Home Regularity Properties for a Class of Non-uniformly Elliptic Isaacs Operators
Article Open Access

Regularity Properties for a Class of Non-uniformly Elliptic Isaacs Operators

  • Fausto Ferrari and Antonio Vitolo EMAIL logo
Published/Copyright: November 13, 2019

Abstract

We consider the elliptic differential operator defined as the sum of the minimum and the maximum eigenvalue of the Hessian matrix, which can be viewed as a degenerate elliptic Isaacs operator, in dimension larger than two. Despite of nonlinearity, degeneracy, non-concavity and non-convexity, such an operator generally enjoys the qualitative properties of the Laplace operator, as for instance maximum and comparison principles, ABP and Harnack inequalities, Liouville theorems for subsolutions or supersolutions. Existence and uniqueness for the Dirichlet problem are also proved as well as local and global Hölder estimates for viscosity solutions. All results are discussed for a more general class of weighted partial trace operators.

1 Introduction and main results

In this paper we investigate the properties of weighted partial trace operators

(1.1) 𝐚 := i = 1 n a i λ i ( X ) ,

where λi(X) are the eigenvalues of X𝒮n, the set of n×n real symmetric matrices, in increasing order, that is

λ 1 ( X ) λ n ( X ) ,

and 𝐚=(a1,,an)n is an n-ple of non-negative coefficients ai such that aj>0 for at least one j{1,,n}.

The class 𝒜¯ of such operators includes the partial trace operators

(1.2) 𝒫 k - ( X ) = i = 1 k λ i ( X ) , 𝒫 k + ( X ) = i = n - k + 1 n λ i ( X ) ,

considered by Harvey and Lawson [49, 50] and Caffarelli, Li and Nirenberg [23, 22].

Here we introduce the subclass 𝒜, characterized by non-negative coefficients ai such that a1>0 and an>0, which in some sense complements the set of operators 𝒫k±(X) with k<n. In fact, the prototype of 𝒜 is the min-max operator

( X ) := λ 1 ( X ) + λ n ( X ) .

As we will see later, can be in fact viewed as a degenerate elliptic Isaacs operator (for n2) whereas 𝒫k±(X) results in a degenerate elliptic Bellman operator (for k<n).

Of course, the case n=2 is by far well known, because reduces to the classical Laplace operator. However, in higher dimension, namely for n>2, the operator ceases to be uniformly elliptic, it becomes a fully nonlinear non-convex degenerate elliptic operator. Nonetheless, we will see, rather surprisingly, that it retains many properties of the Laplace operator.

It also worth noticing that the operators 𝐚 of the smaller subclass 𝒜¯, characterized by weights ai>0 for all i=1,,n, are uniformly elliptic, as we will see in Section 2.

After introducing our main results we shall further come back to the original motivation for studying the operators of 𝒜¯, and in particular the subclass 𝒜.

A good number of results will depend on the dimension n and on the following two quantities, namely the minimum between the coefficients of the smallest and the greatest eigenvalue, and the arithmetic mean of the coefficients, namely

a * := min ( a 1 , a n ) , a ~ := a 1 + + a n n ,

in the sense that the involved constants are uniformly bounded when a positive upper bound of the first one and a finite upper bound of the latter one are avalaible.

The constants which depend only on n, a1, an and a~ will be also called universal constants.

The following result is a revisitation of the bilateral Alexandroff–Bakelman–Pucci (ABP) estimate for the class 𝒜, only depending on n and a*.

Theorem 1.1.

Let ΩRn be a bounded open domain of diameter d. Let fC(Ω) be bounded in Ω. If uC(Ω¯) is a viscosity solution to the equation Ma(D2u)=f in Ω, with MaA, then

(1.3) sup Ω | u | sup Ω u + + C n a * d f L n ( Ω ) ,

where Cn>0 is a positive constant depending only on n.

We emphasize the following difference between estimate (1.3) and the standard ABP estimates, see for instance [47, Theorem 9.1]: the denominator of the right-hand side is a*=min(a1,an) instead of the geometric mean 𝒟*=(a1an)1n, the geometric mean of the coefficients, which would be useless in the non-uniformly elliptic case, as soon as one of the coefficients aj is zero, while a* is positive for the class 𝒜.

The above result is obtained as a consequence of two unilateral ABP estimates for subsolutions (4.6) and supersolutions (4.7).

The ABP estimate stated before also underlies a corresponding Harnack inequality for the equation 𝐚[u]=f, depending on n, a* and a~, instead of the elliptic constants λ and Λλ, which would be ineffective in the degenerate elliptic case in which λ=0. This Harnack inequality cannot be extended to arbitrary degenerate elliptic operators of the class 𝒜¯, and in particular it fails to hold for partial trace equations 𝒫k±[u]=f when k<n.

Theorem 1.2 (Harnack inequality).

Let MaA. Let u be a viscosity solution of the equation Ma(D2u)=f in the unit cube Q1 such that u0 in Q1, where f is continuous and bounded. Then

sup Q 1 / 2 u C ( inf Q 3 / 4 u + f L n ( Q 1 ) ) ,

where C is a positive constant depending only on n, a* and a~.

We prove Theorem 1.2 via two inequalities for subsolutions and non-negative supersolutions, known in literature respectively as the local maximum principle (Theorem 5.1) and the weak Harnack inequality (Theorem 5.2), suitably adapted to this framework, by comparison with Pucci extremal operators.

From the Harnack inequality, the interior Cα estimates of Theorem 5.3 in Section 5 follow with a universal exponent α(0,1), in the same way of the uniformly elliptic case [25].

Here, in Lemma 5.4, we get boundary Hölder estimates assuming for Ω a uniform exterior sphere property, with radius R>0:

  1. for all yΩ there is a ball BR of radius R such that yBR and Ω¯B¯R.

We obtain the following estimates for the Hölder seminorm [u]γ,Ω. See the notation (5.3) in Section 5.

Theorem 1.3 (Global Hölder estimates).

Let uC(Ω¯) be a viscosity solution of the equation Ma(D2u)=f in a bounded domain Ω. We assume that with MaA and f is continuous and bounded in Ω. Let also α(0,1) be the exponent of the interior Cα estimates.

  1. Suppose that Ω satisfies a uniform exterior sphere condition (s) with radius R>0. If u=g on Ω with gCβ(Ω) and β(0,1], then uCγ(Ω¯) with γ=min(α,β2), and

    (1.4) [ u ] γ , Ω C ( g C β ( Ω ) + f L ( Ω ) ) ,

    where C is a positive constants depending only on n, a*, a~n, R, L and β.

  2. Suppose in addition that Ω has a uniform Lipschitz boundary with Lipschitz constant L. If gC1,β(Ω) with β[0,1), then uC(1+β)/2(Ω¯), where γ=min(α,12(β+1)), and

    (1.5) [ u ] 1 2 ( 1 + β ) , Ω C ( g C 1 , β ( Ω ) + f L ( Ω ) ) .

A global estimate for the Hölder norm uC0,γ(Ω)=uL(Ω)+[u]γ,Ω can be obtained combining the above estimates with the uniform estimate of Corollary 3.3.

In some cases, we can obtain an explicit interior Hölder exponent. For instance, in the case of asymmetric distributions of weights, concentrated on the smallest or the largest eigenvalue, as for the upper and lower partial trace operators 𝒫k±, k<n, see Lemma 5.5. The result depends in fact on the smallness of the quotients a^1a1 and a^nan (see Section 3.3), as it can be seen in the statement below.

Theorem 1.4.

Let uC(Ω¯) be a viscosity solution of the equation Ma(D2u)=f in a bounded domain Ω, where f is continuous and bounded. Suppose MaA¯ with a1a^1, resp. with ana^n. Then the global Hölder estimates of Theorem 1.3 hold, namely (1.4) in case (i) and (1.5) in case (ii), with

α = max ( 1 - a ^ 1 a 1 , 1 - a ^ n a n ) .

In particular, we deduce the following Cα estimates.

  1. Suppose that Ω satisfies a uniform exterior sphere condition (s) with radius R>0. If u=g on Ω with gC2α(Ω) and α(0,12], then uCα(Ω¯), and

    (1.6) [ u ] α , Ω C ( g C 2 α ( Ω ) + f L ( Ω ) ) ,

    where C is a positive constant depending only on n, a1, a^1, an, a^n, a~, R and α.

  2. Suppose in addition that Ω has a uniform Lipschitz boundary with Lipschitz constant L. If gC1,2α-1(Ω) with α[12,1], then

    (1.7) [ u ] α , Ω C ( g C 1 , 2 α - 1 ( Ω ) + f L ( Ω ) ) ,

    where C is a positive constant also depending on L.

The optimal regularity of solutions in the case of degenerate, non-uniform ellipticity, is an open problem.

Partial answers are contained for instance in [65] for the special case a1=1, ai=0 for i=2,,n, as mentioned in the sequel, [52, 10] for other kind of singular or degenerate elliptic operators and [37, 40] for non-commutative structures.

Concerning higher regularity, one could borrow the techniques of [11, 54, 75, 24, 19, 73, 27, 28, 56, 53, 57], which however do not seem at the moment directly applicable in the more general non-uniformly elliptic setting.

It is remarkable the particular case of the interior C1,α regularity proved in [65] for the equation λ1[u]=f(x) with C1,α boundary data.

Further aspects of the qualitative theory, like the strong maximum principle and Liouville theorems, will be discussed in the last sections of the paper. New results for operators 𝐚𝒜¯ will be shown there, depending on the relative magnitude of a1 and an, and their complements a^1 and a^n with respect to |𝐚|=a1++an, see Section 3.3.

Turning to the motivations about the importance of this research, we recall that the partial trace operators 𝒫k+(X) are degenerate elliptic operators, which can be represented as Bellman operators

(1.8) 𝒫 k + ( X ) = sup W 𝒢 k Tr ( X W ) , 𝒫 k - ( X ) = inf W 𝒢 k Tr ( X W ) ,

where 𝒢k is the Grassmanian of the k-dimensional subspace W of n and XW is tha matrix of the quadratic form associated to X restricted to W, see [49].

Upper and lower partial trace operators arise in geometric problems of mean partial curvature considered by Wu [80] and Sha [71, 72]. Following the interest generated by the previous works, a number of papers has been devoted to the properties of these operators, we recall for instance [3, 26, 46, 79].

On the other hand, it is also worth noticing that Bellman equations arise in stochastic control problems, see Krylov [58], Fleming and Rishel [42], Fleming and Soner [43], Fleming and Souganidis [44] and the references therein.

As well as the partial trace operators 𝒫k± with k<n constitute a model for degenerate elliptic Bellman operators, the min-max operator provides for n3 a prototype of degenerate elliptic Isaacs operators by the representation

(1.9) ( X ) = sup | ξ | = 1 inf | η | = 1 Tr ( X ξ , η ) ,

where Tr(Xξ,η) is the trace the matrix Xξ,η of the quadratic form associated to X restricted to L(ξ,η), the subspace of n spanned by ξ and η.

The alternative representation

( X ) = max | ξ | = 1 X ξ , ξ + min | ξ | = 1 X ξ , ξ

suggests the relationship between (X) and stochastic zero-sum, two-players differential games and Isaacs equations, for which we refer for instance to [64, 41, 43] and the references therein and to [17, 16] for more recent contributions.

Following the main stream of the mean value properties of solution to linear equations, as well as in the case of the -Laplacian, it is also worth to be remarked that whenever u is C2, the following expansion yields:

u 1 ε ( x ) min | ξ | = 1 u ( x + ε ξ ) + u ( x - ε ξ ) 2 = u ( x ) + ε 2 2 λ 1 ( x ) + o ( ε 2 ) , u 2 ε ( x ) max | η | = 1 u ( x + ε η ) + u ( x - ε η ) 2 = u ( x ) + ε 2 2 λ n ( x ) + o ( ε 2 ) .

As a consequence, if we consider a continuous function u, the operator given by the limit

lim ϵ 0 2 ϵ 2 ( u ε , 1 ( x ) + u ε , 2 ( x ) - 2 u ( x ) ) ,

whenever it exists, may be considered as the weak version of our operator . For an almost compete list of references from this point of view, see [63, 59] for the p-Laplace equation, as well as [38, 39] for further applications to non-commutative fields where a lack of ellipticity occurs.

The paper is organized as follows. In Section 2 we introduce the main definitions about elliptic operators and viscosity solutions. In Section 3 we discuss in detail the properties of the weighted partial trace operators, in particular . We show a comparison principle, an existence and uniqueness theore, and compute the radial solutions. In Section 4 we prove Theorem 4.2. In Section 5 we show the Harnack inequality, interior and boundary Hölder estimates. We also discuss, in Section 6, the strong maximum principle via both the Hopf boundary point lemma and the Harnack inequality, showing suitable counterexamples. Finally, in Section 7, we also prove Liouville theorems and an unilateral Liouville property with the Hadamard’s three circles theorem.

2 General Preliminaries

This section is organized in some subsections, mainly for introducing common notation about viscosity theory of elliptic nonlinear PDEs, see Sections 2.1 and 2.5. In Sections 2.2 and 2.3 we introduce our class of operators 𝒜 and in particular discuss the min-max operator showing by counterexamples that it is nonlinear, non-convex and non-uniformly elliptic. In Section 2.4 we discuss a comparison result with the partial trace operators operator 𝒫k±.

2.1 Ellipticity and Viscosity Solutions

We start recalling some ellipticity notions. Let 𝒮n be the set of n×n symmetric matrices with real entries, partially ordered with the relationship XY if and only if Y-X is semidefinite positive.

A fully nonlinear operator, that is a mapping :𝒮n, is said degenerate elliptic if

(2.1) X Y ( X ) ( Y ) ,

and uniformly elliptic if

(2.2) X Y λ Tr ( Y - X ) ( Y ) - ( X ) Λ Tr ( Y - X ) ,

for positive constants λ and Λ, called ellipticity constants. Note indeed that, by the left-hand side inequality in (2.2), a uniformly elliptic operator satisfies (2.1), and so it is degenerate elliptic. The uniform ellipticity also implies the continuity of the mapping :𝒮n. In what follows we also assume that is a continuous mapping even in the degenerate elliptic case.

Suppose now XY. It is plain that Tr(Y-X)0. Suppose in addition (Y)=(X). If is uniformly elliptic, in view of the left-hand side of (2.2), we also have Tr(Y-X)0, so that Tr(Y-X)=0. Then Y=X. In other words, is strictly increasing on ordered chains of 𝒮n.

The class of uniformly elliptic operators with given ellipticity constants λ and Λ is bounded by two estremal operators, the maximal and minimal Pucci operator, which are in turn uniformly elliptic with the same ellipticity constants, respectively:

λ , Λ + ( X ) = Λ Tr ( X + ) - λ Tr ( X - ) ,
λ , Λ - ( X ) = λ Tr ( X + ) - Λ Tr ( X - ) ,

where X=X+-X- is the unique decomposition of X𝒮n as difference of semidefinite positive matrices X+ and X- such that X+X-=0.

In view of this definition, the uniformly ellipticity (2.2) of can equivalently be stated as

λ , Λ - ( Y - X ) ( Y ) - ( X ) λ , Λ + ( Y - X ) for all  X , Y 𝒮 n .

From this it also follows that, if is uniformly elliptic and (0)=0, then

(2.3) λ , Λ - ( X ) ( X ) λ , Λ + ( X ) for all  X 𝒮 n ,

which shows the extremality of Pucci operators.

Throughout this paper we will assume in fact

( 0 ) = 0 .

Of course, the results can be applied, in the case (0)0, to the operator 𝒢(X)=(X)-(0).

Let Ω be an open set of n. A fully nonlinear operator acts on uC2(Ω) through the Hessian matrix D2u setting

[ u ] ( x ) = ( D 2 u ( x ) ) .

Let f be a function defined in Ω. A solution uC2(Ω) of the equation [u]=f is called a classical solution, as well as classical subsolution or supersolution of F[u]=f if (D2u(x))f(x) or (D2u(x))f(x) for every xΩ, respectively. For instance, if (X)=Tr(X) and f(x) is a continuous function, then [u]=Δu is the Laplacian and the equation [u]=f is the Poisson equation Δu=f.

Let be a degenerate elliptic operator. We can solve the equation [u]=f(x) in a weaker sense, namely in the viscosity sense. We are essentially concerned in this paper with pure second-order operators [u]=(D2u). We refer to [25] and [33] for general operators, also depending on xΩ, u and the gradient Du, and to [49] for a geometric interpretation of viscosity solutions.

We briefly recall what it means to solve the equation [u]=f introducing sub/superjets basic notions. See [32, 33].

Let 𝒪 be a locally compact subset of n, and u:𝒪. The second-order superjet J𝒪2,+u(x0) and subjet J𝒪2,-u(x0) of u at x0𝒪 are respectively the sets

J 𝒪 2 , + u ( x 0 ) = { ( ξ , X ) n × 𝒮 n : u ( x ) u ( x 0 ) + ξ , x - x 0 + 1 2 X ( x - x 0 ) , ( x - x 0 ) + o ( | x - x 0 | 2 )  as  x x 0 }

and

J 𝒪 2 , - u ( x 0 ) = { ( ξ , X ) n × 𝒮 n : u ( x ) u ( x 0 ) + ξ , x - x 0 + 1 2 X ( x - x 0 ) , ( x - x 0 ) + o ( | x - x 0 | 2 )  as  x x 0 } .

We denote by usc(𝒪) and lsc(𝒪) the set of upper and lower semicontinuous functions in 𝒪, respectively. If uusc(𝒪), then u is a viscosity subsolution of a fully nonlinear elliptic equation [u]=f if

( X ) f ( x ) for all  x 𝒪  and all  ( ξ , X ) J 𝒪 2 , + u ( x ) .

If ulsc(𝒪), then u is a viscosity supersolution of the same equation if

( X ) f ( x ) for all  x 𝒪  and all  ( ξ , X ) J 𝒪 2 , - u ( x ) .

A viscosity solution of the equation [u]=f is both a subsolution and a supersolution uC(𝒪).

It is worth noticing that classical solutions are viscosity solutions. Viceversa, viscosity solutions of class C2 are in turn classical solutions. The same holds for subsolutions and supersolutions.

2.2 The Operator Class 𝒜

Let {𝐞1,,𝐞n} be the standard basis in n such that (𝐞i)j=δij for i,j=1,,n, and let λi(X), i=1,,n, be the eigenvalues of X𝒮n in non-decreasing order.

Let 𝐚=(a1,,an)=a1𝐞1++an𝐞n. We consider the class of degenerate elliptic weighted trace operators

𝒜 ¯ = { 𝐚 : a ¯ min i a i 0 , a ¯ max i a i > 0 } ,

where

𝐚 ( X ) = a 1 λ 1 ( X ) + + a n λ n ( X ) (see (1.1)) .

We observe that 𝒜¯ contains both uniformly and non-uniformly elliptic operators. In particular, all previously considered operators belong to this class with a suitable representation:

Tr ( X ) = 𝐞 1 + + 𝐞 n ( X ) , ( X ) = 𝐞 1 + 𝐞 n ( X ) ,
𝒫 + ( X ) = 𝐞 n - k + 1 + + 𝐞 n ( X ) , 𝒫 - ( X ) = 𝐞 1 + + 𝐞 k ( X ) .

Very recently, recalling the pioneeristic paper [66], Blanc and Rossi [15] have shown that it is possible to define a game satisfying a dynamic programming principle (DPP) which leads to the Dirichlet problem

{ 𝐚 [ u ] = 0 in  Ω , u = g ( x ) on  Ω .

Moreover, an associated evolution problem is considered in [14].

We point out that =𝐞1+𝐞n is neither linear nor uniformly elliptic, neither concave nor convex, except when n=2, as it follows from the representation (1.9) and it will be proved in the next section with suitable counterexamples.

Actually, is a model of a larger class of degenerate, possibly non-uniformly elliptic operators

𝒜 = { 𝐚 : a ¯ 0 , a * min ( a 1 , a n ) > 0 } ,

which can be seen as 𝒜=𝒜1𝒜n, where

𝒜 j = { 𝐚 : a ¯ 0 , a j > 0 } .

Setting in addition

𝒜 ¯ = { 𝐚 : a ¯ > 0 } ,

we notice that

𝒜 ¯ 𝒜 = 𝒜 1 𝒜 n 𝒜 ¯ .

We remark for instance that, while the min-max operator belongs to 𝒜, the partial trace operators 𝒫k-𝒜1 and 𝒫k+𝒜n do not belong to 𝒜 for k<n.

On the other hand, every 𝐚𝒜¯ is uniformly elliptic. In fact, if XY, then

(2.4) 𝐚 ( Y ) - 𝐚 ( X ) = i = 1 n a i ( λ i ( Y ) - λ i ( X ) ) a ¯ Tr ( Y - X ) ,

so that every 𝐚A¯ is degenerate elliptic. Since XY also implies

(2.5) 𝐚 ( Y ) - 𝐚 ( X ) = i = 1 n a i ( λ i ( Y ) - λ i ( X ) ) a ¯ Tr ( Y - X ) ,

we conclude that 𝐚A¯ is uniformly elliptic with ellipticity constants λ=a¯miniai and Λ=a¯maxiai.

We also observe that the operators 𝐚𝒜¯ are invariant by rotation, since 𝐚(TX)=(X) for all orthogonal matrices , and are positively homogeneous of degree one:

(2.6) 𝐚 ( ρ X ) = i = 1 n a i λ i ( ρ X ) = ρ i = 1 n a i λ i ( X ) = ρ 𝐚 ( X ) , ρ 0 .

Next, we investigate more closely the peculiar properties of the min-max operator (X)=λ1(X)+λn(X).

2.3 The Min-Max Operator

In the previous subsection, we claimed that is neither linear nor uniformly elliptic, neither concave nor convex, except for n=2. This is intuitive by the representation (1.9):

( X ) = sup | ξ | = 1 inf | η | = 1 Tr ( X ξ , η ) .

Nonetheless, we present a few counterexamples that support the above claim.

Remark 2.1.

Let us consider the matrices

X 1 = 𝐞 1 𝐞 1 - 𝐞 3 𝐞 3 , X 2 = - 𝐞 1 𝐞 1 + 𝐞 2 𝐞 2 , X 3 = 𝐞 1 𝐞 1 - 𝐞 2 𝐞 2 - 𝐞 3 𝐞 3 .

Then λ1(Xi)=-1 and λ3(Xi)=1, so that (Xi)=0 for all i=1,2,3.

  1. The operator is not linear in dimension n3. In fact,

    X 1 - X 2 = 𝐞 1 𝐞 1 - 𝐞 3 𝐞 3 + 𝐞 1 𝐞 1 - 𝐞 2 𝐞 2 = 2 𝐞 1 𝐞 1 - 𝐞 2 𝐞 2 - 𝐞 3 𝐞 3

    and therefore

    λ 1 ( X 1 - X 2 ) = - 1 , λ 3 ( X 1 - X 2 ) = 2 ,

    so that

    ( X 1 ) - ( X 2 ) = 0 1 = ( X 1 - X 2 ) .

  2. The operator is not uniformly elliptic in dimension n3. In fact, we note that X3X1, and

    ( X 3 ) = ( X 1 ) = 0 , but  X 3 X 1 ,

    against the strictly increasing property on ordered chains observed in Section 2.1 for the uniformly elliptic case.

  3. The operator (X) is neither convex nor concave. In fact, for every t[0,1], it turns out that

    t X 1 + ( 1 - t ) X 2 = t ( 𝐞 1 𝐞 1 - 𝐞 3 𝐞 3 ) + ( 1 - t ) ( - 𝐞 1 𝐞 1 + 𝐞 2 𝐞 2 ) = ( 2 t - 1 ) 𝐞 1 𝐞 1 + ( 1 - t ) 𝐞 2 𝐞 2 - t 𝐞 3 𝐞 3

    and therefore

    ( t X 1 + ( 1 - t ) X 2 ) = λ 1 ( t X 1 + ( 1 - t ) X 2 ) + λ 3 ( t X 1 + ( 1 - t ) X 2 )
    = min { 2 t - 1 ; - t } + max { 2 t - 1 ; 1 - t } ,

    so that (tX1+(1-t)X2)=1-2t for t(13,23). From this

    ( t X 1 + ( 1 - t ) X 2 ) { > 0 for  t ( 1 3 , 1 2 ) , < 0 for  t ( 1 2 , 2 3 ) ,

    while it is plain that for every t[0,1],

    t ( X 1 ) + ( 1 - t ) ( X 2 ) = 0 .

    Thus is neither convex nor concave.

Since 𝒜 we already know that it is homogeneous of degree one (2.6): for every ρ0 and for every X𝒮n,

(2.7) ( ρ X ) = ρ ( X ) .

On the other hand,

( - X ) = λ 1 ( - X ) + λ n ( - X ) = - λ n ( X ) - λ 1 ( X ) = - ( X ) ,

and therefore (2.7) continues to hold for ρ<0.

The next remark contains a few comments on the representation (1.9).

Remark 2.2.

The operator (X) can be put in the form

(2.8) ( X ) = sup | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) .

In order to prove this, we start observing that plainly

( X ) = sup | ξ | = 1 inf | η | = 1 ( X ξ , ξ + X η , η ) sup | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) .

To have also the reverse inequality, and so (2.8), we observe that the representative matrix Xξ,η of the quadratic form associated to X restricted to L(ξ,η), the subspace of n spanned by directions ξ and η, has trace

Tr ( X ξ , η ) = X ξ , ξ + X η , η ,

and thus

sup | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) = sup | ξ | = 1 ( X ξ , ξ + inf | η | = 1 η ξ X η , η ) .

To compute the inf in the latter equation, we may assume that X is diagonal, by rotational invariance, with the eigenvalues λ1λn on the diagonal from the top to the bottom. Note also that in this case Xξ,ξ=λ1ξ12++λnξn2 and Xη,η=λ1η12++λnηn2, so that by symmetry we may assume ξi0 and ηi0 for all i=1,,n, that is

sup | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) = sup | ξ | = 1 ξ 0 ( X ξ , ξ + inf | η | = 1 η 0 η ξ X η , η ) .

Using the Lagrange multipliers λ and μ, the inf is obtained in correspondence of a critical point of the function

h ( η , λ , μ ) := X η , η - λ ( η , η - 1 ) - μ ξ , η ,

which solve the system

{ X η = λ η + μ 2 ξ , η , η = 1 , ξ , η = 0

or equivalently

{ λ 1 η 1 = λ η 1 + μ 2 ξ 1 , λ n η n = λ η n + μ 2 ξ n , η 1 2 + + η n 2 = 1 , ξ 1 η 1 + + ξ n η n = 0 .

We can show that μ=0. Otherwise, suppose by contradiction μ0. Let I={i{1,,n}:ξi0}, which is non-empty because |ξ|=1. Then from above (λi-λ)ηi=μ2ξi0, and so λλi for all iI. Inserting ηi=μ2ξiλ-λi in the last row of the system, we get

μ 2 i I ξ i 2 λ - λ i = 0 .

Since ξi>0 and ηi>0 for iI, all the terms of the sum have the same sign (the sign of μ), and this would imply μ=0, against the assumption. Therefore critical points are not affected by the constraint ηξ, and this proves the representation (2.8).

If instead of “sup inf” as in (2.8) we consider “inf sup”, we re-obtain :

(2.9) ( X ) = sup | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) = inf | ξ | = 1 sup | η | = 1 η ξ Tr ( X ξ , η )

with or without the constraint ηξ.

2.4 Comparison with the Partial Trace Operators Operator 𝒫k±

Let us give a comparative look to the partial trace operators (1.2):

𝒫 k - ( X ) = λ 1 ( X ) + + λ k ( X ) , 𝒫 k + ( X ) = λ n - k + 1 ( X ) + + λ n ( X ) .

Remark 2.3.

If in (2.9) we consider “sup sup” or “inf inf” instead of “sup inf” or “inf sup”, it is not difficult to recognize, from (1.8), that we obtain the above partial trace operators with k=2:

𝒫 2 - ( X ) = inf | ξ | = 1 inf | η | = 1 η ξ Tr ( X ξ , η ) ,
𝒫 2 + ( X ) = sup | ξ | = 1 sup | η | = 1 η ξ Tr ( X ξ , η ) .

Next, we list some properties of operators 𝒫k±. By definition, it is plain that 𝒫k-𝒫k+; in addition 𝒫k+ and 𝒫k- are respectively subadditive and superadditive:

𝒫 k - ( X ) + 𝒫 k - ( Y ) 𝒫 k - ( X + Y ) 𝒫 k + ( X + Y ) 𝒫 k + ( X ) + 𝒫 k + ( Y ) .

Moreover, 𝒫k-(X)=-𝒫k+(-X), so that from the left-hand inequality

𝒫 k - ( X + Y ) 𝒫 k - ( X ) - 𝒫 k - ( - Y ) = 𝒫 k - ( X ) + 𝒫 k + ( Y )

and from the right-hand

𝒫 k + ( X + Y ) 𝒫 k + ( X ) - 𝒫 k + ( - Y ) = 𝒫 k + ( X ) + 𝒫 k - ( Y ) .

In particular, since λ1(X)=𝒫1-(X) and λn(X)=𝒫1+(X),

λ 1 ( X ) + λ 1 ( Y ) λ 1 ( X + Y ) λ 1 ( X ) + λ n ( Y )

and

λ 1 ( X ) + λ n ( Y ) λ n ( X + Y ) λ n ( X ) + λ n ( Y ) .

We recall that the inequality stated above for the partial trace operators 𝒫k± continues to hold for the Pucci extremal operators λ,Λ±, that can be in turn regarded as Bellman operators. In fact, setting

𝒮 λ , Λ n = { A 𝒮 n : λ I A Λ I } ,

where I is the n×n identity matrix, we have

λ , Λ + ( X ) = sup A 𝒮 λ , Λ n Tr ( A X ) ,
λ , Λ - ( X ) = inf A 𝒮 λ , Λ n Tr ( A X ) .

2.5 Duality

Let be a fully nonlinear degenerate elliptic operator. If is linear and u is a subsolution of the equation (D2u)=f, then v=-u is a supersolution of the equation (D2v)=-f.

If we deal with an arbitrary fully nonlinear operator and u is a subsolution to F(D2u)=f, then v=-u is a supersolution of an equation ~(D2v)=-f for the dual operator ~,

~ ( X ) = - ( - X ) ,

which is in general different from . Moreover, ¯ is degnerate (uniformly) elliptic if is degenerate (uniformly) elliptic.

Computing the dual of the operators introduced above, we note that by homogeneity for the min-max operator we have ~= as in the case of linear operators, while the upper and lower partial trace operators are each one the dual of the other one, 𝒫~k±=𝒫k±, as well as the maximal and the minimal the Pucci operators, ~λ,Λ±=λ,Λ. In the general case 𝐚𝒜¯, we have ~𝐚=𝐚, where 𝐚=(an,an-1,,a1),.

3 Auxiliary Results

In this section we apply the Perron method, well known in the literature, see for instance [33] and [48], in order to show: weak maximum and comparison principles, existence and uniqueness of solutions, see respectively Sections 3.1 and 3.2. The proofs are based on the properties of our operators, suitably exploited, and an appropriate adaptation of arguments used for the uniformly elliptic case. In Section 3.3 we obtain the radial representation of the operators 𝐚.

3.1 Weak Maximum and Comparison Principles

The following comparison principle holds between viscosity subsolutions and supersolutions of the equation 𝐚[u]=f in a bounded domain Ω, as proved for uniformly elliptic operators in the basic paper by Crandall, Ishii and Lions [33].

Theorem 3.1 (Comparison Principle).

Let uusc(Ω¯) and vlsc(Ω¯) such that Ma(D2u)f and Ma(D2v)f in Ω are satisfied in the viscosity sense, respectively, where Ω is a bounded open set of Rn, MaA¯ and f is a bounded continuous function in Ω. If uv on Ω, then uv in Ω.

Letting v0 and f0, we obtain the following weak maximum principle.

Corollary 3.2 (Weak Maximum Principle).

Let uusc(Ω¯), where Ω is a bounded domain of Rn. If one has Ma(D2u(x))0 in Ω in the viscosity sense for some MaA¯, then

max Ω ¯ u = max Ω u .

On the other hand, if ulsc(Ω¯) is a viscosity solution of the differential inequality Ma(D2u(x))0 in Ω for some MaA¯, then

min Ω ¯ u = min Ω u .

Proof of Theorem 3.1.

The case of f(x)0 is covered in [48, Theorem 6.5]. In fact, considering the Dirichlet set F={X𝒮n:𝐚(X)0} and its dual set F~={X𝒮n:𝐚(X)0} in the geometric setting of Harvey and Lawson [48], then by our assumptions u,-vusc(Ω¯) are of type F and F~ in Ω, and our comparison principle is deduced the subaffinity of u-v established there.

For sake of completeness, we give an analytic proof based on the device contained in the proof of [33, Theorem 3.3] by Crandall, Ishii and Lions. See also [8].

We have to show that, under the given assumptions, the maximum of u-v must be realized on Ω.

(i) Firstly, setting uε(x)=u(x)+12ε|x|2, we prove that uε-v cannot have a positive maximum in Ω, for all fixed ε>0. Actually,

𝐚 ( D 2 u ε ) := i = 1 n a i λ i ( D 2 u ε ) = i = 1 n a i λ i ( D 2 u + ε I ) f ( x ) + | 𝐚 | ε ,
𝐚 ( D 2 v ) := i = 1 n a i λ i ( D 2 v ) f ( x ) ,

where |𝐚|=a1++an>0. Supposing, by contradiction, that uε-v has a positive maximum in Ω and following the proof of [33, Theorem 3.3 ], for all α>0 there exist points xα,yαΩ and matrices Xα,Yα𝒮n such that

(3.1) - 3 α ( I 0 0 I ) ( X α 0 0 - Y α ) 3 α ( I - I - I I )

and

(3.2) i = 1 n a i λ i ( X α ) f ( x α ) + | 𝐚 | ε , i = 1 n a i λ i ( Y α ) f ( y α ) .

Moreover,

(3.3) lim α α | x α - y α | 2 = 0 .

Noting that (3.1) implies XαYα, from (3.2) we get

f ( x α ) + | 𝐚 | ε i = 1 n a i λ i ( X α ) i = 1 n a i λ i ( Y α ) f ( y α ) .

Taking the limit as α and using (3.3), by the continuity of f(x) we have a contradiction: ε0. Therefore uε-v cannot have a positive maximum in Ω.

(ii) From (i) it follows, for all ε>0, that maxΩ¯(uε-v)maxΩ(uε-v). Taking into account that uv on Ω, then we have

u ( x ) + 1 2 ε | x | 2 - v ( x ) 1 2 ε R 2 for  x Ω ,

where R>0 is the radius of a ball BR centered at the origin such that ΩBR. Letting ε0+, we conclude that uv in Ω, as claimed. ∎

From Corollary 3.2 we deduce the following uniform estimates for viscosity solutions of the equation 𝐚[u]=f in a bounded domain Ω

Proposition 3.3 (Uniform Estimate).

Let uusc(Ω¯), where Ω is a bounded domain of Rn. If Ma(D2u(x))f(x) in Ω in the viscosity sense for some MaA¯ and f is bounded below in Ω, then

u ( x ) max Ω u + C d 2 f - L ( Ω ) for all  x Ω ¯ ,

where C is a positive constant, which can be chosen equal to 1/|a|.

On the other hand, if ulsc(Ω¯) is a viscosity solution of the differential inequality Ma(D2u(x))f(x) in Ω for some MaA¯ and f bounded above in Ω, then

u ( x ) min Ω u - C d 2 f + L ( Ω ) for all  x Ω ¯ .

Proof.

Let us prove the first one. Setting K-=f-L(Ω), the function v=u+K-2|𝐚||x|2 is a subsolution of the equation 𝐚[v]=0. By Corollary 3.2 we get v(x)maxΩ¯v, so that

u ( x ) v ( x ) max Ω u + K - 2 | 𝐚 | d 2 ,

which yields the first inequality of the estatement. ∎

3.2 Existence and Uniqueness

As a consequence of the above comparison principle, we can also prove an existence and uniqueness result for the Dirichlet problem in bounded domains Ω via the Perron method, assuming that Ω has a uniform exterior cone condition, see [60] and [21]: there exist θ0(0,π) and r0>0 so that for every yΩ there is a rotation =(y) such that

Ω ¯ B r 0 ( y ) y + Σ θ 0 ,

where

Σ θ 0 = { x n : x n | x | cos θ 0 } .

Theorem 3.4.

Let Ω be a bounded domain of Rn endowed with a uniform exterior cone condition. Let g be a continuous function on the boundary Ω, and let f be a continuous and bounded function in Ω. Then for MaA the Dirichlet problem

(3.4) { 𝐚 ( D 2 u ) = f in  Ω , u = g on  Ω ,

has a unique viscosity solution uC(Ω¯).

Proof.

According to the Perron method [33, Theorem 4.1], we need a comparison principle, and the existence of a subsolution and a supersolution of the equation 𝐚(D2u)=f. Since the comparison principle holds by Theorem 3.1, we only need to look for a viscosity subsolution u¯usc(Ω¯) and a viscosity supersolution u¯lsc(Ω¯) of the equation 𝐚(D2u)=f(x) such that u¯=g=u¯ on Ω.

To do this, we will use the following inequalities, see (2.4) and (2.5):

𝐚 ( X ) = a 1 λ 1 ( X ) + + a n λ n ( X ) = n a 1 n λ 1 ( X ) + + a n λ n ( X ) a 1 n λ 1 ( X ) + i = 2 n ( a 1 n + a i ) λ i ( X ) = : 𝐚 ¯ ( X )

and

𝐚 ( X ) = a 1 λ 1 ( X ) + + a n λ n ( X ) = a 1 λ 1 ( X ) + + n a n n λ n ( X ) i = 1 n - 1 ( a i + a n n ) λ i ( X ) + a n n λ n ( X ) = : 𝐚 ¯ ( X ) .

If 𝐚𝒜1, then 𝐚¯ is uniformly elliptic with ellipticity constants

λ ¯ = a 1 n , Λ ¯ = a 1 n + max 2 i n a i ,

so that

(3.5) 𝐚 ( X ) 𝐚 ¯ ( X ) a 1 n , | 𝐚 | +

and, if 𝐚𝒜n, then 𝐚¯ is uniformly elliptic with ellipticity constants

λ ¯ = a n n , Λ ¯ = a n n + max 1 i n - 1 a i ,

so that

(3.6) 𝐚 ( X ) 𝐚 ¯ ( X ) a n n , | 𝐚 | - ( X ) .

Therefore, if 𝐚𝒜, and λ* and Λ* are positive numbers such that

λ * min ( λ ¯ , λ ¯ ) = a * n min ( a 1 , a n ) n , Λ * max ( Λ ¯ , Λ ¯ ) | 𝐚 | a 1 + + a n ,

by the extremality properties (2.3) of Pucci operators, from (3.5) and (3.6) we have

(3.7) a * n , | 𝐚 | - ( X ) 𝐚 ( X ) a * n , | 𝐚 | + ( X ) .

Next, setting K=supΩ|f|, we solve by [21, Proposition 3.2] the Dirichlet problems

{ a * n , | 𝐚 | - ( D 2 u ¯ ) = K in  Ω , u ¯ = g on  Ω ,

and

{ a * n , | 𝐚 | + ( D 2 u ¯ ) = - K in  Ω , u ¯ = g on  Ω .

Since obviously -Kf(x)K for all xΩ, from (3.7) it follows that u¯ and u¯ provide a subsolution and a supersolution that we were searching for, concluding the proof. ∎

An existence and uniqueness result is provided for all the class 𝒜 by [48, Theorem 6.2] for smooth boundaries. A weaker condition can be obtained from [15], where the authors consider in detail the case 𝐚=𝐞j, namely the equation λj[u]=0, and prove an existence and uniqueness theorem for the Dirichlet problem (3.4) with a sharp geometric condition on the boundary of Ω, depending on j. From there, we take a sufficient condition to solve the Dirichlet problem for any equation λj(D2u)=0, j=1,,n: given yΩ, for every r>0 there exists δ>0 such that, for every xBδ(y) and direction vn (|v|=1),

($G_{1}$) ( x + v ) B r ( y ) Ω .

This condition does not require smooth boundary, but it is nevertheless stronger than the exterior cone property.

Theorem 3.5.

Let Ω be a bounded domain of Rn satisfying condition ($G_{1}$). Let g be a continuous function on the boundary Ω, and f be a continuous and bounded function in Ω. Then for MaA¯ the Dirichlet problem

{ 𝐚 ( D 2 u ) = f in  Ω , u = g on  Ω ,

has a unique viscosity solution uC(Ω¯).

Proof.

Following the same lines of the proof of Theorem 3.4, we only need to look for a viscosity subsolution u¯usc(Ω¯) and a viscosity supersolution u¯lsc(Ω¯) of the equation 𝐚(D2u)=f(x) such that u¯=g=u¯ on Ω. To do this, we observe this time

𝐚 ( X ) = a 1 λ 1 ( X ) + + a n λ n ( X ) | 𝐚 | λ n ( X )

and

𝐚 ( X ) = a 1 λ 1 ( X ) + + a n λ n ( X ) | 𝐚 | λ 1 ( X ) .

Next, setting K=supΩ|f|, we solve by [15, Theorem 1] the Dirichlet problems

{ | 𝐚 | λ 1 ( D 2 u ¯ ) = K in  Ω , u ¯ = g on  Ω ,

and

{ | 𝐚 | λ n ( D 2 u ¯ ) = - K in  Ω , u ¯ = g on  Ω .

As in the proof of Theorem 3.4, u¯ and u¯ provide a subsolution and a supersolution, concluding the proof. ∎

3.3 Radial Solutions

We compute on radial functions u(x)=v(|x|). Suppose v is C2, we recall that for x0,

D u ( x ) = v ( | x | ) x | x | , D 2 u ( x ) = v ′′ ( | x | ) x | x | x | x | + v ( | x | ) | x | ( I - x | x | x | x | ) ,

where x|x|x|x|0, I-x|x|x|x|0 and

x | x | x | x | h , h = x | x | , h 2 , ( I - x | x | x | x | ) h , h = | h | 2 - x | x | , h 2 .

As a consequence, x|x| is eigenvector of x|x|x|x| with eigenvalue 1, and of I-x|x|x|x| with eigenvalue 0. Conversely, all non-zero vectors orthogonal to x|x| are eigenvectors of x|x|x|x| with eigenvalue 0 and of I-x|x|x|x| with eigenvalue 1. It follows that

λ 1 ( D 2 u ( x ) ) + λ n ( D 2 u ( x ) ) = v ′′ ( | x | ) + v ( | x | ) | x | .

From this we deduce useful properties which are collected in the following remark.

Remark 3.6.

  1. The operator is linear on the radial functions u(x)=v(|x|).

  2. Any function of the form

    φ ( x ) = a + b log | x | ,

    with a and b constant, is a solution of [u]=0 in n{0}.

  3. Recall that the k-th Hessian operator, k=1,,n, for radial functions is

    S k ( D 2 u ) = ( n - 1 k - 1 ) ( v | x | ) k - 1 ( v ′′ + n - k k v | x | ) .

    In case n=2k the radial solutions of the equation

    S n 2 ( D 2 u ) = 0

    are just the radial solutions of (D2u)=0.

Recalling that |𝐚|=a1++an, let a^j=|𝐚|-aj, j=1,,n. More generally, for 𝐚𝒜¯ the non-constant radial solutions in n{0}, up to a multiplicative constant, are

(3.8) φ ( x ) = { | x | - γ n if  a ^ n > a n , log | x | - 1 if  a ^ n = a n , | x | γ 1 if  a 1 > a ^ 1 , log | x | if  a 1 = a ^ 1 ,

where γn=a^nan-1.

4 The ABP Estimate

The celebrated ABP estimate provides a uniform estimate for the solution of an elliptic equation F[u]=f with the Ln-norm of f. The original inequality, for linear uniformly elliptic operators in bounded domains, goes back to Alexandroff [1, 2], but it already appears in Bakel’man [5]. A different version has been later obtained by Pucci [68].

In [18] it was also proved for the first time an ABP estimate for solutions in Wloc2,p(Ω) of the equation F[u]=f with fLp and p(n2,n). A result of this kind is known in the framework of Lp-viscosity solutions [21] as the generalized maximum principle, which can be found in [45] and [34] in the fully nonlinear uniformly elliptic case. See also [55] for the case of Lp viscosity solutions.

It is worth noticing that an ABP estimate for degenerate elliptic equations of p-Laplacian type has been proved by Imbert [52].

An extension of this inequality to unbounded domains Ω of cylindrical type for bounded solutions in Wloc2,n(Ω) is due to Cabré [18]. By domains of cylindrical type we intend here a measure-geometric condition, which is satisfied by cylinders and goes back to a famous paper of Berestycki, Nirenberg and Vardhan [9], containing a characterization of the weak maximum principle. In subsequent papers the results of [18] have been generalized to domains of conical type [20, 77, 78] and to viscosity solutions of fully nonlinear uniformly elliptic equations [26], and then to different classes of degenerate elliptic equations [10, 29, 30].

The proof of the ABP estimates of Theorem 1.1 is based on the geometrical argument used in [47, proof of Theorem 9.1] for classical solutions.

We denote by Γu+ the upper convex envelope of u, the smallest concave function greater than u in Ω, and by Γu- the lower convex envelope of u, the largest convex function smaller than u in Ω.

Lemma 4.1.

Let Ω be a bounded domain with diameter d, and MaA1. For every uC2(Ω)C0(Ω¯) such that u0 on Ω we have

(4.1) sup Ω u + 1 a 1 d ω n 1 / n 𝐚 ( D 2 u ( x ) ) - L n ( { Γ u + = u } ) ,

where ωn denotes the Lebesgue measure of the n-dimensional unit ball.

On the other hand, let us assume MaAn. For every uC2(Ω)C0(Ω¯) such that u0 on Ω we have

(4.2) sup Ω u - 1 a n d ω n 1 / n 𝐚 ( D 2 u ( x ) ) + L n ( { Γ u - = u } ) .

Proof.

Let us prove the first estimate (4.1). We argue following [47, proof of Lemma 9.2] and [25, proof of Lemma 3.], denoting by χu:Ωn the normal mapping

χ u ( z ) = { p n : u ( x ) u ( y ) + p , x - z  for all  x Ω } , z Ω .

We remark that on the upper contact set {Γu+=u} the eigenvalues of D2u are non-positive, and the Lebesgue measure of χu can be estimated as

(4.3) | χ u ( Ω ) | Γ u + = u | det D 2 u ( x ) | d x .

If u0 in Ω, then inequality (4.1) is obvious. Suppose then u realizes a positive maximum at a point yΩ, and recall that ΩBd(y).

Let κ be the function whose graph is the cone K with vertex (y,u(y)) and base Bd(y); then χκ(Ω)χu(Ω). Then χu(Ω) and contains all the slopes of Bu(y)/d, so that ωn(u(y)d)n|χu(Ω)| and by (4.3),

(4.4) u + ( y ) d ω n 1 / n ( Γ u + = u | det D 2 u ( x ) | d x ) 1 n .

Since on the contact set we have |λn||λn-1||λ1|, it follows that

(4.5) | det D 2 u | = | λ 1 ( D 2 u ) | | λ n ( D 2 u ) | | λ 1 ( D 2 u ) | n = 1 a 1 n | a 1 λ 1 ( D 2 u ) | n 1 a 1 n | i = 1 n a i λ i ( D 2 u ) | n = 1 a 1 n ( ( 𝐚 ( D 2 u ) ) - ) n .

From (4.4) and (4.5) we obtain the estimate from above (4.1).

For the estimate from below, we can apply (4.1) with v=-u instead of u, observing that by assumption v0 on Ω and by duality

𝐚 ( D 2 v ) ) = - 𝐚 ( D 2 u ) ) .

Then

sup Ω u - = sup Ω v +
1 a 1 d ω n 1 / n 𝐚 ( D 2 v ( x ) ) + L n ( { Γ v + = v } )
= 1 a n d ω n 1 / n 𝐚 ( D 2 u ( x ) ) - L n ( { Γ u - = u } ) .

Theorem 1.1 is obtained combining the two unilateral ABP estimates, which hold separately for subsolutions and supersolutions, contained in the following result.

Theorem 4.2.

Let Ω be a bounded domain of diameter d. Let f be continuous and bounded in Ω. There exist an universal constant Cn>0, depending only on n,

  1. for viscosity subsolutions u usc ( Ω ) of the equation 𝐚 [ u ] = f in Ω with 𝐚𝒜1 and

    (4.6) sup Ω u + sup Ω u + + C n a 1 d f L n ( Ω ) ,

  2. for viscosity supersolutions u lsc ( Ω ) of the equation 𝐚 [ u ] = f in Ω with 𝐚𝒜n and

    (4.7) sup Ω u - sup Ω u - + C n a n d f L n ( Ω ) .

For classical solutions the proof follows directly from Lemma 4.1.

Proof of Theorem 4.2: Classical Solutions.

For subsolutions, supposing 𝐚(D2u)f-f-, we have

𝐚 ( D 2 u ) - f - .

From Lemma 4.1, passing to u-supΩu in (4.1), we get inequality (4.6). For supersolutions, supposing 𝐚(D2u)f-f+, we have

𝐚 ( D 2 u ) + f + .

From Lemma 4.1, passing to u-infΩu in (4.2), we get inequality (4.7). ∎

To consider viscosity subsolutions, we extend u+=max(u,0) and f- to zero outside Ω, keeping the respective notations, and observing that in the viscosity setting 𝐚(D2u+)-f- in n. For viscosity supersolutions we extend u-=max(u,0) and f+ to zero outside Ω so that 𝐚(D2u-)f+ in n.

In what follows we will refer to Γu+ and Γu- as to the upper and the lower convex envelope of u+ and -u-, respectively, relative to the ball B2d concentric with a ball Bd of radius d containing Ω.

The key tool is the following lemma, which allows to apply the classical ABP estimates obtained before to viscosity subsolutions and supersolutions and is the counterpart of [25, Lemma 3.3].

Lemma 4.3.

Let MaA. Let ulsc(B¯δ), where Bδ={|x-x0|<δ} such that

𝐚 ( D 2 u ) f in  B δ

in the viscosity sense, and w be a convex function such that

w ( x 0 ) = u ( x 0 ) , w ( x ) u ( x ) in  B δ .

For sufficiently small ε(0,ε0) and any function f, bounded above, we have

(4.8) ( x ) w ( x ) ( x ) + 1 2 C ε ( sup B δ f + ) | x - x 0 | 2 in  B ε δ ,

where (x) is the supporting hyperplane for w at x0. In particular, there exists a convex paraboloid of opening Cεan touching the graph of w from above.

Here ε0>0 depends on (a positive lower bound of) an and (an upper bound of) a^n defined in Section 3.3; moreover Cε1an as ε0. Therefore, when u is second-order differentiable and f is continuous at x=x0, we get

λ n + ( D 2 w ( x 0 ) ) 1 a n f + ( x 0 ) .

Proof.

The first one inequality in (4.8) depends on the fact that (x) is the supporting hyperplane of w at x0.

Concerning the second one, we may proceed assuming x0=0 and δ=1.

(i) Subtracting (x), we consider the functions v(x)=u(x)-(x) and φ(x)=w(x)-(x), which satisfy in turn the assumptions on u(x) and w(x), respectively. This simplifies the computations, since φ(0)=0 and the supporting hyperplane for v(x) at x=0 is now horizontal, so that φ(x)0 in B1. In this way, we are reduced to show

φ ( x ) 1 2 C ε K | x | 2 in  B ε ,

with K=supB1f+, under the assumptions

(4.9) φ ( 0 ) = v ( 0 ) = 0 , φ ( x ) v ( x ) in  B 1 ,

and

(4.10) a 1 λ 1 ( D 2 v ) + + a n λ n ( D 2 v ) f + ( x ) in  B 1 ,

which implies

a ^ n λ 1 ( D 2 v ) + a n λ n ( D 2 v ) f + ( x ) in  B 1 .

(ii) Let 0<ρ<ε and let Mρ be the maximum of φ on B¯ρ. We may suppose that a maximum point is xρ=(0,,0,ρ)Bρ. Since the supporting hyperplane for φ(x) at xρ is constant on the tangent line to Bρ through xρ, we have

φ ( x ) M ρ for  x = ( x , ρ ) ,

where x=(x1,,xn-1). Let us consider now the cylindrical box

R = { x = ( x , x n ) : | x | < 1 - ρ 2 , - ε ρ < x n < ρ } B 1 ,

and the paraboloid

P ( x ) = 1 2 ( x n + ε ρ ) 2 - 1 2 ( 1 + ε ) 2 1 - ρ 2 ρ 2 | x | 2 .

Evaluating P(x) on R, when xn=-ερ or |x|=1-ρ2, we have P(x)0. On the remaining part of R, xn=ρ, we have P(x)12(1+ε)2ρ2, from which

(4.11) M ρ 1 2 ( 1 + ε ) 2 ρ 2 P ( x ) φ ( x ) on  R .

(iii) Since ρ<ε, it follows that P(x) is solution of the differential inequality

a ^ n λ 1 ( D 2 P ) + a n λ n ( D 2 P ) a n - ( 1 + ε ) 2 1 - ρ 2 a ^ n ρ 2 a n - 1 + ε 1 - ε a ^ n ε 2 a n - a ^ n c ε ,

where cε0 as ε0+, so that an-a^ncε>0 for ε<ε0 small enough, and the function Q(x)Kan-a^ncεP(x) satisfies the differential inequality

(4.12) a ^ n λ 1 ( D 2 Q ) + a n λ n ( D 2 Q ) K f + in  B 1 .

(iv) We claim that

(4.13) M ρ = max B ¯ ρ φ ( x ) 1 2 ( 1 + ε ) 2 a n - a ^ n c ε K ρ 2 .

In fact, arguing by contradiction, suppose that Mρ>12(1+ε)2an-a^ncεKρ2. Then using (4.11) and (4.9),

Q ( x ) = K a n - a ^ n c ε P ( x ) < M ρ 1 2 ( 1 + ε ) 2 ρ 2 P ( x ) φ ( x ) v ( x ) on  R .

By (4.10) and (4.12), the comparison principle would imply Q(x)v(x) in R, and this is a contradiction with v(0)=0<Q(0), which proves the claim.

Setting ρ=|x| in (4.13), as in the proof of [25, Theorem 3.2], we conclude that the statement of the theorem holds with Cε=(1+ε)2an-a^ncε, where cε0 as ε0+. ∎

Proof of Theorem 4.2: Viscosity Solutions.

We follow the lines of the proof of [25, Theorem 3.6], considering subsolutions uusc(Ω¯). The case of supersolutions ulsc(Ω¯) with estimate (4.7) from below can be obtained by duality, passing to -u.

Let 𝐚𝒜. From Lemma 4.3 and duality we deduce a similar conclusion for subsolutions ulsc(B¯δ), where Bδ={|x-x0|<δ} such that

𝐚 ( D 2 u ) f ( x ) in  B δ

in the viscosity sense. Let w be a concave function such that

w ( x 0 ) = u ( x 0 ) , w ( x ) u ( x ) in  B δ .

For sufficiently small ε(0,ε0) and any function f, bounded above, we have

( x ) - 1 2 C ε ( sup B δ f + ) | x - x 0 | 2 w ( x ) ( x ) in  B ε δ ,

where (x) is the supporting hyperplane for w at x0. In particular, there exists a concave paraboloid of opening Cε touching the graph of w from below.

Here ε0>0 depends on a lower bound for a1 and an upper bound for a^1 defined in Section 3.3, and Cε1a1 as ε0. Therefore, when u is second-order differentiable and f is continuous at x=x0, we get

(4.14) λ 1 - ( D 2 w ( x 0 ) ) 1 a 1 f - ( x 0 ) .

Using [25, Lemma 3.5], we deduce from the above that ΓuC1,1(B¯d). Hence Γu is second-order differentiable a.e. in B¯d and (4.4) holds for Γu+ in Bd.

If u0 on Ω, then we have

(4.15) sup B d u + C n d ( Γ u + = u | det D 2 Γ u + ( x ) | d x ) 1 n .

Reasoning as in the proof of [25, Theorem 3.6], that is observing that the upper contact points are in Ω and Γu+ is second-order differentiable a.e. on {Γu+=u}, where f is continuous and therefore, by (4.14)

(4.16) | det D 2 ( Γ u + ( x ) ) | ( λ 1 - ( D 2 Γ u + ( x ) ) ) n 1 a 1 n ( f - ( x ) ) n .

Estimating (4.15) with (4.16), we get the ABP estimate (4.6) for u0 on Ω. Passing to u-supΩu, which is 0 on Ω, we conclude that (4.6) holds. ∎

5 Harnack Inequality and Cα Estimates

The Harnack inequality, classically related to the mean properties of the Laplace operator, is a powerful nonlinear technique for regularity in the framework of fully nonlinear equations. We refer to [47] for solutions of linear uniformly elliptic equations in Sobolev spaces, to [74] for quasi-linear uniformly elliptic equations and to [24, 25] for viscosity solutions of fully nonlinear equations. See also [4, 52, 61] for further contributions.

In order to prove the Harnack inequality for non-negative solutions and the related local estimates for subsolutions and non-negative supersolutions, respectively known in literature (see for instance [47]) as the local maximum principle and the weak Harnack inequality, we could employ the same strategy of [25, Chapter 4].

A quicker way, sufficient for the applications, is based on inequalities (3.7) obtained in Section 3. The results are given in cubes, and here 𝒬 is a cube of n of edge centered at the origin, i.e.

Q = { | x i | < 2 : i = 1 , , n } ,

but they could be equivalently stated in balls.

Theorem 5.1 (Local Maximum Principle).

Let MaA1. Let u be a viscosity subsolution of the equation

𝐚 ( D 2 u ) = f

in Q1, where f is continuous and bounded. Then

(5.1) sup Q 1 / 2 u C p ( u + L p ( Q 2 / 3 ) + f - L n ( Q 1 ) ) ,

where Cp is a constant depending only on n, p, a1 and a~.

Proof.

In view of inequalities (3.7), we have a1/n,a~n+(D2u)𝐚(D2u)f(x)-f-(x), and therefore we can apply [25, Theorem 4.8 (2)] to obtain (5.1). ∎

Theorem 5.2 (Weak Harnack Inequality).

Let MaAn. Let u0 be a viscosity supersolution of the equation

𝐚 ( D 2 u ) = f

in Q1, where f is continuous and bounded. Then

(5.2) u L p 0 ( Q 2 / 3 ) C 0 ( inf Q 3 / 4 u + f + L n ( Q 1 ) ) ,

where p0>0 and C0 are universal constants, depending only on n, p, an and a~.

Proof.

In view of inequalities (3.7), we have an/n,a~n-(D2u)𝐚(D2u)f(x)f+(x), and therefore we can apply [25, Theorem 4.8 (1)] to obtain (5.2). ∎

The proof of Theorem 1.2 (Harnack inequality) follows at once.

Proof of Theorem 1.2.

Let p0>0 be the exponent of Theorem 5.2. From (5.2) and (5.1) it follows that

sup Q 1 / 2 u C p 0 ( u L p 0 ( Q 2 / 3 ) + f - L n ( Q 1 ) ) C p 0 ( C 0 ( inf Q 3 / 4 u + f + L n ( Q 1 ) ) + f - L n ( Q 1 ) ) ,

which yields the result. ∎

From the Harnack inequality, in a standard way, using the technique for the proof of [25, Proposition 4.10] and [47, Lemma 8.23], the following Hölder regularity results and Cα interior estimates can be obtained. We give the result with concentric balls B1 and B1/2 of radius 1 and 12, respectively.

Theorem 5.3 (Interior Hölder Continuity).

Let MaA. Let u be a viscosity solution of the equation

𝐚 ( D 2 u ) = f

in B1, where f is continuous and bounded. Then uCα(B¯1/2) and

u C α ( B ¯ 1 / 2 ) C ( u L ( B 1 ) + f L n ( B 1 ) ) ,

where C is a positive constant depending only on n, a*=min(a1,an) and a~=1n(a1++an).

Global Hölder estimates can be proved for domains with the uniform exterior sphere condition (s), see Section 1, via the boundary Hölder estimates of the lemma below. We adopt the following notations, for the Hölder seminorm (0<γ<1)) of a function h:D in a subset D of n:

(5.3) [ h ] β , D = sup x , y D x y | h ( x ) - h ( y ) | | x - y | β .

Lemma 5.4.

Let MaA¯ and let u be a viscosity solution of the equation Ma(D2u)=f in a bounded domain Ω, where f is continuous and bounded. We assume that Ω satisfies a uniform exterior sphere condition (s) with radius R>0, and u=g on Ω.

  1. If g C β ( Ω ) with β ( 0 , 1 ] , then

    sup x Ω y Ω u ( x ) - u ( y ) | x - y | β 2 C ( [ g ] β , Ω + f - L ( Ω ) )

    with C > 0 depending only on n, a~, R and β.

  2. Assume in addition that Ω has a uniform Lipschitz boundary with Lipschitz constant L. If gC1,β(Ω) with β[0,1), then

    (5.4) sup x Ω y Ω u ( x ) - u ( y ) | x - y | 1 2 ( 1 + β ) C ( [ g ] 1 , Ω + [ D g ] β , Ω + f - L ( Ω ) )

    with C > 0 depending only on n, a~, R, L and β.

If u is a viscosity supersolution, then (i) and (ii) hold with u(y)-u(x) and f+L(Ω) instead of u(x)-u(y) and f-L(Ω), respectively.

Proof.

We treat in detail the case of subsolutions. The result for subsolutions will follow by duality. Therefore, suppose that uusc(Ω¯) is a subsolution of the equation 𝐚(D2u)=f in Ω such that u=g on Ω. Let yΩ and BR a ball of radius R, centered at x0n, such that yBR and Ω¯B¯R, according to (s). Supposing, as we may, y=(0,,0,0) and x0=(0,,0,R); then B¯R is described by the inequality x12++xn-12+(xn-R)2R2. It follows that

(5.5) | x | 2 2 R x n , x Ω ¯ .

Case (i). By assumption on g and (5.5), we have

| g ( x ) | [ g ] β , Ω | x | β , x Ω ¯ .

To simplify, we may suppose g(0)=0, so that in particular

(5.6) g ( x ) [ g ] β , Ω | x | β ( 2 R ) β 2 [ g ] β , Ω x n β 2 , x Ω ¯ .

Next, we define

φ ( x ) = C 1 ( [ g ] β , Ω + ε ) x n β 2 - 1 2 | 𝐚 | f - L ( Ω ) | x | 2 ,

where ε is any positive number and

C 1 ( 2 R ) β 2 + ( 2 R ) 2 - β 2 2 | 𝐚 | f - [ g ] β , Ω + ε .

Thus from (5.6)

u ( x ) = g ( x ) φ ( x ) on  Ω .

Moreover, φ is a supersolution in Ω:

𝐚 ( D 2 φ ) = a 1 C 1 ( [ g ] β , Ω + ε ) β 2 ( β 2 - 1 ) x n β 2 - 2 - f - L ( Ω ) - f - ( x ) in  Ω .

By the comparison principle u(x)φ(x) for all xΩ, from which

u ( x ) | x | β 2 C 2 ( [ g ] β , Ω + f - L ( Ω ) ) .

Then for an arbitrary yΩ we have

u ( x ) - u ( y ) | x - y | β 2 C ( [ g ] β , Ω + f - L ( Ω ) ) ,

from which (5.4) follows.

Case (ii). By assumption on g and on Ω, we have

| g ( x ) - g ( y ) - D g ( y ) , x - y | C 1 [ D g ] β , Ω | x - y | 1 + β , x Ω ¯ ,

where C1 is a positive constant depending on the Lipschitz constant L for Ω. We adopt the above simplifications: y=(0,,0,0), x0=(0,,0,R), g(y)=0, so that in particular

g ( x ) D g ( 0 ) , x + C 1 [ D g ] β , Ω | x | 1 + β , x Ω ¯ .

Therefore

(5.7) g ( x ) D g ( 0 ) , x + C 2 [ D g ] β , Ω x n 1 2 ( 1 + β ) , x Ω ¯ ,

where C2 is a positive constant depending on L, R and β. Next, we define

(5.8) φ ( x ) = D g ( 0 ) , x + C 3 ( [ D g ] β , Ω + ε ) x n 1 2 ( 1 + β ) - 1 2 | 𝐚 | f - L ( Ω ) ,

where ε is any positive number and

C 3 C 2 + ( 2 R ) 1 - β 2 2 | 𝐚 | f - [ D g ] β , Ω + ε .

Therefore by (5.7):

u ( x ) = g ( x ) φ ( x ) on  Ω .

Moreover, φ is a supersolution in Ω:

𝐚 ( D 2 φ ) = a 1 C 3 ( [ D g ] β , Ω + ε ) β + 1 2 β - 1 2 x n β + 1 2 - 2 - f - L ( Ω ) - f - ( x ) in  Ω .

By the comparison principle, we get uφ in Ω, and therefore by (5.8):

u ( x ) | x | 1 2 ( 1 + β ) | D g ( 0 ) | | x | 1 2 ( 1 - β ) + C 4 ( [ D g ] β , Ω + f - L ( Ω ) ) ( 4 R ) 1 2 ( 1 - β ) [ D g ] 0 , Ω + C 4 ( [ D g ] β , Ω + f - L ( Ω ) ) .

Then for arbitrary yΩ we have

u ( x ) - u ( y ) | x - y | 1 2 ( 1 + β ) ( 4 R ) 1 2 ( 1 - β ) [ D g ] 0 , Ω + C 4 ( [ D g ] β , Ω + f - L ( Ω ) )

for all xΩ, from which (5.4) follows. ∎

We are ready to show the global Hölder estimates of Theorem 1.3.

Proof of Theorem 1.3.

Let α(0,1) be the Hölder exponent of Theorem 5.3. From the boundary Hölder estimates of Lemma 5.4 we deduce an estimate of type

(5.9) | u ( x ) - u ( y ) | C ( g , f ) | x - y | γ b , x Ω , y Ω ,

where γb=β2 in case (i) and γb=12(1+β) in case (ii). We want to show a global Hölder estimate with exponent γ=min(α,γb). For proving the result we follow the same lines of [25, Proposition 4.13]. Thus, for x,yΩ we set dx=dist(x,Ω)=|x-x0|, dy=dist(y,Ω)=|y-y0| for x0,y0Ω, and suppose dydx. Here the constants Ci will depend at most on n, a*, a~, R, L and β.

(i) Suppose |x-y|dx2. Since yB¯dx/2(x)Bdx(x)Ω, we can apply Theorem 5.3 properly scaled to the function u(x)-u(x0), and then the Hölder boundary estimate (5.9) obtaining

| u ( x ) - u ( y ) | | x - y | α d x α C 1 u - u ( x 0 ) L ( B d x ( x ) ) C 2 K d x γ b .

Recall that γα. Since dx|x-y|2, from this we get

| u ( x ) - u ( y ) | | x - y | γ d x γ u ( x ) - u ( y ) | x - y | α d x α C 2 K d x γ b .

Since also γγb,

(5.10) | u ( x ) - u ( y ) | | x - y | γ C 2 K d x γ b - γ C 2 K d γ b - γ C 3 K .

(ii) Suppose now |x-y|dx2. Since dydx2|x-y| and |x0-y0|dx+|x-y|+dy, it follows from (5.9) that

(5.11) | u ( x ) - u ( y ) | | u ( x ) - u ( x 0 ) | + | u ( x 0 ) - u ( y 0 ) | + | u ( y 0 ) - u ( y ) | C 4 K ( d x γ b + | x 0 - y 0 | γ b + d y γ b ) C 5 K | x - y | γ .

From (5.10) and (5.11), letting C¯=max(C3,C5), we deduce the desired estimate [u]γ,ΩC¯K. ∎

In some cases, when the weights ai are concentrated near the one of the extremal eigenvalues, we obtain an explicit interior Hölder exponent.

Lemma 5.5.

Let MaA¯ be such that a1a^1 (resp. ana^n). Suppose that uusc(B1) (resp. ulsc(B1)) is a viscosity subsolution (resp. supersolution) of the equation Ma(D2u)=f in B1, a ball of radius 1, and f is continuous and bounded above (resp. below) in B1. Then uCα(B1) and the following interior Cα estimate holds:

(5.12) [ u ] α , B 1 / 2 C ( u L ( B 1 ) + f - L ( B 1 ) )

resp.

[ u ] α , B 1 / 2 C ( u L ( B 1 ) + f + L ( B 1 ) ) ,

where B1/2 is a ball of radius 12 concentric with B1, α=1-a^1a1 (resp. α=1-a^nan), and C a positive constant depending on n, a1 and a^1 (resp. an and a^n).

Proof.

We only treat the case of subsolution, when a1a^1. The case of supersolutions, when ana^n, will follow by duality. We assume that the balls B1 and B1/2 are centered at 0. Then we take x,x′′B1/2, and consider the ball B1/2(x). We note that on B1/2(x):

u ( x ) - u ( x ) 2 u L ( B 1 ) 2 1 + α u L ( B 1 ) | x - x | α .

Next, we define

φ ( x ) = C 1 u L ( B 1 ) | x | α - 1 2 | 𝐚 | f - L ( B 1 ) | x | 2 ,

where C1=21+α+12|𝐚|f-u (in the nontrivial case u0). Thus on B1/2(x):

(5.13) u ( x ) - u ( x ) 2 1 - α u L ( B 1 ) | x - x | α φ ( x - x ) .

On the other hand,

(5.14) 𝐚 ( D 2 φ ( x - x ) ) = C 1 ( a 1 ( α - 1 ) + a ^ 1 ) | x - x | α - 2 - f - L ( Ω ) - f - ( x ) .

By (5.14) and (5.13), using the comparison principle, we get u(x)-u(x)φ(x-x) in B1/2(x), from which in particular

u ( x ′′ ) - u ( x ) C ( u L ( B 1 ) + f - L ( B 1 ) ) | x ′′ - x | α .

Interchanging the role of x and x′′, we get (5.12). ∎

Combining Lemma 5.4 with Lemma 5.5, we obtain the global estimates of Theorem 1.4.

Proof of Theorem 1.4.

To obtain (1.4) and (1.5), it is sufficient to follow the proof of Theorem 1.3. The above estimates (1.6) and (1.7) are in particular obtained from this proof taking γ=α. We use once more the boundary Hölder estimates of Lemma 5.4, with β=2α for (1.6) and β=2α-1 for (1.7), as there. But we use here the interior Cα estimates of Lemma 5.5, instead of Theorem 5.3. ∎

Remark 5.6.

Note that Theorem 1.4 provides Lipschitz estimates only in the case a^1=0 and a^n=0, corresponding to the operators 𝐞1[u]=λ1[u] and 𝐞n[u]=λn[u]. See for instance [12].

Remark 5.7.

Asking for higher regularity of viscosity solutions, we cannot expect viscosity solutions more regular than C2. Indeed, we may consider the function u:=x12+ω(x2)-x32, in 3, where ω is a C2 function but no more regular. The same regularity holds for u. Assuming in addition |ω′′(x2)|<2, by a straightforward computation we get

D 2 u ( x ) = Diag [ ( 2 , ω ′′ ( x 2 ) , - 2 ) ] ,

so that (D2u(x)=0. So we have found a solution uC2, which does not belong to any C2,β space, β(0,1).

6 The Strong Maximum Principle

The strong maximum principle for an elliptic operator F, such that F(0)=0, means that a subsolution of the equation F[u]=0 in an open set Ω cannot have a maximum at a point of Ω unless to be constant. Analogously, the strong minimum principle means that a supersolution u cannot have a minimum at a point of Ω unless u is constant.

One of the most elegant proof of the strong maximum principle, also known for this reason as the celebrated Hopf maximum principle [51], is based on boundary point lemma, which we establishes here below for the class of weigthed partial trace operators 𝐚. To obtain a strong maximum principle, it is sufficient to state this lemma just for a ball.

Lemma 6.1 (Hopf Boundary Point Lemma).

Let uusc(B¯) be a viscosity subsolution of the equation Ma[u]=0 in a ball B, with MaA1. Let x0B. If u(x0)>u(x) for all xB, then the outer normal derivative of u at x0, if it exists, satisfies the strict inequality

(6.1) u ν ( x 0 ) > 0 .

On the other hand, let ulsc(B¯) be a viscosity supersolution of the equation Ma[u]=0 in a ball B, with MaAn. If u(x0)<u(x) for all xB, then the outer normal derivative of u at x0, if it exists, satisfies the strict inequality

(6.2) u ν ( x 0 ) < 0 .

Proof.

We just prove the theorem for subsolutions. We may suppose that B is centered at the origin, i.e. B={|x|<R} for R>0. Arguing as in [47, Section 3.2], and considering 0<ρ<R, we introduce the radial test function v(x)=e-αr2-e-αR2, with r=|x|. By direct computation, see Remark 3.6 in Section 3, we get

𝐚 ( D 2 v ) 2 α ( a 1 ( 2 α ρ 2 - 1 ) - i = 2 n a i ) e - α r 2

for rρ and [v]0 for α>0 large enough. Since u(x0)-u(x)>0 on |x|=ρ, there is a constant ε>0 such that u(x0)-u(x)-εv(x)0 on |x|=ρ, as well as on |x|=R. Therefore εv(x)u(x0)-u(x) on the boundary of the annulus Aρ,R={ρ<|x|<R}. By the comparison principle, the same inequality holds in Aρ,R. In fact, 𝐚[εv]=ε𝐚[v]0, by positive homogeneity, and by duality 𝐚[u(x0)-u]0, so that εv and u(x0)-u are respectively a subsolution and a supersolution in A, and we can apply Theorem 3.1 to deduce that

u ( x 0 ) - u ( x ) ε v ( x ) for all  x A .

Taking x=x0-tx0R in the latter inequality, dividing by t>0 and letting t0+, we get

u ν ( x 0 ) - ε d d r ( e - α r 2 ) | r = R = 2 ε α R e - α R 2 ,

which proves (6.1). ∎

Following [47], we remark that, whether or not the normal derivative exists, we have instead of (6.1) and (6.2), respectively, the inequalities

lim inf x x 0 x Σ u ( x 0 ) - u ( x ) | x - x 0 | > 0

and

lim inf x x 0 x Σ u ( x 0 ) - u ( x ) | x - x 0 | < 0 ,

where Σ is any circular cone of vertex x0 and opening less than π with axis along the normal direction at the boundary point x0.

The Hopf boundary point lemma can be used to prove the strong maximum principle for classical subsolutions or viscosity subsolutions which are differentiable. A strong maximum principle, valid also for nonsmooth viscosity solutions, can be obtained through the weak Harnack inequality of Lemma 5.2.

For a detailed discussion on the strong maximum principle, we refer to the paper [70] and the papers quoted therein. In the case of fully nonlinear elliptic operators, see for instance [6, 7]. For further references see [76, 36, 69, 62].

Theorem 6.2 (Strong Maximum Principle).

Let u be a non-negative continuous viscosity supersolution of the equation Ma(D2u)=0, with MaAn, in a domain Ω of Rn. If u has a minimum m at some point x0Ω, then um in Ω. Similarly, let u be a continuous viscosity subsolution of the equation Ma(D2u)=0, with MaA1. If u has a maximum M at x0Ω, then uM in Ω.

Proof.

For the proof in the case of differentiable solutions u, based on the Hopf lemma, we refer to the proof of [47, Theorem 3.5].

Concerning viscosity supersolutions (strong minimum principle), let A={xΩ:u(x)=m} and B=ΩA, so that AB=Ω, AB= with A and B is open. Moreover, we claim that A is also open. Recalling that Ω is a open connected set, then B=, otherwise we would have a contradiction. Then Ω=A, and the first part of the theorem is proved.

We are left with proving that A is open. Let x0A, that is u(x0)=m, and suppose that the cube Q of side centered at x0 is contained in Ω. By the weak Harnack inequality (5.2), properly scaled and applied to u-m0, we have

u - m L p 0 ( Q 2 / 3 ) C 0 inf Q 3 / 4 ( u - m ) = u ( x 0 ) - m = 0 .

The function u(x)-m is constant in Q2/3 and by continuity u(x)-m=u(x0)-m=0 for all xQ2/3, so that Q2/3A. This shows that A is open, thereby proving the claim and concluding the proof of the first part.

In the case of viscosity subsolutions (strong maximum principle), we argue in a similar manner, considering the set A={xΩ:u(x)=M}. By duality v=M-u is a non-negative supersolution of the equation 𝐚(D2v)=0 such that v(x0)=0. Then by the case of supersolutions v is constant and therefore u(x)=M for all xΩ. ∎

It follows that for elliptic operators 𝐚𝒜 both the strong maximum and minimum principle are satisfied.

It is plain that the strong maximum principle implies the weak maximum principle (see Section 2) in bounded domains. This is no more true in unbounded domains, where the strong maximum principle may hold while the weak maximum principle fails to hold. An elementary example of this fact is given by the function u(x1,x2)=x1x2, which is harmonic in the whole plane, and therefore satisfies the strong maximum principle in all domains of 2, but is positive in the quarter plane Ω=+×+ and zero on Ω, so that the weak maximum principle does not hold in Ω.

Turning to bounded domains, as observed in Section 2, it is sufficient that 𝐚𝒜¯ to have both the weak maximum and minimum principle. Theorem 6.2 requires instead 𝐚𝒜 to have the strong maximum and minimum principle hold together.

Actually, the strong maximum and minimum principle may fail when 𝐚𝒜¯, but 𝐚𝒜. In fact, let us consider the partial trace operator 𝒫k+ defined above for 1kn-1: the non-constant function u(x)=1+sinx1 has a maximum M=2 inside the cube ]0,π[n, even though 𝒫k+(D2u)=0 in ]0,π[n. Similarly, u(x) is non-negative in the cube ]-π,0[n and has a zero inside, even though 𝒫k-(D2u)=0 in the cube ]-π,0[n for 1kn-1.

From the proof of Theorem 6.2, the weak Harnack inequality, which would imply the strong minimum principle, fails to hold in general for the partial trace operator 𝒫k- as soon as k<n. Analogously, the Harnack inequality, which would imply both the strong maximum and minimum principle, fails to hold in general for the partial trace operators 𝒫k± as soon as k<n.

7 Liouville Theorems

A direct application of the Harnack inequality yields in a standard fashion the following Liouville result for entire solutions, defined in the whole n. See for instance [4].

Theorem 7.1 (Liouville Theorem).

Let MaA. If u is an entire viscosity solution of the equation M(D2u)=0 which is bounded above or below, then u is constant.

It is well known that the above Liouville theorem holds in a stronger unilateral version for the Laplace operator in dimension n=2, where instead of solutions, bounded above or below, we may consider subsolutions bounded above and supersolutions bounded below. This is due to the fact that the fundamental solutions are of logarithmic type. See [67, Theorem 29].

On the other hand, this is no longer true in higher dimension. For instance, the function

(7.1) u ( x ) = { - 1 8 ( 15 - 10 | x | 2 + 3 | x | 4 ) for  | x | 1 , - 1 | x | for  | x | > 1 ,

is a non-constant subharmonic, bounded function in 3. We refer to [67, Chapter 2, Section 12].

As well, the unilateral Liouville theorem does not hold for general elliptic operators even in dimension n=2. Actually, as soon as λ<Λ we can find subsolutions u, bounded above, of the equation λ,Λ+(D2u)=0 in 2. For instance, the function (7.1), regarded as a function of x2, is a subsolution of the equation λ,2λ+(D2u)=0 in 2.

Therefore, the uniform ellipticity is not sufficient by itself to guarantee such an unilateral Liouville property, even in dimension 2.

However, for particular uniformly elliptic operators as the minimal Pucci operators λ,Λ-, which are suitably smaller than the Laplace operator, precisely when n1+Λλ, the Liouville property still holds for subsolutions, bounded above (see [35]). We thank Dr. Goffi for drawing our attention to the latter issue during a workshop where the results of this paper have been announced for the first time.[1]

We notice here that the same is true for the min-max operator (X)=λ1(X)+λn(X), and more generally for the operators 𝐚𝒜¯ such that a1=a^1. See also [13].

Theorem 7.2 (Hadamard Three Circles Theorem).

Let uusc(Rn{0}) be a subsolution of the equation

𝐚 ( D 2 u ) = 0

with MaA¯ such that a1=a^1. Setting M(r)=maxBru for r>0, we have that M(r) is a convex function of logr, namely for 0<r1<r2,

(7.2) M ( r ) M ( r 1 ) log ( r 2 r ) + M ( r 2 ) log ( r r 1 ) log ( r 2 r 1 ) , r 1 r r 2 .

Proof.

Actually, by Remark 3.6 the function

φ ( x ) = M ( r 1 ) log ( r 2 | x | ) + M ( r 2 ) log ( | x | r 1 ) log ( r 2 r 1 )

satisfies the equation 𝐚(D2φ)=0 as linear combination of a constant and log|x| with non-negative coefficients, by positive homogeneity. Moreover, u(x)φ(x) on the boundary of the annulus Ar1,r2={r1<|x|<r2}. From the comparison principle (Theorem 3.1) in Ar1,r2 then we obtain (7.2). ∎

Note that =𝐞1+𝐞n satisfies the condition a1=a^1 and therefore the Hadamard Three Circles Theorem holds for min-max operator .

From Theorem 7.2 it follows that such operators satisfy the same Liouville unilateral property which holds for the Laplace operator in dimension n=2: if a1=a^1, the constant functions are the only viscosity subsolutions, bounded above, of the equation 𝐚=0 in n.

Theorem 7.3 (Unilateral Liouville Property).

Let MaA¯ be such that a1=a^1. Let u be a viscosity subsolution of the equation M(D2u)=0 in Rn{0}, which is bounded above. Then u is constant. If u is a subsolution in the whole Rn bounded above, then the same conclusion holds if a1>a^1.

On the other hand, suppose MaA¯ such that an=a^n. Let v be a viscosity supersolution of the equation M(D2v)=0 in Rn{0} which is bounded below. Then v is constant. If u is a supersolution in the whole Rn, bounded below, then the same conclusion holds if an>a^n.

Proof.

Let 𝐚𝒜¯ with a1=a^1. Reasoning as in [67, Section 12], we take alternatively the limits as r10+ and r2 in (7.2). So we get

M ( r ) M ( r 2 ) for  r r 2 , M ( r ) M ( r 1 ) for  r r 1

concluding that M(r1)=M(r2) for arbitrary pairs of positive numbers r1,r2. Then M(r) is constant, and by the strong maximum principle u is in turn a constant function.

Supposing a1>a^1, for any arbitrary x0n we set

v ( x ) = u ( x 1 ) + C | x - x 1 | γ 2 ,

with γ2=1-a^1a1 and C0 to be determined, recalling that by (3.8) we have 𝐚(D2v)=0 in n{0}. We will compare the entire subsolution bounded above, say u(x)M, in every punctured ball BR(x1){x1}, noting that u(x1)=v(x1) and on BR(x1) we have

u ( x ) M u ( x 1 ) + C | x - x 1 | γ 2

choosing C=(M-u(x1)R-γ2. Using the comparison principle (Theorem 3.1), we infer that this inequality holds in BR(x1). Letting R, we will have u(x)u(x1) for all xn. The same holds true for any other x2n so that u(x1)=u(x2) for all x1,x2n.

Concerning supersolutions v, bounded below, of the same equation in n, it is sufficient to note that by duality the function u=-v is a subsolution, bounded above, of the equation 𝐚[u]=0, where 𝐚=(an,an-1,,a1), in n, and then to use the result proved before for subsolutions. ∎


Communicated by Enrico Valdinoci


Award Identifier / Grant number: project 2018

Award Identifier / Grant number: project 2019

Funding statement: Fausto Ferrari is partially funded by INDAM-GNAMPA project 2018: Costanti critiche e problemi asintotici per equazioni completamente non lineari and INDAM-GNAMPA project 2019: Proprietà di regolarità delle soluzioni viscose con applicazioni a problemi di frontiera libera.

References

[1] A. D. Aleksandrov, Certain estimates for the Dirichlet problem (in Russian), Dokl. Akad. Nauk. SSSR 134 (1960), 1001–1004; translation in Soviet Math. Dokl. 1 (1960), 1151–1154. Search in Google Scholar

[2] A. D. Aleksandrov, Uniqueness conditions and bounds for the solution of the Dirichlet problem (in Russian), Vestnik Leningrad. Univ. Ser. Mat. Meh. Astronom. 18 (1963), no. 3, 5–29; translation in Amer. Math. Soc. Transl. (2) 68 (1968), 89–119. Search in Google Scholar

[3] M. E. Amendola, G. Galise and A. Vitolo, Riesz capacity, maximum principle, and removable sets of fully nonlinear second-order elliptic operators, Differential Integral Equations 26 (2013), no. 7–8, 845–866. 10.57262/die/1369057820Search in Google Scholar

[4] M. E. Amendola, L. Rossi and A. Vitolo, Harnack inequalities and ABP estimates for nonlinear second-order elliptic equations in unbounded domains, Abstr. Appl. Anal. 2008 (2008), Article ID 178534. 10.1155/2008/178534Search in Google Scholar

[5] I. J. Bakel’man, On the theory of quasilinear elliptic equations (in Russian), Sibirsk. Mat. Z. 2 (1961), 179–186. Search in Google Scholar

[6] M. Bardi and F. Da Lio, Propagation of maxima and strong maximum principle for viscosity solutions of degenerate elliptic equations. I. Convex operators, Nonlinear Anal. 44 (2001), no. 8, 991–1006. 10.1016/S0362-546X(99)00317-XSearch in Google Scholar

[7] M. Bardi and F. Da Lio, Propagation of maxima and strong maximum principle for viscosity solutions of degenerate elliptic equations. II. Concave operators, Indiana Univ. Math. J. 52 (2003), no. 3, 607–627. 10.1512/iumj.2003.52.2147Search in Google Scholar

[8] M. Bardi and P. Mannucci, On the Dirichlet problem for non-totally degenerate fully nonlinear elliptic equations, Commun. Pure Appl. Anal. 5 (2006), no. 4, 709–731. 10.3934/cpaa.2006.5.709Search in Google Scholar

[9] H. Berestycki, L. Nirenberg and S. R. S. Varadhan, The principal eigenvalue and maximum principle for second-order elliptic operators in general domains, Comm. Pure Appl. Math. 47 (1994), no. 1, 47–92. 10.1002/cpa.3160470105Search in Google Scholar

[10] I. Birindelli, I. Capuzzo Dolcetta and A. Vitolo, ABP and global Hölder estimates for fully nonlinear elliptic equations in unbounded domains, Commun. Contemp. Math. 18 (2016), no. 4, Article ID 1550075. 10.1142/S0219199715500753Search in Google Scholar

[11] I. Birindelli and F. Demengel, 𝒞1,β regularity for Dirichlet problems associated to fully nonlinear degenerate elliptic equations, ESAIM Control Optim. Calc. Var. 20 (2014), no. 4, 1009–1024. 10.1051/cocv/2014005Search in Google Scholar

[12] I. Birindelli, G. Galise and H. Ishii, A family of degenerate elliptic operators: maximum principle and its consequences, Ann. Inst. H. Poincaré Anal. Non Linéaire 35 (2018), no. 2, 417–441. 10.1016/j.anihpc.2017.05.003Search in Google Scholar

[13] I. Birindelli, G. Galise and F. Leoni, Liouville theorems for a family of very degenerate elliptic nonlinear operators, Nonlinear Anal. 161 (2017), 198–211. 10.1016/j.na.2017.06.002Search in Google Scholar

[14] P. Blanc, C. Esteve and J. D. Rossi, The evolution problem associated with eigenvalues of the Hessian, preprint (2019), https://arxiv.org/abs/1901.01052. 10.1112/jlms.12363Search in Google Scholar

[15] P. Blanc and J. D. Rossi, Games for eigenvalues of the Hessian and concave/convex envelopes, J. Math. Pures Appl. (9) 127 (2019), 192–215. 10.1016/j.matpur.2018.08.007Search in Google Scholar

[16] R. Buckdahn, P. Cardaliaguet and M. Quincampoix, Some recent aspects of differential game theory, Dyn. Games Appl. 1 (2011), no. 1, 74–114. 10.1007/s13235-010-0005-0Search in Google Scholar

[17] R. Buckdahn and J. Li, Stochastic differential games and viscosity solutions of Hamilton–Jacobi–Bellman–Isaacs equations, SIAM J. Control Optim. 47 (2008), no. 1, 444–475. 10.1137/060671954Search in Google Scholar

[18] X. Cabré, On the Alexandroff–Bakel’man–Pucci estimate and the reversed Hölder inequality for solutions of elliptic and parabolic equations, Comm. Pure Appl. Math. 48 (1995), no. 5, 539–570. 10.1002/cpa.3160480504Search in Google Scholar

[19] X. Cabré and L. A. Caffarelli, Regularity for viscosity solutions of fully nonlinear equations F(D2u)=0, Topol. Methods Nonlinear Anal. 6 (1995), no. 1, 31–48. 10.12775/TMNA.1995.030Search in Google Scholar

[20] V. Cafagna and A. Vitolo, On the maximum principle for second-order elliptic operators in unbounded domains, C. R. Math. Acad. Sci. Paris 334 (2002), no. 5, 359–363. 10.1016/S1631-073X(02)02267-7Search in Google Scholar

[21] L. Caffarelli, M. G. Crandall, M. Kocan and A. Świȩch, On viscosity solutions of fully nonlinear equations with measurable ingredients, Comm. Pure Appl. Math. 49 (1996), no. 4, 365–397. 10.1002/(SICI)1097-0312(199604)49:4<365::AID-CPA3>3.0.CO;2-ASearch in Google Scholar

[22] L. Caffarelli, Y. Li and L. Nirenberg, Some remarks on singular solutions of nonlinear elliptic equations III: viscosity solutions including parabolic operators, Comm. Pure Appl. Math. 66 (2013), no. 1, 109–143. 10.1002/cpa.21412Search in Google Scholar

[23] L. Caffarelli, Y. Y. Li and L. Nirenberg, Some remarks on singular solutions of nonlinear elliptic equations. I, J. Fixed Point Theory Appl. 5 (2009), no. 2, 353–395. 10.1007/s11784-009-0107-8Search in Google Scholar

[24] L. A. Caffarelli, Interior a priori estimates for solutions of fully nonlinear equations, Ann. of Math. (2) 130 (1989), no. 1, 189–213. 10.2307/1971480Search in Google Scholar

[25] 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

[26] I. Capuzzo Dolcetta, F. Leoni and A. Vitolo, On the inequality F(x,D2u)f(u)+g(u)|Du|q, Math. Ann. 365 (2016), no. 1–2, 423–448. 10.1007/s00208-015-1280-2Search in Google Scholar

[27] I. Capuzzo Dolcetta and A. Vitolo, C1,α and Glaeser type estimates, Rend. Mat. Appl. (7) 29 (2009), no. 1, 17–27. Search in Google Scholar

[28] I. Capuzzo Dolcetta and A. Vitolo, Glaeser’s type gradient estimates for non-negative solutions of fully nonlinear elliptic equations, Discrete Contin. Dyn. Syst. 28 (2010), no. 2, 539–557. 10.3934/dcds.2010.28.539Search in Google Scholar

[29] I. Capuzzo Dolcetta and A. Vitolo, The weak maximum principle for degenerate elliptic operators in unbounded domains, Int. Math. Res. Not. IMRN 2018 (2018), no. 2, 412–431. 10.1093/imrn/rnw249Search in Google Scholar

[30] I. Capuzzo Dolcetta and A. Vitolo, Directional ellipticity on special domains: Weak maximum and Phragmén–Lindelöf principles, Nonlinear Anal. 184 (2019), 69–82. 10.1016/j.na.2019.02.004Search in Google Scholar

[31] S. Cho and M. Safonov, Hölder regularity of solutions to second-order elliptic equations in nonsmooth domains, Bound. Value Probl. 2007 (2007), Article ID 57928. 10.1155/2007/57928Search in Google Scholar

[32] M. G. Crandall, Viscosity solutions: A primer, Viscosity Solutions and Applications, Lecture Notes in Math. 1660, Springer, Berlin (1997), 1–47. 10.1007/BFb0094294Search in Google Scholar

[33] 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

[34] M. G. Crandall and A. Świȩch, A note on generalized maximum principles for elliptic and parabolic PDE, Evolution Equations, Lecture Notes Pure Appl. Math. 234, Dekker, New York (2003), 121–127. 10.1201/9780429187599-8Search in Google Scholar

[35] A. Cutrì and F. Leoni, On the Liouville property for fully nonlinear equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 17 (2000), no. 2, 219–245. 10.1016/s0294-1449(00)00109-8Search in Google Scholar

[36] F. Da Lio and B. Sirakov, Symmetry results for viscosity solutions of fully nonlinear uniformly elliptic equations, J. Eur. Math. Soc. (JEMS) 9 (2007), no. 2, 317–330. 10.4171/JEMS/81Search in Google Scholar

[37] F. Ferrari, An application of the theorem on Sums to viscosity solutions of degenerate fully nonlinear equations, Proc. Roy. Soc. Edinburgh Sect. A (2019), 10.1017/prm.2018.147. 10.1017/prm.2018.147Search in Google Scholar

[38] F. Ferrari, Q. Liu and J. Manfredi, On the characterization of p-harmonic functions on the Heisenberg group by mean value properties, Discrete Contin. Dyn. Syst. 34 (2014), no. 7, 2779–2793. 10.3934/dcds.2014.34.2779Search in Google Scholar

[39] F. Ferrari and A. Pinamonti, Characterization by asymptotic mean formulas of q-harmonic functions in Carnot groups, Potential Anal. 42 (2015), no. 1, 203–227. 10.1007/s11118-014-9430-9Search in Google Scholar

[40] F. Ferrari and E. Vecchi, Hölder behavior of viscosity solutions of some fully nonlinear equations in the Heisenberg group, Topol. Methods Nonlinear Anal., to appear. 10.12775/TMNA.2019.073Search in Google Scholar

[41] W. H. Fleming and M. Nisio, Differential games for stochastic partial differential equations, Nagoya Math. J. 131 (1993), 75–107. 10.1017/S0027763000004554Search in Google Scholar

[42] W. H. Fleming and R. W. Rishel, Deterministic and Stochastic Optimal Control, Appl. Math. 1, Springer, Berlin, 1975. 10.1007/978-1-4612-6380-7Search in Google Scholar

[43] W. H. Fleming and H. M. Soner, Controlled Markov Processes and Viscosity Solutions, Appl. Math. (New York) 25, Springer, New York, 1993. Search in Google Scholar

[44] W. H. Fleming and P. E. Souganidis, On the existence of value functions of two-player, zero-sum stochastic differential games, Indiana Univ. Math. J. 38 (1989), no. 2, 293–314. 10.1512/iumj.1989.38.38015Search in Google Scholar

[45] K. Fok, A nonlinear Fabes–Stroock result, Comm. Partial Differential Equations 23 (1998), no. 5–6, 967–983. 10.1080/03605309808821375Search in Google Scholar

[46] G. Galise and A. Vitolo, Removable singularities for degenerate elliptic Pucci operators, Adv. Differential Equations 22 (2017), no. 1–2, 77–100. 10.57262/ade/1484881286Search in Google Scholar

[47] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[48] F. R. Harvey and H. B. Lawson, Jr., Dirichlet duality and the nonlinear Dirichlet problem, Comm. Pure Appl. Math. 62 (2009), no. 3, 396–443. 10.1002/cpa.20265Search in Google Scholar

[49] F. R. Harvey and H. B. Lawson, Jr., Existence, uniqueness and removable singularities for nonlinear partial differential equations in geometry, Surveys in Differential Geometry. Geometry and Topology, Surv. Differ. Geom. 18, International Press, Somerville (2013), 103–156. 10.4310/SDG.2013.v18.n1.a3Search in Google Scholar

[50] F. R. Harvey and H. B. Lawson, Jr., Removable singularities for nonlinear subequations, Indiana Univ. Math. J. 63 (2014), no. 5, 1525–1552. 10.1512/iumj.2014.63.5398Search in Google Scholar

[51] E. Hopf, Elementare Bemerkungen über die Lösungen partieller Differentialgleichungen zweiter Ordnung vom elliptischen Typus, Sitzungsberichte Akad. Berlin 1927 (1927), 147–152. 10.1515/crll.1927.158.80Search in Google Scholar

[52] C. Imbert, Alexandroff–Bakelman–Pucci estimate and Harnack inequality for degenerate/singular fully non-linear elliptic equations, J. Differential Equations 250 (2011), no. 3, 1553–1574. 10.1016/j.jde.2010.07.005Search in Google Scholar

[53] C. Imbert and L. Silvestre, C1,α regularity of solutions of some degenerate fully non-linear elliptic equations, Adv. Math. 233 (2013), 196–206. 10.1016/j.aim.2012.07.033Search in Google Scholar

[54] H. Ishii and P.-L. Lions, Viscosity solutions of fully nonlinear second-order elliptic partial differential equations, J. Differential Equations 83 (1990), no. 1, 26–78. 10.1016/0022-0396(90)90068-ZSearch in Google Scholar

[55] S. Koike and A. Świȩch, Maximum principle for fully nonlinear equations via the iterated comparison function method, Math. Ann. 339 (2007), no. 2, 461–484. 10.1007/s00208-007-0125-zSearch in Google Scholar

[56] J. Kovats, The minmax principle and W2,p regularity for solutions of the simplest Isaacs equations, Proc. Amer. Math. Soc. 140 (2012), no. 8, 2803–2815. 10.1090/S0002-9939-2012-11610-7Search in Google Scholar

[57] J. Kovats, On the second order derivatives of solutions of a special Isaacs equation, Proc. Amer. Math. Soc. 144 (2016), no. 4, 1523–1533. 10.1090/proc/12956Search in Google Scholar

[58] N. V. Krylov, Controlled Diffusion Processes, Springer, New York, 1977. Search in Google Scholar

[59] J. J. Manfredi, M. Parviainen and J. D. Rossi, An asymptotic mean value characterization for p-harmonic functions, Proc. Amer. Math. Soc. 138 (2010), no. 3, 881–889. 10.1090/S0002-9939-09-10183-1Search in Google Scholar

[60] K. Miller, Extremal barriers on cones with Phragmén–Lindelöf theorems and other applications, Ann. Mat. Pura Appl. (4) 90 (1971), 297–329. 10.1007/BF02415053Search in Google Scholar

[61] A. Mohammed, G. Porru and A. Vitolo, Harnack inequality for nonlinear elliptic equations with strong absorption, J. Differential Equations 263 (2017), no. 10, 6821–6843. 10.1016/j.jde.2017.07.035Search in Google Scholar

[62] A. Mohammed and A. Vitolo, On the strong maximum principle, Complex Var. Elliptic Equ. (2019), 10.1080/17476933.2019.1594207. 10.1080/17476933.2019.1594207Search in Google Scholar

[63] I. Netuka and J. Veselý, Mean value property and harmonic functions, Classical and Modern Potential Theory and Applications (Chateau de Bonas 1993), NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci. 430, Kluwer Academic, Dordrecht (1994), 359–398. 10.1007/978-94-011-1138-6_29Search in Google Scholar

[64] M. Nisio, Stochastic differential games and viscosity solutions of Isaacs equations, Nagoya Math. J. 110 (1988), 163–184. 10.1017/S0027763000002932Search in Google Scholar

[65] 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

[66] Y. Peres, O. Schramm, S. Sheffield and D. B. Wilson, Tug-of-war and the infinity Laplacian, J. Amer. Math. Soc. 22 (2009), no. 1, 167–210. 10.1007/978-1-4419-9675-6_18Search in Google Scholar

[67] M. H. Protter and H. F. Weinberger, Maximum Principles in Differential Equations, Springer, New York, 1984. 10.1007/978-1-4612-5282-5Search in Google Scholar

[68] C. Pucci, Limitazioni per soluzioni di equazioni ellittiche, Ann. Mat. Pura Appl. (4) 74 (1966), 15–30. 10.1007/BF02416445Search in Google Scholar

[69] P. Pucci and V. D. Rădulescu, The maximum principle with lack of monotonicity, Electron. J. Qual. Theory Differ. Equ. 2018 (2018), Paper No. 58. 10.14232/ejqtde.2018.1.58Search in Google Scholar

[70] P. Pucci and J. Serrin, The strong maximum principle revisited, J. Differential Equations 196 (2004), no. 1, 1–66. 10.1016/j.jde.2003.05.001Search in Google Scholar

[71] J.-P. Sha, p-convex Riemannian manifolds, Invent. Math. 83 (1986), no. 3, 437–447. 10.1007/BF01394417Search in Google Scholar

[72] J.-P. Sha, Handlebodies and p-convexity, J. Differential Geom. 25 (1987), no. 3, 353–361. 10.4310/jdg/1214440980Search in Google Scholar

[73] A. Świech, W1,p-interior estimates for solutions of fully nonlinear, uniformly elliptic equations, Adv. Differential Equations 2 JYea1997, no. 6, 1005–1027. 10.57262/ade/1366638681Search in Google Scholar

[74] N. S. Trudinger, Local estimates for subsolutions and supersolutions of general second order elliptic quasilinear equations, Invent. Math. 61 (1980), no. 1, 67–79. 10.1007/BF01389895Search in Google Scholar

[75] N. S. Trudinger, Hölder gradient estimates for fully nonlinear elliptic equations, Proc. Roy. Soc. Edinburgh Sect. A 108 (1988), no. 1–2, 57–65. 10.1017/S0308210500026512Search in Google Scholar

[76] J. L. Vázquez, A strong maximum principle for some quasilinear elliptic equations, Appl. Math. Optim. 12 (1984), no. 3, 191–202. 10.1007/BF01449041Search in Google Scholar

[77] A. Vitolo, On the maximum principle for complete second-order elliptic operators in general domains, J. Differential Equations 194 (2003), no. 1, 166–184. 10.1016/S0022-0396(03)00193-1Search in Google Scholar

[78] A. Vitolo, A note on the maximum principle for second-order elliptic equations in general domains, Acta Math. Sin. (Engl. Ser.) 23 (2007), no. 11, 1955–1966. 10.1007/s10114-007-0976-ySearch in Google Scholar

[79] A. Vitolo, Removable singularities for degenerate elliptic equations without conditions on the growth of the solution, Trans. Amer. Math. Soc. 370 (2018), no. 4, 2679–2705. 10.1090/tran/7095Search in Google Scholar

[80] H. Wu, Manifolds of partially positive curvature, Indiana Univ. Math. J. 36 (1987), no. 3, 525–548. 10.1512/iumj.1987.36.36029Search in Google Scholar

Received: 2019-09-11
Revised: 2019-10-14
Accepted: 2019-10-15
Published Online: 2019-11-13
Published in Print: 2020-02-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 2.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2019-2069/html?srsltid=AfmBOopdp1sdyn86Jx_Fo_V62lI4MolFB6_j27ueEzmLxBXgckeBy5iY
Scroll to top button