Home A sharp global estimate and an overdetermined problem for Monge-Ampère type equations
Article Open Access

A sharp global estimate and an overdetermined problem for Monge-Ampère type equations

  • Ahmed Mohammed EMAIL logo and Giovanni Porru
Published/Copyright: February 26, 2022

Abstract

We consider Monge-Ampère type equations involving the gradient that are elliptic in the framework of convex functions. Through suitable symmetrization we find sharp estimates to solutions of such equations. An overdetermined problem related to our Monge-Ampère type operators is also considered and we show that such a problem may admit a solution only when the underlying domain is a ball.

MSC 2010: 35B45; 35B51; 35J96; 35N25; 49J40

1 Introduction

About forty years ago, G. Talenti [14] pioneered a method for establishing the following result. Let Ω be a convex domain in ℝ2, and let u = u(x, y) be a (negative) convex solution to the problem

uxxuyyuxy2=1inΩ,u=0onΩ.

Moreover, let Ω(1)* be the disc centered at the origin, having the same perimeter as Ω, and let v = v(x, y) be the negative solution to the problem

vxxvyyvxy2=1inΩ(1)*,v=0onΩ(1)*.

Then

u*(x)v(x)inΩ(1)*,

where u* is a suitable symmetrization of u which is defined in Ω(1)* . Clearly, the solution u depends on the shape of Ω, and cannot be written explicitly in general, whereas v is radially symmetric and can be computed explicitly as a solution to an ordinary differential equation. The previous result has been extended to arbitrary dimension n by K. Tso [18, 19] and by N. Trudinger [15, 16], by using a suitable (depending on the dimension n) symmetrization of u. See also [1, 10, 17]. One purpose of the present paper is to extend a similar result to equations which also depend on the gradient Du.

The second problem we consider in this paper is inspired by the following well known result. Let Ω be a bounded smooth domain in ℝn, and let u be the solution to the problem

Δu=1inΩ,u=0onΩ.

If there is a constant c > 0 such that |Du| = c on ∂Ω, then it is well-known that Ω must be a ball (see [13]). As far as we know, a problem of this kind, usually referred to as an overdetermined problem, was first considered by J. Serrin, [13]. The above result is a special case of Serrin's problem. The paper [13] has inspired numerous investigations involving linear, nonlinear and fully nonlinear operators. We refer to [3, 4, 7, 20] and the references therein.

We now introduce the Monge-Ampère type equations considered in this paper. Let Ω ⊂ ℝn be a bounded convex domain with a smooth boundary ∂Ω. For xΩ, we write x = (x1, ⋯, xn). We use subscripts to denote partial differentiation. For example, we write ui=uxi,uij=2uxixj , etc. The familiar Monge-Ampère operator can be written as

det(D2u)=1n(T(n1)ij(D2u)ui)j,

where D2u denotes the Hessian matrix of the function u, det(D2u) is the determinant of D2u, and T(n−1) = T(n−1)(D2u) is the (n − 1)-order Newton tensor of D2u ([11, 12]). Here and in what follows, the summation convention from 1 to n over repeated indices is in effect. Note that T(n−1) corresponds to the cofactor matrix of D2u.

Let g : [0, ∞) (0, ∞) be a smooth real function satisfying

G(s2)g(s2)+2s2g(s2)>0s0.

Since

G(s2)=dds(g(s2)s),

the function g(s2)s is positive and strictly increasing for s > 0. Note that we are assuming that g(0) and G(0) are positive numbers.

We define the g-Monge-Ampère operator as

1n(T(n1)ij(D2u)gn(|Du|2)ui)j.

Recalling that T(n−1)(D2u) is divergence free (see [11, 12]), we have

(1) (T(n1)ij(D2u))j=0,i=1,,n.

Moreover, since T(n−1)(D2u) is the cofactor matrix of D2u we have

(2) T(n1)(D2u)D2u=Idet(D2u),

where I denotes the n × n identity matrix. By using (1) and (2), we find

(3) 1n(T(n1)ij(D2u)gn(|Du|2)ui)j=gn1(|Du|2)G(|Du|2)det(D2u).

Since the operator det(D2u) is elliptic (in the framework of convex functions), our g-Monge-Ampère operator is elliptic.

A motivation for the definition of the g-Monge-Ampère operator is the following. Using the Kronecker delta δi𝓁, define the n × n matrix 𝒜 = [𝒜ij] with

𝒜ij=(g(|Du|2)ui)j=[g(|Du|2)δi𝓁+2g(|Du|2)uiu𝓁]u𝓁j.

The trace of the matrix 𝒜 is the familiar operator (g(|Du|2)ui)i. We claim that the determinant of the matrix 𝒜 coincides with our operator (3). Indeed, the eigenvalues Λi, i = 1, ⋯ , n, of the n × n matrix

=[g(|Du|2)δi𝓁+2g(|Du|2)uiu𝓁]

are the following

Λ1==Λn1=g(|Du|2),Λn=G(|Du|2).

Since det𝒜=detℬ·det(D2u), we find

det𝒜=gn1(|Du|2)G(|Du|2)det(D2u).

The claim follows from the latter equation and (3).

In this paper we show that many results which hold for Monge-Ampère equations can be extended to the corresponding g-Monge-Ampère equations. To state our main results we begin by introducing two assumptions. We need the following notations.

(4) 𝒦(s2)=gn1(s2)G(s2),and𝒢(s)gn(s2)sn+1,s>0.

We make the following assumptions.

There is a constants γ > 0 such that

(5) tγn𝒦(s2)𝒦(t2s2)(t,s)(0,1)×(0,).

Let f (t) be a non-decreasing function. With γ as in (5), suppose there is 0 ≤ q < γ such that

(6) f(t)tqisnon-increasingon(0,).

In stating our results we will consider functions from the class

Φ(Ω){uC2(Ω)C0,1(Ω¯):uisconvexinΩandu=0onΩ}.

We now state our first result. We refer to Section 2 for the definitions of (n − 1)-symmetrand of a function on Ω, the one-mean radius η(1)(Ω), and the Gauss curvature ℋ(n−1) of ∂Ω.

Theorem 1.1.

Let Ω ⊂ ℝn be a convex and smooth domain, and suppose that 𝒢 defined as in (4) is convex in (0, ∞). Suppose f is positive, non-decreasing. We further assume that conditions (5) and (6) hold. Let vΦ (Ω) be a super-solution of

(7) 𝒦(|Du|2)det(D2u)=f(u)inΩ,

and let v* be its (n − 1)-symmetrand. Let zΦ(BR) be a sub-solution of (7) in the ball BR, where Rη(1)(Ω). If v*(r) = v*(x) and z(r) = z(x) for r = |x|, then we have

v*(r)z(r),0<r<R.

For our second main result we consider an overdetermined system that may admit a convex solution only when the underlying domain is a ball. More specifically we have the following. Recall that functions in Φ(Ω) vanish on the boundary ∂Ω.

Theorem 1.2.

Let c > 0 be a given constant. If there is a solution uΦ(Ω) of the system

(8) 1n(T(n1)ij(D2u)gn(|Du|2)ui)j=cinΩ,

(9) (n1)(g(|Du|2)|Du|)n1=cn1nonΩ,

then Ω is a ball.

For results of existence and regularity of Monge-Ampère equations we refer to [2, 5] and the references therein.

We organize the rest of the paper as follows. Section 2 provides a brief preliminary in which basic notions are recalled. In Section 3 we give an estimate of (n − 1)-symmetrands of supersolutions vΦ(Ω) to (7). This will be followed by a comparison principle for the equation (7), which may be of independent interest. In the same section we will also present the proof of Theorem 1.1. As an application of Theorem 1.1, we give an estimate of an Hessian integral. In Section 4, we will give the proof of Theorem 1.2.

2 Preliminaries

Let κ1, ⋯ , κn−1 be the principal curvatures of ∂Ω. For j = 1, ⋯ , n − 1 we define the j-th mean curvature of ∂Ω by

(j)(Ω)=S(j)(κ1,,κn1),

where S(j) denotes the elementary symmetric function of order j of κ1, ⋯ , κn−1. Let uΦ(Ω). From Sard's theorem it follows that for almost all t ∈ (m0, 0), m0 = minΩ u, the sub-level sets

Ωt={xΩ:u(x)<t}

will have smooth boundary t given by the level surface

Σt={xΩ:u(x)=t}.

Clearly, Ω0 = Ω. Furthermore, we denote with ωn the volume of the unit ball in ℝn.

For an open set 𝒪 ⊂ ℝn, the quermassintegral V(1) = V(1)(𝒪) is defined by

(10) V(1)(𝒪)=1n(n1)𝒪(n2)(𝒪)dσ,

where denotes the (n − 1)-dimensional Hausdorff measure in ℝn.

Following [16] we define the 1-mean radius of 𝒪, denoted by η(1)(𝒪), as

η(1)(𝒪)=V(1)(𝒪)ωn.

In case 𝒪 is a ball, η(1)(𝒪) is the radius of 𝒪. For a general 𝒪, we denote by 𝒪(n1)* the ball with radius η(1)(𝒪).

As usual, we denote by |E| the Lebesgue measure of a subset E ⊂ ℝn. The following isoperimetric inequality is well known.

(|𝒪|ωn)1nV(1)(𝒪)ωn.

It follows that

|𝒪||𝒪(n1)*|.

We define the rearrangement of u with respect to the quermassintegral V(1) as

(11) u(s)=sup{t0:V(1)(Ωt)s,0sV(1)(Ω)}.

The function u(s) is negative, increasing and satisfies

u(0)=minΩu(x),u(V(1)(Ω))=0.

We also define

(12) u*(x)=u(ωn|x|),0|x|η(1)(Ω).

The function u*(x) is called the (n − 1)-symmetrand of u, and can also be defined by (see [16])

u*(x)=sup{t0:η(1)(Ωt)|x|,0|x|η(1)(Ω)}.

Since u*(x) is radially symmetric we often write u*(x) = u*(r) for |x| = r. We have u*(0) = min Ω u(x) and u*(η(1)(Ω)) = 0.

3 Estimates via symmetrization

Lemma 3.1.

Let uΦ (Ω). For t ∈ (m0, 0), m0 = minΩ u(x), let

Σt={xΩ:u(x)=t}.

For almost every t ∈ (m0, 0) we have

(13) T(n1)ij(D2u)uiuj=(n1)|Du|n+1onΣt.

Furthermore we have

(14) ddrΣt(n2)dσ=(n1)Σt(n1)|Du|1dσ.

Proof

For a proof of (13) we refer to [11, 12]. For a proof of (14) see equation (2.20) of [16].

Theorem 3.2.

Let vΦ(Ω), let f > 0 be non-decreasing and let

(15) 1n(T(n1)ij(D2v)gn(|Dv|2)vi)jf(v)inΩ.

If the function 𝒢 defined as in (4) is convex then

(16) (g((v˙*)2)v˙*)nn0rtn1f(v*(t))dt,

where v* = v*(r) is the (n − 1)-symmetrand of v, and v˙*=dv*dr .

Proof

We note that for g(s2) = 1 this result is proved in [10]. Integrating (15) over Ωt we find

1nΣtT(n1)ij(D2v)gn(|Dv|2)|Dv|1vivjdσΩtf(v)dx.

On using (13), this inequality yields

Σt(n1)gn(|Dv|2)|Dv|ndσnΩtf(v)dx.

In terms of 𝒢, the latter inequality reads as

(17) Σt(n1)𝒢(|Dv|)|Dv|1dσnΩtf(v)dx.

Now we use Jensen's inequality in the following form

𝒢(Σt|Dv|dmΣtdm)Σt𝒢(|Dv|)dmΣtdm.

With dm = ℋ(n−1)|Dv|−1 we find

𝒢(Σt(n1)dσΣt(n1)|Dv|1dσ)Σt𝒢(|Dv|)(n1)|Dv|1dσΣt(n1)|Dv|1dσ.

Recalling that

Σt(n1)dσ=nωn,

by the previous inequality and (17) we find

(18) 𝒢(nωnΣt(n1)|Dv|1dσ)nΩtf(v)dxΣt(n1)|Dv|1dσ.

On using (14) and putting V(1)(Ωt) = V(1)(t) we have

Σt(n1)|Dv|1dσ=nV(1)(t).

Insertion of this equation into (18) yields

(19) 𝒢(ωnV(1)(t))Ωtf(v)dxV(1)(t).

On the other hand we have (see [6], Theorem 3.36 and Proposition 6.23; or [10], pag 190)

Ωtf(v)dxnωn1nm0tf(τ)V(1)n1(τ)V(1)(τ)dτ.

Putting V(1)(τ) = ρ, and recalling that v(ρ) is the inverse of V(1)(τ) (see (11)) we find

Ωtf(v)dxnωn1n0V(1)(t)f(v(ρ))ρn1dρ.

Inserting this into (19) we get

𝒢(ωnV(1)(t))nωn1n0V(1)(t)f(v(ρ))ρn1dρV(1)(t).

Putting V(1)(t) = s we have

V(1)(t)=(dv(s)ds)1.

Hence,

(dv(s)ds)1𝒢(ωndv(s)ds)nωn1n0sρn1f(v(ρ))dρ.

Putting sωnr, recalling (12) and writing dv*(r)drv˙* , the latter inequality yields

(20) (v˙*)1𝒢(v˙*)nωnn0ωnrρn1f(v(ρ))dρ.

Finally, putting ρωnt and recalling the definition of 𝒢, (20) yields

(21) (g((v˙*)2)v˙*)nn0rtn1f(v*(t))dt.

The theorem is proved.

If Ω is a ball and v is a (radial) solution of (15) with equality sign, then, all the inequalities used in the proof of the latter theorem are equalities. In this situation we have v*(r) = v(r) and

(22) (g(v˙2)v˙)n=n0rtn1f(v(t))dt.

Now we prove a comparison principle.

Lemma 3.3.

Let Ω ⊂ ℝn be a convex domain. Let f = f (t) be positive, non-decreasing and such that conditions (5) and (6) hold. Let u, vΦ(Ω) satisfy

(23) 𝒦(|Du|2)det(D2u)f(u)inΩ,

(24) 𝒦(|Dv|2)det(D2v)f(v)inΩ.

Then uv in Ω.

Proof

Since u is convex, and hence sub-harmonic on Ω, it follows from Hopf's lemma that, for some constant c1 > 0, −u(x) ≥ c1d(x) in Ω, where d(x) denotes the distance of xΩ to the boundary ∂ Ω, and c1 is a suitable positive constant. Moreover, since vC0,1(Ω¯) , there is a constant c2 > 0 such that −v(x) ≤ c2d(x) on Ω. Consequently,

u(x)c1c2v(x)xΩ¯.

We can take c1 > 0 sufficiently small such that c1 < c2. To produce a contradiction, suppose that uv in Ω does not hold. Let

S{λ[0,1]:u(x)λv(x)xΩ}.

Let Λ ≔ sup S. Note that 0 < Λ < 1, and u(x) ≤ Λ v(x) for all xΩ . By condition (6) we have

(25) f(Λv(x))Λqf(v(x))xΩ¯,

Since 0 ≤ q < γ, we can choose ϵ > 0 sufficiently small that Λ + ϵ < 1 and Λq > (Λ + ϵ )γ. Let w ≔ (Λ + ϵ)v. Then we find

𝒦(|Du|2)det(D2u)f(u)(by(23))f(Λv)(sincefisnondecreasing)Λqf(v)(by(25))>(Λ+ɛ)γf(v)(sinceΛq>(Λ+ɛ)γ)(Λ+ɛ)γnK(|Dv|2)(Λ+ɛ)ndet(D2v)(by(4))K(|Dw|2)det(D2w)(by(5)).

In conclusion, we find that

(26) 𝒦(|Du|2)det(D2u)>𝒦(|Dw|2)det(D2w)xΩ.

We claim that

u(x)w(x)inΩ.

If not, let x¯Ω such that 0<(uw)(x¯) is the maximum of uw in Ω. Then Du(x¯)=Dw(x¯) and, recalling that u and w are convex in Ω, we have

det(D2u(x¯))det(D2w(x¯)).

On noting that 𝒦(|Du(x¯)|2)=𝒦(|Dw(x¯)|2) we find

𝒦(|Du(x¯)|2)det(D2u(x¯))𝒦(|Dw(x¯)|2)det(D2w(x¯)),

which contradicts the strict inequality in (26). Hence,

u(x)w(x)=(Λ+ɛ)v(x),

which implies the contradictory conclusion Λ + ϵS. The lemma is proved.

Proof of Theorem 1.1. Let w < 0 be a (radial) solution of

(27) 𝒦(|Dw|2)det(D2w)=f(v*)inBR,w=0onBR.

Recalling the definition of 𝒦 given in (4), and since

det(D2w)=w¨(w˙r)n1,

from (27) we find

gn1(w˙2)G(w˙2)w˙n1w¨=rn1f(v*),(g(w˙2)w˙)n1ddr(g(w˙2)w˙)=rn1f(v*),1nddr(g(w˙2)w˙)n=rn1f(v*).

Integration over (0, r) yields

(g(w˙2)w˙)n=n0rtn1f(v*(t))dt.

Comparing this equation with inequality (16) we find

(g((v˙*)2)v˙*)n(g(w˙2)w˙)n,

which implies

v˙*(r)w˙(r),0<r<R.

Integrating over (r, R) and making use of the conditions v*(R) = w(R) = 0, we find

v*(r)w(r).

With v*(x) = v*(r) and w(x) = w(r) for r = |x| we have

(28) v*(x)w(x).

By (27) and (28) we find

𝒦(|Dw|2)det(D2w)=f(v*)f(w)inBR.

In summary, we see that w and z satisfy

𝒦(|Dw|2)det(D2w)f(w)inBR,w=0onBR,𝒦(|Dz|2)det(D2z)f(z)inBR,z=0onBR.

By Lemma 3.3 we have,

w(x)z(x)inBR.

Thus, from (28) and the latter inequality we conclude

v*(x)z(x)inBR.

The theorem follows.

Corollary 3.4.

Let h : [0, ∞) [0, ∞) be a non-decreasing smooth function such that h(t) > 0 for t > 0. Under the assumptions and notation of Theorem 1.1, we have

Ωh(v)dxΩ(n1)*h(z)dx.

Proof

Let

μ(t)=|{xΩ:v(x)<t}|,μ*(t)=|{xΩ(n1)*:v*(x)<t}|.

We know that (see (2.24) of [16])

μ(t)μ*(t)t(m0,0),m0=infΩv(x).

Recalling that m0 = v(0) = v*(0) we find

Ωh(v)dx=m00h(t)μ(t)dt=h(0)μ(0)+m00h(t)μ(t)dth(0)μ*(0)+m00h(t)μ*(t)dt=m00h(t)(μ*)(t)dt=Ω(n1)*h(v*)dx.

Since by Theorem 1.1 we have −v*(x) ≤ −z(x), and h is non-decreasing, the corollary follows.

Let β be a non-negative real number. For uΦ(Ω), we define the Hessian integral

I(Ω,β,v)=Ω(v)β𝒦(|Dv|2)det(D2v)dx.

Special cases.

If β = 1, recalling (4) and (3), we find

I(Ω,1,v)=Ω(v)𝒦(|Dv|2)det(D2v)dx=1nΩT(n1)ijgn(|Dv|2)vivjdx.

If β = 0, recalling (4), (3) and (13), we find

I(Ω,0,v)=Ω𝒦(|Dv|2)det(D2v)dx=1nΩ(n1)gn(|Dv|2)|Dv|ndσ.

Proposition 3.5.

Assume the conditions of Theorem 1.1, and let vΦ (Ω) be a super-solution of the equation

𝒦(|Du|2)det(D2u)=f(u)inΩ.

If zΦ(Ω(n1)*) is a sub-solution of the above equation in the ball Ω(n1)* , we have

I(Ω,β,v)I(Ω(n1)*,β,z).

Proof

By using Corollary 3.4 we find

I(Ω,β,v)=Ω(v)β𝒦(|Dv|2)det(D2v)dxΩ(v)βf(v)dx(sincevisasuper-solution)Ω(n1)*(z)βf(z)dx(byCorollary3.4)Ω(n1)*(z)β𝒦(|Dz|2)det(D2z)dx(sincezisasub-solution)=I(Ω(n1)*,β,z).

The proposition is proved.

We end this section with the following remark. In [5] (Section 5) and in [8] (Chapter 17), the following equation has been investigated.

det(D2v)=f(v)Q(|Dv|2),Q(s2)>0s0.

We note that this equation coincides with our equation (7) with

𝒦(s2)=gn1(s2)G(s2)=Q(s2).

Since

gn1(s2)G(s2)=gn1(s2)dds(g(s2)s)=s1n1ndds(g(s2)s)n,

we find

s1n1ndds(g(s2)s)n=Q(s2)

and

gn(s2)=nsn0stn1Q(t2)dt.

Using the latter function g(s2) we can apply our previous results to the present equation. For instance, (16) reads as

0v*(r)tn1Q(t2)dt0rtn1f(v*(t))dt.

4 Overdetermined problem

Proof of Theorem 1.2. We note that in case of g(s2) = 1 and c = 1, this theorem is proved in [4]. On using (13) we find

Ω[cn1nn(g(|Du|2)uj)j1n(T(n1)ij(D2u)gn(|Du|2)ui)j]dx=1nΩ(cn1ng(|Du|2)ujT(n1)ij(D2u)gn(|Du|2)ui)jdx=1nΩ(cn1ng(|Du|2)ujT(n1)ij(D2u)gn(|Du|2)ui)uj|Du|dσ=1nΩ(cn1ng(|Du|2)|Du|T(n1)ij(D2u)gn(|Du|2)|Du|1uiuj)dσ=1nΩ(cn1ng(|Du|2)|Du|(n1)gn(|Du|2)|Du|n)dσ.

From this and (9) we find

(29) Ω[cn1nn(g(|Du|2)uj)j1n(T(n1)ij(D2u)gn(|Du|2)ui)j]dx=1nΩg(|Du|2)|Du|(cn1n(n1)(g(|Du|2)|Du|)n1)dσ=0.

Now we prove that for a solution u to (8) we have

(30) cn1nn(g(|Du|2)uj)j1n(T(n1)ij(D2u)gn(|Du|2)ui)j0.

Indeed, in view of (8), inequality (30) can be written as

(31) 1n(g(|Du|2)uj)jc1nnc=c1n.

Recall that, if 𝒜 = [𝒜ij] with 𝒜ij = g(|Du|2ui)j, we have

1n(T(n1)ij(D2u)gn(|Du|2)ui)j=det𝒜.

By the latter equation and (8), (31) can be written as

(32) 1n(g(|Du|2)uj)j(det𝒜)1n.

To prove (32), we first apply the arithmetic-geometric mean inequality

(33) 1n(g(|Du|2)uj)j((g(|Du|2)u1)1(g(|Du|2)un)n)1n.

We have equality in (33) if and only if

(34) (g(|Du|2)u1)1==(g(|Du|2)un)n.

Now we apply the following Hadamard inequality to the positive definite matrix 𝒜 (see Theorem 7.8.1 of [9])

(35) 𝒜11𝒜nndet𝒜,

with equality sign if and only if

(36) 𝒜ij=0,ij.

With 𝒜ij = g(|Du|2)ui)j, (35) reads as

(37) (g(|Du|2)u1)1(g(|Du|2)un)ndet𝒜.

Conditions (36) become

(38) (g(|Du|)2ui)j=0,ij.

Inequality (32) follows by (33) and (37). Hence (30) holds. Then, by (29) we must have

(39) cn1nn(g(|Du|2)uj)j1n(T(n1)ij(D2u)gn(|Du|2)ui)j=0.

But, as equality holds in (30), also equations (34) and (38) must hold.

To finish the proof, we note that by (39) and (8) we find

cn1nn(g(|Du|2)uj)j=c.

Recall that the summation convention is in effect in the latter equation. Therefore, on using (34), for j fixed we find

(g(|Du|2)uj)j=c1n,j=1,,n.

Integrating and using (38) we get

(40) g(|Du|2)uj=c1n(xjx¯j),j=1,,n,

where x¯=(x¯1,,x¯n) is a point of minimum for u(x). It follows that

g2(|Du|2)uj2=c2n(xjx¯j)2,j=1,,n.

Adding from j = 1 to j = n we find

g2(|Du|2)|Du|2=c2nr2,r2=1n(xjx¯j)2,

and

g(|Du|2)|Du|=c1nr.

Since g(s2)s is strictly increasing, |Du| must be radial around the point of minimum of u(x). Then, from (40) we get

(41) uj=F(r)(xjx¯j),j=1,,n,

where F is a suitable function. Note that (41) can be written as

(42) uj=(K(r))j,j=1,,n,

with

K(r)=0rF(t)tdt.

By (42) we find

u(r)=K(r)+c.

Hence, u is radially symmetric. Since u is strictly convex, Ω must be a ball centered at x¯ . The theorem is proved.

In [21], the Monge-Ampère equation det(D2u) = 1 in two dimensions has been discussed. In particular, it is proved that the P-function P(x) = |Du|2 − 2u attains its maximum value on ∂Ω.

Appendix

If g(s2) = 1, we find G(s2) = 1.

If g(s2)=(1+s2)α2 , α ≤ 1, then G(s2)=(1+s2)α+22(1+(1α)s2)>0 .

Comments on the convexity of the function 𝒢(s) defined in (4).

If g(s2) = 1, we find

𝒢(s)=sn+1,

which is convex for s > 0.

If g(s2)=(1+s2)α2 , then

𝒢(s)=(1+s2)nα2sn+1,s>0.

We have

𝒢(s)=(1+s2)nα+22[((1α)n+1)sn+2+(n+1)sn],

and

𝒢(s)=(1+s2)nα+42{(nα+2)[((1α)n+1)sn+3+(n+1)sn+1]+(n+2)((1α)n+1)sn+1+n(n+1)sn1+(n+2)((1α)n+1)sn+3+n(n+1)sn+1}.

Simplification yields

𝒢(s)=n(1+s2)nα+42{((1α)n+1)(1α)sn+3+(2n(1α)+23α)sn+1+(n+1)sn1}.

By computations one finds

𝒢(s)>0forα<8n+88n+9.

Comments on the condition (5).

If g(s2) = 1, condition (5) with γ = 2n reads as

tnsn(ts)n,

which is trivially satisfied.

If g(s2)=(1+s2)α2 with n ≥ 3, 0 < α < (n − 2)/n, condition (5) with

γ=n(1α)2,

reads as

t(nα+2)(1+s2)nα+22(1+(1α)s2)(1+t2s2)nα+22(1+(1α)t2s2).

Simplifying we find

(t2+s2)nα+22(1+(1α)s2)(1+s2)nα+22(1+(1α)t2s2),

which is clearly satisfied for 0 < t < 1.

  1. Conflict of Interest:

    The authors declare that they have no conflict of interest.

References

[1] B. Brandolini and C. Trombetti, A symmetrization result for Monge-Ampère type equations, Math. Nachr. 280 (2007), 467–478.10.1002/mana.200410495Search in Google Scholar

[2] L. Caffarelli, L. Nirenberg and J. Spruck, The Dirichlet problem for nonlinear second order elliptic equations III: Functions of eigenvalues of the Hessian, Acta Math., 155 (1985), 261–301.10.1007/BF02392544Search in Google Scholar

[3] C. Chen, X. Ma and S. Shi, Curvature estimates for the level sets of solutions to the Monge-Ampère equation D2u = 1, Chin. Ann. Math., Series B (2014), 895–906.10.1007/s11401-014-0864-6Search in Google Scholar

[4] C. Enache, Maximum principles and symmetry results for a class of fully nonlinear elliptic PDEs, Nonlinear Differ. Equ. Appl. 17 (2010), 591–600.10.1007/s00030-010-0070-5Search in Google Scholar

[5] A. Figalli, The Monge-Ampère Equation and Its Applications, Zurich Lectures in Advanced Mathematics.10.4171/170Search in Google Scholar

[6] G.B. Folland, Real analysis. Modern techniques and their applications, Second edition. Pure and Applied Mathematics (New York). A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1999.Search in Google Scholar

[7] N. Garofalo and J. L. Lewis, A symmetry result related to some overdetermined boundary value problems, American Journal of Mathematics, 111 (1989), 9–33.10.2307/2374477Search in Google Scholar

[8] D. Gilbarg and N.S. Trudinger, Elliptic partial differential equations of second order , 2.nd edn. Springer, Berlin (1983).Search in Google Scholar

[9] R.A. Horn and C.R. Johnson, Matrix analysis. Second edition. Cambridge University Press, Cambridge, 2013.Search in Google Scholar

[10] A. Mohammed, G. Porru and A. Safoui, Isoperimetric inequalities for k-Hessian equations, Adv. Nonlinear Anal. 1 (2012), 181–203. DOI 10.1515/anona-2011-0006 © de Gruyter 2012.10.1515/anona-2011-0006Search in Google Scholar

[11] R.C. Reilly, On the Hessian of a function and the curvatures of its graph, Michigan Math. J. 20 (1973), 373–383.10.1090/pspum/027.1/9926Search in Google Scholar

[12] R.C. Reilly, Variational properties of functions of the mean curvatures for hypersurfaces in space forms, J. Differential Geometry, 8 (1973), 465–477.10.4310/jdg/1214431802Search in Google Scholar

[13] J. Serrin, A symmetry problem in potential theory, Arch. Rat. Mech, Anal. 43, (1971), 304–318.10.1007/BF00250468Search in Google Scholar

[14] G. Talenti, Some estimates of solutions to Monge-Ampère type equations in dimension two, Ann. Scuola Norm. Pisa Cl Sci. 8 (4) (1981), no. 2, 183–230.Search in Google Scholar

[15] N.S. Trudinger, Isoperimetric inequalities for quermassintegrals, Ann. Inst. H. Poincaré Anal. non linéaire, 11 (1994), 411–425.10.1016/s0294-1449(16)30181-0Search in Google Scholar

[16] N.S. Trudinger, On new isoperimetric inequalities and symmetrization, J. Reine Angew. Math., 488 (1997), 203–220.10.1515/crll.1997.488.203Search in Google Scholar

[17] N.S. Trudinger, Xu-Jia Wang, Hessian measures II, Ann. of Math. 150 (1999), 579–604.10.2307/121089Search in Google Scholar

[18] K. Tso, On symmetrization and Hessian equations, J. Analyse Math., 52 (1989), 94–106.10.1007/BF02820473Search in Google Scholar

[19] K. Tso, On a real Monge-Ampère functional, Invent. Math., 101 (1990), 425–448.10.1007/BF01231510Search in Google Scholar

[20] H.F. Weinberger, Remark on the preceeding paper of Serrin, Arch. Rat. Mech. Anal. 43 (1971), 319–320.10.1007/BF00250469Search in Google Scholar

[21] Ma Xi-Nan, A necessary condition of solvability for the capillary boundary of Monge-Ampère equations in two dimensions, Proc. Amer. Math. Soc. 233, (1999), 257–266.Search in Google Scholar

Received: 2021-08-18
Accepted: 2022-01-14
Published Online: 2022-02-26

© 2022 Ahmed Mohammed et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Editorial
  3. A sharp global estimate and an overdetermined problem for Monge-Ampère type equations
  4. Non-degeneracy of bubble solutions for higher order prescribed curvature problem
  5. On fractional logarithmic Schrödinger equations
  6. Large solutions of a class of degenerate equations associated with infinity Laplacian
  7. Chemotaxis-Stokes interaction with very weak diffusion enhancement: Blow-up exclusion via detection of absorption-induced entropy structures involving multiplicative couplings
  8. Asymptotic mean-value formulas for solutions of general second-order elliptic equations
  9. Weighted critical exponents of Sobolev-type embeddings for radial functions
  10. Existence and asymptotic behavior of solitary waves for a weakly coupled Schrödinger system
  11. On the Lq-reflector problem in ℝn with non-Euclidean norm
  12. Existence of normalized solutions for the coupled elliptic system with quadratic nonlinearity
  13. Normalized solutions for a class of scalar field equations involving mixed fractional Laplacians
  14. Multiplicity and concentration of semi-classical solutions to nonlinear Dirac-Klein-Gordon systems
  15. Multiple solutions to multi-critical Schrödinger equations
  16. Existence of solutions to contact mean-field games of first order
  17. The regularity of weak solutions for certain n-dimensional strongly coupled parabolic systems
  18. Uniform stabilization for a strongly coupled semilinear/linear system
  19. Existence of nontrivial solutions for critical Kirchhoff-Poisson systems in the Heisenberg group
  20. Existence of ground state solutions for critical fractional Choquard equations involving periodic magnetic field
  21. Least energy sign-changing solutions for Schrödinger-Poisson systems with potential well
  22. Lp Hardy's identities and inequalities for Dunkl operators
  23. Global well-posedness analysis for the nonlinear extensible beam equations in a class of modified Woinowsky-Krieger models
  24. Gradient estimate of the solutions to Hessian equations with oblique boundary value
  25. Sobolev-Gaffney type inequalities for differential forms on sub-Riemannian contact manifolds with bounded geometry
  26. A Liouville theorem for the Hénon-Lane-Emden system in four and five dimensions
  27. Regularity of degenerate k-Hessian equations on closed Hermitian manifolds
  28. Principal eigenvalue problem for infinity Laplacian in metric spaces
  29. Concentrations for nonlinear Schrödinger equations with magnetic potentials and constant electric potentials
  30. A general method to study the convergence of nonlinear operators in Orlicz spaces
  31. Existence of ground state solutions for critical quasilinear Schrödinger equations with steep potential well
  32. Global existence of the two-dimensional axisymmetric Euler equations for the Chaplygin gas with large angular velocities
  33. Existence of two solutions for singular Φ-Laplacian problems
  34. Existence and multiplicity results for first-order Stieltjes differential equations
  35. Concentration-compactness principle associated with Adams' inequality in Lorentz-Sobolev space
Downloaded on 12.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2022-0001/html
Scroll to top button