Abstract

We are concerned with fully nonlinear uniformly elliptic operators with a superlinear gradient term. We look for local estimates, such as weak Harnack inequality and local maximum principle, and their extension up to the boundary. As applications, we deduce ABP-type estimates and weak maximum principles in general unbounded domains, a strong maximum principle, and a Liouville-type theorem.

1. Introduction

The qualitative theory of second-order elliptic equations received a strong effort from Harnack inequalities. Here, we will make use of this powerful technique to study continuous viscosity solutions of fully nonlinear elliptic equations :in unbounded domains of , where is a real function of , , and in the set of real symmetric matrices.

We recall that is (degenerate) elliptic if is nondecreasing in and uniformly elliptic if there exist (ellipticity) constants and such that andfor (1.2), that is (1.2) is semidefinite positive, where (1.2) denotes the trace of the matrix (1.2).

In the class of uniformly elliptic operators, there are two extremal ones, well known as Pucci maximal and minimal operators, respectively:where are the positive and negative parts of , which can be decomposed in a unique way as with and . Other examples of fully nonlinear uniformly elliptic operators can be found in [13].

Throughout this paper, we will consider elliptic operators with the structure conditionswhere are the extremal Pucci operators, is a continuous function and the exponent , so that the gradient term can have a superlinear, at most quadratic growth.

Remark 1.1. The above structure conditions are exactly equivalent to the uniform ellipticity when is linear in the variable . In the nonlinear case they allow a slight generalization. Let us consider, for and , the functionthen the operator is elliptic and satisfies both the conditions (1.4) and (1.5), even that it is not uniformly elliptic.
However, if (1.5) (resp., (1.4)) holds, then subsolutions (resp., supersolutions) of the equation are subsolutions (resp., supersolutions) of uniformly elliptic equations, and this is needed to prove our results.

We will be concerned principally with the following topics in unbounded domains; see [46] for classical results.

[MP] maximum principle for u.s.c. subsolutions of in the viscosity sense (v.s.), in the form

[LT] Liouville theorem for continuous solutions of v.s., in the formConcerning MP, it is worth to note that the condition from above on the size of can be weakened in the framework of the Phragmén-Lindelöf theory (see, e.g., [79]) but not omitted at all, even for classical subsolutions (see, e.g., [4, 10]). It is also well known that MP fails to hold in general in exterior domains. In fact, due to the boundedness of the fundamental solution of the Laplace equation , the function provides a counterexample to MP in . Thus we introduce a local measure-geometric condition in at , which depends on the real parameter : there exists a ball such thatwhere is the connected component of containing .

If is satisfied in at all , we simply say that is a domain (with parameter ). This is a generalization of condition of Cabré [10], which ultimately goes back to Berestycki et al. [11].

Let denote the radius of the ball provided by condition . We will call domains of cylindrical and conical type the domains such that and as , respectively. Examples of the first kind are domains with finite measure, cylinders, slabs, complements of a periodic lattice of balls, whereas cones, and complements, in the plane, of logarithmic spirals, are examples of the second kind.

In [12], it is shown that MP holds true in a domain for strong solutions of a linear second-order uniformly elliptic operator ; see also [13, 14] for earlier results and [15, 16] for viscosity solutions of a fully nonlinear operator with linear and quadratic growth in the gradient (i.e., in the case of the structure condition (1.5) with and ) provided that and as , respectively.

With the aim to find conditions on the coefficient such that MP holds in domains when , our result is the following.

Theorem 1.2 (MP). Let and . Let be a domain of satisfying condition or alternatively such that, for a closed subset of , (i) holds in each connected component of ;(ii)condition is satisfied in at each . Suppose that is a viscosity solution of and structure condition (1.5) holds with , such that .
If on and in , then in .

This yields indeed MP in a wider class of domains than , for example, the cut plane and more generally the complement of continuous semi-infinite curves in and their generalizations to hypersurfaces in .

We also outline that the limit cases and of the above mentioned papers are obtained by continuity from the intermediate cases , as it follows from Theorem 1.2. Nonetheless, there are technical improvements with respect to the previous works even in the limit cases.

Consider in particular a parabolic shaped domain , satisfying condition with , ; the limit cases and correspond to domains of cylindrical and conical types, respectively.

Based on an argument of [12], eventually passing to a smaller , we can suppose that condition is satisfied with exactly. We get the new following variant of ABP estimate.

Theorem 1.3 (ABP). Let , , , and . Let be a domain, such that condition in is fulfilled at each with , . Assume that satisfies the structure condition (1.5), with and .
If is a viscosity solution of such that in and on , thenwhere is a positive constant depending on , , , , , , , , , .

Note that in the case of a domain of cylindrical type (), it is sufficient to have , for all , as well as in the case of a quadratic growth in the gradient variable (), for all .

This result extends the previous ones contained in [10, 14] for the linear case, and [8, 16], dealing with fully nonlinear equations, in the limit situations of cylindrical/conical domains and linear/quadratic gradient terms.

Remark 1.4. In general, unless , the above ABP type estimate is different from the so-called ABP maximum principles since depends on the upper bound of if and . For ABP-type estimates of this kind in bounded domains we refer to [17]. Counterexamples to the ABP maximum principle can been found in [1719].

Consider now . The classical Liouville theorem says that harmonic functions in the entire , which are bounded either above or below, are constant. This result continues to hold for strong solutions of quasilinear uniformly elliptic equations; see [20]. For viscosity solutions of fully nonlinear uniformly elliptic equations with an additive gradient term having linear growth, we refer to [21, 22]. Our result is the following.

Theorem 1.5 (Liouville theorem). Let be such that in the viscosity sense, and assume structure conditions (1.4) and (1.5), with such that as . If is bounded either above or below, then is constant.

Remark 1.6. Under some additional assumptions, Liouville-type results also hold in un-bounded domains of containing balls of arbitrary large radius; see [23].

Our main tools are Krylov-Safonov Harnack inequalities and local MP; see [20] for strong solutions of quasilinear uniformly elliptic equations. For viscosity solutions and satisfying the structure condition (1.4), they can be found in [3] if and in [24] if ; see also [25]. In the case of linear or superlinear, almost quadratic, growth in the gradient (), weak Harnack (wH) inequality and local MP can be deduced using arguments of [17], in which a (full) Harnack inequality has been established for viscosity solutions; see also [26].

Nevertheless, for convenience of the reader we believed that it is worth to report systematically on this kind of inequalities in Section 3.

As the previous ones, our approach follows the lines of [3], based on the methods of [27, 28] and on the ABP maximum principle for viscosity solutions in bounded domains, due to Caffarelli [29].

Remark 1.7. In deriving wH inequality and local MP, we only need the Alexandroff-Bakelman-Pucci (ABP) estimate with and continuous, so [30, Proposition 2.12] and also [17, Theorem 4.1] in the case of linear growth in the gradient term, are sufficient to our purpose. But we notice that new ABP-type estimates have been established for -viscosity solutions of equations with discontinuous coefficients by Koike and Święch [19, 27] for and .

Remark 1.8. In the case of a superlinear first-order term, wH inequality and local MP are obtained by interpolation between the linear and quadratic cases, eliminating the square gradient term by means of an exponential transformation used before by Trudinger [24], see Lemmas 3.1 and 3.2 below. This kind of ideas have been also considered by Sirakov in [31].

What we definitely need are, for MP, the scaled boundary wH inequality (3.16), derived in Section 3 by means of typical viscosity methods, and, for technical reasons, its version in annular regions (3.24), and, for LT, the scaled Harnack inequality (3.11). Moreover, using the interior wH inequality (3.7) and assuming the structure condition (1.5), we also state a strong MP theorem, according to which a subsolution of equation cannot achieve a positive maximum inside any domain (open connected set) of unless it is constant; see Theorem 5.1 below. For a different approach, based on Hopf lemma, and more general versions see [32].

The paper is organized as follows. In Section 2, we recall some basic results of elliptic theory for viscosity solutions of second-order fully nonlinear equations with a linear gradient term; in Section 3, we extend local maximum principle and weak Harnack inequality, even up to the boundary, to the case of a superlinear gradient term; these results are applied in Section 4 to get Alexandroff-Bakelman-Pucci-type estimates and maximum principles, with the proof of Theorems 1.2 and 1.3; finally, a strong maximum principle is derived and the proof of Liouville theorem (Theorem 1.5) is given in Section 5. In the appendix, for the sake of completeness, we show the basic weak Harnack inequality and local MP for a uniformly elliptic operator with an additive first-order term having linear growth in the gradient.

2. Basic Estimates (Linear Gradient Term)

Let be a domain of , and denote by and , respectively, the sets of the upper and lower semicontinuous functions in . The function is said to be a viscosity subsolution of ifat any point and for all such that has a local minimum in . Similarly, a viscosity supersolution of satisfiesat any point and for all such that has a local minimum in .

We may also assume that in the above definition, that is the graph of the test function touches that one of from above for subsolutions and from below for supersolutions [3]. Moreover, if is continuous in the matrix-variable, as for uniformly elliptic operators, then we may assume that is a paraboloid, that is a quadratic polynomial.

We will make use of the following version of the ABP estimate, in which denotes the upper contact setof the graph of the function . Using [30, Proposition 2.12] or [17, Theorem 4.1], we have the following.

Lemma 2.1 (ABP estimate). Let be a viscosity supersolution of the equationin a ball of unit radius, such that on , where , for some constant . Thenfor a positive constant . Similarly, if is a viscosity subsolution of the equationsuch that on , then

From Lemma 2.1, we obtain the following results, see the appendix, which extend [3, Theorem 4.8, (1) and (2)]; see also [15].

Here we denote by a ball centered at of radius .

Lemma 2.2 (wH inequality). Let and . Suppose that is a viscosity supersolution of (2.4), with , and in . Thenwhere and are positive numbers, depending on and .

Lemma 2.3 (local MP). Let and . Suppose that is a viscosity subsolution of (2.6) with . Then for all where is a positive constant, depending on and .

3. Interior and Boundary Harnack Estimates and Local MP (Superlinear Gradient Term)

Firstly, we extend interior estimates (2.8) and (2.9) to fully nonlinear operators with a superlinear first-order term, such that, respectively, (1.4) and (1.5) hold.

Lemma 3.1 (wH inequality). Let , and . Suppose that is a viscosity solution of , under structure condition (1.4) with , , and in . Then (2.8) holds with positive constants and , depending on and .

Lemma 3.2 (local MP). Let , and . Suppose that is a viscosity solution of , under structure condition (1.5), with , , and . Then (2.9) holds for all with a positive constant , depending on and .

Proof of Lemmas 3.1 and 3.2. . We only show the proof of Lemma 3.2, since that one of Lemma 3.1 is similar. By the structure condition (1.5), we haveand also, in the viscosity sense,From this, by Young's inequality, it follows thatwithUsing the transformation , then the USC function satisfies the differential inequalityin . Therefore, we can apply Lemma 2.3 to the subsolution . To conclude the proof of Lemma 3.2, it is sufficient to observe that

Rescaling variables and functions, we highlight the dependence on geometric parameters.

Theorem 3.3 (scaled wH inequality). Let , , and . Suppose that is a viscosity solution of , under structure condition (1.4), with , , and in . Thenwith positive constants and , depending on and .

Proof. Considering, for , the function , defined by , we haveThus, applying Lemma 3.1, we getwith , from which the assert follows.

Note that constants and of the above wH inequality depend in general on the upper bound for the supersolution and on the radius of the ball, but in the case there is no dependence on and in the case no dependence on .

In the same manner as in Theorem 3.3 for wH inequality, we make the dependence on the geometric constants explicit in the following local MP.

Theorem 3.4 (scaled local MP). Let , , and . Suppose that is a viscosity solution of , under structure condition (1.5), with , , and . Then for all with a positive constant , depending on , and .

Combining Theorems 3.3 and 3.4, we get the full Harnack inequality for solutions.

Theorem 3.5 (Harnack inequality). Let , , and . Suppose that is a viscosity solution of in , under structure conditions (1.4) and (1.5), with , , and . Thenwith a positive constant .

We wish to extend the above estimates up to the boundary, that is, to balls intersecting the boundary of the domain , where the solutions are defined. For this purpose we will need suitable extensions of such solutions outside . Precisely, take concentric balls such that and . For a nonnegative viscosity supersolution of equation in , we putwhere . Similarly, for a viscosity subsolution , we putDenote also by and the continuations of and vanishing outside , respectively. Following [3, Proposition 2.8] and using the structure conditions (1.4) and (1.5), we havein for a viscosity supersolution , andin for a viscosity subsolution .

Observe that, if on , then is continuous, and then we can apply Theorem 3.3 to get a boundary wH inequality. Similarly, if on , we can use Theorem 3.4 to deduce a boundary local MP.

Nevertheless, even in the general case, when and are not necessarily continuous, we can get boundary estimates by means of an approximation process, as shown here below, where we use the notations defined just above.

Theorem 3.6 (boundary wH inequality). Let , , and . Suppose that is a viscosity solution of , under structure condition (1.4), with , , and in . Thenwith positive constants and , depending on and .

Proof. For setand, for ,whereIt is easy to check that , , andin . Therefore, we can apply Theorem 3.3 with instead of and instead of to getNote that and in , outside . Also observing that, by lower semicontinuity,and therefore, by Fatou's lemma,from inequality (3.21) we get the assert.

In the sequel, we will make also use of a version of boundary wH inequality for annular regions , , which can be deduced by Theorem 3.6 reasoning as in [10, Theorem 3.1].

In this case , where , , .

Corollary 3.7 (boundary wH inequality). Let , , and . Suppose that is a viscosity solution of , under structure condition (1.4), with , and in . Thenwith positive constants and , depending on , and .

In a similar manner, we extend the local MP up to the boundary.

Theorem 3.8 (boundary local MP). Let , , and . Suppose that is a viscosity solution of , under structure condition (1.5), with and in . Then for all with a positive constant , depending on , and .

4. ABP-Type Estimates and Maximum Principles

Here we use boundary estimates of previous section to obtain MP in unbounded domains of for viscosity subsolutions , bounded above, of equation under structure condition (1.5).

We will make use of the measure-geometric condition , , given in the introduction. By a continuity argument, see [12], eventually passing to a smaller , which we will call , we can assume that condition is satisfied with exactly.

We also recall that is a domain (with parameter ) if each point satisfies condition in . In particular, if is the radius of the ball provided by condition , we define domains of cylindrical and conical type as domains such that and , respectively as .

4.1. Domains of Cylindrical Type

We start with the condition of Cabré [10]. Let , and be positive real numbers. We say that an open connected set of is a domain if to each we can associate a ball of radius such thatwhere is the connected component of containing .

Since , then a domain of is of cylindrical type, like domains of finite Lebesgue -dimensional measure, subdomains of , where has finite Lebesgue -dimensional measure, the complement of the spiral of equation in polar coordinates of .

Given a differential operator with structure conditions, like (1.4) and (1.5), will be called a narrow domain when, for given and , condition is satisfied for suitably close to , depending on the structure constants and the remaining geometric constants.

A straightforward application of Theorem 3.8 yields MP in narrow domains. Indeed, assume that and in . Then, by (1.5), we haveSuppose that on and set . Applying Theorem 3.8 in with , we obtainFrom this, taking the supremum over , we get , that is in , provided , and hence MP holds in this case.

In order to pass from narrow domains to arbitrary cylindrical domains we will use Theorem 3.6, from which the following ABP-type estimate follows.

Theorem 4.1 (ABP estimate). Let , let and be positive real numbers, and . Let be a cylindrical domain such that condition in is satisfied at each with .
Suppose that is a viscosity solution of , under the structure condition (1.5), with and .
If in and on , thenwhere depends on and .

Proof. It is enough to show the result for .
Set and . Let and of radius , provided by condition in . We choose such that see the beginning of this section. We also denote by the concentric ball of radius .
Now we apply Theorem 3.6 to in with instead of and close enough to in such a way that and . Since on , then , hence we getfrom which, for , we obtain the pointwise inequalitywith . On the other hand, setting and , by virtue of (4.6) we have and therefore, by our choice of and ,A further application of Theorem 3.6 to in yieldssince in this case . From this we deduce that, for with . From the definition of , see also (4.6), it follows thatwith , for all and hence also for .
Finally, passing to the supremum over , we get the result.

4.2. General Domains

Firstly, we consider domains , such that condition in holds at each without bounds for the radii of the balls provided by .

Note that in general the ABP-type estimate of Theorem 4.1 is useless unless , see [13], since the the constant of ABP estimate blows up when . This is why we assume as in the structure condition (1.5). Moreover, to take advantage from the decay of , it is convenient to use the boundary wH inequality for annular regions of Corollary 3.7 rather than Theorem 3.6.

Reasoning as in the proof of Theorem 4.1, but quite more carefully with the aid of (3.24) instead of (3.16), see [16], we get the following ABP-type estimate.

Theorem 4.2 (ABP). Let and be positive real numbers and . Let be a domain (with parameter ).
Suppose that is a viscosity solution of , under the structure condition (1.5), with such thatfor all small enough, all sufficiently close to and some .
If in and on , thenfor possibly smaller and larger , depending on and .
Here and are positive constants depending on , , , , , , , , , while also depends on .

Proof of Theorem 1.2. In the case of domains, Theorem 1.2 follows at once letting in Theorem 4.2. Suppose now that can be split by a closed set in components where MP holds and each satisfies condition in . By MP in the components, since we assume that on , then for we haveReasoning as above for (4.10), but using Corollary 3.7 instead of Theorem 3.6 as before to obtain Theorem 4.2, from condition we deduce for thatwhere is independent of . Inserting this inequality in the former one, and taking the supremum over , we get the result.

Examples
Provided that as , this last result yields MP in very general domains such as, for instance: (i) domains, like a proper cone such that and in general a domain of conical type, like the complement in of , where is a logarithmic spiral of equation in polar coordinates, or also complement of a larger spiral of equation , with a positive increasing function.(ii)Domains which can be split in subdomains by a suitable closed set of , like the cut plane in or in general the complement in of a graph such that for positive constants and .
As a further example, we show a repeated application of Theorem 1.2. Look at the complement in of a sequence of balls , , with . Consider the nonnegative -axis as , then is connected. If is the half-line of equations , , then we have the following: (i) has two components, which are domains of conical type, where MP holds;(ii)each point of satisfies condition in . Thus MP holds in by Theorem 1.2. Also, each point of satisfies condition in for some depending on . Therefore, again by Theorem 1.2, we conclude that MP holds in .

4.3. Parabolic Shaped Domains

For a parabolic shaped domain, condition at holds with as , for some , the limit cases and representing, respectively, the cylindrical and the conical cases. Hence for all with positive constants and . Then, choosing sufficiently small in Theorem 4.2, if thenso that the supremum in the second term of the right-hand side of (4.12) is taken over a bounded subset of , in which for some positive constant . Thus

Proof of Theorem 1.3. Since condition holds with , , the assumption as implies the finiteness of in (4.11). Taking account of (4.16), by continuity of the estimate (1.10) follows letting .

5. Strong Maximum Principle and Liouville Theorem

The weak Harnack inequality of Theorem 3.3 can be used to show the following strong MP.

Theorem 5.1 (strong MP). Let be a domain of . Let be such that in the viscosity sense, and assume structure condition (1.5), with . If and for all , then in .

Proof. Following [33], set and . By assumption . By continuity of , it turns out that is an open subset of . Moreover, plainly, and .
Recall that is an open connected set. Thus it is sufficient to show that is in turn an open subset to have , as claimed in the statement of the theorem. Indeed, let , that is, , and set , then is a nonnegative viscosity solution of . Applying (3.7) in a ball , we getfrom which, by continuity, in . This shows that is an open subset of and concludes the proof.

The Liouville type result of Theorem 1.5 is instead based on Harnack inequality (3.11) of Theorem 3.5. It is convenient to consider its version in annular regions to take advantage of the decay of , obtained in standard way, using inequality (3.11) in a chain of linked balls. This yields, for continuous solutions of equation , , under the structure conditions (1.4) and (1.5), with , the following inequality:with a positive constant .

Proof of Theorem 1.5. By the strong MP of Theorem 5.1, we know that can achieve neither a maximum nor a minimum at a point of unless it is constant, in which case we should be done.
Suppose for instance that . Let be an increasing sequence of positive numbers such that . Set and . By weak maximum principle, is increasing and is decreasing; thusThen, using Harnack inequality (5.2), with , we getfrom whichand, letting , we get , as we wanted to show.

Appendix

Proof of Lemmas 2.2 and 2.3

Although the proof of Lemma 2.2 is already contained in previous papers also in the case of an almost-quadratic gradient term, see for instance [17], here, for the sake of completeness, we give a sketch of the simple version in the case of linear gradient term, following [3], where the fundamental case of a second-order uniformly elliptic operator is treated, with no lower-order terms.

However, it seems useless to repeat the nice proof of [3], to which we refer for the ideas and details. We only outline the steps which are influenced by the first-order term. For this reason, we keep the same notations of [3].

Also, for the sake of brevity, we will refer to constants depending only on as to structural constants.

Firstly, we introduce a test function; see [3, Lemma 4.1].

Lemma 5. There exist positive structural constants , and a function such thatwhere , , .

Proof. We search for a function of type , for , where and are positive constants to be chosen in order that and . Next, we extend to in such a way that . By calculations, choosing we havefor . Also, for ,and therefore (A.3) holds taking a cut-off function such that in and outside .

Next, we get a lower bound for the size of level sets of supersolutions. Denoting by a cube of side , consider a nonnegative viscosity solution of the differential inequality .

Setting and observing thata positive lower boundwith a structural positive constant , see (A.2), follows, in the same way as in [3, Lemma 4.5], using ABP estimate (2.5), provided that and for a positive structural constant . Moreover, under the same assumptions, [3, Lemma 4.6] says thatfor all , which for agrees with (A.7). Then we point out that (A.8) follows by an induction process, based on the Calderón-Zygmund decomposition of the cube , centered at the origin, supposing (A.8) to hold for .

To perform the induction step it is crucial that, for a supersolution of (2.4), the rescaled function , where runs in the dyadic cube , centered at , is in turn a supersolution of (2.4) with a correspondingly scaled , namelyfor . From (A.8) it follows thatfor all , with and positive structural constants. Then, following the proof of Theorem 4.8(1) of [3], we use (A.8) in the identitysee [1], with and, by rescaling, remove the normalization conditions and to getFrom this, with a covering argument as in [10, Theorem 3.1], we obtain (2.8).

We argue in the same manner for Lemma 2.3. Suppose again that is centered at the origin. Following the proof of Theorem 4.8(2) of [3], firstly we consider a subsolution of (2.6) such that and to get, even in this case, (A.10). Then, arguing as in [3, Lemma 4.7], there exist structural constants and such that, for all large enough,where and . As above, to get this result we use the invariance of equation by scale transformations, namely that the function , where runs in the small cube , is in turn a supersolution of (2.4) with a correspondingly scaled , that is,for , provided that .

On the base of (A.13), reasoning as in the proof of Lemma 4.4 of [3], we infer that , from which, by rescaling to remove normalization conditions and , we getas in the proof of Theorem 4.8 (2) of [3]. By a covering argument, as above for supersolutions, we get (2.9) for . Note that (A.10) a fortiori holds replacing with . Thus (2.9) follows for all . Finally, by Hölder inequality, we obtain (2.9) for all .