, u(p, t) = u(p, t) = C for all p ∈ H outside a compact set and for all t ∈ [0, T ]. Proof. Suppose there exists Br such that u0 (p) = C for any p ∈ H\Br . Then for any pˆ ∈ H\Br and t ≥ 0, we use the strategy of Player I introduced in Lemma 5.2, we get ζ N ∈ H\Br regardless of Player II’s choices, which implies that uε (p, t) ≤ u0 (ζ N ) = C. Similarly, we may use the strategy of Player II to deduce that uε (p, t) ≥ C. Hence, uε = C and u = u = C in H\Br . Proposition 5.2. Assume that u0 is uniformly continuous function in H and is constant outside a compact set. Let uε be the value function defined by (5.2). Then u(p, 0) ≤ u0 (p) and u(p, 0) ≥ u0 (p) for all p ∈ H. In order to prove this result, we first need to regularize the initial data with the smooth gauge G in (3.13). We define ψ L (p) = sup {u0 (q) − LG(p−1 · q)}

(5.10)

ψL (p) = inf {u0 (q) + LG(p−1 · q)}

(5.11)

q∈H

and q∈H

for any p ∈ H and fixed L > 0. These two functions are called the sup-convolution and inf-convolution of u0 respectively. Our definitions here are slightly different from 1350027-24

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

those in [33] in that we plug p−1 · q instead of q · p−1 in G. However, the properties remain the same. One of the important properties is presented in Lemma A.1. Proof of Proposition 5.2. We arbitrarily fix pˆ ∈ H. By Lemma A.1 in Appendix A, for any δ > 0, there exists L > 0 such that p) ≤ u0 (ˆ p) + δ, ψ L (ˆ which implies that u0 (p) ≤ u0 (ˆ p) + δ + LG(ˆ p−1 · p). ˆ in (5.9) with c = u0 (ˆ Let us use the right-hand side, which is exactly G p) + δ, as ε the objective function of the games. Suppose the game value is w . Then by using the special strategy of Player I given in Lemma 5.3 and Remark 5.2, we obtain a game estimate p) + δ + LG(ˆ p−1 · ζ N ) wε (p, t) ≤ u0 (ˆ p) + δ + L(|p1 − pˆ1 |2 + |p2 − pˆ2 |2 + 6N ε2 )2 ≤ u0 (ˆ  2   1  + 16L p3 − pˆ3 + (p1 pˆ2 − p2 pˆ1 ) 2 no matter what choices are made by Player II during the game. On the other hand, since it is clear that uε ≤ wε and N ε2 ≤ t, we get p) + δ + L(|p1 − pˆ1 |2 + |p2 − pˆ2 |2 + 6t)2 uε (p, t) ≤ u0 (ˆ  2   1  + 16L p3 − pˆ3 + (p1 pˆ2 − p2 pˆ1 ) . 2 Taking the relaxed limit of uε at (ˆ p, 0) as ε → 0, we have u(ˆ p, 0) ≤ u0 (ˆ p) + δ. We finally send δ → 0 and get u(ˆ p, 0) ≤ u0 (ˆ p) for any pˆ ∈ H. The proof for the statement that u(p, 0) ≥ u0 (p) for all p ∈ H is symmetric. In fact, the key is to use the strategy of Player II introduced in Lemma 5.3 and Remark 5.2 to deduce p) − δ − L(|p1 − pˆ1 |2 + |p2 − pˆ2 |2 + 6t)2 uε (p, t) ≥ u0 (ˆ  2   1 − 16L p3 − pˆ3 + (p1 pˆ2 − p2 pˆ1 ) . 2 Proposition 5.3 (Axial symmetry of the game values). Suppose that u0 is uniformly continuous on H and is spatially axisymmetric with respect to the vertical axis. Let uε be the value function defined as in (5.2). Then uε , u and u are also spatially axisymmetric about the vertical axis. 1350027-25

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

Proof. We argue by induction. Assume that uε (p, t) = uε (p , t) for some t ≥ 0 and for any p, p ∈ H such that p21 + p22 = (p1 )2 + (p2 )2

and p3 = p3 .

(5.12)

We aim to show uε (p, t + ε2 ) = uε (p , t + ε2) for all p, p ∈ H satisfying the condition (5.12). Since the dynamic programming principle (5.3) gives √ uε (p, t + ε2 ) = min1 max uε (p · ( 2εbv), t), v∈Sh b=±1

there exists v ∈ Sh1 such that

√ uε (p, t + ε2 ) = max uε (p · ( 2εbv), t). b=±1

(5.13)

√ We that there is v  ∈ Sh1 such that the coordinates of p · ( 2εbv) and p · √ claim ( 2εbv  ) satisfy (5.12) as well. Indeed, as   √ √ √ 1√ 2εb(p1 v2 − p2 v1 ) p · ( 2εbv) = p1 + 2εbv1 , p2 + 2εbv2 , p3 + 2 and √ p · ( 2εbv  ) = 

  √ √ 1√          p1 + 2εbv1 , p2 + 2εbv2 , p3 + 2εb(p1 v2 − p2 v1 ) , 2

we are looking for v1 , v2 ∈ Sh1 such that  √ √ √ √   2   2 2 2  (p1 + 2εbv1 ) + (p2 + 2εbv2 ) = (p1 + 2εbv1 ) + (p2 + 2εbv2 ) 1√ 1√  p3 + 2εb(p1 v2 − p2 v1 ) = p3 + 2εb(p1 v2 − p2 v1 ). 2 2 Since p and p satisfy (5.12), it suffices to solve the linear system    p1 v1 + p2 v2 = p1 v1 + p2 v2 , −p2 v1 + p1 v2 = −p2 v1 + p1 v2 . The problem is trivial if p21 + p22 = (p1 )2 + (p2 )2 = 0. When p21 + p22 = (p1 )2 + (p2 )2 = 0, we get a unique pair of solutions v1 =

1 ((p1 p1 + p2 p2 )v1 + (p1 p2 − p1 p2 )v2 ), (p1 )2 + (p2 )2

v2 =

1 ((p1 p2 − p1 p2 )v1 + (p1 p1 + p2 p2 )v2 ). (p1 )2 + (p2 )2

Thanks to the relation (5.12), it is easy to verify that v  = (v1 , v2 , 0) ∈ Sh1 , i.e. (v1 )2 + (v2 )2 = 1. We complete the proof of the claim. 1350027-26

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

In view of the induction hypothesis, we obtain √ √ uε (p · ( 2εbv  ), t) = uε (p · ( 2εbv), t) for both b = ±1, which, together with the dynamic programming (5.3) and (5.13), yields √ uε (p , t + ε2 ) ≤ max uε (p · ( 2εbv  ), t) ≤ uε (p, t + ε2 ). b=±1

We may similarly prove that uε (p , t + ε2 ) ≥ uε (p, t + ε2 ) and therefore uε (p , t + ε2 ) = uε (p, t + ε2 ) for all p, p ∈ H satisfying (5.12). It follows from the definitions (1.5)–(1.6) of half relaxed limits that the same results for u and u hold. Proposition 5.4. Assume that uε satisfies the dynamic programming principle (5.3). Let u be the upper relaxed limit defined as in (1.5). Then u is a subsolution of (1.2). Proof. Assume that there exists (ˆ p, tˆ) ∈ H × (0, ∞) and φ ∈ C 2 (H × (0, ∞)) such p, tˆ). Then by definitions of u, we may take that u − φ attains a strict maximum at (ˆ ε ε p, tˆ) and a sequence, still indexed by ε, (p , t ) ∈ H × (0, ∞) such that (pε , tε ) → (ˆ p, tˆ) as ε → 0 and uε (pε , tε ) → u(ˆ uε (pε , tε ) − φ(pε , tε ) = max (uε − φ). Br (p, ˆ tˆ)

(5.14)

Applying the dynamic programming principle (5.3) with (p, t) = (pε , tε ), we have √ uε (pε , tε ) = min max uε (pε · ( 2εbv), tε − ε2 ), v

b

which, combined with (5.14), implies that √ φ(pε , tε ) ≤ min max φ(pε · ( 2εbv), tε − ε2 ). v

b

We next use the Taylor expansion for the right-hand side at (pε , tε ) and obtain √ 2 ε2 φt (pε , tε ) − min max( 2εbv, ∇φ(pε , tε ) + ε2 (∇H φ)∗ (pε , tε )vh , vh ) ≤ o(ε2 ), v

b

(5.15) where vh is the horizontal projection of v, i.e. vh = (v 1 , v 2 ) for any v = (v 1 , v 2 , v 3 ). Since v = (v 1 , v 2 , 0), we may rewrite (5.15) as √ 2 φ)∗ (pε , tε )vh , vh ) ε2 φt (pε , tε ) − min max( 2εbvh , ∇H φ(pε , tε ) + ε2 (∇H v

b

≤ o(ε2 ).

(5.16)

We discuss the following two cases. 1350027-27

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

Case A. ∇H φ(pε , tε ) = 0 for all sufficiently small ε > 0. Letting v˜ =

1 (X2 φ(pε , tε ), −X1 φ(pε , tε ), 0) |∇H φ(pε , tε )|

with v˜h =

1 (X2 φ(pε , tε ), −X1 φ(pε , tε )) |∇H φ(pε , tε )|

we have from (5.15) 2 φ)∗ (pε , tε )˜ vh , v˜h ) ≤ o(1). φt (pε , tε ) − (∇H

(5.17)

Noticing that v˜h ⊗ v˜h = I −

∇H φ(pε , tε ) ⊗ ∇H φ(pε , tε ) |∇H φ(pε , tε )|2

we are thus led from (5.17) to    ∇H φ(pε , tε ) ⊗ ∇H φ(pε , tε ) ε ε 2 ∗ ε ε φt (p , t ) − tr I − (∇H φ) (p , t ) ≤ o(1). |∇H φ(pε , tε )|2

(5.18)

Sending ε → 0, we get    p, tˆ) ⊗ ∇H φ(ˆ p, tˆ) ∇H φ(ˆ 2 ∗ ˆ φt (ˆ p, tˆ) − tr I − φ) (ˆ p , t ) ≤ 0. (∇ H |∇H φ(ˆ p, tˆ)|2 Case B. There exists a subsequence εj such that ∇H φ(pεj , tεj ) = 0 for all j, then it follows from (5.15) that 2 φ)∗ (pεj , tεj )vh , vh ) ≤ o(1) φt (pεj , tεj ) − (∇H

for some v,

which implies p, tˆ) ≤ 0, φt (ˆ

(5.19)

as the limit when εj → 0. Proposition 5.5. Assume that uε satisfies the dynamic programming principle (5.3). Let u be the lower relaxed limit defined as in (1.6). Then u is a supersolution of (1.2). In order to facilitate the proof, let us present an elementary result. Lemma 5.4 ([18, Lemma 4.1]). Suppose ξ is a unit vector in R2 and X is a real symmetric 2 × 2 matrix. Then there exists a constant M > 0 that depends only on the norm of X, such that for any unit vector v ∈ R2 , | Xξ ⊥ , ξ ⊥  − Xv, v| ≤ M | ξ, v|, where ξ ⊥ denotes a unit orthonormal vector of ξ. 1350027-28

(5.20)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Proof. Let cos θ = ξ ⊥ , v and sin θ = ξ, v. Then we have | Xξ ⊥ , ξ ⊥  − Xv, v| = |tr(X(ξ ⊥ ⊗ ξ ⊥ − v ⊗ v))| ≤ Xξ ⊥ ⊗ ξ ⊥ − (ξ sin θ + ξ ⊥ cos θ) ⊗ (ξ sin θ + ξ ⊥ cos θ) = X sin2 θξ ⊥ ⊗ ξ ⊥ − sin θ cos θ(ξ ⊗ ξ ⊥ + ξ ⊥ ⊗ ξ) − sin2 θξ ⊗ ξ ≤ M | sin θ|, where M > 0 depends on X. We refer the reader to [23, Lemma 2.3] for a higher-dimensional extension of this lemma. Proof of Proposition 5.5. Assume that there exists (ˆ p, tˆ) ∈ H × (0, ∞) and φ ∈ p, tˆ). We may again take C 2 (H×(0, ∞)) such that u−φ attains a strict minimum at (ˆ p, tˆ) and uε (pε , tε ) → u(ˆ p, tˆ) a sequence (pε , tε ) ∈ H × (0, ∞) such that (pε , tε ) → (ˆ as ε → 0 and uε (pε , tε ) − φ(pε , tε ) = min (uε − φ). Br (p, ˆ tˆ)

(5.21)

Applying the dynamic programming principle (5.3) with (p, t) = (pε , tε ), we have √ uε (pε , tε ) = min max uε (pε · 2εbv, tε − ε2 ). v

b

It then follows from (5.21) that φ(pε , tε ) ≥ min max φ(pε · v

b

√ 2εbv, tε − ε2 ).

As an analogue of (5.16), the Taylor expansion at (pε , tε ) yields   1 √ ε ε ε ε 2 ∗ ε ε

2bvh , ∇H φ(p , t ) + (∇H φ) (p , t )vh , vh  ≥ o(1), φt (p , t ) − min max v b ε (5.22) as ε → 0. We again divide our discussion into two situations. Case A. ∇H φ(pε , tε ) = 0 for all sufficiently small ε > 0. We adopt Lemma 5.4 and get   1 √ ε ε 2 ∗ ε ε

2bvh , ∇H φ(p , t ) + (∇H φ) (p , t )vh , vh  max b ε  √  2 2 ∗ ε ε ≥ (∇H φ) (p , t )˜ | vh , ∇H φ(pε , tε )| vh , v˜h  + −C + ε 1350027-29

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

for some C ≥ 0 depending only on the sup-norm of (∇2H φ)∗ on Br (ˆ p, tˆ), where v˜h =

1 (X2 φ(pε , tε ), −X1 φ(pε , tε )), |∇H φ(pε , tε )|

as given in the proof of Proposition 5.4. It is now clear that   1 √ 2 min max

2bvh , ∇H φ(pε , tε ) + (∇H φ)∗ (pε , tε )vh , vh  v b ε 2 ≥ (∇H φ)∗ (pε , tε )˜ vh , v˜h ,

when ε > 0 is sufficiently small. This implies through (5.22) that 2 φt (pε , tε ) − (∇H φ)∗ (pε , tε )˜ vh , v˜h  ≥ o(1).

Letting ε → 0, we obtain    p, tˆ) ⊗ ∇H φ(ˆ p, tˆ) ∇H φ(ˆ 2 ∗ ˆ ˆ φt (ˆ p, t ) − tr I − p, t ) ≥ 0. (∇H φ) (ˆ |∇H φ(ˆ p, tˆ)|2 Case B. There is a subsequence εj such that ∇H φ(pεj , tεj ) = 0. By (5.22), we have on this occasion 2 φt (pεj , tεj ) − (∇H φ)∗ (pεj , tεj )vh , vh ) ≥ o(1)

for some v.

Sending ε → 0, we get φt (ˆ p, tˆ) ≥ 0.

(5.23)

We are now in a position to prove Theorem 5.1. Proof of Theorem 5.1. In terms of Propositions 5.6–5.8, u and u are respectively a subsolution and a supersolution of (1.2) that are axisymmetric with respect to the vertical axis. For any T > 0, u(p, t) and u(p, t) are constant outside a compact set of H for all t ∈ [0, T ], owing to Proposition 5.1. Also, since u(p, 0) ≤ u0 (p) and u(p, 0) ≥ u0 (p) for all p ∈ H, we may apply Theorem 1.1 to get u ≤ u in H × [0, T ]. As it is obvious that u ≥ u, we get u = u in H × [0, T ] with u(·, 0) = u0 (·). In conclusion, u = u = u is the unique continuous solution of (MCF) and the locally uniform convergence uε → u follows immediately. 6. Stability The following stability result is standard in the theory of viscosity solutions. Theorem 6.1 (Stability under the uniform convergence). Let uε be solutions of (1.2) and uε → u locally uniformly in H × [0, ∞). Then u is also a solution of (1.2). 1350027-30

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Lemma 6.2. If uε is a subsolution (respectively, supersolution) of (1.2) for all small ε > 0, then   u = lim sup∗ uε respectively, u = lim inf∗ uε ε→∞

ε→∞

is also a subsolution (respectively, supersolution) of (1.2). Proof. We adopt Definition 3.3 to prove the lemma. We only show the subsolution case, since the supersolution part can be proved by using a symmetric argument. It is easily seen that u is upper semicontinuous. Suppose there exist φ ∈ C 2 (H × p, tˆ). [0, ∞)) and (ˆ p, tˆ) ∈ H × (0, ∞) such that u − φ attains a strict maximum at (ˆ Then by the convergence of maximizers as shown in [17, Lemma 2.2.5], we can p, tˆ) as take subsequences of pε , tε and uε , still indexed by ε, satisfying (pε , tε ) → (ˆ ε → 0 and (uε − φ)(pε , tε ) = max (uε − φ). H×[0,∞)

ε

Applying Definition 3.3 to u , we get 2 φ)∗ (pε , tε )) ≤ 0. φt (pε , tε ) + F (∇H φ(pε , tε ), (∇H

Letting ε → 0 and using the lower semicontinuity of F , we end up with 2 φt (ˆ p, tˆ) + F (∇H φ(ˆ p, tˆ), (∇H φ)∗ (ˆ p, tˆ)) ≤ 0.

Proof of Theorem 6.1. Let u = lim sup∗ uε ε→0

and u = lim inf∗ uε . ε→0

Then in virtue of Lemma 6.2, u is a subsolution of (1.2) and u is a supersolution of (1.2). Noting that uε → u locally uniformly, we must have u = u = u and therefore u is a solution of (1.2). 7. Properties of the Evolution We have shown that there is a unique solution u of (MCF) for any given continuous function u0 which is axisymmetric with respect to the vertical axis and attains constant value outside a compact set. Let us turn to discuss the surface evolution described by the level-set equation (MCF). More precisely, given an axisymmetric compact surface Γ0 ⊂ H, we choose u0 ∈ C(H) such that it is axisymmetric constant outside a compact set and satisfies Γ0 = {p : H : u0 (p) = 0}.

(7.1)

We then solve (MCF) for the unique solution u and get the surface Γt = {p ∈ H : u(p, t) = 0} for any t ≥ 0. 1350027-31

(7.2)

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

We may show, by following [13, 8], that the surface represented by the levelset Γt of u does not depend on the particular choice of u0 (see Theorem A.2 and Corollary A.3 in Appendix A). We give a simple geometric property of the mean curvature flow. The following result shows that an axisymmetric compact surface evolving by its mean curvature shrinks and disappears in finite time. Theorem 7.1 (Finite time extinction for bounded evolution). Suppose that of the mean curvature flow. If {Γt }t≥0 denotes an axisymmetric surface evolution √ Γ0 ⊂ Br , for r > 0, then Γt = ∅ when t > r2 / 12. Proof. We may take an axisymmetric u0 ∈ C(H) with a constant value C > 0 outside Br satisfying (7.1) and u0 ≥ min{|p|4 − r4 , C}. Taking w(p, t) = (p21 + p22 )2 + 12t(p21 + p22 ) + 16p23 + 12t2 as in (3.14), we easily see that wC (p, t) := min{w(p, t) − r4 , C} is a solution of (1.2) with initial data wC (p, 0) = min{|p|4 − r4 , C}, by Theorem A.2 with θ(x) = 4 by Theorem 1.1. min{x, C}. We are therefore led to u ≥ w − r√ C 2 that It is clear that w√ (p, t) > 0 when t > r / 12 for all p ∈ H, which implies √ u > 0 when t > r2 / 12. Hence Γt defined in (7.2) is empty when t > r2 / 12. Note that the conclusion does not depend on the particular choice of u0 , as explained in Corollary A.3. Remark 7.1. Theorem 7.1 indicates that a bounded axisymmetric mean curvature flow encounters singularities at a certain time T > 0. Remark 7.2. The following result stronger than Theorem 7.1 holds: For any continuous solution u of (MCF) with zero level √ set Γt for any t ≥ 0, if Γ0 ⊂ Br with some r > 0, then Γt = ∅ when t > r2 / 12. Here we do not need to assume the axial symmetry of Γ0 but we must specify the solution u since it is not known in general whether or not Γt depends on the choice of u0 . Definition 7.1. We say T ≥ 0 is the extinction time of the mean curvature flow Γt in the Heisenberg group, if Γt = ∅ when t ≤ T and Γt = ∅ when t > T . We next proceed to investigate the asymptotic profile after normalization for a sphere in the Heisenberg group. It is well known that in the Euclidean space any normalized compact convex surface converges to a sphere as t tends to the extinction time [19]. However, the normalized curvature flow from a sphere of radius r in the 1350027-32

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Heisenberg group looks like an ellipsoid ET := {P ∈ H : 12T (P12 + P22 ) + 16P32 = 1} √ at the extinction time T = r2 / 12.

(7.3)

Proposition 7.2. Suppose that Γt ⊂ H (t ≥ 0) is the horizontal mean curvature flow as defined in (7.2) with Γ0 = {p ∈ H : |p| = r}, where r > 0 is a given radius. √ √ Then the extinction time T = r2 / 12 and the normalized flow Γt / r4 − 12t2 → ET as t → T, where ET is given in (7.3). Proof. We take wC (p, t) = min{(p21 + p22 )2 + 12t(p21 + p22 ) + 16p23 + 12t2 − r4 , C} with C > 0. It is easily seen that wC (p, 0) = 0 if and only if p ∈ Γ0 . We have also Γt = {p ∈ shown that wC is a solution of (1.2). We track the evolution by setting √ H : wC (p, t) =√0} for all t ≥ 0. It is clear that Γt = ∅ when t > r2 / 12 and Γt = ∅ when t ≤ r2 / 12. For any p(t) = (p1 (t), p2 (t), p3 (t)) ∈ Γt , we have 12t(p21 (t) + p22 (t)) + 16p23 (t) ≤ r4 − 12t2 . (7.4) √ 4 2 We normalize the flow by letting P (t) = p(t)/ r − 12t for any p(t) ∈ Γt . Then (7.4) is written as 12t(P12 (t) + P22 (t)) + 16P32 (t) ≤ 1.

(7.5)

By setting U (P, t) = (P12 + P22 )2 (r4 − 12t2 ) + 12t(P12 + P22 ) + 16P32 − 1 we get

 1 C w ( r4 − 12t2 P (t), t) = U (P (t), t). r4 − 12t2 Sending the limit as t → T with (7.5) taken into account, we obtain 0=

12T (P12(T ) + P22 (T )) + 16P32 (T ) = 1 for the limit P (T ) of any subsequence of P (t) √ as t → ∞. The consequence above amounts to saying that the limit of the set Γt / r4 − 12t2 is contained in ET . On the other hand, for any P = (P1 , P2 , P3 ) ∈ ET , we have  wC ( r4 − 12t2 λP, t) = (r4 − 12t2 )W (λ, P, t), where λ > 0 and W (λ, P, t) = (λ4 (P12 + P22 )2 (r4 − 12t2 ) + λ2 − 1 + 12λ2 (t − T )(P12 + P22 ). √ C 4 2 One may take √ λ(t) > 0 such that w ( r − 12t λ(t)P, t) = 0; in other words, 4 2 λ(t)P ∈ Γt / r − 12t . Moreover, λ(t) → 1 as t √→ T , which implies that P 4 2 belongs to the limit of a sequence √ of elements in Γt / r − 12t . 4 2 In conclusion, we obtain Γt / r − 12t → ET as t → T . 1350027-33

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

We stress that this result is very different from that in the Euclidean space. The normalized asymptotic shape of horizontal mean curvature flow in the Heisenberg group starting from a Kor´ anyi ball is a Euclidean ellipsoid. Moreover, the shape of the ellipsoid depends on the extinction time T and therefore the size of the initial surface. It would be interesting to show this result for a general compact and convex initial surface. Acknowledgments We thank the anonymous referee whose careful reading and thoughtful comments have enabled us to improve the manuscript. F.F. is supported by MURST and by the University of Bologna (Italy), by EC project CG-DICE, and by the ERC starting grant project 2011 EPSILON (Elliptic PDEs and Symmetry of Interfaces and Layers for Odd Nonlinearities). F.F. wishes to thank the Department of Mathematics at the University of Pittsburgh for their kind hospitality. Q.L. and J.M. are supported by NSF award DMS-1001179. Appendix A. Several Useful Results and Proofs We append several results and proofs used in this paper. Lemma A.1 (Approximation by semi-convolutions). Assume that u0 is uniformly continuous on H and is constant outside a compact set. Let ψ L and ψL be respectively defined as in (5.10) and (5.11). Then ψ L and ψL converge to u0 uniformly in H as L → ∞. Proof. We only show the statement for ψ L . The proof for the statement on ψL is symmetric. It is easily seen that ψ L ≥ u0

in H.

(A.1)

On the other hand, since u0 is uniformly continuous, for any p ∈ H, we may find qL ∈ H such that ψ L (p) = sup {u0 (p) − LG(p−1 · q)} = u0 (qL ) − LG(p−1 · qL ). q∈H

By (A.1), we have LG(p−1 · qL ) ≤ u0 (qL ) − u0 (p),

(A.2)

which, by the boundedness of u0 , implies that  1/4 2K0 , |p−1 · qL | ≤ L where K0 = supH |u0 |. By the uniform continuity of u0 , for any δ > 0, there exists ε > 0 such that |u0 (p) − u0 (q)| ≤ δ for any p, q ∈ H satisfying |p−1 · q| ≤ ε. Then 1350027-34

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

we may let L > 0 be sufficiently large such that (2K0 /L)1/4 ≤ ε and therefore u0 (qL ) − u0 (p) ≤ δ, which, combined with (A.1), yields |ψ L (p) − u0 (p)| ≤ δ

for all p ∈ H.

Theorem A.2 (Invariance). Assume that θ : R → R is continuous. If u is a solution of (1.2). Then w = θ ◦ u is also a solution of (1.2). Proof. We prove the theorem in several steps. Step 1. We first give the proof in the case that θ ∈ C 2 (R) and θ > 0. Suppose that p, tˆ) ∈ H × (0, ∞) such that θ ◦ u − φ attains there exist φ ∈ C 2 (H × [0, ∞)) and (ˆ a maximum at (ˆ p, tˆ). Then it is clear that u − h(φ) attains a maximum at (ˆ p, tˆ), −1 2  where h = θ ∈ C (R) with h > 0. Denote ψ = h(φ). Since u is a subsolution of (1.2), we have    ∇H ψ ⊗ ∇H ψ 2 ∗ (∇H ψ) ≤ 0 at (ˆ p, tˆ). ψt − tr I − |∇H ψ|2 Note that ψt = h φt , ∇H ψ = h ∇H φ and 2 2 (∇H ψ)∗ = h ∇H φ ⊗ ∇H φ + h (∇H φ)∗ .

It follows that

   ∇H φ ⊗ ∇H φ 2 ∗ φt − tr I − (∇H φ) ≤ 0 at (ˆ p, tˆ), |∇H φ|2

which shows that θ ◦ u is a subsolution of (1.2). An analogue of this argument yields that θ ◦ u is also a supersolution. We also claim that θ ◦ u remains being a solution when θ ∈ C 2 (R) and θ < 0. Indeed, when θ is a decreasing function, −θ is increasing. We obtain that −θ ◦ u is a solution of (1.2). Thanks to the fact that the mean curvature flow is orientation-free or (1.2) is homogeneous in all of the derivatives, we easily see that θ ◦ u is a solution as well. In particular, we note that −u is a solution when u is a solution. Step 2. We generalize the consequence obtained in Step 1 for a continuous nondecreasing or nonincreasing function. Indeed, for any continuous nondecreasing function θ, we may take θn ∈ C 2 (R) with θn > 0 for all n = 1, 2, . . . such that lim sup∗ θn ◦ u = θ ◦ u. n→∞

We refer the reader to [17, Lemma 4.2.3] for details about the construction of θn . Since θn ◦ u is a solution of (1.2) for all n, as shown in Step 1, θ ◦ u is a subsolution, due to Lemma 6.2. ˜ To show that θ ◦ u is a supersolution, we define θ(x) = θ(−x) for any x ∈ R ˜ and observe that θ(u) = θ(−u). Since θ˜ is nonincreasing and −u is a solution, 1350027-35

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

˜ we may apply a symmetric version of [17, Lemma 4.2.3] to get θ(u) = θ(−u) is a supersolution. When θ is a continuous nonincreasing function, −θ is nondecreasing. We apply again the homogeneity of (1.2) to obtain that θ◦u is a solution. Since the verification of definition of (1.2) is pointwise, one can further relax the monotonicity condition on θ to a local monotonicity condition. To conclude this step, we notice that max{min{u, C} − C} is a solution for any C > 0 provided that u is a solution. Step 3. We finally discuss the situation when θ is assumed to be continuous only. By Theorem 6.1, it suffices to discuss the bounded function max{min{u, C} − C} instead of u for arbitrarily large C > 0. We approximate θ uniformly by polynomials θm in [−C − 1, C + 1]. Since polynomials only have finitely many maximizers and minimizers, we may also assume each θm is constant near all of its local maximizers and minimizers. In fact, if, for instance, θm attains a local maximum at x0 ∈ R, we take min{θm (x), θ(x0 ) − εm }, where εm > 0 is sufficiently small (εm → 0 as m → ∞) such that θm is continuous. Now θm is locally nonincreasing or nondecreasing. We apply the result in Step 2 and find that θm ◦ u is a solution of (1.2). Since θm → θ uniformly, by the stability result given in Theorem 6.1, we see that θ ◦ u is a solution by sending m → ∞.

An immediate consequence of the theorem above is that our generalized surface evolution does not depend on the choice of the initial level-set function u0 . Corollary A.3 (Independence of the choice of the initial function). Sup˜0 are continuous functions in H axisymmetric about the vertical pose that u0 and u axis and are constant outside a compact set K ⊂ H. Let Γ0 = {p ∈ H : u0 (p) = 0} = ˜ be the unique continuous solutions of {p ∈ H : u ˜0 (p) = 0} be bounded. Let u and u (1.2) with the initial conditions u0 and u˜0 respectively. For any t ≥ 0, set Γt = {p ∈ H : u(p, t) = 0}

and

˜ t = {p ∈ H : u Γ ˜(p, t) = 0}.

˜ t for all t ≥ 0. Then Γt = Γ Proof. We follow the proof of [13, Theorem 5.1]. It is obvious, from Theorem 1.1 and Theorem 5.1, that u and u ˜ are axisymmetric about the vertical axis. We may assume u0 ≥ 0 without changing the zero level set of u0 , since |u| is a solution of (1.2) with the initial condition u(p, 0) = |u0 | by Theorem A.2. Similarly, ˜ ≥ 0. let us also assume that u ˜0 ≥ 0 and u For any k = 1, 2, . . . let E0 = ∅ and Ek = {p ∈ H : u0 (p) > l/k} such that Ek is  nondecreasing and H ⊂ Γ0 = k Ek . Define ˜0 (k = 1, 2, . . .). ak = max u H\Ek−1

1350027-36

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

Then we have limk→∞ ak = 0. We then construct a continuous function θ satisfying θ(0) = 0, θ(1/k) = ak for all k and θ = a1 in [1, ∞). Now it is clear that θ ◦ u is an axisymmetric solution of (1.2) with initial data θ ◦ u0 , again due to Theorem A.2. By our construction of θ, we easily see that θ ◦ u0 ≥ u˜0 . Applying Theorem 1.1 for all T > 0, we get θ ◦ u ≥ u˜. This means that ˜ t for any t ≥ 0. Indeed, for any p ∈ Γt , we have u(p, t) = 0, which implies Γt ⊂ Γ that θ ◦ u(p, t) = 0 and therefore u ˜(p, t) = 0. ˜ t ⊂ Γt for any t ≥ 0. We conclude the proof by similarly showing the inclusion Γ

Appendix B. Viscosity Solutions and Semijets in the Heisenberg Group B.1. Subelliptic jets Let u be an upper-semicontinuous real function in H × [0, ∞). The second-order (parabolic) superjet of u at (ˆ p, tˆ) is defined as  p, tˆ) = (τ, η, X ) ∈ R × R3 × S2 (R) such that J 2,+ u(ˆ 1 u(p, t)  u(ˆ p, tˆ) + τ (t − tˆ) + η, pˆ−1 · p + X h, h 2  + o(|ˆ p−1 · p|2 + |t − tˆ|) . Similarly, for lower-semincontinuous u, we define the second-order subjet  J 2,− u(ˆ p, tˆ) = (τ, η, Y) ∈ R × R3 × S 2 (R) such that 1 u(p) ≥ u(ˆ p, tˆ) + τ (t − tˆ) + η, pˆ−1 · p + Yh, h 2  + o(|ˆ p−1 · p|2 + |t − tˆ|) . One easy way to get jets is by using smooth functions that touch u from above or below: 2 p, tˆ) = {(ϕt (ˆ p, tˆ), ∇ϕ(ˆ p, tˆ), (∇H ϕ(ˆ p, tˆ))∗ ) : ϕ ∈ C 2 in p1 , p2 , ϕ ∈ C 1 in p3 , K 2,+ u(ˆ

ϕ(ˆ p, tˆ) = u(ˆ p, tˆ), ϕ(p, t) ≥ u(p, t), (p, t) in a neighborhood of (ˆ p, tˆ)}. As in the Euclidean case every jet can be obtained by this method. Lemma B.1. We always have K 2,+ u(ˆ p, tˆ) = J 2,+ u(ˆ p, tˆ) and p, tˆ) = J 2,− u(ˆ p, tˆ). K 2,− u(ˆ 1350027-37

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

We also define the “closure” of the second-order superjet of an upperp, tˆ), as the set of (τ, η, X ) ∈ semicontinuous function u at (ˆ p, tˆ), denoted by J¯2,+ u(ˆ 3 2 R×R ×S (R) such that there exist sequences of points (pm , tm ) and (τm , ηm , Xm ) ∈ J 2,+ u(pm , tm ) such that (pm , tm , u(pm , tm ), τm , ηm , Xm ) → (ˆ p, tˆ, u(ˆ p, tˆ), τ, η, X ) as m → ∞. The closure of the second-order subjet of a lower-semicontinuous function v at (ˆ p, tˆ), denoted by J¯2,− (u, pˆ, tˆ) is defined in an analogous manner. B.2. Crandall–Ishii lemma in the Heisenberg group Theorem B.2. Assume u(p, t) ∈ USC(Ω) and v(p, t) ∈ LSC(Ω) with Ω a bounded domain in H × (0, ∞). Suppose that for any ε > 0   1 Mε = sup u(p, t) − v(q, s) − ψ(p · q −1 ) − b(t, s) ε Ω×Ω attains a maximum at (pε , tε , q ε , sε ) ∈ Ω × Ω with ψ(p1 , p2 , p3 ) = (p21 + p22 )2 + p23 and any smooth function b in Ω. Then there exist (τ, ξ, X ) ∈ J¯2,+ u(pε , tε ) and (ζ, η, Y) ∈ J¯2,− v(q ε , sε ) such that τ = bt (tε , sε ),

ζ = −bs (tε , sε ),

X w, w − Y, w, w ≤ a(ε)|w|2

ξ=η

and

for any w ∈ R2 ,

where a(ε) → 0 as ε → 0. Proof. By semicontinuity Mε < ∞ and Mε is attained at a point (pε , tε , q ε , sε ). Note that Mε is increasing in ε and uniform bounded. We have M2ε ≥ u(pε , tε ) − v(q ε , sε ) −

1 ψ(pε · (q ε )−1 ) − b(tε , sε ), 2ε

1 ψ(pε · (q ε )−1 ) ≥ Mε , 2ε 1 M2ε − Mε ≥ ψ(pε · (q ε )−1 ). 2ε First conclude that 1 lim ψ(pε · (q ε )−1 ) = 0. ε→0 ε M2ε −

(B.1)

We apply now the Crandall–Ishii lemma in the Euclidean space [10] for the parabolic problem [17, Theorem 3.3.3]. There exist 3 × 3 symmetric matrices Xε , Yε so that 2,+

(τ, ξε , Xε ) ∈ J eucl.u(pε , tε ) and 2,−

(ζ, ηε , Yε ) ∈ J eucl.v(q ε , sε ) 1350027-38

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

with the property

Xε γ, γ − Yε χ, χ  Cγ ⊕ χ, γ ⊕ χ, ε

ε

where τ = bt (t , s ), ζ = −bs (tε , sε ),   1 p −1  ξε = ∇ (ψ(p · q )) , ε ε ε p=p ,q=q

  1 q −1  ηε = − ∇ (ψ(p · q )) , ε p=pε ,q=qε

3

the vectors γ, χ ∈ R , and C=

1 2 (A + A) ε

and A = ∇2p,q (ψ(p · q −1 )) are 6 × 6 matrices. It is not difficult to see that (τ, DLpε · ξε , (DLpε · Xε · (DLpε )t )2×2 ) ∈ J

2,+

(ζ, DLqε · ηε , (DLqε · Yε · (DLqε )t )2×2 ) ∈ J

u(pε , tε ),

2,−

v(q ε , sε ),

where the subindex 2 × 2 indicates the principal 2 × 2 minor of a 3 × 3 matrix and the mapping Lp is just left multiplication by p in H. One can easily see that its differential is given by  −p2  1 0  2     p 1  DLp = 0 1 .   2   0 0 1 Our choice of ψ(p · q −1 ) implies that DLpε · ξε = DLqε · ηε . Denote it by ξ. Set   1 wpε = (DLpε )t (w1 , w2 , 0)t = w1 , w2 , (pε1 w2 − pε2 w1 ) , 2   1 wqε = (DLpε )t (w1 , w2 , 0)t = w1 , w2 , (q1ε w2 − q2ε w1 ) , 2 X = (DLpε · Xε · (DLpε )t )2×2

and Y = (DLqε · Yε · (DLqε )t )2×2 .

Then

X w, w − Yw, w = Xε wpε , wpε  − Yε wqε , wqε  ≤ C(wpε ⊕ wqε ), wpε ⊕ wqε  2  M pε2 q1ε − pε1 q2ε + pε3 − q3ε , = |w|2 ε 2 where M is a positive constant. We conclude by (B.1) that

X w, w − Yw, w ≤ o(1)|w|2 . 1350027-39

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

F. Ferrari, Q. Liu & J. J. Manfredi

References [1] S. J. Altschuler, S. B. Angenent and Y. Giga, Mean curvature flow through singularities for surfaces of rotation, J. Geom. Anal. 5 (1995) 293–358. [2] N. Arcozzi, F. Ferrari and F. Montefalcone, CC-distance and metric normal of smooth hypersurfaces in sub-Riemannian Carnot groups, preprint (2009); arXiv:0910.5648v1. [3] G. Barles and C. Georgelin, A simple proof of convergence for an approximation scheme for computing motions by mean curvature, SIAM J. Numer. Anal. 32 (1995) 484–500. [4] G. Barles, H. M. Soner and P. Souganidis, Front propagation and phase field theory, SIAM J. Control Optim. 31 (1993) 439–469. [5] T. Bieske, On ∞-harmonic functions on the Heisenberg group, Comm. Partial Differential Equations 27 (2002) 727–761. [6] L. Capogna and G. Citti, Generalized mean curvature flow in Carnot groups, Comm. Partial Differential Equations 34 (2009) 937–956. [7] L. Capogna, D. Danielli, S. Pauls and J. Tyson, An Introduction to the Heisenberg Group and the Sub-Riemannian Isoperimetric Problem, Progress in Mathematics, Vol. 259 (Birkh¨ auser-Verlag, Basel, 2007). [8] Y.-G. Chen, Y. Giga and S. Goto, Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations, J. Differential Geom. 33 (1991) 749–786. [9] P. Cheridito, H. M. Soner, N. Touzi and N. Victoir, Second-order backward stochastic differential equations and fully nonlinear parabolic PDEs, Comm. Pure Appl. Math. 60 (2007) 1081–1110. [10] M. Crandall, H. Ishii and P. L. Lions, User’s guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc. 27 (1992) 1–67. [11] F. Da Lio, N. Forcadel and R. Monneau, Convergence of a non-local eikonal equation to anisotropic mean curvature motion. Application to dislocation dynamics, J. Eur. Math. Soc. 10 (2008) 1061–1104. [12] N. Dirr, F. Dragoni and M. von Renesse, Evolution by mean curvature flow in subRiemannian geometries: A stochastic approach, Commun. Pure Appl. Anal. 9 (2010) 307–326. [13] L. C. Evans and J. Spruck, Motion of level sets by mean curvature. I, J. Differential Geom. 33 (1991) 635–681. , Motion of level sets by mean curvature. III, J. Geom. Anal. 2 (1992) [14] 121–150. [15] F. Ferrari, Q. Liu and J. J. Manfredi, On the characterization of p-harmonic functions on the Heisenberg group by mean value properties, preprint (2012); arXiv:1210.2881. , Some properties of the mean curvature of axisymmetric surfaces in the [16] Heisenberg group, preprint (2012). [17] Y. Giga, Surface Evolution Equations, A Level Set Approach, Monographs in Mathematics, Vol. 99 (Birkh¨ auser-Verlag, Basel, 2006). [18] Y. Giga and Q. Liu, A remark on the discrete deterministic game approach for curvature flow equations, in Nonlinear Phenomena with Energy Dissipation: Mathematical Analysis, Modeling and Simulation, Gakuto International Series Mathematical Sciences and Applications, Vol. 29 (Gakk¯ otosho, 2008), pp. 103–115. [19] G. Huisken, Flow by mean curvature of convex surfaces into spheres, J. Differential Geom. 20 (1984) 237–266. [20] H. Ishii, Perron’s method for Hamilton–Jacobi equations, Duke Math. J. 55 (1987) 369–384.

1350027-40

May 14, 2014 15:25 WSPC/S0219-1997

152-CCM

1350027

Horizontal mean curvature flow in the Heisenberg group

[21] R. V. Kohn and S. Serfaty, A deterministic-control-based approach to motion by curvature, Comm. Pure Appl. Math. 59 (2006) 344–407. , A deterministic-control-based approach to fully nonlinear parabolic and ellip[22] tic equations, Comm. Pure Appl. Math. 63 (2010) 1298–1350. [23] Q. Liu, Fattening and comparison principle for level-set equations of mean curvature type, SIAM J. Control Optim. 49 (2011) 2518–2541. [24] J. J. Manfredi, Analysis and geometry in metric spaces, Note for the course on “Nonlinear Subelliptic Equations on Carnot Groups” (2003). [25] J. J. Manfredi, M. Parviainen and J. D. Rossi, An asymptotic mean value characterization for p-harmonic functions, Proc. Amer. Math. Soc. 138 (2010) 881–889. , An asymptotic mean value characterization for a class of nonlinear parabolic [26] equations related to tug-of-war games, SIAM J. Math. Anal. 42 (2010) 2058–2081. [27] R. Montgomery, A Tour of Subriemannian Geometries, Their Geodesics and Applications, Mathematical Surveys an Monographs, Vol. 91 (American Mathematical Society, Providence, RI, 2002). [28] Y. Peres, O. Schramm, S. Sheffield and D. Wilson, Tug-of-war and the infinity Laplacian, J. Amer. Math. Soc. 22 (2009) 167–210. [29] Y. Peres and S. Sheffield, Tug of war with noise: A game theoretic view of the pLaplacian, Duke Math. J. 145 (2008) 91–120. [30] H. M. Soner, Motion of a set by the curvature of its boundary, J. Differential Equations 101 (1993) 313–372. [31] H. M. Soner and P. E. Souganidis, Singularities and uniqueness of cylindrically symmetric surfaces moving by mean curvature, Comm. Partial Differential Equations 18 (1993) 859–894. [32] H. M. Soner and N. Touzi, A stochastic representation for mean curvature type geometric flows, Ann. Probab. 31 (2003) 1145–1165. [33] C. Wang, Viscosity convex functions on Carnot groups, Proc. Amer. Math. Soc. 133 (2005) 1247–1253.

1350027-41

On the horizontal mean curvature flow for axisymmetric ...

Aug 7, 2013 - We study the horizontal mean curvature flow in the Heisenberg group by using the level- set method. We prove the uniqueness, existence and stability of axisymmetric viscosity solutions of the level-set equation. An explicit solution is given for the motion starting from a subelliptic sphere. We also give ...

511KB Sizes 0 Downloads 219 Views

Recommend Documents

pdf-1294\regularity-theory-for-mean-curvature-flow ...
Connect more apps... Try one of the apps below to open or edit this item. pdf-1294\regularity-theory-for-mean-curvature-flow-by-klaus-ecker-birkhauser.pdf.

pdf-1295\regularity-theory-for-mean-curvature-flow-by ...
pdf-1295\regularity-theory-for-mean-curvature-flow-by-klaus-ecker.pdf. pdf-1295\regularity-theory-for-mean-curvature-flow-by-klaus-ecker.pdf. Open. Extract.

The mean curvature integral is invariant under bending ...
The author would like to thank Dick Canary, Ed Taylor, and Brian Bowditch for ...... Rob Kirby, Frances Kirwan, Dieter Kotschick, Peter Kronheimer, Ib Madsen,.

Translation surfaces with constant mean curvature in 3 ... - Springer Link
Gauss curvature in 3-dimensional Euclidean space E 8 and 3-dimensional Minkowski space. E~. 1991 Mathematics subject Classifications. 53 C 42, 53 C 40, 53 C 50. Keywords and phrases. Mean curvature, translation surface, spacelike surface, timelike su

3D Mesh Segmentation Using Mean-Shifted Curvature
meaningful and pleasing results, several ingredients are introduced into the .... function definition which uses the product of kernels to filter two feature components ..... Pan, X., Ye, X. and Zhang, S.: 3D Mesh segmentation using a two-stage ...

Translation surfaces with constant mean curvature in 3 ... | Springer Link
of finite-type harmonic maps. For the .... 9"(x) (1 + g'(x) 2 + h'(y) 2) - 4g'(x)g"(x) 2 = O. .... [11] A. TREIBERGS, Gauss map of spacelike constant mean curvature ...

A Level Set Formulation of Geodesic Curvature Flow on ...
2009; published online 19 Aug. 2009. Recommended for acceptance ...... and the PhD degree in applied mathematics from Jyvaskalya University in Finland in ...

Deterministic Performance Bounds on the Mean Square Error for Near ...
the most popular tool [11]. However ... Date of publication November 27, 2012; ... of this manuscript and approving it for publication was Dr. Benoit Champagne.

Deterministic Performance Bounds on the Mean Square Error for Near ...
mean square error applied to the passive near field source localization. More precisely, we focus on the ... Index Terms—Deterministic lower bounds, mean square error, near field source localization, performance analysis ..... contained in the samp

Life on the Mean Reefs
“It's hot-bunking,” says Friedlander, referring to the Navy custom ... Still, life is literally no picnic at the top of the food chain. ... stocks,” Sandin says. “That's the key ...

Visual Curvature
extreme points, the sharper the contour is at the point, i.e. point A, the higher is the ... window, thus they in essence estimate the curvature locally. The size of the ...

Ribbed elastic structures under a mean flow
systems. When the rib array becomes non-equispaced the pass bands are destroyed ...... Research Council, BAE Systems, the ONR and the Isaac Newton Trust.

Curvature effects on optical response of Si nanocrystals ...
Aug 26, 2009 - At planar interfaces (or surfaces) the electronic charge density is shown to ... method for measuring interface effects due to Si suboxides in Si NC in SiO2 is also ...... 8 O. Madelung, Semiconductors: Data Handbook (Springer-.

Appendix - The Riemann curvature tensor.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Appendix - The ...

Deterministic Performance Bounds on the Mean ...
IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 0, NO. , 2012. 1. Deterministic Performance Bounds on the Mean Square. Error for Near Field Source Localization. Mohammed Nabil El Korso, Alexandre Renaux, Rémy Boyer, and. Sylvie Marcos. Abstract—This

On-line index maintenance using horizontal partitioning - People
Nov 6, 2009 - In step (9), the sub-index at root node is split, and the infre- quent part ..... incremental indexing for full-text information retrieval. In VLDB '94: ...

On-line index maintenance using horizontal partitioning
Nov 6, 2009 - tion Storage and Retrieval]: Content Analysis and Index- ing—Indexing ... eral, a query to an IR system is a list of terms along with ... not made or distributed for profit or commercial advantage and that copies ...... In Proceedings

On-line index maintenance using horizontal partitioning - People
Nov 6, 2009 - on-disk inverted index into two types of sub-indexes. In- verted lists of the terms ... Inverted Index is an important data structure used in ... Permission to make digital or hard copies of all or part of this work for personal or ...

Horizontal Equity
designing an optimal income tax. ... illegal behaviour of the individuals that a given policy addresses (e.g. tax ..... John M. Olin Center for Law, Economics.

On surface flow rheology
granular flows by accounting for nonlocal effects,6–10 by adapting kinetic theory .... lomb's one for which the basic features are as follows: The friction force lies in ... LMGC90 software.48 On a SGI Origin 3800 with 16 proces- sors, about 20 h .

See Visual Flow Diagram on the following pages -
Prioritze KME's based on Strategic Direction Sequence. 2. Prioritized KME's assigned to small team workgroups. 3. Small teams brainstorm activities necessary ...

On the Flow Anonymity Problem in Network Coding
proposed Source Authentication Scheme using Network Coding. (SANC). .... propound a signature mechanism that provides data integrity for the packet mixed ...

Horizontal Vertical Lines.pdf
Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Horizontal Vertical Lines.pdf. Horizontal Vertical Lines.pdf. Open.