Home Generalized principal eigenvalues of convex nonlinear elliptic operators in ℝ N
Article Open Access

Generalized principal eigenvalues of convex nonlinear elliptic operators in ℝ N

  • Anup Biswas EMAIL logo and Prasun Roychowdhury
Published/Copyright: November 25, 2020

Abstract

We study the generalized eigenvalue problem in N for a general convex nonlinear elliptic operator which is locally elliptic and positively 1-homogeneous. Generalizing [H. Berestycki and L. Rossi, Generalizations and properties of the principal eigenvalue of elliptic operators in unbounded domains, Comm. Pure Appl. Math. 68 2015, 6, 1014–1065], we consider three different notions of generalized eigenvalues and compare them. We also discuss the maximum principles and uniqueness of principal eigenfunctions.

MSC 2010: 35J60; 35P30; 35B50

1 Introduction

This article contributes to the study of eigenvalue problem of the form

F ( D 2 ψ , D ψ , ψ , x ) = λ ψ in  N ,

where F is a fully nonlinear, convex, positively 1-homogeneous elliptic operator with measurable coefficients. We establish the existence of half-(or demi-)eigenvalues and characterize the set of all eigenvalues with positive and negative eigenfunctions. This generalizes a recent work of Berestycki and Rossi [12] which considers linear elliptic operators. We also derive necessary and sufficient conditions for the validity of maximum principles in N and discuss the uniqueness of principal eigenfunctions.

It has long been known that certain types of positively homogeneous operators possess two principal eigenvalues (one corresponds to a positive eigenfunction and the other one corresponds to a negative eigenfunction). In fact, it first appeared in the work of Pucci [24] who computed these eigenvalues explicitly for certain extremal operators in the unit ball. Later it also appeared in a work of Berestycki [8] while studying the bifurcation phenomenon for some nonlinear Sturm–Liouville problem and Berestycki referred them as half eigenvalues. In connection to this work of Berestycki, Lions used a stochastic control approach in [22] to characterize these eigenvalues (he called it demi-eigenvalues) of operators which are the supremum of linear operators with 𝒞 1 , 1 -coefficients, and relate them to certain bifurcation problem. In their seminal work [10] Berestycki, Nirenberg and Varadhan introduced the notion of Dirichlet generalized principal eigenvalue for linear operators in non-smooth bounded domains and also established a deep connection between sign of the principal eigenvalue and validity of maximum principles. This work serves as a founding stone of the modern eigentheory and has been used to study eigenvalue problems for general nonlinear operators, including degenerate ones. We are in particular, attracted by the works [5, 6, 9, 13, 14, 15, 19, 20, 23, 25]. We owe much to the work of Quass and Sirakov [25] who study the Dirichlet principal eigenvalue problem for convex, fully nonlinear elliptic operators in bounded domains.

All the above-mentioned works deal with bounded domains. It is then natural to ask how the eigentheory changes for unbounded domains. In fact, the study of eigenvalue problems in N becomes important to understand the existence and uniqueness of solutions for certain semilinear elliptic operators. See, for instance, the discussion in [11, 12] and references therein. Principal eigenvalue is a key ingredient to find the rate functional for the large deviation estimate of empirical measures of diffusions [17, 18, 21]. Recently, eigenvalue problems in N have got much attention due to its application in the theory of risk-sensitive controls [4, 2, 1] (some discussions are left to Section 2.3). Our present work is motivated by a recent study of Berestycki and Rossi [12] where the authors consider non-degenerate linear elliptic operators and develop an eigentheory for unbounded domains. Monotonicity property of the principal eigenvalue (with respect to the potentials) in N and its relation with the stability property of the twisted process is established in [4]. The paper [1] considers a class of semilinear elliptic operators in N and obtains a variational representation of the principal eigenvalue under the assumption of geometric stability. The chief goal of this article is to develop an eigentheory for fully nonlinear positively homogeneous operators. Though the results of this article are obtained in the whole space N , one can mimic the arguments for any unbounded domains (see Remark 3.1 for more details).

The rest of the article is organized as follows: In the next section we introduce our model and state our main results. We also motivate the model by providing a discussion in Section 2.3. Proofs of the main results are given in Section 3.

2 Statement of main results

In this section we introduce our model and state the main results. We also provide a motivation in Section 2.3 for considering these eigenvalue problems.

2.1 Model and assumptions

Let λ , Λ : N ( 0 , ) be two locally bounded functions with the property that for any compact set K N we have

0 < inf x K λ ( x ) sup x K Λ ( x ) < .

Choosing K = { x } , it follows from above that 0 < λ ( x ) Λ ( x ) for all x N . These two functions will be treated as the bounds of the ellipticity constants at point x. By 𝒮 N we denote the set of all N × N real symmetric matrices. The extremal Pucci operators corresponding to λ , Λ are defined as follows. For M 𝒮 N the extremal operators at x N are defined to be

λ , Λ + ( x , M ) = sup λ ( x ) I A Λ ( x ) I trace ( A M ) = Λ ( x ) β i 0 β i + λ ( x ) β i < 0 β i ,
λ , Λ - ( x , M ) = inf λ ( x ) I A Λ ( x ) I trace ( A M ) = λ ( x ) β i 0 β i + Λ ( x ) β i < 0 β i ,

where β 1 , , β n denote the eigenvalues of the matrix M.

Our operator F is a Borel measurable function F : 𝒮 N × N × × N with the following properties:

  1. F is positively 1-homogeneous in the variables ( M , p , u ) 𝒮 N × N × , i.e., for every t > 0 we have

    F ( t M , t p , t u , x ) = t F ( M , p , u , x ) for all  x N .

    In particular, F ( 0 , 0 , 0 , x ) 0 .

  2. F is convex in the variables ( M , p , u ) 𝒮 N × N × .

  3. There exist locally bounded functions γ , δ : N [ 0 , ) satisfying

    λ , Λ - ( x , M - N ) - γ ( x ) | p - q | - δ ( x ) | u - v | F ( M , p , u , x ) - F ( N , q , v , x )
    λ , Λ + ( x , M - N ) + γ ( x ) | p - q | + δ ( x ) | u - v |

    for all M , N 𝒮 N , p , q N , u , v and x N .

  4. The function ( M , x ) 𝒮 N × N F ( M , 0 , 0 , x ) is continuous.

Throughout this article we assume conditions (H1)(H4) without any further mention. Also, observe that due to our hypotheses the operator F satisfies the conditions in [25] which studies the Dirichlet eigenvalue problem for F in bounded domains. Therefore the results of [25] holds for F in smooth bounded domains.

Let us now define the principal eigenvalues of F in a smooth domain Ω N , possibly unbounded. For any real number λ we define the following sets:

Ψ + ( F , Ω , λ ) = { ψ 𝒲 loc 2 , N ( Ω ) : F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  and  ψ > 0  in  Ω } ,
Ψ - ( F , Ω , λ ) = { ψ 𝒲 loc 2 , N ( Ω ) : F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  and  ψ < 0  in  Ω } .

By sub- or supersolution we always mean L N -strong solution. The (half-)eigenvalues are defined to be

λ 1 + ( F , Ω ) = sup { λ : Ψ + ( F , Ω , λ ) } ,
λ 1 - ( F , Ω ) = sup { λ : Ψ - ( F , Ω , λ ) } .

Using the convexity of F and [25, Proposition 4.2] it follows that λ 1 + ( F , Ω ) λ 1 - ( F , Ω ) < . For F linear we also have λ 1 + ( F , Ω ) = λ 1 - ( F , Ω ) . In this article we would be interested in the case Ω = N and for notational economy we denote λ 1 ± ( F , N ) = λ 1 ± ( F ) .

Remark 2.1.

We can replace the L N -strong super and subsolutions in Ψ ± ( F , Ω , λ ) by L N -viscosity super and subsolutions, respectively.

2.2 Main results

We now state our main results. Most of the results obtained here are generalization of its linear counterpart in [12]. Recall from [12, Theorem 1.4] that for F linear and λ ( - , λ 1 ( F ) ] there exists a positive function φ 𝒲 loc 2 , p ( N ) , p > N , satisfying F ( D 2 φ , D φ , φ , x ) + λ φ = 0 in N . Thus there is a continuum of eigenvalues with the largest one being the principal eigenvalue. This leads us to the following sets of eigenvalues.

Definition 2.1.

We say λ is an eigenvalue with a positive eigenfunction if there exists ϕ 𝒲 loc 2 , p ( N ) , p > N , such that

F ( D 2 ϕ , D ϕ , ϕ , x ) = - λ ϕ in  N    and    ϕ > 0 in  N .

We denote the collection of all eigenvalues with positive eigenfunctions by + . Analogously, we define - as the collection of all eigenvalues with negative eigenfunctions.

Our first result generalizes [12, Theorem 1.4].

Theorem 2.1.

We have E + = ( - , λ 1 + ( F ) ] and E - = ( - , λ 1 - ( F ) ] .

It is well known that for bounded domains it is also possible to define principal eigenvalues through subsolutions (cf. [25, Theorem 1.2]). However, this situation is bit different for unbounded domains. To explain, we introduce the following quantities:

λ 1 , + ( F ) inf { λ : there exists  ψ 𝒲 loc 2 , N ( N ) L ( N ) , ψ > 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } ,
λ 1 , - ( F ) inf { λ : there exists  ψ 𝒲 loc 2 , N ( N ) L ( N ) , ψ < 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } ,
λ 1 ′′ , + ( F ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( N ) , inf N ψ > 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } ,
λ 1 ′′ , - ( F ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( N ) , sup N ψ < 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } .

We remark that in case of bounded domains one has

λ 1 + ( F , Ω ) = λ 1 , + ( F , Ω ) = λ 1 ′′ , + ( F , Ω )

and

λ 1 - ( F , Ω ) = λ 1 , - ( F , Ω ) = λ 1 ′′ , - ( F , Ω ) ,

provided we required the subsolution (supersolution) to vanish on Ω in the definition of λ 1 , + ( λ 1 , - , respectively) (cf. [25]). But the same might fail to hold in unbounded domains (counter-example in [11, p. 201]). However, we could prove the following relation which generalizes [12, Theorem 1.7].

Theorem 2.2.

The following hold:

  1. We have λ 1 , + ( F ) λ 1 + ( F ) and λ 1 , - ( F ) λ 1 - ( F ) .

  2. Suppose that

    (2.1) sup N δ ( x ) < , lim sup | x | γ ( x ) | x | < , 𝑎𝑛𝑑 lim sup | x | Λ ( x ) | x | 2 < .

    Then we have λ 1 ′′ , + ( F ) λ 1 , + ( F ) and λ 1 ′′ , - ( F ) λ 1 , - ( F ) .

In view of Theorem 2.2 we see that λ 1 ′′ , + ( F ) λ 1 , + ( F ) λ 1 + ( F ) and λ 1 ′′ , - ( F ) λ 1 , - ( F ) λ 1 - ( F ) , provided (2.1) holds. Again, due to the convexity of F we have λ 1 + ( F ) λ 1 - ( F ) . One might wonder if there is any natural relation between “plus” and “minus” eigenvalues. We now argue that this might not be possible, in general. If we consider F to be linear, then we have λ 1 , + ( F ) = λ 1 , - ( F ) , and therefore if (2.1) holds, then λ 1 + ( F ) λ 1 ′′ , - ( F ) , by Theorem 2.2. We now produce an example where the reverse inequality holds.

Example 2.1.

Consider two linear elliptic operators of the form

L α u = Δ u + b α ( x ) D u + c α ( x ) u ,

for α { 1 , 2 } with the properties that

λ 1 ′′ ( L 2 , N ) > λ 1 ′′ ( L 1 , N ) and λ 1 ′′ ( L 1 , N ) = λ 1 ( L 1 , N ) = λ 1 ( L , N ) .

Now define a nonlinear operator

F ( D 2 u , D u , u , x ) Δ u + max α { 1 , 2 } { b α ( x ) D u + c α ( x ) u } .

It is then easily seen that

λ 1 ′′ , - ( F ) max { λ 1 ′′ ( L 1 , N ) , λ 1 ′′ ( L 2 , N ) }

and

λ 1 + ( F ) min { λ 1 ( L 1 , N ) , λ 1 ( L 2 , N ) } .

Combining we obtain

λ 1 ′′ , - ( F ) λ 1 ′′ ( L 2 , N ) > λ 1 ′′ ( L 1 , N ) = λ 1 ( L 1 , N ) λ 1 + ( F ) .

Next we list a few class of operators for which these three eigenvalues coincide (compare them with [12, Theorem 1.9]). We only provide the result for “plus” eigenvalues and the analogous result for “minus” eigenvalues are easy to guess.

Theorem 2.3.

The equality λ 1 + ( F ) = λ 1 ′′ , + ( F ) holds in each of the following cases:

  1. F = F ~ + γ ~ ( x ) , where F ~ is a nonlinear operator with an additional property λ 1 + ( F ~ , N ) = λ 1 ′′ , + ( F ~ , N ) , and γ ~ L ( N ) is a non-negative function satisfying lim | x | γ ~ ( x ) = 0 .

  2. λ 1 + ( F ) - lim sup | x | F ( 0 , 0 , 1 , x ) .

  3. Assume that λ 0 λ ( x ) Λ ( x ) Λ 0 for all x N , lim | x | γ ( x ) = 0 and for all r > 0 and all β such that β < lim sup | x | F ( 0 , 0 , 1 , x ) there exists r ( x 0 ) satisfying inf r ( x 0 ) F ( 0 , 0 , 1 , x ) > β .

  4. There exists a V 𝒞 2 ( N ) with inf N V > 0 and

    F ( D 2 V , D V , V , x ) - λ 1 + ( F ) V for all  x c ,

    for some ball .

Now we turn our attention towards maximum principles. It was observed in the seminal work of Berestycki, Nirenberg and Varadhan [10] that the sign of the principle eigenvalue determines the validity of maximum principles in bounded domains. Extension of this result for nonlinear operators are obtained by Quaas and Sirakov [25] and Armstrong [5]. Further generalization in smooth bounded domains for a class of degenerate, nonlinear elliptic operators are obtained by Berestycki, Capuzzo Dolcetta, Porretta and Rossi [9] and Birindelli and Demengel [13]. Recently, Berestycki and Rossi [12] establish the maximum principles in unbounded domains for linear elliptic operators. Here we extend their results to our nonlinear setting.

Definition 2.2 (Maximum principles).

We say that the operator F satisfies β + -MP with respect to a positive function β if for any function u 𝒲 loc 2 , N ( N ) satisfying

F ( D 2 u , D u , u , x ) 0 in  N    and    sup N u β < ,

we have u 0 in N . For β = 1 , we simply mention this property as + MP.

We say that the operator F satisfies β - -MP with respect to a negative function β if for any function u 𝒲 loc 2 , N ( N ) satisfying

F ( D 2 u , D u , u , x ) 0 in  N    and    sup N u β < ,

we have u 0 in N . For β = - 1 , we simply mention this property as - MP.

Note that β 1 corresponds to the well-known maximum principle. We would be interested in a function β : N ( 0 , ) which satisfies either

(2.2) there exists  σ > 0  such that  lim sup | x | β ( x ) | x | - σ = 0

or

(2.3) there exists  σ > 0  such that  lim sup | x | β ( x ) exp ( - σ | x | ) = 0 .

Generalizing [12, Definition 1.2], we now consider the following quantities.

Definition 2.3.

Given a positive function β : N , we define

λ β ′′ , + ( F ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( N )  such that  ψ β , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } ,
λ β ′′ , - ( F ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( N )  such that  ψ - β , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } .

Our maximum principles would be established under the following growth conditions on the coefficients:

(2.4) sup N δ ( x ) < , lim sup | x | γ ( x ) | x | < , lim sup | x | Λ ( x ) | x | 2 < ,
(2.5) sup N δ ( x ) < , sup N γ ( x ) < , sup N Λ ( x ) < .

Next we state our maximum principle

Theorem 2.4.

Suppose that either (2.2) and (2.4) or (2.3) and (2.5) hold. Then the following hold:

  1. The operator F satisfies β + - MP in N if λ β ′′ , + ( F ) > 0 .

  2. The operator F satisfies ( - β ) - - MP in N if λ β ′′ , - ( F ) > 0 .

As a consequence of Theorem 2.4 we obtain the following corollaries.

Corollary 2.1.

Suppose that either (2.4) or (2.5) holds. Then we have:

  1. The operator F satisfy + MP in N if λ 1 ′′ , + ( F ) > 0 .

  2. The operator F satisfy - MP in N if λ 1 ′′ , - ( F ) > 0 .

  3. Suppose that λ 1 ′′ , + ( F ) > 0 (therefore λ 1 ′′ , - ( F ) > 0 ). Let u 𝒲 loc 2 , N ( N ) L ( N ) satisfy F ( D 2 u , D u , u , x ) = 0 in N . Then u 0 .

Corollary 2.2.

Suppose that F satisfies β + -MP. Let u , v W loc 2 , N ( R N ) be such that

F ( D 2 u , D u , u , x ) 0 , F ( D 2 v , D v , v , x ) 0 in  N    𝑎𝑛𝑑    sup N u - v β < .

Then we have u v in R N .

Proof.

Denote by w = u - v . By using the convexity of F, it follows that

F ( D 2 w , D w , w , x ) F ( D 2 u , D u , u , x ) - F ( D 2 v , D v , v , x ) 0 in  N .

Hence the result follows from β + -MP. ∎

Generalizing λ 1 , + ( F ) and λ 1 , - ( F ) , we define the following quantities. Let β be a positive valued function and

λ β , + ( F ) inf { λ : there exists  ψ 𝒲 loc 2 , N ( N )  such that  β ψ > 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } ,
λ β , - ( F ) inf { λ : there exists  ψ 𝒲 loc 2 , N ( N )  such that  - β ψ < 0 , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  N } .

As a necessary condition for the validity of maximum principles we deduce the following.

Theorem 2.5.

The following hold:

  1. If F satisfies the β + - MP, then λ β , + ( F ) 0 . In particular, if F satisfies + MP, then we have λ 1 , + ( F ) 0 .

  2. If F satisfies the ( - β ) - - MP, then λ β , - ( F ) 0 . In particular, if F satisfies the - MP, then we have λ 1 , - ( F ) 0 .

Finally, we discuss about simplicity of the principal eigenvalues. For linear F uniqueness of principal eigenfunctions can be established imposing Agmon’s minimal growth condition at infinity [12, Definition 8.2] on the eigenfunctions. But such criterion does not seem to work well for nonlinear F. Recently, in [3, Theorem 2.1] it is shown that Agmon’s minimal growth criterion is equivalent to monotonicity of the principal eigenvalue on the right. Our next result establish simplicity of principal eigenvalue under certain monotonicity condition of principal eigenvalue at infinity.

Theorem 2.6.

Suppose that there exists a positive V W loc 2 , N ( R N ) satisfying

(2.6) F ( D 2 V , D V , V , x ) - ( λ 1 + ( F ) + ε ) V for all  x K c ,

for some compact ball K and ε > 0 . Then λ 1 + ( F ) is simple, i.e., the positive principal eigenfunction is unique up to a multiplicative constant.

We remark that (2.6) is equivalent to

λ 1 + ( F ) < lim r λ 1 + ( F , ¯ r c ) .

Our next result is about simplicity of λ 1 - ( F ) .

Theorem 2.7.

Suppose that there exists a positive V W loc 2 , N ( R N ) satisfying

(2.7) F ( D 2 V , D V , V , x ) - ( λ 1 - ( F ) + ε ) V for all  x K c ,

for some compact ball K and ε > 0 . Then λ 1 - ( F ) is simple.

2.3 Motivation

One of the important examples of F comes from the control theory. In particular, we may consider

F ( D 2 ϕ , D ϕ , ϕ , x ) = sup α { trace ( a α ( x ) D 2 ϕ ( x ) ) + b α ( x ) D ϕ ( x ) + c α ( x ) ϕ ( x ) } = sup α { L α ϕ + c α ϕ } ,

where α varies over some index set , λ ( x ) I a α ( x ) Λ ( x ) I , and sup α | b α ( x ) | , sup α | c α ( x ) | are locally bounded. The eigenvalue problem corresponding to the operator F appears in the study of risk-sensitive controls. See, for instance, [2, 4] and references therein. To elaborate, suppose that is a compact subset of m . Let 𝔘 be the collection of Borel measurable maps α : N . Note that constant functions are also included in 𝔘 . This set 𝔘 represents the collection of all Markov controls. Given α 𝔘 , suppose that X α is the Markov diffusion process with generator L α . Denote the law of X α by α and 𝔼 α [ ] is the expectation operator associated with it. Consider the maximization problem

Λ = sup α 𝔘 lim sup T 1 T log 𝔼 α [ e 0 T c α ( X t ) d t ] .

Then under reasonable hypothesis, one can show that Λ is an eigenvalue of F (i.e., Λ + ) and for many practical reasons it is desirable that Λ = λ 1 + ( F ) . Also, simplicity of λ 1 + ( F ) is important to find an optimal strategy or control. We refer the readers to [2, 4] for more details on this problem.

3 Proofs of main results

In this section we prove Theorems 2.12.7. Let us start by recalling the following Harnack inequality from [25, Theorem 3.6] which will be crucial for our proofs. The result in [25, Theorem 3.6] is stated for L N -viscosity solutions and also applies to L N -strong solutions due to [16, Lemma 2.5].

Theorem 3.1.

Let Ω R N be bounded. Let u C ( Ω ¯ ) W loc 2 , N ( Ω ) and f L N ( Ω ) satisfy u 0 in Ω and

λ , γ + ( x , D 2 u ) + γ | D u | + δ u f in  Ω ,
λ , γ - ( x , D 2 u ) - γ | D u | - δ u f in  Ω .

Then for any compact set K Ω we have

sup K u C [ inf K u + f L N ( Ω ) ]

for some constant C dependent on K , Ω , N , γ , δ , min Ω λ and max Ω Λ .

Next we prove Theorem 2.1. The idea is the following: we show using the Harnack inequality and stability estimate that the Dirichlet principal eigenpair in n converges to a principal eigenpair in N . For any λ < λ 1 + ( F ) or λ < λ 1 - ( F ) we use the refined maximum principle in bounded domains and then stability estimate to pass the limit. We spilt the proof of Theorem 2.1 in Lemmas 3.1 and 3.2.

Lemma 3.1.

It holds that E + = ( - , λ 1 + ( F ) ] .

Proof.

Let λ 1 + ( F , n ) be the Dirichlet principal eigenvalue in n corresponding to the positive principal eigenfunction. Existence of λ 1 + ( F , n ) follows from [25, Theorem 1.1]. For notational economy we denote λ 1 + ( F , n ) = λ 1 , n + and λ 1 + ( F ) = λ 1 + . We also set E p ( Ω ) = 𝒲 loc 2 , p ( Ω ) 𝒞 ( Ω ¯ ) . We divide the proof into two steps.

Step 1. We show that lim n λ 1 , n + = λ 1 + and λ 1 + + . It is obvious from the definition that λ 1 , n + is decreasing in n and bounded below by λ 1 + . Thus if lim n λ 1 , n + = - , we also have λ 1 + = - and there is nothing to prove. So we assume lim n λ 1 , n + λ ~ > - . It is then obvious that λ ~ λ 1 + . From [25, Theorem 1.1] we have ψ 1 , n + E p ( n ) for all p < such that ψ 1 , n + > 0 in n , ψ 1 , n + = 0 on n and satisfies

(3.1) F ( D 2 ψ 1 , n + , D ψ 1 , n + , ψ 1 , n + , x ) = - λ 1 , n + ψ 1 , n + in  n

for all n 1 . Normalize each ψ 1 , n + by choosing ψ 1 , n + ( 0 ) = 1 . Fix any compact K N such that 0 K and choose n 0 large so that K m for all m n 0 . Applying Theorem 3.1 on (3.1), we find a constant C = C ( n 0 ) satisfying

sup K ψ 1 , n + C inf K ψ 1 , n + C ψ 1 , n + ( 0 ) = C .

Thus applying [25, Theorem 3.3], we obtain, for p > N , that

ψ 1 , n + 𝒲 2 , p ( K ) C for all  n > n 0 .

Since K is arbitrary, using a standard diagonalization argument we can find a non-negative φ + E p ( N ) , p < , such that ψ 1 , n + φ + in 𝒲 loc 2 , p ( N ) , up to a subsequence. Hence by [16, Theorem 3.8 and Corollary 3.7] we obtain

{ F ( D 2 φ + , D φ + , φ + , x ) = - λ ~ φ + in  N , φ + ( 0 ) = 1 .

Again, applying Theorem 3.1, we have φ + > 0 . Thus, λ ~ λ 1 + . This shows λ ~ = λ 1 + and λ 1 + + .

Step 2. We show that + = ( - , λ 1 + ] . It is obvious that + ( - , λ 1 + ] . To show the reverse relation, we consider λ < λ 1 + . We choose a sequence { f n } n 1 of continuous, non-positive, non-zero functions satisfying

support ( f n ) n ¯ n - 1 for all  n .

Denote by F ~ = F + λ . Then λ 1 + ( F ~ , n ) = λ 1 , n + - λ λ 1 + - λ > 0 . Therefore, by [25, Theorems 1.5 and 1.8], there exists a unique non-negative u n E p ( B n ) , p N , which satisfies

(3.2) { F ~ ( D 2 u n , D u n , u n , x ) = f n in  n , u n = 0 on  n .

By the strong maximum principle [25, Lemma 3.1] it follows that u n > 0 in n . For natural number n 2 we define

v n ( x ) u n ( x ) u n ( 0 ) .

Clearly, v n E p ( n - 1 ) , p < , positive in n - 1 and v n ( 0 ) = 1 . Also, by (3.2),

F ( D 2 v n , D v n , v n , x ) = - λ v n in  n - 1 .

Now we continue as in Step 1 and extract a subsequence of v n that converges in 𝒲 loc 2 , p ( N ) to some positive φ E p ( N ) , p < , and satisfies

F ( D 2 φ , D φ , φ , x ) = - λ φ in  N .

This gives us λ + . Thus + = ( - , λ 1 + ] . ∎

Next lemma concerns the eigenvalues with negative eigenfunctions.

Lemma 3.2.

It holds that E - = ( - , λ 1 - ( F ) ] .

Proof.

Idea of the proof is similar to Lemma 3.1. Let λ 1 - ( F , n ) be the Dirichlet principal eigenvalue in n corresponding to the negative principal eigenfunction [25, Theorem 1.1]. For simplicity we denote

λ 1 - ( F , n ) = λ 1 , n - and λ 1 - ( F ) = λ 1 - .

We divide the proof of into two steps.

Step 1. We show that lim n λ 1 , n - = λ 1 - and λ 1 - - . It is obvious from the definition that λ 1 , n - in decreasing in n and bounded below by λ 1 - . Thus if lim n λ 1 , n - = - , we also have λ 1 - = - and there is nothing to prove. So we assume lim n λ 1 , n - λ ^ > - . It is then obvious that λ ^ λ 1 - . From [25, Theorem 1.1], for all n , we have ψ 1 , n - E p ( n ) for all p < such that ψ 1 , n - < 0 in n , ψ 1 , n - = 0 in n , and

(3.3) F ( D 2 ψ 1 , n - , D ψ 1 , n - , ψ 1 , n - , x ) = - λ 1 , n - ψ 1 , n - in  n .

Normalize each ψ 1 , n - by fixing ψ 1 , n - ( 0 ) = - 1 . Denoting G ( M , p , u , x ) = - F ( - M , - p , - u , x ) we find from (3.3)

G ( D 2 ϕ 1 , n - , D ϕ 1 , n - , ϕ 1 , n - , x ) = - λ 1 , n - ϕ 1 , n - in  n ,

for ϕ 1 , n - = - ψ 1 , n - 0 . Since G satisfies conditions (H1), (H3) and (H4), Theorem 3.1 applies. Then using (3.3) and [25, Theorem 3.3], we can obtain locally uniform 𝒲 loc 2 , p bounds on ϕ 1 , n - . Now apply the arguments of Step 1 in the proof of Lemma 3.1 to show that lim n λ 1 , n - = λ 1 - and λ 1 + - .

Step 2. As discussed in Lemma 3.1, it is enough to show that for any λ < λ 1 - we have λ - . Consider a sequence { f n } n 1 of continuous, non-negative, non-zero functions satisfying

support ( f n ) n ¯ n - 1 for all  n .

Denote by F ~ = F + λ . Then λ 1 - ( F ~ , n ) = λ 1 , n - - λ λ 1 - - λ > 0 . Therefore, by [25, Theorem 1.9], there exists a non-zero, non-positive u n E p ( n ) , p N , satisfying

{ F ~ ( D 2 u n , D u n , u n , x ) = f n in  n , u n = 0 in  n .

Since G satisfies (H3) we can apply the strong maximum principle [25, Lemma 3.1] to obtain that u n < 0 in n . Now repeat the arguments of Step 2 in the proof of Lemma 3.1 to conclude that λ - . This completes the proof. ∎

Proof of Theorem 2.1.

The proof follows from Lemmas 3.13.2. ∎

The following (standard) existence result will be required.

Lemma 3.3.

Suppose that u ¯ , u ¯ E p ( Ω ) for some p N and Ω is a smooth bounded domain, and u ¯ ( u ¯ ) is a supersolution (subsolution) of F ( D 2 u , D u , u , x ) = f ( x , u ) in Ω for some f L loc ( Ω ¯ × R ) . Assume that f is locally Lipschitz in its second argument uniformly (almost surely) with respect to the first argument and u ¯ 0 , u ¯ 0 on Ω . Then there exists u E p ( Ω ) with u ¯ u u ¯ in Ω and satisfies

{ F ( D 2 u , D u , u , x ) = f ( x , u ) in  Ω , u = 0 on  Ω .

Proof.

The proof is based on monotone iteration method. See also [25, Lemma 4.3] for a similar argument. Define the operator F ~ = F - θ in such a way that F ~ is proper, i.e., decreasing in u. We may choose θ large enough so that

θ > Lip ( f ( x , )  on  [ inf Ω u ¯ , sup Ω u ¯ ] ) almost surely for  x Ω .

Also, note that F ~ satisfying (H1)(H4). Now we define the monotone sequence. Denote by v 0 = u ¯ , and for each n 0 , we define

{ F ~ ( D 2 v n + 1 , D v n + 1 , v n + 1 , x ) = f ( x , v n ) - θ v n in  Ω , v n + 1 = 0 on  Ω .

Existence of v n + 1 E p follows from [25, Theorem 3.4]. Also, since F ~ is proper, we can apply comparison principle [25, Theorem 3.2] to obtain v 0 v 1 v 2 u ¯ . It is then standard to show that v n u in 𝒞 ( Ω ¯ ) for some u E p ( Ω ) and u is our required solution (see for instance, [25, Lemma 4.3]). This completes the proof. ∎

Applying Lemma 3.3, we obtain the following.

Theorem 3.2.

It holds that λ 1 , + ( F ) λ 1 + ( F ) and λ 1 , - ( F ) λ 1 - ( F ) .

Proof.

We divide the proof into two steps.

Step 1. We show that λ 1 , + ( F ) λ 1 + ( F ) . Replacing F by F - λ 1 + ( F ) we may assume that λ 1 + ( F ) = 0 . Considering any λ satisfying λ > 0 we show that λ 1 , + ( F ) λ . Recall from Lemma 3.1 that λ 1 + ( F , n ) λ 1 + ( F ) as n . Thus we can find k large enough satisfying λ > λ 1 + ( F , k ) > λ 1 + ( F ) = 0 . Let ψ k + E p ( k ) , p < , satisfy

{ F ( D 2 ψ k + , D ψ k + , ψ k + , x ) = - λ 1 , k + ψ k + in  k , ψ 1 + > 0 in  k , ψ k + = 0 in  k ,

where λ 1 + ( F , k ) = λ 1 , k + . Let δ ~ = sup k δ where δ is given by (H3). Normalize ψ k + so that

ψ k + L ( k ) = min { 1 , λ - λ 1 , k + λ + δ ~ } .

Now we plan to find a bounded, positive solution of

(3.4) F ( D 2 u , D u , u , x ) = ( λ + c + ( x ) ) u 2 - λ u in  N ,

where c ( x ) = F ( 0 , 0 , 1 , x ) L loc ( N ) . This would imply F ( D 2 u , D u , u , x ) - λ u , and therefore, λ 1 , + ( F ) λ . Thus to complete the proof of Step 1 we only need to establish (3.4).

Let u ¯ = 1 and u ¯ = ψ k + . Note that u ¯ is a supersolution in N and u ¯ is a subsolution in k . Now fix any ball containing k . Since 0 is a subsolution, by Lemma 3.3, we find v E p ( ) , p < , with 0 v 1 and satisfies

F ( D 2 v , D v , v , x ) = ( λ + c + ( x ) ) v 2 - λ v in  , v = 0 on .

The proof of Lemma 3.3 also reveals that v ψ k + in k . Now choosing a sequence of increasing to N , and the interior estimate [25, Theorem 3.3] we can find a subsequence locally converging to a solution u of (3.4). Positivity of u follows from Theorem 3.1.

Step 2. We next show that λ 1 , - ( F ) λ 1 - ( F ) . Replacing F by F - λ 1 - ( F ) we may assume that λ 1 - ( F ) = 0 . Considering any λ satisfying λ > 0 we show that λ 1 , - ( F ) λ . As done in Step 1, we can choose k large enough so that λ > λ 1 - ( F , k ) λ 1 , k - and there exists ψ k - E p ( k ) satisfying

{ F ( D 2 ψ k - , D ψ k - , ψ k - , x ) = - λ 1 , k - ψ k - in  k , ψ k - < 0  in  k , ψ k - = 0 in  k .

Normalize ψ k - so that

ψ k - L ( k ) = min { 1 , λ - λ 1 , k - λ + δ ~ } ,

where δ ~ is same as in Step 1. Then

F ( D 2 ψ k - , D ψ k - , ψ k - , x ) - ( λ + c - ( x ) ) ( ψ k - ) 2 - λ ψ k - in  k .

Thus, using Lemma 3.3 and the arguments of Step 1, we obtain a negative, bounded solution u 𝒲 loc 2 , p ( N ) , p < , to

F ( D 2 u , D u , u , x ) = - ( λ + c - ( x ) ) u 2 - λ u - λ u .

This, of course, implies λ 1 , - ( F ) λ . Hence the theorem. ∎

Theorem 2.2 (ii) will be proved using Theorem 2.4. Thus we prove Theorem 2.4 first.

Theorem 3.3.

Suppose that either (2.2) and (2.4) or (2.3) and (2.5) hold. Then F satisfies β + -MP in R N provided λ β ′′ , + ( F ) > 0 .

Proof.

Let u 𝒲 loc 2 , N ( N ) be a function satisfying

F ( D 2 u , D u , u , x ) 0 in  N    and    sup N u β < .

Also, since λ β ′′ , + ( F ) > 0 , there exist λ > 0 and ψ 𝒲 loc 2 , N ( N ) with the property that ψ β and

F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0 in  N .

Multiplying ψ with a suitable constant, we may assume that ψ u .

For this proof we follow the idea of [12, Theorem 4.2]. Choose a smooth positive function χ : N such that, for | x | > 1 ,

χ ( x ) = { | x | σ if  β  satisfies (2.2) , exp ( σ | x | ) if  β  satisfies (2.3) .

Using (H3) and an easy computation, we obtain, for x 1 c ,

F ( D 2 χ , D χ , χ , x ) { [ ( σ 2 + N σ - 2 σ ) Λ ( x ) | x | 2 + σ γ ( x ) | x | + δ ( x ) ] χ if  β  satisfies (2.2) , [ σ ( σ + N - 1 | x | ) Λ ( x ) + σ γ ( x ) + δ ( x ) ] χ if  β  satisfies (2.3) .

Hence for both the cases, using (2.4) and (2.5) accordingly on ¯ 1 c , there exists a positive constant C such that

(3.5) F ( D 2 χ , D χ , χ , x ) C χ .

Modifying C, if required, we can assume (3.5) to hold in N . Now set ψ n = ψ + 1 n χ and define κ n = sup N u ψ n . If κ n 0 , then there is nothing to prove. Thus we assume κ n > 0 to reach a contradiction. Since ψ u , it follows that κ n 1 and κ n κ n + 1 for all n 1 . Moreover, by (2.2) and (2.3),

lim sup | x | u ( x ) ψ n ( x ) n sup N u β lim sup | x | β ( x ) χ ( x ) = 0 .

Hence there exist x n N such that κ n = u ( x n ) ψ n ( x n ) .

Let us now estimate the term χ ( x n ) n . Note that

1 κ 2 n ψ 2 n ( x n ) u ( x n ) = 1 κ n - χ ( x n ) 2 n u ( x n ) ,

which implies

χ ( x n ) n 2 ( 1 κ n - 1 κ 2 n ) u ( x n ) 2 ( 1 κ n - 1 κ 2 n ) ψ ( x n ) .

Hence for each natural number n there exist a small positive η n such that

(3.6) χ ( x ) n ( 1 κ n - 1 κ 2 n ) ψ ( x ) in  η n ( x n ) .

On the other hand, using convexity of F with (3.5) and (3.6), we get

F ( D 2 ψ n , D ψ n , ψ n , x ) F ( D 2 ψ , D ψ , ψ , x ) + 1 n F ( D 2 χ , D χ , χ , x ) [ - λ + C ( 1 κ n - 1 κ 2 n ) ] ψ ( x ) in  η n ( x n ) .

Since { κ n } is a convergent sequence, we can choose m large enough so that

(3.7) F ( D 2 ψ m , D ψ m , ψ m , x ) < 0 in  η m ( x m ) .

Now note that w = κ m ψ m - u is non-negative and by (H3), there exist positive a , b such that in η m ( x m ) we have

λ , Λ - ( x , D 2 w ) - a | D w | - b w κ m F ( D 2 ψ m , D ψ m , ψ m , x ) - F ( D 2 u , D u , u , x ) < 0 .

By the strong maximum principle [25, Lemma 3.1] we then obtain w 0 in η m ( x m ) . But this contradicts (3.7) as

0 F ( D 2 u , D u , u , x ) = κ m F ( D 2 ψ m , D ψ m , ψ m , x ) < 0 in  η m ( x m ) .

Therefore, κ n 0 for large n and hence u 0 . ∎

In the same spirit of Theorem 3.3 we can also prove β - -MP.

Theorem 3.4.

Suppose that either (2.2) and (2.4) or (2.3) and (2.5) hold for the function β. Then F satisfies ( - β ) - -MP in R N provided λ β ′′ , - ( F ) > 0 .

Proof.

As done in Theorem 3.3, we choose λ ( 0 , λ β ′′ , - ( F ) ) and ψ 𝒲 loc 2 , N ( N ) satisfying ψ - β and

F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0 in  N .

Let u 𝒲 loc 2 , N ( N ) be a function satisfying

F ( D 2 u , D u , u , x ) 0 in  N    and    sup N u ( - β ) < .

We need to show that u 0 . To the contrary, we suppose that u is negative somewhere in N . Multiplying ψ with a suitable positive constant, we may assume ψ u . Consider the function χ from Theorem 3.3 and note that (3.5) holds. Set ψ n ( x ) = ψ ( x ) - 1 n χ ( x ) and κ n sup N u ψ n . It can easily be checked that ( κ n ) n is positive, increasing and bounded by 1. Furthermore, κ n = u ( x n ) ψ n ( x n ) for some x n N . Then repeating a similar calculation we find that for each natural number n there exist a small positive η n satisfying

- χ ( x ) n ( 1 κ n - 1 κ 2 n ) ψ ( x ) in  η n ( x n ) .

Then using convexity, (3.5) and above estimate, we obtain

F ( D 2 ψ n , D ψ n , ψ n , x ) F ( D 2 ψ , D ψ , ψ , x ) - 1 n F ( D 2 χ , D χ , χ , x )
[ - λ ψ ( x ) - C χ ( x ) n ]
[ - λ + C ( 1 κ n - 1 κ 2 n ) ] ψ ( x )

in η n ( x n ) . As ψ ( x ) is negative and { κ n } is convergent, we can choose m large enough such that

(3.8) F ( D 2 ψ m , D ψ m , ψ m , x ) > 0 in  η m ( x m ) .

Note that w κ m ψ n - u is a non-positive function vanishing at x m . Repeating the arguments of Theorem 3.3, we find positive constants a 1 , b 1 satisfying

λ , Λ + ( x , D 2 w ) + a 1 | D w | - b 1 w 0 in  η m ( x m ) .

This, of course, implies w 0 in η m ( x m ) which is a contradiction to (3.8). Thus it must hold that u 0 . ∎

Proof of Theorem 2.4.

The proof follows by combining Theorems 3.3 and 3.4. ∎

Now we prove Theorem 2.5.

Proof of Theorem 2.5.

First we consider (i). To the contrary, suppose that λ β , + ( F ) < 0 . Then there exists λ < 0 such that λ β , + ( F ) < λ < 0 and there exists ψ 𝒲 loc 2 , N ( N ) satisfying

0 < ψ β , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0 .

This, of course, implies F ( D 2 ψ , D ψ , ψ , x ) - λ ψ > 0 and sup ψ β 1 . This clearly violates β + -MP.

Next we consider (ii). Again, we suppose that λ β , - ( F ) < 0 . Then there exists λ < 0 such that λ β , - ( F ) < λ < 0 and there exists ψ 𝒲 loc 2 , N ( N ) satisfying

0 > ψ - β , F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0 .

This gives F ( D 2 ψ , D ψ , ψ , x ) - λ ψ < 0 and sup ψ ( - β ) 1 . This clearly violates ( - β ) - -MP. ∎

Now we can prove Theorem 2.2 (ii).

Theorem 3.5.

Assume that either (2.4) or (2.5) holds. Then we have

λ 1 ′′ , + ( F ) λ 1 , + ( F ) 𝑎𝑛𝑑 λ 1 ′′ , - ( F ) λ 1 , - ( F ) .

Proof.

Let us first show that λ 1 ′′ , + ( F ) λ 1 , + ( F ) . To the contrary, suppose that there exists λ with λ < λ 1 ′′ , + ( F ) and λ 1 , + ( F ) < λ . Then there exists positive ψ 𝒲 loc 2 , N ( N ) L ( N ) such that F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0 . Also, note that λ 1 ′′ , + ( F + λ ) = λ 1 ′′ , + - λ > 0 . By Theorem 3.3, the operator F + λ satisfies + MP. Therefore, ψ 0 which contradicts the fact ψ > 0 . Hence we must have λ 1 ′′ , + ( F ) λ 1 , + ( F ) .

We prove the second claim. To the contrary, suppose that there exists λ with λ < λ 1 ′′ , - ( F ) and λ 1 , - ( F ) < λ . Then there exists a negative function ψ 𝒲 loc 2 , N ( N ) L ( N ) such that F ( D 2 ψ , D ψ , ψ , x ) + λ ψ  0 . Also, we have λ 1 ′′ , - ( F + λ ) = λ 1 ′′ , - ( F ) - λ > 0 , and therefore, the operator F + λ satisfies - MP. This gives ψ 0 which contradicts the fact ψ < 0 . Hence we must have λ 1 ′′ , - ( F ) λ 1 , - ( F ) . ∎

Proof of Theorem 2.2.

The proof follows by combining Theorems 3.2 and 3.5. ∎

Our next result should be compared with [12, Theorem 7.6]. Recall that for a smooth domain Ω,

λ 1 ′′ , + ( F , Ω ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( Ω )  such that  inf Ω ψ > 0  and  F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  Ω } ,
λ 1 ′′ , - ( F , Ω ) sup { λ : there exists  ψ 𝒲 loc 2 , N ( Ω )  such that  sup Ω ψ < 0  and  F ( D 2 ψ , D ψ , ψ , x ) + λ ψ 0  in  Ω } .

Theorem 3.6.

It holds that

λ 1 ′′ , + ( F ) = min { λ 1 + ( F ) , lim r λ 1 ′′ , + ( F , ¯ r c ) } .

Proof.

Notice that the function λ 1 ′′ , + ( r ) λ 1 ′′ , + ( F , ¯ r c ) is an increasing function with respect to r and

λ 1 ′′ , + ( F ) lim r λ 1 ′′ , + ( r ) .

Also, from definition we already have λ 1 ′′ , + ( F ) λ 1 + ( F ) . Combining these two, we obtain

λ 1 ′′ , + ( F ) min { λ 1 + ( F ) , lim r λ 1 ′′ , + ( F , ¯ r c ) } .

Let us now show that the above inequality cannot be strict. That is, for every

λ < min { λ 1 + ( F ) , lim r λ 1 ′′ , + ( F , ¯ r c ) } ,

we have λ 1 ′′ , + ( F ) λ . To do this, we need to construct a positive supersolution of the operator F + λ in the admissible class of λ 1 ′′ , + ( F ) . Choose a positive number R so that λ < λ 1 ′′ , + ( R ) . Then there exists positive function ϕ 𝒲 loc 2 , N ( ¯ R c ) with inf R c ϕ > 0 and F ( D 2 ϕ , D ϕ , ϕ , x ) + λ ϕ 0 in ¯ R c . We claim that there exists a function φ 𝒲 loc 2 , p ( R + 1 c ) , p > N , with inf R + 1 c φ 1 and satisfies

(3.9) F ( D 2 φ , D φ , φ , x ) + λ φ 0 in  R + 1 c .

Let us first complete the proof assuming (3.9). By Morrey’s inequality we see that φ 𝒞 1 ( ¯ R + 1 c ) . Consider a positive eigenfunction ψ 𝒲 loc 2 , N ( N ) associated to λ 1 + ( F ) . Choose a non-negative function χ 𝒞 2 ( N ) such that χ = 0 in R + 2 and χ = 1 in R + 3 c . For ϵ > 0 , define u ψ + ϵ χ φ . Using convexity of F, we can write

F ( D 2 u , D u , u , x ) F ( D 2 ψ , D ψ , ψ , x ) + ϵ F ( D 2 ( χ φ ) , D ( χ φ ) , ( χ φ ) , x ) .

From the construction we can immediately say that F ( D 2 u , D u , u , x ) + λ u 0 in R + 2 R + 3 c . We are left with the annuals region ¯ R + 3 R + 2 . In this compact set we have

F ( D 2 u , D u , u , x ) + λ u ( λ - λ 1 + ( F ) ) ψ + ϵ [ F ( D 2 ( χ φ ) , D ( χ φ ) , ( χ φ ) , x ) + λ χ φ ]
= ( λ - λ 1 + ( F ) ) ψ + ϵ [ F ( χ D 2 φ + 2 D χ D φ + φ D 2 χ , χ D φ + φ D χ , χ φ , x ) + λ χ φ ]
( λ - λ 1 + ( F ) ) ψ + ϵ χ [ F ( D 2 φ , D φ , φ , x ) + λ φ ] + ϵ F ( 2 D χ D φ + φ D 2 χ , φ D χ , 0 , x )
( λ - λ 1 + ( F ) ) ψ + ϵ C < 0

for ϵ small enough, where we have again used convexity of F. This, of course, implies λ 1 ′′ , + ( F ) λ , as required.

To complete the proof, we only need to show (3.9). To this end, we may assume that inf ϕ 2 . Let c ( x ) = F ( 0 , 0 , 1 , x ) + λ and define f ( x , u ) = | c ( x ) | f ( u ) , where f : ( - , 0 ] is a Lipschitz function with the property that f ( 1 ) = - 1 , f ( t ) = 0 for t 2 . Then u ¯ = ϕ is supersolution to

F ( D 2 u , D u , u , x ) + λ u = f ( x , u ) in  R c ,

and u ¯ = 1 is a subsolution. The existence of a solution to (3.9) follows by constructing solutions (squeezed between u ¯ and u ¯ ) in an increasing sequence of bounded domains in R c and the passing to the limit using local stability bound [25, Theorem 3.3]. To construct a solution in any smooth bounded domain, we may follow the idea of Lemma 3.3 with the help of general existence results from [27, Theorem 4.6] which deals with non-zero boundary condition. ∎

Now we would like to see if a result analogous to Theorem 3.6 holds for λ 1 ′′ , - ( F ) . Denote by

G ( M , p , u , x ) = - F ( - M , - p , - u , x ) .

It is easily seen that G is a concave operator and λ 1 ′′ , - ( F ) = λ 1 ′′ , + ( G ) . But we cannot apply the arguments of Theorem 3.6 for concave operators. To obtain the results, we impose a mild condition at infinity.

Theorem 3.7.

Suppose that

(3.10) lim r λ 1 ′′ , - ( F , ¯ r c ) = lim r λ 1 ′′ , - ( G , ¯ r c ) .

Then we have

λ 1 ′′ , - ( F ) = min { λ 1 - ( F ) , lim r λ 1 ′′ , - ( F , ¯ r c ) } .

Proof.

It is easy to see that

λ 1 ′′ , - ( F ) min { λ 1 - ( F ) , lim r λ 1 ′′ , - ( F , ¯ r c ) } .

As done in Theorem 3.6, we show that the above inequality can be strict. So we consider any

(3.11) λ < min { λ 1 - ( F ) , lim r λ 1 ′′ , - ( F , ¯ r c ) } ,

and show that λ 1 ′′ , - ( F ) λ . We now construct a subsolution of the operator F + λ in the admissible class of λ 1 ′′ , - ( F ) . Using (3.10) and (3.11), we find a positive R so that

λ < λ 1 ′′ , - ( G , ¯ R c ) .

Hence repeating the arguments of Theorem 3.6, we can find φ 𝒲 loc 2 , p ( R + 1 c ) , p > N , with sup R + 1 c φ < 0 and G ( D 2 φ , D φ , φ , x ) + λ φ 0 in R + 1 c . By Morrey’s inequality φ 𝒞 1 ( B R + 1 c ) . Also, consider a negative eigenfunction ψ 𝒲 loc 2 , N ( N ) associated to λ 1 - ( F ) . Let χ be the cut-off function in Theorem 3.6 and define u = ψ + ϵ χ φ for ϵ > 0 . Since, by convexity,

F ( D 2 u , D u , u , x ) F ( D 2 ψ , D ψ , ψ , x ) + ϵ G ( D 2 ( χ ϕ ) , D ( χ ϕ ) , ( χ ϕ ) , x ) ,

repeating a calculation analogous to Theorem 3.6, we find that for some ϵ small F ( D 2 u , D u , u , x ) + λ u 0 in N . Thus we get λ 1 ′′ , - ( F ) λ . ∎

To this end, we define c ( x ) = F ( 0 , 0 , 1 , x ) and d ( x ) = F ( 0 , 0 , - 1 , x ) . Our next result is a generalization to [12, Proposition 1.11].

Proposition 3.1.

Define ζ = lim sup | x | c ( x ) and ξ = lim sup | x | d ( x ) . Then the following hold:

  1. Suppose that ζ < 0 , and either ( 2.4 ) or ( 2.5 ) holds. Then F satisfies the + MP if and only if λ 1 + ( F ) > 0 .

  2. Suppose that ξ > 0 , and either ( 2.4 ) or ( 2.5 ) holds. Furthermore, assume ( 3.10 ). Then F satisfies the - MP if and only if λ 1 - ( F ) > 0 .

We need a small lemma to prove Proposition 3.1.

Lemma 3.4.

The following hold for any smooth domain Ω:

  1. - sup Ω c ( x ) inf Ω d ( x ) ,

  2. - sup Ω c ( x ) λ 1 ′′ , + ( F , Ω ) ,

  3. inf Ω d ( x ) λ 1 ′′ , - ( F , Ω ) .

Proof.

Part (i) follows from convexity property of F. Note that for λ = - sup Ω c ( x ) , ψ = 1 is an admissible function for λ 1 ′′ , + ( F , Ω ) . This gives us (ii). In a similar fashion we get (iii). ∎

Now we prove Proposition 3.1

Proof of Proposition 3.1.

First consider (i). Assume that λ 1 + ( F ) > 0 . Using Lemma 3.4, we obtain

(3.12) 0 < - ζ = lim r ( - sup ¯ r c c ( x ) ) lim r λ 1 ′′ , + ( F , ¯ r c ) .

By Theorem 3.6, we obtain λ 1 ′′ , + ( F ) > 0 , and therefore, using Theorem 3.3, we see that F satisfies the + MP. To show the converse direction, we assume that F satisfies + MP. Then Theorem 2.5 implies that λ 1 , + ( F ) 0 . Using Theorem 3.2, we then have λ 1 + ( F ) 0 . If possible, suppose that λ 1 + ( F ) = 0 . We show that there exists a bounded principal eigenfunction φ which would give a contradiction to the validity of + MP, and hence we must have λ 1 + ( F ) > 0 . Consider a smooth positive function ϕ satisfying ϕ = 1 in r c for some large r. Since ζ < 0 , we have a compact set K satisfying

c ( x ) < ζ 2 ϕ ( x ) , ϕ ( x ) = 1 , x K c .

Recall the Dirichlet principal eigenfunction ψ 1 , n + from (3.1). Choose κ n = max n ψ 1 , n + ϕ and define v n = κ n - 1 ψ 1 , n + . Observe that ϕ - v n must vanish in K. Indeed, in n K we have

λ , Λ - ( x , D 2 ( ϕ - v n ) ) - γ ( x ) | D ϕ - v n | - δ ( x ) ( ϕ - v n ) F ( 0 , 0 , ϕ , x ) - F ( D 2 v n , D v n , v n , x )
ζ 2 + λ 1 , n + v n ζ 2 + λ 1 , n + < 0

for all large n, and therefore, by strong maximum principle [25, Lemma 3.1], ϕ - v n cannot vanish in n K . Now applying Harnack’s inequality and standard 𝒲 2 , p estimates we can extract a convergent subsequence of v n converging to a positive eigenfunction φ. This completes the proof.

Proof for (ii) would be analogous. ∎

Next we prove Theorem 2.3.

Proof of Theorem 2.3.

(i) From the definition it follows that

λ 1 ′′ , + ( F , ¯ r c ) λ 1 ′′ , + ( F ~ , ¯ r c ) - sup ¯ r c γ ~ ( x ) ,

and then letting r towards infinity we have

lim r λ 1 ′′ , + ( F , ¯ r c ) lim r λ 1 ′′ , + ( F ~ , ¯ r c ) λ 1 ′′ , + ( F ~ ) = λ 1 + ( F ~ ) .

Since γ ~ ( x ) 0 , it gives us λ 1 + ( F ~ ) λ 1 + ( F ) . Combining it with above calculation, we find

lim r λ 1 ′′ , + ( F , ¯ r c ) λ 1 + ( F ) .

Applying Theorem 3.6, we obtain λ 1 + ( F ) = λ 1 ′′ , + ( F ) .

(ii) Using Lemma 3.4 and the given hypothesis, we find

λ 1 + ( F ) - lim sup | x | c ( x ) = lim r ( - sup ¯ r c c ( x ) ) lim r λ 1 ′′ , + ( F , ¯ r c ) .

Hence, by Theorem 3.6, we get λ 1 + ( F ) = λ 1 ′′ , + ( F ) .

(iii) We show that under the given condition we have (ii). It is enough to show that if σ < lim sup | x | c ( x ) , then λ 1 + ( F ) - σ . Now define a positive function

ψ ( x ) = exp ( - 1 1 - | ε x | 2 )

on the ball 1 ε where an appropriate ε will be chosen later. It is easily checked that

D x i ψ = - 2 ε 2 x i ( 1 - | ε x | 2 ) 2 ψ , D x i x j ψ = [ 4 ε 4 ( 1 - | ε x | 2 ) 4 x i x j - 2 ε 2 ( 1 - | ε x | 2 ) 2 δ i j - 8 ε 4 ( 1 - | ε x | 2 ) 3 x i x j ] ψ .

For x 0 N , define ϕ ( x ) = ψ ( x - x 0 ) . We will choose ε and x 0 such that

(3.13) F ( D 2 ϕ , D ϕ , ϕ , x ) - σ ϕ > 0 in  1 ε ( x 0 ) .

Since all the notions of eigenvalues of F coincide in bounded domains (cf. [25]), using (3.13) we deduce

- σ λ 1 , + ( F , 1 ε ( x 0 ) ) = λ 1 + ( F , 1 ε ( x 0 ) ) λ 1 + ( F ) .

Thus we only need to establish (3.13). For a different way to construct such subsolutions we refer to [26]. Using (H3), we see that

F ( D 2 ϕ , D ϕ , ϕ , x ) - σ ϕ = F ( D 2 ϕ , D ϕ , ϕ , x ) - F ( 0 , 0 , ϕ , x ) + F ( 0 , 0 , 1 , x ) ϕ - σ ϕ
λ , Λ - ( x , D 2 ϕ ) - γ ( x ) | D ϕ | + c ( x ) ϕ - σ ϕ
[ 4 λ 0 ε 2 | ε ( x - x 0 ) | 2 ( 1 - | ε ( x - x 0 ) | 2 ) 4 - 2 N Λ 0 ε 2 ( 1 - | ε ( x - x 0 ) | 2 ) 2 - 8 Λ ε 2 | ε ( x - x 0 ) | 2 ( 1 - | ε ( x - x 0 ) | 2 ) 3
(3.14) - 2 ϵ 2 | x - x 0 | γ ( x ) ( 1 - | ϵ ( x - x 0 ) | 2 ) 2 + c ( x ) - σ ] ϕ .

Given ε, we choose R such that | γ ( x ) | ε for | x | R and then choose x 0 N satisfying | x 0 | R + 2 ε - 1 . Furthermore, due to our hypothesis, we can choose x 0 such that

(3.15) inf 1 ε ( x 0 ) c ( x ) > σ .

We now compute (3.14) in two steps.

Step 1. Suppose 1 - δ < | ε ( x - x 0 ) | 2 < 1 , where δ is very close to zero and will be chosen later. It then follows from (3.14) that

F ( D 2 ϕ , D ϕ , ϕ , x ) - σ ϕ ε 2 ( 1 - | ε ( x - x 0 ) | 2 ) 4 [ 4 λ ( 1 - δ ) - 2 N Λ δ 2 - 8 Λ ( 1 - δ ) δ - 2 δ 2 ] ϕ + ( c ( x ) - σ ) ϕ .

Now we can choose small positive δ, independent of ε, so that 4 λ ( 1 - δ ) - 2 N Λ δ 2 - 8 Λ ( 1 - δ ) δ - 2 δ 2 > 0 . This proves (3.13) in the annulus.

Step 2. Now we are left with the part 0 | ε ( x - x 0 ) | 2 1 - δ , where δ is already chosen in Step 1. An easy calculation reveals

F ( D 2 ϕ , D ϕ , ϕ , x ) - σ ϕ [ ( c ( x ) - σ ) - 2 N Λ ε 2 δ 2 - 8 Λ ( 1 - δ ) ε 2 δ 3 - 2 ε 2 δ 2 ] ϕ .

Using (3.15), we can choose ε small enough so that the right-hand side becomes positive.

Combining the above steps, we obtain (3.13), completing the proof of part (iii).

(iv) This follows from Theorem 3.6. Let us also provide a more direct proof. Let φ * be an eigenfunction corresponding to λ 1 + ( F ) = λ 1 + . For δ , ε > 0 we define ϕ ε = φ * + ε V . Choose ε small enough so that

(3.16) δ min ¯ φ * > ε max ¯ [ F ( D 2 V , D V , V , x ) + λ 1 + V ] .

By using convexity and homogeneity, it follows that

F ( D 2 ϕ ε , D ϕ ε , ϕ ε , x ) F ( D 2 φ * , D φ * , φ * , x ) + ε F ( D 2 V , D V , V , x )
= - λ 1 + φ * + ε 𝟙 ( x ) F ( D 2 V , D V , V , x ) - ε λ 1 +  1 c ( x ) V ( x )
- λ 1 + ϕ ε + ε max ¯ [ F ( D 2 V , D V , V , x ) + λ 1 + V ]
- ( λ 1 + - δ ) ϕ ε ,

using (3.16). Hence λ 1 ′′ , + ( F ) λ 1 + ( F ) - δ and from the arbitrariness of δ the result follows. ∎

Thus it remains to prove Theorems 2.6 and 2.7. Let us first attack Theorem 2.6.

Proof of Theorem 2.6.

Without any loss of generality, we assume that λ 1 + ( F ) = 0 . Recall from Lemma 3.1 that the pair ( ψ 1 , n + , λ 1 , n + ) solving the Dirichlet eigenvalue problem with positive eigenfunction in n . That is,

(3.17) { F ( D 2 ψ 1 , n + , D ψ 1 , n + , ψ 1 , n + , x ) = - λ 1 , n + ψ 1 , n + in  n , ψ 1 , n + > 0 in  n , ψ 1 , n + = 0 on  n .

Let κ n > 0 be such that κ n ψ 1 , n + V in n and it touches V at some point in n . We claim that κ n ψ 1 , n + has to touch V inside K. Note that, by (H3), if w = V - κ n ψ 1 , n + , then

λ , Λ - ( x , w ) - γ | D w | - δ w - ε V + λ 1 , n + ( κ n ψ 1 , n + ) ( - ε + λ 1 , n + ) ( κ n ψ 1 , n + ) 0 in  K c n

for large n using (2.6) and (3.17). Thus, if w vanishes in K c n , then it must be identically 0 in K c n , by the strong maximum principle [25, Lemma 3.1]. And this is not possible since w > 0 on n . Now onwards we denote κ n ψ 1 , n + by ψ 1 , n + . By the above normalization, ψ 1 , n + would converge, up to a subsequence, to a positive function φ 𝒲 loc 2 , p ( N ) , p < , an eigenfunction corresponding to λ 1 + ( F ) = 0 . See, for instance, the argument in Lemma 3.1.

We now show that any other principal eigenfunction is a multiple to φ. For η, a small positive number, we define Ξ η = ψ 1 , n + - η V . Using convexity of F, we note that, in n K c ,

F ( D 2 Ξ η , D Ξ η , Ξ , x ) F ( D 2 ψ 1 , n + , D ψ 1 , n + , ψ 1 , n + , x ) - η F ( D 2 V , D V , V , x )
( - λ 1 , n + ψ 1 , n + + η ε V )
( - λ 1 , n + + η ε ) V > 0 ,

provided we choose n large (depending on η). Let ψ be any principal eigenfunction satisfying

F ( D 2 ψ , D ψ , ψ , x ) = 0 in  N .

Define

δ = δ ( η ) = min K ψ Ξ η .

Then δ Ξ η ψ on K. Since, by the Harnack inequality,

0 < inf n inf K ψ 1 , n + sup n sup K ψ 1 , n + < ,

we can choose η 0 small enough (independent of n) so that

0 < inf η ( 0 , η 0 ] inf n inf K Ξ η sup η ( 0 , η 0 ] sup n sup K Ξ η < ,

Thus, δ remains bounded and positive as n and η 0 . Since F ( D 2 ψ , D ψ , ψ , x ) = 0 in n K c and λ 1 + ( F , n K c ) > 0 , it follows from [25, Theorem 1.5] that

δ Ξ η ψ in  n .

Furthermore, there exists x η K so that δ Ξ η ( x η ) = ψ ( x η ) . Now letting n first, and then η 0 , we can extract a subsequence so that δ θ > 0 , and x η x ^ K and θ φ ( x ^ ) = ψ ( x ^ ) with θ φ ψ in N . Let u = ψ - θ φ . It is easy to see that

λ , Λ - ( x , u ) - γ | D u | - δ u 0 in  N .

By the strong maximum principle we must have u = 0 and hence the proof. ∎

Finally, we prove Theorem 2.7.

Proof of Theorem 2.7.

The main idea of the proof is the same as that of the proof of Theorem 2.6. Without any loss in generality, we assume that λ 1 - ( F ) = 0 . Let ( ψ 1 , n - , λ 1 , n - ) be the pair satisfying the Dirichlet eigenvalue problem in the ball n , i.e.,

(3.18) { F ( D 2 ψ 1 , n - , D ψ 1 , n - , ψ 1 , n - , x ) = - λ 1 , n - ψ 1 , n - in  n , ψ 1 , n - < 0 in  n , ψ 1 , n - = 0 on  n .

By Lemma 3.2, ψ 1 , n - 0 as n . Recall that G ( M , p , u , x ) - F ( - M , - p , - u , x ) . Denote by ϕ n = - ψ 1 , n - . Then we get from (3.18) that

{ G ( D 2 ϕ n , D ϕ n , ϕ n , x ) = - λ 1 , n - ϕ n in  n , ψ n < 0 in  n , ϕ n = 0 on  n .

Note that G satisfies (H1), (H2) and (H3) but it is a concave operator. So need some extra care to apply the proof of Theorem 2.6. Since F is convex, it follows from (2.7) that

(3.19) G ( D 2 V , D V , V , x ) F ( D 2 V , D V , V , x ) - ( λ 1 - ( F ) + ε ) V for all  x K c .

As done in Theorem 2.6, using (3.19), we can normalize ϕ n to touch V from below and it would touch V somewhere in K. Therefore, we can apply the Harnack inequality (see Lemma 3.2) to find a positive function φ such that ϕ n φ in 𝒲 loc 2 , p ( N ) , p > N , along some subsequence and

0 = - λ 1 - ( F ) φ = G ( D 2 φ , D φ , φ , x ) = - F ( - D 2 φ , - D φ , - φ , x ) in  N .

It is enough to show that φ agrees with any other positive eigenfunction (up to a multiplicative constant) of G with eigenvalue 0.

Next we define Ξ η ( x ) = ϕ n - η V . Since ϕ n - φ L ( K ) 0 , it is evident that Ξ η > 0 for all η small, independent of n. Using (2.7) and (3.18), we see that, in K c n ,

F ( - D 2 Ξ η , - D Ξ η , - Ξ , x ) F ( - D 2 ϕ n , - D ϕ n , - ϕ n , x ) + η F ( D 2 V , D V , V , x )
( λ 1 , n - ϕ n - η ε V )
(3.20) ( | λ 1 , n - | - η ε ) V < 0

for all large n. Now consider any positive eigenfunction ψ 𝒲 loc 2 , p ( N ) satisfying

F ( - D 2 ψ , - D ψ , - ψ , x ) = 0 ,

and let

δ = δ ( η ) = min K ψ Ξ η .

Then - δ Ξ η - ψ on K n for all n. From (2.7) it follows that λ 1 + ( F , K c ) ε . Since

λ 1 + ( F , K c n ) λ 1 + ( F , K c ) > 0 as  n ,

we can apply the maximum principle [25, Theorem 1.5] in K c n for all large n. From (3.20) we therefore get ψ δ Ξ η and δ Ξ η touches ψ at some point in K. Now we can follow the arguments in Theorem 2.6 we show that φ = t ψ for some t > 0 . Hence the proof. ∎

We conclude the paper with a remark on the eigenvalue problem in a general smooth unbounded domain.

Remark 3.1.

For the case of an unbounded domain with smooth boundary all the results developed here hold true and the proofs would be somewhat similar. As mentioned in [12], in case of general unbounded domains, one needs the boundary Harnack property to control the behavior of eigenfunctions near the boundary. For the operator F, the boundary Harnack property has been obtained recently by Armstrong, Sirakov and Smart in [7, Appendix A]. Therefore one can easily adopt the techniques of [12] along with our results to deal with general unbounded domains.


Communicated by Hitoshi Ishii


Funding statement: The research of Anup Biswas was supported in part by DST-SERB grants EMR/2016/004810 and MTR/2018/000028. Prasun Roychowdhury was supported in part by Council of Scientific & Industrial Research (File no. 09/936(0182)/2017-EMR-I).

References

[1] A. Arapostathis and A. Biswas, A variational formula for risk-sensitive control of diffusions in d , SIAM J. Control Optim. 58 (2020), no. 1, 85–103. 10.1137/18M1218704Search in Google Scholar

[2] A. Arapostathis, A. Biswas, V. Borkar and K. Suresh Kumar, A variational characterization of the risk-sensitive average reward for controlled diffusions on d , preprint (2019), https://arxiv.org/abs/1903.08346. Search in Google Scholar

[3] A. Arapostathis, A. Biswas and D. Ganguly, Certain Liouville properties of eigenfunctions of elliptic operators, Trans. Amer. Math. Soc. 371 (2019), no. 6, 4377–4409. 10.1090/tran/7694Search in Google Scholar

[4] A. Arapostathis, A. Biswas and S. Saha, Strict monotonicity of principal eigenvalues of elliptic operators in d and risk-sensitive control, J. Math. Pures Appl. (9) 124 (2019), 169–219. 10.1016/j.matpur.2018.05.008Search in Google Scholar

[5] S. N. Armstrong, Principal eigenvalues and an anti-maximum principle for homogeneous fully nonlinear elliptic equations, J. Differential Equations 246 (2009), no. 7, 2958–2987. 10.1016/j.jde.2008.10.026Search in Google Scholar

[6] S. N. Armstrong, The Dirichlet problem for the Bellman equation at resonance, J. Differential Equations 247 (2009), no. 3, 931–955. 10.1016/j.jde.2009.03.007Search in Google Scholar

[7] S. N. Armstrong, B. Sirakov and C. K. Smart, Singular solutions of fully nonlinear elliptic equations and applications, Arch. Ration. Mech. Anal. 205 (2012), no. 2, 345–394. 10.1007/s00205-012-0505-8Search in Google Scholar

[8] H. Berestycki, On some nonlinear Sturm–Liouville problems, J. Differential Equations 26 (1977), no. 3, 375–390. 10.1016/0022-0396(77)90086-9Search in Google Scholar

[9] H. Berestycki, I. Capuzzo Dolcetta, A. Porretta and L. Rossi, Maximum principle and generalized principal eigenvalue for degenerate elliptic operators, J. Math. Pures Appl. (9) 103 (2015), no. 5, 1276–1293. 10.1016/j.matpur.2014.10.012Search in Google Scholar

[10] 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

[11] H. Berestycki and L. Rossi, On the principal eigenvalue of elliptic operators in N and applications, J. Eur. Math. Soc. (JEMS) 8 (2006), no. 2, 195–215. 10.4171/JEMS/47Search in Google Scholar

[12] H. Berestycki and L. Rossi, Generalizations and properties of the principal eigenvalue of elliptic operators in unbounded domains, Comm. Pure Appl. Math. 68 (2015), no. 6, 1014–1065. 10.1002/cpa.21536Search in Google Scholar

[13] I. Birindelli and F. Demengel, First eigenvalue and maximum principle for fully nonlinear singular operators, Adv. Differential Equations 11 (2006), no. 1, 91–119. 10.57262/ade/1355867725Search in Google Scholar

[14] J. Busca, Existence results for Bellman equations and maximum principles in unbounded domains, Comm. Partial Differential Equations 24 (1999), no. 11–12, 2023–2042. 10.1080/03605309908821493Search in Google Scholar

[15] J. Busca, M. J. Esteban and A. Quaas, Nonlinear eigenvalues and bifurcation problems for Pucci’s operators, Ann. Inst. H. Poincaré Anal. Non Linéaire 22 (2005), no. 2, 187–206. 10.1016/j.anihpc.2004.05.004Search in Google Scholar

[16] 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

[17] M. D. Donsker and S. R. S. Varadhan, On a variational formula for the principal eigenvalue for operators with maximum principle, Proc. Natl. Acad. Sci. USA 72 (1975), 780–783. 10.1073/pnas.72.3.780Search in Google Scholar

[18] M. D. Donsker and S. R. S. Varadhan, Asymptotic evaluation of certain Markov process expectations for large time. III, Comm. Pure Appl. Math. 29 (1976), no. 4, 389–461. 10.1002/cpa.3160290405Search in Google Scholar

[19] N. Ikoma and H. Ishii, Eigenvalue problem for fully nonlinear second-order elliptic PDE on balls, Ann. Inst. H. Poincaré Anal. Non Linéaire 29 (2012), no. 5, 783–812. 10.1007/s13373-015-0071-0Search in Google Scholar

[20] H. Ishii and Y. Yoshimura, A note on demi-eigenvalues for uniformly elliptic Isaacs operators, Viscosity Solution Theory of Differential Equations and its Developments, Kyoto University, Kyoto (2006), 106–114. Search in Google Scholar

[21] H. Kaise and S.-J. Sheu, On the structure of solutions of ergodic type Bellman equation related to risk-sensitive control, Ann. Probab. 34 (2006), no. 1, 284–320. 10.1214/009117905000000431Search in Google Scholar

[22] P.-L. Lions, Bifurcation and optimal stochastic control, Nonlinear Anal. 7 (1983), no. 2, 177–207. 10.1016/0362-546X(83)90081-0Search in Google Scholar

[23] S. Patrizi, Principal eigenvalues for Isaacs operators with Neumann boundary conditions, NoDEA Nonlinear Differential Equations Appl. 16 (2009), no. 1, 79–107. 10.1007/s00030-008-7042-zSearch in Google Scholar

[24] C. Pucci, Maximum and minimum first eigenvalues for a class of elliptic operators, Proc. Amer. Math. Soc. 17 (1966), 788–795. 10.1090/S0002-9939-1966-0199576-1Search in Google Scholar

[25] A. Quaas and B. Sirakov, Principal eigenvalues and the Dirichlet problem for fully nonlinear elliptic operators, Adv. Math. 218 (2008), no. 1, 105–135. 10.1016/j.aim.2007.12.002Search in Google Scholar

[26] L. Rossi, Non-existence of positive solutions of fully nonlinear elliptic equations in unbounded domains, Commun. Pure Appl. Anal. 7 (2008), no. 1, 125–141. 10.3934/cpaa.2008.7.125Search in Google Scholar

[27] N. Winter, W 2 , p and W 1 , p -estimates at the boundary for solutions of fully nonlinear, uniformly elliptic equations, Z. Anal. Anwend. 28 (2009), no. 2, 129–164. 10.4171/ZAA/1377Search in Google Scholar

Received: 2020-04-13
Accepted: 2020-10-28
Published Online: 2020-11-25
Published in Print: 2022-10-01

© 2022 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2020-0035/html
Scroll to top button