Skip to content
BY 4.0 license Open Access Published by De Gruyter December 7, 2017

Superposition of p-superharmonic functions

  • Karl K. Brustad EMAIL logo

Abstract

The dominative p-Laplace operator is introduced. This operator is a relative to the p-Laplacian, but with the distinguishing property of being sublinear. It explains the superposition principle in the p-Laplace equation.

MSC 2010: 35J15; 35J60; 35J70

1 Introduction

The p-Laplace equation

(1.1)Δpu:=div(|u|p-2u)=0

is the Euler–Lagrange equation for the variational integral Ω|u|pdx. When p=2, we have Dirichlet’s integral and Laplace’s equation Δu=0. For p= we define

Δu:=i,j=1n2uxixjuxiuxj=0.

The object of our work is a superposition principle, originally discovered by Crandall and Zhang in [5].

Although the p-Laplace equation is nonlinear when p2, a principle of superposition for the fundamental solutions (note the sign change of the scaling constant as p crosses n)

(1.2)wn,p(x):={-p-1p-n|x|p-np-1,pn,-ln|x|,p=n,-|x|,p= or n=1,

is valid. That is, if p2 and ci0, then the superposition

(1.3)V(x):=i=1Nciwn,p(x-yi)

satisfies

ΔpV0

away from the poles y1,,yN. Moreover, the sum is p-superharmonic in the whole of n according to Definition 3 in Section 4.[1] In [1] an explicit formula for ΔpV(x) was derived. There it was shown that an arbitrary concave function also may be added to the sum (1.3). The result can be extended to infinite sums, and via Riemann sums one obtains that the potentials

V(x)=nρ(y)wn,p(x-y)dy,ρ0,

are p-superharmonic functions, provided that V(x). See [14] and [6].

It has been a little mystery why the sum (1.3) is p-superharmonic. It has not been clear what the underlying reasons are, or how far the superposition could be extended. It turns out that a class of functions called dominative p-superharmonic functions plays a central rôle in these questions. We introduce them through the sublinear operator

(1.4)𝒟pu:=λ1++λn-1+(p-1)λn,p2,

where λ1λn are the eigenvalues of the Hessian matrix

u:=(2uxixj)i,j=1n.

The fundamental solutions are members of this class and for C2 functions we have:

Proposition.

Let uC2(Ω). Then

𝒟pu0Δpu0in Ω.

In general, the inequality 𝒟pu0 must be interpreted in the viscosity sense, see Section 4. As we shall see, the superposition principle is governed by the equation 𝒟pu0.

Needless to say, the eigenvalues of u have been much studied. In [16] the equation λ1=0 is found. The related equation λn=0 is produced by the dominative operator in the limit p:

1p𝒟puλn=:𝒟u.

The supersolutions 𝒟u0 are, in fact, the concave functions. We also mention the papers [11] and [18] where symmetric functions of the eigenvalues are investigated. So far as we know, the dominative p-Laplace equation is new for p.

Our main results are Theorem 1 and Theorem 2 below. Theorem 1 gives sufficient and necessary conditions for a sum to be p-superharmonic. In short: a generic sum is p-superharmonic if and only if its terms are dominative p-superharmonic functions. Furthermore, it will be demonstrated that one cannot expect a sum of p-harmonic functions (like (1.3)) to be p-superharmonic unless the functions involved have a high degree of symmetry (Definition 1).

In Theorem 2 we extend the superposition principle for the fundamental solutions to arbitrary radial p-superharmonic functions. Its proof is obtained by showing that important properties of the dominative p-Laplace operator in the smooth case, also hold in the viscosity sense.

Throughout the paper, we restrict ourselves to the case 2p and n2. An open subset of n is denoted by Ω.

Theorem 1.

Let 2p. The following conditions hold pointwise in Rn.

  1. Let u1,,uN be dominative p-superharmonic C2 functions. Then

    Δp[i=1Nui]0.
  2. Let u be C2. Then the following claims are equivalent.

    1. For every linear function l(x)=aTx,an,

      Δp[u+l]0.
    2. For all constants c0 and every translation T(x)=x-x0,

      Δp[u+cwn,pT]0.
    3. For every isometry T:nn,

      Δp[u+uT]0.
    4. u is a dominative p-superharmonic function.

    If u, in addition, is p-harmonic and 2<p<, then

    1. u is locally a cylindrical fundamental solution (see Definition 1).

Theorem 2.

Let 2p and let u1,,uN be radial p-superharmonic functions in Rn. Then the sum

i=1Nui(x-yi)+K(x),yin,

is p-superharmonic in Rn for any concave function K.

To keep things simple, we have not extended Theorem 2 to cover cylindricalp-superharmonic functions.

A function f in n is radial if there exists a one-variable function F so that f(x)=F(|x|). As usual,

|x|:=x12++xn2

denotes the Euclidean norm of x. An equivalent definition is the symmetry condition f(Qx)=f(x) for every n×n orthogonal matrix Q. Radial functions have concentric spherical level-sets, and so do the translated ones f(x-x0). We generalize this notion to functions having concentric cylindrical level-sets:

Definition 1.

A function f in n is cylindrical (or k-cylindrical) if there exists a one-variable function F, an integer 1kn, and an n×k matrix Q with orthonormal columns, i.e. QTQ=Ik, so that

f(x)=F(|QT(x-x0)|)

for some x0n.

We say that a function u in n is a cylindrical fundamental solution (to the p-Laplace equation) if u is in the form

u(x)=C1wk,p(QT(x-x0))+C2,C10,

for some k, Q and x0 as above, and where wk,p is given by (1.2).

Notice that a 1-cylindrical fundamental solution

u(x)=-C1|qT(x-x0)|+C2=aTx+b

is affine in the regions where it is differentiable, while an n-cylindrical fundamental solution

u(x)=C1wn,p(QT(x-x0))+C2=C1wn,p(x-x0)+C2

is a translated radial function.

A calculation shows (Proposition 1, part b) that the cylindrical fundamental solutions solve the p-Laplace equation, except on the (n-k)-dimensional affine space {x:QT(x-x0)=0}, where the functions become singular. When p<, a Dirac delta is produced. For example, setting Q=(𝐞1,,𝐞k), x0=0 and splitting x=(y,z)n in yk, zn-k yields

nΔpwk,p(QTx)ϕ(x)dx=n-kkΔpwk,p(y)ϕ(y,z)dydz
=-Cn-kϕ(0,z)dz,C=C(k,p)>0,

for every ϕC0(n). Moreover, we shall see that these solutions have an essential property that is not shared by any other p-harmonic function:

Property.

The gradient of a cylindrical fundamental solution is an eigenvector corresponding to the largest eigenvalue of the Hessian matrix.

2 The dominative p-Laplace operator

2.1 Preliminary basics and notation

For a function uC2(Ω), we denote by u=u(x) the Hessian matrix of u at x. We list some elementary and useful facts about this matrix.

  1. u is a symmetric n×n matrix: uT=u.

  2. u has n real eigenvalues, which we label in increasing order:

    λ1λn.

    The largest eigenvalue, λn, has special importance and is denoted by λu to indicate its origin. Sometimes we are inconsistent with the notation and write λX for the largest eigenvalue of a symmetric matrix X.

  3. The eigenvectors, ξ1,,ξn, of u can be chosen to be orthonormal: ξiTξj=δij. A unit eigenvector corresponding to the largest eigenvalue λu is labeled ξu.

  4. tru=ux1x1++uxnxn=λ1++λn=Δu. In general

    Δu=i=1nziTuzi

    for every orthonormal set {z1,,zn}n.

  5. We have for any vector zn,

    λ1|z|2zTuzλu|z|2,

    and

    λu=max|z|=1zTuz.

    Conversely, if zn, |z|=1, satisfies

    zTuz=λu,

    then

    uz=λuz.

We adapt the convention that vectors/points in space are column vectors, except gradients which are to be read as row vectors.

2.2 Definition and fundamental properties

Definition 2.

We define the Dominative[2]p-Laplace operator, 𝒟p, as

𝒟pu:={(p-2)λu+Δu,when 2p<,λu,when p=.

A C2 function u is dominative p-superharmonic if 𝒟pu0 at each point in its domain.

The expression (1.4) is, of course, an alternative representation when p<. Observe that 𝒟2=Δ2=Δ.

Remark 1.

The operator can be written as

𝒟pu=max|ξ|=1trAp(ξ)u,Ap(ξ):=(p-2)ξξT+I,

and is, in this form, sublinear and degenerate elliptic all the way down to p=1. For p2 it reads

𝒟pu=(p-2)λ1+Δu.

Nevertheless, to avoid complicating the exposition (e.g., Proposition 1 (3) will not hold if p<2), we shall let p2 throughout the paper.

In low dimensions it is possible to express 𝒟pu in terms of the second-order partial derivatives uxixj. In 2 it can be calculated to be

𝒟pu=p2(uxx+uyy)+p-22(uxx-uyy)2+4uxy2,
𝒟u=12(uxx+uyy)+12(uxx-uyy)2+4uxy2.

We clearly see the nonlinearity introduced when p>2.

The motivation behind Definition 2 came from the following observations: By carrying out the differentiation in (1.1), we arrive at the identity

(2.1)Δpu=|u|p-2((p-2)uuuT|u|2+Δu).

The normalized variant of the -Laplacian appearing in (2.1) satisfies

(2.2)ΔNu:=Δu|u|2=uuuT|u|2λu=𝒟u.

Thus the normalized p-Laplacian also satisfies

(2.3)ΔpNu:=Δpu|u|p-2=(p-2)ΔNu+Δu(p-2)λu+Δu=𝒟pu

when 0p-2<. When u is a fundamental solution, we have equality in (2.2) and (2.3). Since 𝒟p is sublinear and invariant under translations, a very simple proof of the superposition principle for the fundamental solutions (1.3) is produced. However, the above calculations are, for the moment, not valid at the poles or at critical points.

Proposition 1 (Fundamental properties of Dp. Smooth case).

Let u,vC2(Ω). Then the following holds pointwise for 2p.

  1. Domination: We have

    Δpu{|u|p-2𝒟pu,2p<,|u|2𝒟u,p=.

  2. Sublinearity: We have

    1. 𝒟p[u+v]𝒟pu+𝒟pv, and

    2. 𝒟p[αu]=α𝒟pu, α0.

  3. Cylindrical Equivalence: Let 1kn and let u be k-cylindrical.

    1. Assume that the corresponding one-variable function U=U(r) satisfies UrU′′. If k<n, we also require that U′′0. Then

      (2.4)Δpu={|u|p-2𝒟pu,2p<,|u|2𝒟u,p=.

    2. If u is a k-cylindrical fundamental solution, then

      𝒟pu=0=Δpu.

    3. If 2<p<, k2, and 𝒟pu=0=Δpu, then u is a k-cylindrical fundamental solution.

  4. Nesting Property: The following hold:

    1. If 𝒟pu0, then 𝒟qu0 for every 2qp.

    2. If 𝒟pu0, then 𝒟qu0 for every pq.

  5. Invariance: We have

    𝒟p[uT]=(𝒟pu)T

    in T-1(Ω) for all isometries T:nn.

Proof.

We prove (1)–(5) separately.

(1) Domination: Since

Δu=uuuT|u|2λu=|u|2𝒟u,

we also get, when u0,

Δpu=|u|p-2((p-2)ΔNu+Δu)
|u|p-2((p-2)λu+Δu)
=|u|p-2𝒟pu

for 0p-2<. If u=0 or p=2, the claim is trivial: The p-Laplacian is zero at critical points when p>2.

(2) Sublinearity: Since

𝒟[u+v]=λu+v
=ξu+vT[u+v]ξu+v
=ξu+vT(u+v)ξu+v
=ξu+vTuξu+v+ξu+vTvξu+v
λu+λv
=𝒟u+𝒟v,

we also get

𝒟p[u+v]=(p-2)λu+v+Δ[u+v]
(p-2)(λu+λv)+Δu+Δv
=𝒟pu+𝒟pv

for 0p-2<. Also, if α0 and λu is the largest eigenvalue of u, then αλu is the largest eigenvalue of [αu]=αu. Thus λαu=αλu. This means that

𝒟[αu]=λαu=αλu=α𝒟u

and

𝒟p[αu]=(p-2)λαu+Δ[αu]=α𝒟pu

for p<.

(3) Cylindrical Equivalence: (a) Let 1kn, x0n, let Q be an n×k matrix with QTQ=Ik, and let

u(x)=U(|QT(x-x0)|).

Write y:Ωnk,

y(x):=QT(x-x0).

Assume first that u is C2 at a point x1Ω, where y(x1)=0. Then, for a number h and vector q, we have y(x1+hq)=hQTq and

±u(x1)q=limh0u(x1±hq)-u(x1)h
=limh0U(|hQTq|)-U(0)h,

which is independent of the sign. Therefore, u(x1)=0 and the equality in (2.4) is trivial.

The Jacobian matrix of y is Dy=QT, and

|y|=yT|y|Dy=y^TQT,y^:=y|y|.

Moreover,

1|y|=-yT|y|3Dy

so

Dy^=Dy|y|+y1|y|=1|y|(Dy-yyT|y|2Dy)=1|y|(QT-y^y^TQT).

Now, u(x)=U(|y(x)|) and

u=U|y|=Uy^TQT.

The Hessian matrix is

u=D[uT]
=D[UQy^]
=Qy^U′′|y|+UQDy^
=U′′Qy^y^TQT+U|y|Q(QT-y^y^TQT)
=Q{U′′y^y^T+U|y|(Ik-y^y^T)}QT

with trace

Δu=U′′+(k-1)U|y|.

There are n-k perpendicular constant eigenvectors in the null-space of QT with zero eigenvalues. The (transposed) gradient is in the column-space of Q and is an eigenvector:

uuT=Q{U′′y^y^T+U|y|(Ik-y^y^T)}QT(UQy^)
=UQ{U′′y^y^T+U|y|(Ik-y^y^T)}y^
=UQ{U′′y^+0}
=U′′uT.

Finally, there are k-1 eigenvectors ξ=ξ(x)n in the column-space of Q that are perpendicular to u, i.e. y^TQTξ=0 and ξ=Qξ~ for some ξ~k:

uξ=Q{U′′y^y^T+U|y|(Ik-y^y^T)}QTξ
=Q{0+U|y|(QTξ-0)}
=U|y|QQTξ
=U|y|QQTQξ~=U|y|Qξ~=U|y|ξ.

Thus the n eigenvalues of u are U|y| with multiplicity k-1, 0 with multiplicity n-k and U′′ with multiplicity 1.

By the assumption UrU′′ and U′′0, if k<n, it is clear that the largest eigenvalue is λu=U′′ and it follows that

Δu=uuu=|u|2U′′=|u|2𝒟u

and

Δpu=|u|p-2((p-2)uuuT|u|2+Δu),u0,
=|u|p-2((p-2)λu+Δu)
=|u|p-2𝒟pu.

Again, the equality is trivial if u=0.

(b) Now assume that u(x)=U(|y|) is a C2k-cylindrical fundamental solution:

U(r)={-C1p-1p-krp-kp-1+C2,pk,-C1lnr+C2,p=k,-C1r+C2,p= or k=1,

with C10, C2. Then

U(r)={-C1r1-kp-1,2p<,-C1,p=,

and

U′′(r)={-C11-kp-1r2-p-kp-1,2p<,0,p=.

We see that U′′0Ur in every case, so 𝒟u=λu=U′′=0 if p=, and

Δu=|u|2𝒟u=0.

Also,

𝒟pu=(p-2)U′′+U′′+(k-1)U|y|
=(p-1)U′′+(k-1)U|y|
=-C1((1-k)|y|2-p-kp-1+(k-1)|y|1-kp-1-1)=0

and

Δpu=|u|p-2𝒟pu=0

when p<.

(c) Finally, assume that u is k-cylindrical, 2kn, 2<p< and Δpu=0=𝒟pu. The ODE for U produced by Δpu=0 is

(2.5)0=ΔpNu=(p-1)U′′+(k-1)Ur

with general solution U satisfying

U(r)=Cr-k-1p-1,C.

By the definition of the cylindrical fundamental solutions, we only need to show that C0. The equation 𝒟pu=0 gives

(2.6)0=(p-2)max{U′′,Ur}+U′′+(k-1)Ur,k=n,
(2.7)0=(p-2)max{U′′,Ur,0}+U′′+(k-1)Ur,2k<n.

Subtract (2.5) from (2.6) or (2.7) and divide by p-2 to obtain the condition

0=max{U′′,Ur}-U′′.

That is,

-Ck-1p-1r-k+p-2p-1=U′′Ur=Cr-k+p-2p-1

which is true only if C0.

(4) Nesting Property: Let 2qp and assume 𝒟pu0. For each xΩ we consider two cases. If 𝒟u=λu<0, then every eigenvalue is negative and

𝒟qu=λ1++λn-1+(q-1)λu<0.

If 𝒟u=λu0, then also

𝒟qu=(q-2)λu+Δu(p-2)λu+Δu=𝒟pu0.

Now let 2pq and assume

𝒟pu=(p-1)λu+λ1++λn-10.

Then, obviously, λu0 and

𝒟qu=(q-2)λu+Δu(p-2)λu+Δu=𝒟pu0.

(5) Invariance: An isometry in n is on the form T(x)=QT(x-x0) for some x0n and some constant orthogonal n×n matrix Q: QTQ=I. Define v(x):=u(T(x)) on T-1(Ω), i.e. T(x)Ω. Write y:=T(x). Then v(x)=u(y)QT and

v(x)=Qu(y)QT.

So λv(x)=λu(y), proving the case p=, since

λv(x)=max|z|=1zTv(x)z=max|z|=1zTQu(y)QTz=max|z|=1zTu(y)z=λu(y).

Also

Δv(x)=trv(x)=trQu(y)QT=tru(y)=Δu(y).

Therefore

(𝒟p[uT])(x)=(𝒟pv)(x)
=(p-2)λv(x)+Δv(x)
=(p-2)λu(y)+Δu(y)
=(𝒟pu)(T(x)).

3 Proof of Theorem 1

That dominative p-superharmonicity of the terms is a sufficient condition for the sum to be p-superharmonic is now an immediate consequence of Proposition 1:

Proof of Theorem 1 (i).

Let 2p and let u1,,uN be dominative p-superharmonic C2 functions. That is, 𝒟pui0. Write

V(x):=i=1Nui(x)

to denote the sum. Then V is C2 and

ΔpV|V|p-2𝒟p[i=1Nui](by Proposition 1 (1))
|V|p-2i=1N𝒟pui(by Proposition 1 (2))
0.

The calculations are the same when p=. ∎

Notice that a sum of fundamental solutions, V(x)=i=1Nciwn,p(x-yi) in a domain not containing the singularities, is just a special case by (3) and (5) of Proposition 1.

We restate and prove the first part of Theorem 1 (ii).

Proposition 2.

Let 2p and let uC2(Ω). Then the following properties are equivalent.

  1. u(x)+aTx is p-superharmonic for every linear function aTx.

  2. u(x)+cwn,p(x-y) is p-superharmonic in Ω{y} for every c0 and every yn.

  3. u+uT is p-superharmonic in ΩT-1(Ω) for every isometry T:Ωn.

  4. u is a dominative p-superharmonic function.

Proof.

We show

The upward implications are immediate from the fundamental properties of 𝒟p. As for the downward implications, assume that u is not dominative p-superharmonic. Then there is a point x0Ω so that 𝒟pu(x0)>0.

(a)  (d) By the contrapositive assumption 𝒟p(x0)>0, the implication is proved if we can find a constant an so that Δp[u+aTx]>0 at x0. To this end, let ξu=ξu(x0) be a unit eigenvector of u(x0) corresponding to the largest eigenvalue λu and let v(x):=aTx be the linear function with

a:=ξu-uT(x0).

Then, at x0,

u+v=u+aT=ξuT,

so |(u+v)|=1 and

Δ[u+v]=ΔN[u+v]
=(u+v)(u+v)(u+v)T
=ξuT(u+0)ξu
=λu
=𝒟u(x0)>0

if p=. Also, if p<,

Δp[u+v]=ΔpN[u+v]
=(p-2)ΔN[u+v]+Δ[u+v]
=(p-2)λu+Δu+0
=𝒟pu(x0)>0.

(b)  (d) The implication is proved if we can find a yn and a c0 so that

Δp[u(x)+cwn,p(x-y)]>0

at x=x0. To this end, let ξu=ξu(x0) be a unit eigenvector of u(x0) corresponding to the largest eigenvalue λu, and denote by q:=uT(x0) the gradient of u at x0. The idea is to consider a fundamental solution with centre far away from x0 in the proper direction, and then scale it in order to achieve a convenient cancellation in the sum of the gradients.

Introduce a (large) parameter s and let the centre of the scaled fundamental solution

fs(x):=cswn,p(x-ys)

be at ys:=x0-q+sξu. Let zs be the point

zs:=x0-ys=q-sξu.

Then

fs(x0)=cswn,p(zs)=csWn,p(|zs|)zsT|zs|,Wn,p(|x|):=wn,p(x),

which equals -zsT=-(q-sξu)T if we choose the scale cs to be

cs:=-|zs|Wn,p(|zs|)={|q-sξu|n+p-2p-1,2p<,|q-sξu|,p=.

We may read the fraction n+p-2p-1 as 1 if p=.

We now get, at x0, u+fs=qT-(q-sξu)T=sξuT and

ΔpN[u+fs]=(p-2)(u+fs)(u+fs)(u+fs)T|u+fs|2+Δu+Δfs
=(p-2)ξuT(u+fs)ξu+Δu+Δfs
=𝒟pu+(p-2)ξuTfsξu+Δfs

if p< and

ΔN[u+fs]=𝒟u+ξuTfsξu

if p=. Since 𝒟pu(x0)>0 and does not depend on s, we finish the proof by making the remaining term(s) arbitrarily close to zero.

The Hessian matrix of the fundamental solution is

wn,p(x)=-Wn,p(|x|)|x|(n+p-2p-1xxT|x|2-I),n+-2-1:=1,

so at zs we get

fs(x0)=cswn,p(zs)=-csWn,p(|zs|)|zs|(n+p-2p-1zszsT|zs|2-I)=n+p-2p-1zszsT|zs|2-I,
Δfs(x0)=n+p-2p-1-n.

Thus, when p<,

(p-2)ξuTfsξu+Δfs=(p-2)(n+p-2p-1(zsTξu)2|zs|2-1)+n+p-2p-1-n
=(p-2)n+p-2p-1((zsTξu)2|zs|2-1)0

as s since

lims(zsTξu)2|zs|2=lims(qTξu-s)2|q-sξu|2=1.

Likewise, when p=,

ξuTfsξu=(zsTξu)2|zs|2-10.

To summarize: If 𝒟pu(x0)>0 and if s is large enough, then for

zs:=uT(x0)-sξu(x0),fs(x):=-|zs|Wn,p(|zs|)wn,p(x-x0+zs)

we have[3]

Δp[u+fs](x0)=sαpΔpN[u+fs](x0)>0

and the sum u+fs is not p-superharmonic.

(c)  (d) Without loss of generality we may assume x0 to be the origin. i.e. 0Ω and 𝒟pu(0)>0. We shall prove the implication by finding an isometry T and a point y0Ω, equal or close to 0, so that T(y0)Ω and Δp[u+uT]>0 at y0. To this end, let yn, |y|=1 be a fixed direction in space defining the line

:={αy:α}.

The projection onto is given by the 1-rank matrix

P:=yyT.

We have, as for every projection,

PxandPP=P.

The reflection about is now given by Rx:=Px-(x-Px). That is,

R=2P-I.

A reflection satisfies

R|=idandRR=I.

After carefully choosing y, T(x):=Rx will be our isometry.

Define the superposition

V(x):=u(x)+u(Rx)2.

The main idea of the proof is that, on , V will be pointing in the y-direction: The chain rule gives V(x)=12(u(x)+u(Rx)R) and V(x)=12(u(x)+Ru(Rx)R) and ΔV(x)=12(Δu(x)+Δu(Rx)). For x we have Rx=x, and

V=uI+R2=uP,
V=u+RuR2,
ΔV=Δu.

This gives, when V=uP0,

ΔpNV|=(p-2)VVVT|V|2+ΔV
=(p-2)1|uP|2uPu+RuR2PuT+Δu
=(p-2)uPuPuT|uP|2+Δu
=(p-2)yTuy+Δu

since RP=P and 0PuT is parallel to y. Similarly

ΔNV|=yTuy.

Now choose y:=ξu, where ξu=ξu(0) is a unit eigenvector of u(0) corresponding to the largest eigenvalue λu. The isometry T is then

T(x):=Rx=(2P-I)x=(2ξuξuT-I)x.

Since 0, and unless u(0)ξu=0, it follows that

12ΔpN[u+uT]|x=0=ΔpNV(0)=𝒟pu(0)>0

and Δp[u+uT]=2αp+1|uP|αpΔpN[u+uT]>0 at x=0 since u(0)P=u(0)ξuξuT0.

If u(0)ξu=0, we complete the proof with a continuity argument: Since λu(0)>0,[4]𝒟pu(0)>0 and Ω is open, and since the Hessian is continuous, there must be a common ϵ>0 so that

  1. ξuTu(tξu)ξu>0,

  2. (p-2)ξuTu(tξu)ξu+Δu(tξu)>0 if p<,

  3. T(tξu)=tξuΩ for all t[0,ϵ].

A Taylor expansion of the gradient about 0 in the ξu-direction then gives

u(ϵξu)=u(0)+ϵξuTu(t0ξu)

for some t0[0,ϵ]. So

u(ϵξu)ξu=0+ϵξuTu(t0ξu)ξu>0,

and again, since ϵξu,

12Δp[u+uT]|x=ϵξu=|2u(ϵξu)P|p-2((p-2)yTu(ϵξu)y+Δu(ϵξu))
=(2u(ϵξu)ξu)p-2((p-2)ξuTu(ϵξu)ξu+Δu(ϵξu))>0

if p< and

12Δ[u+uT]|x=ϵξu=(2u(ϵξu)ξu)2ξuTu(ϵξu)ξu>0

if p=. Thus the sum u+uT is not p-superharmonic in ΩT-1(Ω). ∎

We finish the proof of Theorem 1 by showing the equivalence of (d) and (e). The nontrivial implication is (d)  (e). Namely that if 2<p< and uC2(Ω) is both p-harmonic and dominative p-superharmonic, then u is locally a cylindrical fundamental solution. Since the hypothesis and the domination implies

0=Δpu|u|p-2𝒟pu0,

the claim follows from Proposition 3 below. It is partially the converse of the Cylindrical Equivalence (part (3) of Proposition 1).

Proposition 3.

Let 2<p< and let uC2(Ω).[5] If

(3.1)Δpu=0=𝒟puin Ω,

then u is locally a cylindrical fundamental solution.

The proof relies on a rather deep result in differential geometry. We refer to [3], [19] and [17] for the details of the following exposition.

A nonconstant smooth function u:M on a Riemannian manifold M is called isoparametric if there exist functions f and g so that

12|u|2=f(u)andΔu=g(u).

A regular level-set of an isoparametric function is called an isoparametric hypersurface. The investigation of isoparametric functions and hypersurfaces was originally motivated by the study of wave propagation.

The isoparametric hypersurfaces in the Euclidean case Mn have been classified completely. Apparently, this was first done by Segre (see [17]) in 1938:

Theorem 3 (Segre).

A connected isoparametric hypersurface in Rn is, upon scaling and an Euclidean motion, an open part of one of the following hypersurfaces:

  1. a hyperplane n-1,

  2. a sphere Sn-1,

  3. a generalized cylinder Sk-1×n-k, k=2,,n-1.

Moreover, the family of cylinders u-1(c) is concentric. Thus u is a function of the distance to the common “axis” of the cylinders, the axis being an (n-k)-dimensional affine subspace, k=1,,n, in n. Call this subset 𝒜k.

The axis 𝒜k is isomorphic to n-k,

𝒜kn-k{(000In-k)x:xn}=:𝒜k~,

where, obviously,

dist(x,𝒜k~)=x12++xk2=|(Ik  0)x|,0k×n-k.

Translating and rotating back to 𝒜k via an isometry QnT(x-x0), Qnn×n orthogonal, we find that

u(x)=U(dist(x,𝒜k))=U(|(Ik 0)QnT(x-x0)|)=U(|QkT(x-x0)|),

where Qk is the n×k matrix consisting of the first k columns of Qn.

Thus an isoparametric function is cylindrical.

Proof of Proposition 3.

If u is affine, it can locally be written as a 1-cylindrical fundamental solution. If u is not an affine function, let x0Ω be a point with a neighbourhood ΩΩ where u has connected level-sets and where u0, u0. By the discussion above and part (c) of Proposition 1, it is sufficient to show that u is a smooth isoparametric function.

As u0 and p>2, our equation (3.1)

|u|p-2((p-2)uuuT|u|2+Δu)=0=(p-2)λu+Δu

implies uuuT=λu|u|2 and the gradient is therefore an eigenvector of the Hessian:

uuT=λuuT.

Let 𝐜 be a differentiable curve in a level-set of u. Then 0=ddtu(𝐜(t))=ud𝐜dt and

ddt12|u(𝐜(t))|2=uud𝐜dt=λuud𝐜dt=0.

Thus the length of the gradient is constant on the level-sets and can be written as a function only of u. Say,

(3.2)12|u(x)|2=f(u(x))>0.

We need to show that f is differentiable, because if so, differentiation of (3.2) yields

(3.3)uu=f(u)u

and λu=f(u). Using (3.1) once more, we then find that also Δu is a function of u:

(3.4)Δu=-(p-2)λu=-(p-2)f(u)=:g(u).

Fix xΩ and let 𝐱(t) now be an integral curve of the gradient field starting from x:

d𝐱dt(t)=uT(𝐱(t)),𝐱(0)=x.

Then define the function h as h(t):=u(𝐱(t)). We see that h is C2 and

h(t)=u(𝐱(t))uT(𝐱(t))=2f(u(𝐱(t)))=2f(h(t))>0.

Thus h is strictly monotone and

12h′′(0)h(0)=12limt0h(t)-h(0)h(t)-h(0)=limt0f(h(t))-f(h(0))h(t)-h(0)=f(h(0))=f(u(x)).

This is enough to conclude that (3.3), and thus (3.4), is valid.

As for the regularity of u, observe that if F is an anti-derivative of (2f)p-22 and ϕ(x):=F(u(x)), then ϕ is C2 and

ϕ=F(u)u=|u|p-2u.

That is, ϕ is harmonic by (3.1). It follows that ϕ is real-analytic, and so is u since u=|ϕ|-p-2p-1ϕ. ∎

This concludes the proof of Theorem 1.

It is worth noting that Proposition 3 is not true for p=2 and p=. The case p=2 is obvious since 𝒟2Δ and every harmonic function satisfies (3.1). When p=, a counter-example is provided by a function

u(x)=dist(x,)in Ω,

where is a non-spherical solid ellipse in, say, 2 and where Ω is a small ball sufficiently close to the boundary of the ellipse. Then u is neither affine nor a circular cone (i.e. not a cylindrical -fundamental solution). But u is smooth with |u|=1, so uu=0 and λu=0 is the larger eigenvalue as u is concave (see [7, Lemma 1 and Lemma 2, Appendix]). It follows that

𝒟u=λu=0=uuuT=Δuin Ω.

4 Viscosity solutions

The equation 𝒟pu=0 needs to be interpreted in the viscosity sense (v.s.). We refer to [4], [10] and [9] for the general theory of viscosity solutions. For our purpose, only the basic notions of the concept are needed.

A PDE in the form F(u,u)=0 is said to be degenerate elliptic if for any two symmetric matrices X and Y such that Y-X is positive semi-definite, i.e. XY, we have

F(q,X)F(q,Y)

for all qn.[6]

In our case

0=𝒟pu=Fp(u),

where

Fp(X):={(p-2)λX+trX,p<,λX,p=.

So if zTXzzTYz for all zn, then

λX=ξXTXξXξXTYξXλY

and F(X)F(Y). Also, for any orthonormal set {z1,,zn},

trX=i=1nziTXzii=1nziTYzi=trY,

so Fp(X)Fp(Y) when 0p-2<. Thus the dominative p-Laplace equation 𝒟pu=0 is degenerate elliptic.

It is known that the p-Laplace equation is degenerate elliptic for 2p.

4.1 Definitions and fundamental properties

Consider a degenerate elliptic equation

(4.1)F(u,u)=0.

Definition 3.

We say that u:Ω(-,] is a viscosity supersolution of the PDE (4.1) if the following hold:

  1. u is lower semi-continuous (l.s.c.).

  2. u< in a dense subset of Ω.

  3. If x0Ω and ϕC2 touches u from below at x0, i.e.

    ϕ(x0)=u(x0),ϕ(x)u(x)for x near x0,

    we require that

    F(ϕ(x0),ϕ(x0))0.

The viscosity subsolutions v:Ω[-,) are defined in a similar way: they are upper semi-continuous and the test functions touch from above. Finally, a function w:Ω is a viscosity solution if it is both a viscosity supersolution and a viscosity subsolution. Necessarily, wC(Ω).

We say that u is dominative p-superharmonic if u is a viscosity supersolution to the equation 𝒟pw=0. We write this as 𝒟pu0 v.s. (viscosity sense). The p-superharmonic functions were traditionally defined by the comparison principle and weak integral formulations – and not by viscosity. According to [8], however, the two concepts are equivalent and we may therefore define the p-superharmonic functions in terms of viscosity as well. The comparison principle then becomes a theorem:

Theorem 4 (Comparison Principle).

Let 2p. Assume that v is p-subharmonic and that u is p-superharmonic in Ω. Let DΩ. Then

v|Du|Dvuin D.

Before we extend the fundamental properties of the dominative operator (Proposition 1) to the setting of viscosity, we establish that a dominative -superharmonic function is the same as a concave function:

Proposition 4.

Let ΩRn be open and convex. Then

𝒟u0 v.s. in Ωu is concave in Ω.

This is [15, Proposition 4.1] or, alternatively, [16, Theorem 2.2] in disguise. Note that continuity is automatically given by either direction: It is well known that concave functions are continuous in open domains. Also, if 𝒟u0 v.s., then u is -superharmonic by Proposition 5 and is therefore continuous by [13, Lemma 6.7].

Proposition 5 (Fundamental properties of Dp. Viscosity sense).

The following hold for 2p.

  1. Domination: If u is dominative p-superharmonic, then u is p-superharmonic.

  2. Sublinearity: If 𝒟pu0 and 𝒟pv0 v.s. and u is radial, then 𝒟p[u+v]0 v.s.

  3. Radial Equivalence: If u is a radial p-superharmonic function, then 𝒟pu0 v.s.

  4. Nesting Property: The following hold:

    1. If 𝒟pu0 v.s., then 𝒟qu0 v.s. for every 2qp. In particular, if u is locally concave, then 𝒟qu0 v.s. for all 2q.

    2. If 𝒟pu0 v.s., then 𝒟qu0 v.s. for every pq. In particular, if u is subharmonic, then 𝒟qu0 v.s. for all 2q.

  5. Invariance: If 𝒟pu0 v.s. in n, then 𝒟p[uT]0 v.s. in n for all isometries T:nn.

As shown below, claims (1), (4) and (5) follow easily from the corresponding properties in the smooth case (Proposition 1). The proofs of (2) and (3) are more difficult and are postponed until the survey of Radial p-superharmonic functions in Section 5.

Proof.

As said, we prove (1), (4) and (5).

(1) Domination: Assume 𝒟pu0 v.s. Let x0 be a point in the domain of u, and assume that ϕ is a test function of u from below at x0. Then 𝒟pϕ(x0)0 and

Δpϕ(x0){|ϕ(x0)|p-2𝒟pϕ(x0)0,2p<|ϕ(x0)|2𝒟ϕ(x0)0,p=,

by the smooth case Domination. Hence Δpu0 v.s. and u is p-superharmonic.

(4) Nesting Property: Let 2qp and suppose 𝒟pu0 v.s. Let x0 be a point in the domain of u, and assume that ϕ is a test function of u from below at x0. Then 𝒟pϕ(x0)0 and

𝒟qϕ(x0)0

by the smooth case Nesting Property and 𝒟qu0 v.s. The additional claim follows from Proposition 4.

Let pq and suppose 𝒟pu0 v.s. Let x0 be a point in the domain of u, and assume that ϕ is a test function of u from above at x0. Then 𝒟pϕ(x0)0 and

𝒟qϕ(x0)0

by the smooth case Nesting Property and 𝒟qu0 v.s.

(5) Invariance: Assume 𝒟pu0 v.s. in n and let T:nn be an isometry. Let x0n, and assume that ϕ is a test function for uT from below at x0. Then the function ϕ^:=ϕT-1 is a test function for u from below at y0:=T(x0), since

ϕ^(y0)=ϕT-1(T(x0))=ϕ(x0)=u(y0)

and, for y near y0, T-1(y) is near x0 and

ϕ^(y)=ϕ(T-1(y))(uT)(T-1(y))=u(y).

Thus 𝒟pϕ^(y0)0 and

𝒟pϕ(x0)=(𝒟p[ϕ^T])(x0)
=(𝒟pϕ^)(T(x0))(by the smooth case 𝐼𝑛𝑣𝑎𝑟𝑖𝑎𝑛𝑐𝑒)
=𝒟pϕ^(y0)
0

and 𝒟p[uT]0 v.s. in n. ∎

4.2 The superposition principle for radial p-superharmonic functions

Proposition 5 contains everything we need in order to show that radial p-superharmonic functions can be added and perturbed by a concave function:

Proof of Theorem 2.

Let 2p and let u1,,uN be radial p-superharmonic functions in n. We must show that the sum

i=1Nui(x-yi)+K(x),yin,

is p-superharmonic in n for any concave function K.

Observe first that if 𝒟pu0, 𝒟pv0 v.s., where u is radial, then 𝒟p[vT-1]0 v.s. for every isometry T by the Invariance (5). By the Sublinearity (2) it then follows that 𝒟p[u+vT-1]0 v.s. and

𝒟p[uT+v]=𝒟p[(u+vT-1)T]0 v.s.,

again by (5).

From the Radial Equivalence (3), 𝒟pui0 v.s. for each i=1,,N and also 𝒟pK0 v.s. by the Nesting Property (4). Denoting the translations by Ti(x):=x-yi, and adding the functions uiTi one by one, starting with K, we obtain that

𝒟p[i=1NuiTi+K]0 v.s.

We now use the Domination (1) to conclude that the sum

i=1NuiTi+K

is p-superharmonic in n. ∎

5 Radial p-superharmonic functions

A radial p-superharmonic function u:n(-,] is of the form u(x)=U(|x|) for some l.s.c. one-variable function U:[0,)(-,]. By comparing with constant functions (which are p-harmonic), it is clear that U must be decreasing (= non-increasing). Also, since the set {x:u(x)=} has measure zero [12], the origin is the only possible pole of u. Therefore, u is bounded in every annulus

Aab:=B(0,b)B(0,a)¯,0<a<b<.

Equivalently, U is bounded on the interval (a,b).

Let

wn,p(x)=Wn,p(|x|),2p,n2,

denote the fundamental solution to the p-Laplace equation in n, see (1.2) in the Introduction. In this section, the important properties of the fundamental solution is that any scaled version, C1wn,p+C2 (C1,C2), is still a radial Cp-harmonic function in n{0}. And when C10, it is also dominative p-harmonic by part (b) of Proposition 1.

Furthermore, we shall frequently use the fact that the function Wn,p:[0,)(-,] is strictly decreasing. A simple calculation shows that a scaled fundamental solution is uniquely determined by its values at two different positive radii.

Given a radial p-superharmonic function u(x)=U(|x|) in n and two numbers 0<a<b, we define hab on n{0} as the scaled fundamental solution hab(x)=Hab(|x|), where

(5.1)Hab(r):=Cab[Wn,p(r)-Wn,p(b)]+U(b),
(5.2)Cab:=U(a)-U(b)Wn,p(a)-Wn,p(b).

Figure 1 We have Ha⁢b{H_{ab}}≤U{\leq U} on [a,b]{[a,b]}, while Ha⁢b{H_{ab}}≥U{\geq U} outside the interval.
Figure 1

We have HabU on [a,b], while HabU outside the interval.

The point of this is that hab is p-harmonic, smooth and it satisfies

Hab(a)=U(a)andHab(b)=U(b).

(See Figure 1.) Thus hab=u on the boundary of the annulus Aab and by the comparison principle we must have

habuin Aab.

Equivalently,

HabUin (a,b).

We now deduce other immediate properties of Hab and the scaling constant Cab.

Lemma 1.

Let 2p and let u(x)=U(|x|) be a given radial p-superharmonic function in Rn. For numbers a and b with 0<a<b define the scaled fundamental solution hab(x)=Hab(|x|) with scaling constant Cab as in (5.1) and (5.2).

  1. We have the opposite inequality outside the annulus Aab:

    HabUin (0,a][b,).
  2. 0CabCbc< whenever 0<a<b<c.

  3. The mappings aCab and cCbc are increasing.

  4. The one-sided limits

    Cb-:=limab-Cab,Cb+:=limcb+Cbc

    exist and

    0Cb-Cb+<.

Observe that the existence of the limits (4) implies that U(r) has one-sided derivatives at every r0. For example,

U(a)-U(b)a-b=U(a)-U(b)Wn,p(a)-Wn,p(b)Wn,p(a)-Wn,p(b)a-b

which goes to Cb-Wn,p(b) as ab-.

Proof.

(1) Suppose for the sake of contradiction that there is a number d>b so that Hab(d)<U(d). See Figure 2. The function had(x)=Had(|x|) satisfies

Had(a)=U(a)=Hab(a)(by definition),
Had(d)=U(d)>Hab(d)(by assumption),
Had(b)U(b)=Hab(b)(by the comparison principle).

By the Intermediate Value Theorem, there is an c[b,d) so that Had(c)=Hab(c). Since also Had(a)=Hab(a), the two functions are identical. This is the contradiction to the assumption Had(d)>Hab(d). The proof for 0<r<a is symmetric.

Figure 2 Impossible situation. If there is a d with d>b{d>b} such that Ha⁢b⁢(d){H_{ab}(d)}<U⁢(d){<U(d)}, then there exists a c∈[b,d){c\in[b,d)} so that Ha⁢d⁢(c)=Ha⁢b⁢(c){H_{ad}(c)=H_{ab}(c)}. Thus Ha⁢d≡Ha⁢b{H_{ad}}\equiv{H_{ab}}.
Figure 2

Impossible situation. If there is a d with d>b such that Hab(d)<U(d), then there exists a c[b,d) so that Had(c)=Hab(c). Thus HadHab.

(2) We have

0Cab:=U(a)-U(b)Wn,p(a)-Wn,p(b)<

for all 0<a<b since U is decreasing and Wn,p is strictly decreasing. Moreover,

Hbc(b+ϵ)U(b+ϵ)Hab(b+ϵ)

whenever b+ϵ is between b and c, see Figure 3. The first inequality follows from the comparison principle, and the second inequality follows from (1). Therefore

CabWn,p(b)=Hab(b)
=limϵ0+Hab(b+ϵ)-Hab(b)ϵ
limϵ0+Hbc(b+ϵ)-Hbc(b)ϵ
=Hbc(b)
=CbcWn,p(b)

and CabCbc since Wn,p(b)<0.

Figure 3 Proof of (2): We have Hb⁢c≤Ha⁢b{H_{bc}\leq H_{ab}} on (b,c){(b,c)} soHb⁢c′⁢(b)≤Ha⁢b′⁢(b){H_{bc}^{\prime}(b)\leq H_{ab}^{\prime}(b)}.
Figure 3

Proof of (2): We have HbcHab on (b,c) soHbc(b)Hab(b).

Figure 4 Proof of (3): We have Ha′⁢b≥Ha⁢b{H_{a^{\prime}b}\geq H_{ab}} on (0,b){(0,b)} soHa′⁢b′⁢(b)≤Ha⁢b′⁢(b){H_{a^{\prime}b}^{\prime}(b)\leq H_{ab}^{\prime}(b)}.
Figure 4

Proof of (3): We have HabHab on (0,b) soHab(b)Hab(b).

(3) Let 0<a<a<b. By the comparison principle,

Hab(a)=U(a)Hab(a).

Since Hab(b)=Hab(b), the two functions are either identical or Hab>Hab on (0,b). It follows that

CabWn,p(b)=Hab(b)
=limϵ0+Hab(b-ϵ)-Hab(b)-ϵ
limϵ0+Hab(b-ϵ)-Hab(b)-ϵ
=Hab(b)
=CabWn,p(b)

and CabCab since Wn,p(b)<0. The proof for b<c<c is symmetric. (See Figure 4.)

(4) The claims in (4) are immediate consequences of (2) and (3). ∎

We are now able to reveal a crucial fact about radial p-superharmonic functions: They have a smooth p-harmonic test function touching from above at every finite value.[7]

Lemma 2.

Let 2p. Moreover, let u be a radial p-superharmonic function and let x0Rn. If u(x0)<, then there exists a fundamental solution

h(x):=C1wn,p(x)+C2,C10,

touching u from above at x0:

h(x0)=u(x0)  𝑎𝑛𝑑  h(x)u(x) near x0.

Proof.

If x0=0 and u is bounded at the origin, the constant function h(x)u(0) will do.

Write u(x)=U(|x|) and b:=|x0|>0. Let Cb[Cb-,Cb+], where the end-points of the, possibly singleton but non-empty, interval are defined in part (4) of Lemma 1. We claim that the scaled fundamental solution h(x)=H(|x|) given by

H(r):=Cb[Wn,p(r)-Wn,p(b)]+U(b)

touches u from above at x0.

Obviously, H(b)=U(b) and if 0<a<b, then CabCb by parts (3) and (4) of Lemma 1 and

H(a)=Cb[Wn,p(a)-Wn,p(b)]+U(b)
Cab[Wn,p(a)-Wn,p(b)]+U(b)  (Wn,p(a)-Wn,p(b)>0)
=Hab(a)=U(a).

If b<c, then similarly CbCbc and

H(c)=Cb[Wn,p(c)-Wn,p(b)]+U(b)
Cbc[Wn,p(c)-Wn,p(b)]+U(b)  (Wn,p(c)-Wn,p(b)<0)
=U(c).

The lemma is proved. ∎

Observe that H is uniquely determined if and only if U is differentiable at r=b. That is, if and only if Cb+=Cb-.

With these new tools, we restate and prove the Radial Equivalence (3) and the Sublinearity (2) of Proposition 5.

Proposition 6 (Radial Equivalence).

Let 2p. If u is a radial p-superharmonic function in Rn, then

𝒟pu0in n

in the viscosity sense.

Proof.

Let x0n and assume that u has a test function ϕ from below at x0. We need to show 𝒟pϕ(x0)0. Obviously, u(x0)<. By Lemma 2 there is a scaled fundamental solution h touching u from above at x0. Thus,

ϕ(x0)=u(x0)=h(x0)

and, for x close to x0,

ϕ(x)u(x)h(x),

which implies the Hessian matrix inequality ϕ(x0)h(x0). It follows that

𝒟pϕ(x0)𝒟ph(x0)=0

from the fact that 𝒟p is degenerate elliptic and from the smooth case Cylindrical Equivalence. ∎

Proposition 7 (Sublinearity).

Let 2p. Assume Dpu0 and Dpv0 in the viscosity sense in Rn, where u is radial. Then

𝒟p[u+v]0

in the viscosity sense in Rn.

Proof.

Let x0n and assume that ϕ is a test function to u+v from below at x0. Clearly, u(x0)< since otherwise u+v= at x0 and there would be no test function there. Also, u is p-superharmonic by the Domination (1) of Proposition 5. Hence, by Lemma 2, there exists a scaled fundamental solution h touching u from above at x0:

h(x0)=u(x0),u(x)h(x)near x0.

Again, 𝒟ph=0 by the smooth case Cylindrical Equivalence.

Define ψ(x):=ϕ(x)-h(x). Then ψ is C2 and

ψ(x0)=u(x0)+v(x0)-u(x0)=v(x0)

and

ψ(x)u(x)+v(x)-u(x)=v(x)near x0,

so ψ is a test function for v from below at x0. This means that 𝒟pψ(x0)0, and it follows that

𝒟pϕ=𝒟p[ψ+h]𝒟pψ+𝒟ph0

at x0 by the smooth case Sublinearity. Hence,

𝒟p[u+v]0

in the viscosity sense in n. ∎

The proof of Proposition 5, and hence Theorem 2, is now completed.


Communicated by Juan Manfredi


Acknowledgements

I thank Peter Lindqvist for useful discussions and for naming the operator. I thank Fredrik Arbo Høeg for checking calculations. Also, I thank Juan Manfredi for pointing out the work [16].

References

[1] K. K. Brustad, Superposition in the p-Laplace equation, Nonlinear Anal. 158 (2017), 23–31. 10.1016/j.na.2017.04.004Search in Google Scholar

[2] L. A. Caffarelli and X. Cabré, Fully Nonlinear Elliptic Equations, Amer. Math. Soc. Colloq. Publ. 43, American Mathematical Society, Providence, 1995. 10.1090/coll/043Search in Google Scholar

[3] T. E. Cecil and P. J. Ryan, Isoparametric Hypersurfaces, Geometry of Hypersurfaces, Springer, New York (2015), 85–184. 10.1007/978-1-4939-3246-7_3Search in Google Scholar

[4] M. G. Crandall, H. Ishii and P.-L. Lions, User’s guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc. (N.S.) 27 (1992), no. 1, 1–67. 10.1090/S0273-0979-1992-00266-5Search in Google Scholar

[5] M. G. Crandall and J. Zhang, Another way to say harmonic, Trans. Amer. Math. Soc. 355 (2003), no. 1, 241–263. 10.1090/S0002-9947-02-03055-6Search in Google Scholar

[6] N. Garofalo and J. T. Tyson, Riesz potentials and p-superharmonic functions in Lie groups of Heisenberg type, Bull. Lond. Math. Soc. 44 (2012), no. 2, 353–366. 10.1112/blms/bdr102Search in Google Scholar

[7] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Grundlehren Math. Wiss. 224, Springer, Berlin, 1977. 10.1007/978-3-642-96379-7Search in Google Scholar

[8] P. Juutinen, P. Lindqvist and J. J. Manfredi, On the equivalence of viscosity solutions and weak solutions for a quasi-linear equation, SIAM J. Math. Anal. 33 (2001), no. 3, 699–717. 10.1137/S0036141000372179Search in Google Scholar

[9] N. Katzourakis, An Introduction to Viscosity Solutions for Fully Nonlinear PDE with Applications to Calculus of Variations in L, Springer Briefs Math., Springer, Cham, 2015. 10.1007/978-3-319-12829-0Search in Google Scholar

[10] S. Koike, A Beginner’s Guide to the Theory of Viscosity Solutions, MSJ Mem. 13, Mathematical Society of Japan, Tokyo, 2004. Search in Google Scholar

[11] S.-Y. Li, On the Dirichlet problems for symmetric function equations of the eigenvalues of the complex Hessian, Asian J. Math. 8 (2004), no. 1, 87–106. 10.4310/AJM.2004.v8.n1.a8Search in Google Scholar

[12] P. Lindqvist, On the definition and properties of p-superharmonic functions, J. Reine Angew. Math. 365 (1986), 67–79. 10.1515/crll.1986.365.67Search in Google Scholar

[13] P. Lindqvist, Notes on the Infinity Laplace Equation, Springer Briefs Math., Springer, Cham, 2016. 10.1007/978-3-319-31532-4Search in Google Scholar

[14] P. Lindqvist and J. J. Manfredi, Note on a remarkable superposition for a nonlinear equation, Proc. Amer. Math. Soc. 136 (2008), no. 1, 133–140. 10.1090/S0002-9939-07-09142-3Search in Google Scholar

[15] P. Lindqvist, J. Manfredi and E. Saksman, Superharmonicity of nonlinear ground states, Rev. Mat. Iberoam. 16 (2000), no. 1, 17–28. 10.4171/RMI/269Search in Google Scholar

[16] A. M. Oberman and L. Silvestre, The Dirichlet problem for the convex envelope, Trans. Amer. Math. Soc. 363 (2011), no. 11, 5871–5886. 10.1090/S0002-9947-2011-05240-2Search in Google Scholar

[17] G. Thorbergsson, A survey on isoparametric hypersurfaces and their generalizations, Handbook of Differential Geometry. Vol. 1, North-Holland, Amsterdam (2000), 963–995. 10.1016/S1874-5741(00)80013-8Search in Google Scholar

[18] N. S. Trudinger and X.-J. Wang, Hessian measures. II, Ann. of Math. (2) 150 (1999), no. 2, 579–604. 10.2307/121089Search in Google Scholar

[19] Q. M. Wang, Isoparametric functions on Riemannian manifolds. I, Math. Ann. 277 (1987), no. 4, 639–646. 10.1007/BF01457863Search in Google Scholar

Received: 2017-05-22
Revised: 2017-11-13
Accepted: 2017-11-16
Published Online: 2017-12-07
Published in Print: 2020-04-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/acv-2017-0030/html
Scroll to top button