Regularity for Double Phase Problems at
Nearly Linear Growth
Abstract
Minima of functionals of the type
with , have locally Hölder continuous gradient provided .
1 Introduction
In this paper we prove the first Schauder type results for minima of nonuniformly elliptic integrals with nearly linear growth, and under optimal bounds on the rate of nonuniform ellipticity. This is a non-trivial task, as in nonuniformly elliptic problems Schauder type estimates change their nature and cannot be achieved via perturbation methods. This was first shown by the counterexamples in sharp ; FMM . A relevant model functional we have in mind is
(1.1) |
Here, as also in the following, denotes an open subset, and ; as our results will be local in nature, without loss of generality we assume that is also bounded. The functions are bounded and non-negative with which is also bounded away from zero; a crucial point here is that is instead allowed to vanish. For the definition of local minimizer see (1.15). We are interested in the case are only Hölder continuous, when proving gradient Hölder regularity of minima in fact corresponds to establish a nonlinear version of Schauder estimates (originally also due to Hopf and Caccioppoli). The functional in (1.1) appears to be a combination of the classical nearly linear growth one
(1.2) |
treated when for instance in FM and by Marcellini & Papi in MP , with a -power growth term weighted by a non-negative coefficient . The terminology “nearly linear” accounts for an integrand whose growth in the gradient variable is superlinear, but it is still slower than that of any power type integrand as with . When the functional can be in fact considered as a limiting case () of the by now largely studied double phase integral
(1.3) |
This last one is a basic prototype of a nonautonomous functional featuring nonstandard polynomial growth conditions and a soft kind of nonuniform ellipticity (see discussion below). Originally introduced by Zhikov Z0 ; Z1 ; Zhom in the setting of Homogenization of strongly anisotropic materials, the functional in (1.3) was first studied in CM ; sharp ; FMM and can be thought as a model for composite media with different hardening exponents and . The geometry of the mixture of the two materials is in fact described by the zero set of the coefficient , where the transition from -growth to -growth takes place. In turn, the functional in (1.2) appears for instance in the theory of plasticity with logarithmic hardening, another borderline case between plasticity with power hardening () and perfect plasticity () FrS ; FS1 . Here, we are mostly interested in the theoretical issues raised by functionals as (1.1) in the context of regularity theory. In fact, provides a basic example of nonuniformly and nonautonomous elliptic functional for which the currently available techniques do not allow to prove sharp Schauder type estimates. In particular, although being a limiting case of (1.3), there is no way to adapt the methods available for (1.3) to treat , already when . We recall that general nonuniformly elliptic problems have been the object of intensive investigation over the last two decades BS0 ; BS ; BS2 ; bbgi ; bildhauer ; BF ; BF2 ; FM , starting from the more classical case of minimal surface type functionals LU ; Simon and those from the 60s ivanov0 ; ivanov1 ; IO ; serrin ; trudinger . The monograph ivanov2 gives a valuable account of the early age of the theory and we recall that nonuniformly elliptic energies have been considered several times in the setting of nonlinear elasticity BFM ; FMal ; FMa ; M0 ; ma5 . As for linear and nearly linear growth conditions in the vectorial case, we mention the recent extensive work of Kristensen & Gmeineder gme1 ; gme2 ; gmek1 ; gmek2 . But let us first recall the situation for (1.3). The bound
(1.4) |
guarantees the local Hölder continuity of the gradient of minima of (1.3) BCM ; CM ; DM . Condition (1.4) is optimal in the sense that its failure generates the existence of minimizers that do not even belong to sharp and develop singularities on fractals with almost maximal dimension FMM . In (1.4) notice the delicate interaction between ambient dimension, the -growth conditions in the gradient and the regularity of coefficient . Despite the simple form of (1.3), all this already happens in the scalar case. In the autonomous case bounds of the type in (1.4), with no appearance of , play a key role in the regularity theory of functionals with nonstandard growth as built by Marcellini in his by now classical papers M0 ; M1 ; M2 ; M3 . The key property of the functional (1.3) is that it is uniformly elliptic in the classical (pointwise) sense. This means that, with , we have
(1.5) |
where denotes the (pointwise) ellipticity ratio of , i.e.,
Nevertheless, the functional exhibits a weaker form of nonuniform ellipticity, detectable via a larger quantity called nonlocal ellipticity ratio ciccio . This is defined by
where is any ball. Indeed, note that
(1.6) |
holds whenever is nonempty, and it is therefore unbounded with respect to . We refer to ciccio for a discussion on these quantities; we just note that they do coincide in the autonomous case. It is precisely the occurrence of both (1.5) and (1.6) to imply the regularity of minima of the double phase functional in (1.3) when assuming (1.4). Indeed (1.4) and (1.6), allow to implement a delicate perturbation method based on the fact that minimizers of the frozen integral enjoy good decay estimates, which is a consequence of (1.5). A similar scheme, based on the occurrence of pointwise uniform ellipticity (1.5), is the starting point for treating larger classes of problems featuring non-polynomial ellipticity conditions, see for instance the recent interesting papers by Hästo & Ok HO1 ; HO2 . These also include integrals of the type
where has superlinear growth (these have been treated in byun0 ; byun1 ). Such schemes completely break down in the case of functionals as since the boundedness of fails for . Indeed, already in the case of the functional in (1.2), which is nonuniformly elliptic in the classical sense (i.e., (1.5) fails), no Lipschitz continuity result for minima is available but when considering differentiablity assumptions on (see dm ). In other words, the validity of Hopf-Caccioppoli-Schauder estimates is an open issue. We fix the current situation in the following:
Theorem 1.1 (Nearly linear Hopf-Caccioppoli-Schauder)
Let be a local minimizer of the functional in (1.1), with
(1.7) |
where and . Then is locally Hölder continuous in and moreover, for every ball , , the inequality
holds with and .
Of course the condition in (1.7)1 is the sharp borderline version of (1.4) (let ), that guarantees gradient Hölder continuity of minima for the classical double phase functional (1.3). Let us only observe that the equality case in (1.4), in comparison to (1.7)1 assumed here, is typically linked to superlinear growth conditions in (1.3). The equality in (1.4) is indeed obtained via Gehring type results and it is ultimately again an effect of the uniform ellipticity in the sense of (1.5) (see BCM ; dm0 ). Such effects are missing in the nearly linear growth case. Once local gradient Hölder continuity is at hand, in the non-singular/degenerate case we gain
Corollary 1 (Improved exponents)
We do not expect that, at least with the currently available techniques, the exponent in Corollary 1 can be improved manth1 ; manth2 . The loss of in the exponent is typical when dealing with functionals with subquadratic growth.
1.1 General statements
Theorem 1.1 and Corollary 1 are special cases of a class of results covering general functionals of the type
(1.9) |
with being arbitrarily close to have linear growth and . This time we are thinking of integrands of the type
(1.10) |
where is as in (1.7)2. For situations like this, we prepare a more technical theorem. We consider continuous integrands such that for every and satisfying
(1.11) |
for every choice of , , with , and being fixed constants. Here is a non-decreasing, concave and unbounded function such that is convex. Moreover, we assume that for every there exists a constant such that
(1.12) |
The above assumptions are sufficient to get Lipschitz continuity of minimizers. In order to get gradient local Hölder continuity we still need a technical one (see also Remark 1 below):
(1.13) |
For functionals as in (1.9) we have
Theorem 1.2
Let be a local minimizer of the functional in (1.9) under assumptions (1.7)1 and (1.11)-(1.12). There exists , depending only , such that, if , then
(1.14) |
holds whenever , , where and . Moreover, assuming also (1.13) implies that is locally Hölder continuous is . Finally, if in addition to all of the above we also assume that and that is continuous, then , with .
The notion of local minimizer we are using in this paper is classical and prescribes that a function is a local minimizer of if, for every ball , it happens that
(1.15) |
This implies, in particular, that is finite whenever is an open subset. In Theorem 1.2, we are using the shorthand notation
(1.16) |
that we adopt for the rest of the paper. For further notation we refer to Section 2 below. We just remark that, in order to simplify the reading, in the following we shall still denote by a constant actually depending only on a subset of the parameters listed in (1.16).
Theorem 1.2 has a technical nature, especially as far as the size of is concerned, which is not allowed to be too far from one (a natural limitation in view of the condition ). This is anyway sufficient to cover the model examples considered here, including (1.10); see Section 5. We note that in the autonomous, nearly linear case, the literature reports several interesting results for functionals , satisfying (1.11)1,2 with restrictions as , when . These bounds are in fact of the type we also need to impose when selecting (this can be checked by tracking the constant dependence in the proofs). Starting by FM conditions (1.11)2 have been used in several papers devoted to functionals with linear or nearly linear growth; see for instance bildhauer ; BF ; gmek2 and related references.
Remark 1
The technical assumption (1.13) is automatically implied, for a different constant , by ; this follows using the second inequality in (1.11)2. In turn, when is independent of we can always assume that . Indeed, we can switch from to observing that such a replacement in the integrand in (1.9) does not change the set of local minimizers. Assumption (1.13) is anyway verified by the model cases we are considering and in fact it deals with the behaviour of at the origin, while in nonuniformly elliptic equations the main problems usually come from the behaviour at infinity.
1.2 Techniques
The techniques needed for Theorem 1.2 are entirely different from those used in the case of classical superlinear double phase functionals in (1.3) BCM ; CM and for pointwise uniformly elliptic problems with nonstandard growth conditions HO1 ; HO2 . This is essentially due to the failure of uniform ellipticity (1.5) for functionals as and . On the other hand, the techniques based on De Giorgi and Moser’s methods used in the literature to deal with nearly linear growth functionals bildhauer ; BF2 ; FrS ; FM are not viable here. In fact, as in our case coefficients are Hölder continuous, it is impossible to differentiate the Euler-Lagrange equation to get gradient regularity. This is in fact the classical obstruction to direct regularity proofs one encounters when proving Schauder estimates, eventually leading to the use of perturbation methods. In turn, perturbation methods are again not working in the nonuniformly elliptic case as the available a priori estimates are not homogeneous and iteration schemes fail. Our approach is new to the context and relies on a number of ingredients, partially already introduced and exploited in piovra . Most notably, we make a delicate use of nonlinear potential theoretic methods allowing to work under the sharp bound
(1.17) |
We employ a rather complex scheme of renormalized fractional Caccioppoli inequalities on level sets (10). These are aimed at replacing the standard Bernstein technique, and encode the essential data of the problem via a wide range of parameters; see (5.37)-(5.39). Bernstein technique is traditionally based on the fact that functions as are subsolutions to linear elliptic equations for certain positive values of . This follows differentiating the original equation, something that is impossible in our setting. We therefore use a fractional and anisotropic version of the method according to which certain nonlinear functions of the gradient, namely
(1.18) |
belong to suitable fractional De Giorgi’s classes, i.e., they satisfy inequalities (10). The term renormalized accounts for the fact that inequalities (10) are homogeneous with respect to , despite the functional is not. This costs the price of getting multiplicative constants depending on . Such constants must be carefully kept under control all over the proof and reabsorbed at the very end. Notice that inequalities (10) are obtained via a dyadic/atomic decomposition technique, finding its roots in KM , that resembles the one used for Besov spaces AH ; triebel in the setting of Littlewood-Paley theory; see Section 5.6 and Remark 3. Iterating inequalities (10) via a potential theoretic version of De Giorgi’s iteration (Lemma 5) then leads to establish local Lipschitz bounds in terms of the nonlinear potentials defined in (2.11); see Proposition 2. A crucial point is that the fractional nature of the inequalities (10) allows to sharply quantify how the rate of Hölder continuity of coefficients interacts with the growth of the terms they stick to. Using this last very fact and combining the various parameters in (5.37)-(5.39) leads to determine the optimal nonlinear potentials allowing to prove Lipschitz continuity of local minimizers under the sharp bound in (1.17), but no matter how small the exponent is. Throughout the whole process, the structure of the functional needs to be carefully exploited at every point. This reflects in the peculiar choice of (1.18) as the leading quantity to iterate. Lipschitz regularity is as usual the focal point in nonuniformly elliptic problems, since once is (locally) bounded condition (1.5) gets automatically satisfied when . Indeed at this stage gradient Hölder continuity can be achieved by exploiting and extending some hidden facts from more classical regularity theory; see Section 5.10. In particular, we shall use some of the iteration schemes employed in the proof of nonlinear potential estimates for singular parabolic equations presented in KuM . The arguments developed for this point are general and can be used in other settings; see Remark 4. We finally observe that in this paper we preferred to concentrate on model cases in order to highlight the main ideas. Nevertheless we believe that the approach included here can be exported to a variety of different settings where it might yield results that are either sharp or unachievable otherwise.
2 Preliminaries
2.1 Notation
In the following we denote by a general constant larger than . Different occurrences from line to line will be still denoted by . Special occurrences will be denoted by or likewise. Relevant dependencies on parameters will be as usual emphasized by putting them in parentheses. We denote by the open ball with center and radius ; we omit denoting the center when it is not necessary, i.e., ; this especially happens when various balls in the same context share the same center. We also denote
when the ball in question is centred at the origin. Finally, with being a given ball with radius and being a positive number, we denote by the concentric ball with radius and by . Moreover, denotes the inner hypercube of , i.e., the largest hypercube, with sides parallel to the coordinate axes and concentric to , that is contained in , . The sidelength of equals . The vector valued version of function spaces like , i.e., when the maps considered take values in , , will be denoted by ; when no ambiguity will arise we shall still denote and so forth. For the rest of the paper we shall keep the notation
(2.1) |
for , and . In particular, when , we have a function defined on . With being a measurable subset with bounded positive measure , and with , , being an integrable map, we denote
Moreover, given a (scalar) function and a number , we denote
(2.2) |
Finally, whenever and is any set, we define
As usual, with we denote
2.2 Auxiliary results
We shall use the vector fields , defined by
(2.3) |
whenever (compare (ha, , (2.1)) for such extended range of ). They satisfy
(2.4) |
that holds for every , and ; for this we refer to (ha, , Lemma 2.1). We shall also use the following, trivial
(2.5) |
that holds for every and whenever . When restricting from above the range of , the inequality becomes double-sided, i.e.
(2.6) |
that instead holds for every ; see ha , (giamod, , Section 2). With being a ball and , we let
(2.7) |
and define
(2.8) |
where , . A consequence of assumptions (1.11) is the following:
Lemma 1
Proof
This is a version of (M1, , Lemma 2.1) and the proof mimics that of Marcellini. Specifically, there we take and, after using (1.11)1 as in M1 , we estimate (here denotes the standard basis of ). Here we have used that is non-decreasing and sublinear (since it is concave and non-negative) and finally (1.12) to get .
Finally, a classical iteration lemma (giu, , Lemma 6.1).
Lemma 2
Let be a non-negative and bounded function, and let be non-negative numbers. Assume that the inequality holds whenever . Then , holds too.
2.3 Fractional spaces
The finite difference operator is defined as for , for every and where . With , , , , and with being an open subset, the space is defined as the set of maps such that
When considering regular domains, as for instance the ball (this is the only case we are going to consider here in this respect), the embedding inequality
(2.9) |
holds provided , and . For this and basic results concerning Fractional Sobolev spaces we refer to guide and related references. A characterization of fractional Sobolev spaces via finite difference is contained in the following lemma, that can be retrieved for instance from dm .
Lemma 3
Let be concentric balls with , , and assume that, for , , there holds
Then
(2.10) |
holds for every , with .
2.4 Nonlinear potentials
We shall use a family of nonlinear potentials that, in their essence, goes back to the fundamental work of Havin & Mazya HM . We have recently used such potentials in the setting of nonuniformly elliptic problems in BM ; camel ; camel2 ; ciccio ; piovra and in fact we shall in particular use (piovra, , Section 4) as a main source of tools in this respect. With , being fixed parameters, and being such that , where is a ball, the nonlinear Havin-Mazya-Wolff type potential is defined by
(2.11) |
Suitable integrability properties of guarantee that is bounded, as stated in
Lemma 4
Let be numbers such that
(2.12) |
Let be concentric balls with and be a function such that , for some . Then
(2.13) |
and with .
Note that in (2.13) we are not requiring that and any value is allowed. Next lemma is a version of (piovra, , Lemma 2.3) and takes information from various results from Nonlinear Potential Theory starting by the seminal paper kilp . It is essentially a nonlinear potential theoretic version of De Giorgi’s iteration. Notice that the very crucial point in the next statement is the tracking, in the various inequalities, of the explicit dependence on the constants .
Lemma 5
Let be a ball, , and, for , consider functions , , and constants , and , . Assume that is such that for all , and for every ball , the inequality
(2.14) |
holds (recall the notation in (2.2)). If is a Lebesgue point of , then
(2.15) |
holds with .
3 Auxiliary integrands and their eigenvalues.
With reference to a general functional of the type in (1.9), considered under the assumptions of Theorem 1.2 (in particular implying that ), we define the integrand
(3.1) |
for , . We fix an arbitrary ball centred at and, with and , recalling (2.1) and (2.7), we define
(3.2) |
It follows that
(3.3) |
Note that abbreviations as , used in (3.2), will take place whenever there will be no ambiguity concerning the identity of the ball in question. A similar shortened notation will be adopted for the case of further quantities depending on a single ball . We next introduce two functions, , bound to describe the behaviour of the eigenvalues of . These are (recall that )
(3.4) |
It then follows that
(3.5) |
hold for all , , where (recall the definition of in (2.8) and (2.3)). Notice that (3.5)1,2 directly follow from (1.11)1,2, respectively, and (1.12) used with . The inequality in is a straightforward consequence of (2.4)-(2.5) and (3.5)2 (see Remark 2 below). Finally, (3.5)4 is a consequence of (1.11)3. When , by (1.12) and (3.5)2 the integrand also satisfies, for all , and
(3.6) |
with and Note that, on the contrary of the previous displays, the constant in (3.6) depends on and that for (3.6)2 we also need Lemma 1 and (3.5)2. As for , by defining
(3.7) |
we have that
(3.8) |
holds whenever , with ; for (3.8)2 we have used Lemma 1. In particular, it holds that
(3.9) |
(recall that ). As in (3.6), when the function satisfies
(3.10) |
for all , with Two quantities that will have an important role in the forthcoming computations are, for and
(3.11) |
It follows that (recall that )
(3.12) |
and this leads to define
(3.13) |
We record a few basic properties of these functions in the following:
Lemma 6
The following holds about the functions in (3.11), whenever is a ball:
-
•
For every
(3.14) -
•
For all
(3.15) -
•
There exists constant such that
(3.16) hold for all , .
-
•
There exists a constant such that
(3.17) hold for every and .
Proof
4 Lipschitz estimates with small anisotropicity
Let be a ball with and let . With , we define as the unique solution to the Dirichlet problem
(4.1) |
where, according to the notation in (3.2)2, it is . In particular, we have
(4.2) |
Existence, and uniqueness, follow by Direct Methods of the Calculus of Variations and the Euler-Lagrange equation
(4.3) |
holds for all by (3.10). Again (3.10) and standard regularity theory (giu, , Chapter 8), manth1 ; manth2 imply
(4.4) |
Proposition 1
With as in (4.1), for every integer there exists a continuous and non-decreasing function such that, for every , if (1.11) holds with , then
(4.5) |
holds too, with , where has been defined in (3.2)3. Moreover, whenever is a number such that
(4.6) |
where is a ball (not necessarily concentric to ), the Caccioppoli type inequality
(4.7) |
holds whenever , again with . Here, according to the notation in (3.11)2, it is .
Proof
The final shape of will be determined in the course of the proof via successive choices of (to be taken closer and closer to one, depending on ). All the constants in the forthcoming estimates will be independent of , so we shall omit indicating the dependence on such parameters, simply denoting , , , , and so on. Properties (4.4) allow to differentiate (4.3), i.e., replacing by for and integrate by parts. Summing the resulting equations over , we get
(4.8) |
that, again thanks to (4.4), holds for all such that . Let be a ball, be any non-negative number and be a cut-off function such that and . By (4.4) the functions are admissible in (4.8), therefore via (3.7)-(3.8) and Young’s inequality we obtain
(4.9) |
with . See (BM, , Lemmas 4.5-4.6) for more details and a similar inequality. Using (3.9) and a few elementary manipulations, we again find
and taking as in (4.6) we arrive at
again with . In turn, using (1.12) with and recalling that , we find
(4.10) |
again with . This gives (4.7) recalling that on and provided and are such that
(4.11) |
We pass to the proof of (4.5). With being the constant in (3.16), we can assume without loss of generality that
(4.12) |
holds, otherwise (4.5) follows trivially. By (4.10) and Sobolev embedding theorem we obtain
(4.13) |
where is the exponent coming from the Sobolev embedding exponent, and . We now want to apply Lemma 5 to using (4.13) to satisfy (2.14) (here all the terms involving the functions in (2.14) are not present). We fix parameters and related balls
this time all concentric to the initial ball . Take arbitrary and set , so that , choose and take in (4.13). This yields
(4.14) |
where . Note that the above choice of fits (4.6) via (4.12). By (4.14) we are able to apply Lemma 5 with , and , to get
(4.15) |
with . Being arbitrary in , (4.15) implies
and we have used the actual definition of . Recalling (4.12), we continue to estimate as
(4.16) |
for . By choosing and such that
(4.17) |
holds (note that this function is increasing in both and and that ), we can apply Young’s inequality in (4.16), thereby obtaining
where
Notice that and that this is an increasing function of its arguments, therefore we find and such that
(4.18) |
is satisfied in addition to (4.17). Lemma 2, applied with
which is always finite by (4.4)2, now gives
so that, recalling (4.2), we arrive at
(4.19) |
Setting for , we find, also thanks to (3.16), that
(4.20) |
where . Recalling (1.11)1, the definitions in (3.2) imply
so that inequality (4.19) together with (3.3) and (4.20) now gives
with , provided we further take such that
(4.21) |
Note that in the above inequality we have incorporated the dependence on in the dependence on as this last quantity influences the choice of via (4.17) and (4.18). Using the very definition (3.2)3 we then continue to estimate
and, by again using (1.12) with , we conclude with
(4.22) |
with . Since is monotone increasing, convex and such that , we deduce that its inverse is increasing, concave and , therefore it is subadditive and, for any given constant it is , see for instance (DM, , Remark 10 in Section 4). Using this last property, we can apply to both sides of (4.22) in order to conclude with (4.5). Finally, the choice of the function mentioned in the statement, that we can always take such that , comes by the choices made in (4.11), (4.17), (4.18) and (4.21) and a standard continuity argument (for this recall that both and are increasing functions of their arguments and that depends on ). Notice that by the above construction it follows that .
5 Proof of Theorem 1.2
The proof will take eleven different steps, distributed along Sections 5.1-5.11 below. We shall concentrate on the singular case , then giving remarks on how to deal with the (actually simpler) case at the very end of Section 5.9. We shall start assuming that (1.11)-(1.12) for some . We shall make further restrictions on the size of in the course of the proof until we finally come to determine the value mentioned in the statement of Theorem 1.2.
5.1 Absence of Lavrentiev phenomenon
We have the following approximation-in-energy result that actually implies the absence of Lavrentiev phenomenon:
Lemma 7
The proof follows (dm, , Section 5) almost verbatim (see also (sharp, , Lemma 13)). Notice that this is essentially the only point where the assumed convexity of is used (this implies that is convex as is also non-decreasing). For more related results on the absence of Lavrentiev phenomenon we refer to the recent papers AFM ; balci ; balci2 ; buli ; koch1 ; koch2 and related references.
5.2 Auxiliary Dirichlet problems and convergence
In the following denotes a local minimizer of , as in Theorem 1.2. By we denote two decreasing sequences of positive numbers such that , and ; we will several times extract subsequences and these will still be denoted by (for this reason we drop the pendice ). We denote by a quantity such that as . Similarly, we denote by a quantity, depending both on and , such that as for each fixed . The exact value of such quantities might change on different occurences and only the aforementioned asymptotic properties will matter. Let be a ball with ; by Lemma 7, there exists a sequence so that
(5.1) |
We define the sequence
(5.2) |
(uniformly with respect to ). Then we consider as the unique solution to the Dirichlet problem
(5.3) |
where
(5.4) |
Recall that the integrand has been defined in (3.2), with . The solvability of (5.3) follows by Direct Methods and standard convexity arguments. By (3.6), we can apply the by now classical regularity theory contained in (giu, , Chapter 8) and manth1 ; manth2 , therefore
(5.5) |
Mean value theorem and (5.2) imply
(5.6) |
Similarly, noting that by (1.11)1 and (follows from (1.7)1), we find, using also mean value theorem
(5.7) |
In turn, using in order: the minimality of , (5.1) and (5.6), we have
(5.8) |
so that, using the content of the last two displays we gain
(5.9) |
Estimate (5.8) and (3.5)1 imply that for every the sequence is uniformly bounded in , therefore, up to not relabelled subsequences, we have
(5.10) |
as . Letting in (5.9) and using standard weak lower semicontinuity theorems, yields
for every fixed . Using (5.8) we conclude with
(5.11) |
and again this holds for every . By (1.11)1 and (5.11) the sequence is uniformly bounded in . Recalling that the assumptions on imply that as , by classical results of Dunford & Pettis and de la Vallée Poussin, there exists such that weakly in and . Letting in (5.11) weak lower semicontinuity (see (giu, , Theorem 4.3)) and (5.2) yield , while the opposite inequality follows by the minimality of . We conclude with , so that, by strict convexity of the functional we find that in and we deduce that
(5.12) |
5.3 Blow-up
We fix and as in (5.3). With being a ball not necessarily concentric to , we take such that
(5.13) |
where is defined in (3.13). The above quantities are finite by (5.5). We rescale and on defining
(5.14) |
with . Note that, obviously, is still an integrand of the type in (3.1) (see also (5.15) below) and therefore the content of Section 3 applies to as well. Since solves (5.3), recalling the notation fixed in (3.2), it follows that is a local minimizer on of the functional
where, recalling the notation in (2.7)1, with it is
(5.15) |
From now on, keeping fixed the choice of made at the beginning, in order to simplify the notation we shall omit to specify dependence on such parameters, simply abbreviating
(5.16) |
for . The minimality of implies the validity of the Euler-Lagrange equation
(5.17) |
By (3.5) the integrand satisfies
(5.18) |
for any and all , where and, according to the definitions in (3.4) and the notation in (5.16), we are denoting
(5.19) |
5.4 Minimal integrands
Here we are going to play with auxiliary functionals whose integrands are of the type in (3.2)2, and therefore featuring no explicit dependence on (these are usually called “frozen” integrands). The results of Section 4 can be therefore applied. Let us fix a number , to be determined in a few lines, and a vector such that
(5.20) |
We take and fix a ball centered at with radius , denoted by . By (5.20) we have . We set
(5.21) |
According to the notation established in (2.7), (3.2) and (5.15)-(5.16), we define
(5.22) |
Note that . As in (3.7), with
from (1.12) and (3.8) it follows that
(5.23) |
hold for all , with . Consistently with (2.8), here we are denoting
(5.24) |
Recalling (3.11) and (5.19), we define, for and
(5.25) |
with related explicit expressions as in (3.12) and
By (3.15) we have
(5.26) |
and (5.13) implies
(5.27) |
Looking at (5.23), Direct Methods and strict convexity provide us with a unique minimizer defined by
(5.28) |
so that, (5.23)1 obviously implies
(5.29) |
and minimality gives
(5.30) |
Lemma 8
Proof
The integrand is of the type considered in Proposition 1 (compare with (5.22) and (5.23)), and we apply this last result to , with . It follows that for every
(5.33) | |||||
holds provided , where . Here, as in (3.2)3, it is for and in fact in (5.33) we have used (3.3). From (5.33) we can derive (5.31) using (1.12) with . Next, we apply (4.7), that gives
so that (5.32) follows using (5.31) in the above inequality and observing that .
5.5 A comparison estimate
This is in the following:
Lemma 9
Proof
From (5.21) and (5.27), and yet recalling (3.16), we deduce
(5.35) |
where in (5.35)1. We have
where . We now estimate the four integrals stemming from the second-last line in the above display. As , we have
and in the last line we have used (5.35)2. Similarly, this time using (5.35)1, we find
Next, note that and (follows from (1.7)1) implies and therefore we can estimate
where we again used (5.35)2. By means of (5.35)1 we find
Finally, observe that as so that
Merging the content of the last six displays yields (9) with the asserted dependence of the constant . Notice that the dependence of the constants on comes from (1.12) (that has been used with ). Notice also that (9) holds whenever we are assuming (1.11) with .
5.6 A fractional Caccioppoli inequality via nonlinear
atomic type decompositions
Lemma 10 (Fractional Caccoppoli inequality)
Proof
This will be obtained by a technique that works as an analogue of dyadic decomposition in Besov spaces, but using the functions in (5.28) as “atoms”; see Remark 3 below. We therefore divide the proof in two steps.
Step 1: Estimates on a single ball . Here we again use the notation and the results in Sections 5.4-5.5. In particular, here we again argue on a fixed ball . Our goal here is to prove estimate (5.43) below. We recall that basic properties of difference quotients yield
where is any number (recall that ). Using this in connection with (5.32) we have
(5.40) |
with and moreover (5.31) holds by (5.38)1; recall that is defined in (5.21). Use of (5.32) is legitimate here as (5.38)1 is assumed. Let us recall that here it is . Let us now record a couple of auxiliary estimates. The first is obtained as follows:
(5.41) |
for . For the second auxiliary inequality, we estimate
where and in the last line we have also used (5.35)1. Using (5.35)2, we gain
Using this last estimate with (9) we conclude with
(5.42) |
where , that is the second auxiliary estimate we were aiming at. Triangle inequality now yields
Using the standard property of translations
valid for every , (for this note that as ), and also the Lipschitz continuity of truncations, that is for every , we continue to estimate as follows
where . Note that we have used the elementary estimate
that follows from the very definition of in (5.25)2. By then using (5.6) and (5.6) to estimate the last two integrals in the above display, respectively, and again (5.35)2, we come to
(5.43) |
with , where , and are as in (5.39) and we have taken such that
Step 2: Patching estimates (5.43) on different balls. In this second and final step we are now going to recover estimates for on by patching up estimates (5.43) via a dyadic covering argument. This goes as follows: we take a lattice of cubes with sidelength equal to , centered at points , with sides parallel to the coordinate axes, and such that
(5.44) |
This family of cubes corresponds to a family of balls in the sense that and , as defined above. By construction, and in particular by (5.20), it is for all and , where the implied constant depends on . Moreover, by (5.44), each of the dilated balls intersects the similar ones fewer than times, that is a number depending only on (uniform finite intersection property). This implies that
(5.45) |
holds for every Borel measure defined on . We then write estimates (5.43) on balls and sum up in order to obtain
(5.46) |
Thanks to the inequality in the last display we apply Lemma 3. This yields for all with the related a priori bound
(5.47) |
that holds for every as in (5.37), where . Notice that here we have used, in order, first (2.10), as a consequence of (5.46), and then (2.9). Scaling back from to in (5.47) via (5.14), (5.19) and (5.25), squaring the resulting inequality, and restoring the original notation, we arrive at (10) and the proof is complete.
Remark 3 (Nonlinear atoms)
Here we briefly expand on the possible analogy between the classic atomic decompositions in fractional spaces (see for instance (AH, , Section 4.6) and (triebel, , Chapter 2)) and the construction made in the proof of Lemma 10. Atomic decompositions of a Besov function usually go via decompositions of the space in dyadic grids with mesh , , corresponding to the annuli (in the frequency space) considered in Littllewood-Paley theory. On each cube of the grid one considers an atom , i.e., a smooth function with certain control on its derivatives, up to the maximal degree of regularity one is interested in describing for . Specifically, one requires that
(5.48) |
hold for sufficiently large multi-indices . Summing up (over ) such atoms multiplied by suitable modulating coefficients, and then yet over all possible grids , allows to give a precise description of the smoothness of the function . Such “linear” decompositions, although very efficient, are of little use when dealing with nonlinear problems as those considered in this paper. The idea in Lemma 10 is then, given a grid of size , and therefore a certain “height” in the frequency space, to consider atoms that are in a sense close to the original solution in that they are themselves solutions to nonlinear problems (with frozen coefficients). In other words we attempt a decomposition of the type
(5.49) |
where is defined as in (5.28), with as in Step 2 from Lemma 10 ( is the indicator function of the ball ). Notice in fact the analogy with the second information in (5.48), describing the maximal smoothness of a classical atom, with the Caccioppoli inequality (5.32), from which one infers its fractional version for , that is (10).
5.7 -bounds via nonlinear potentials
Here we deliver
Proposition 2
Proof
In the following we are going to use Lemma 10 with whose size will determined in a few lines as a function of , and with by now fixed by
(5.51) |
Note that follows as a consequence of the assumed bound (1.17) and that the choice in (5.51) makes depending only on . Without loss of generality we can assume that
(5.52) |
otherwise (5.50) is obvious. We consider concentric balls and set . It follows that holds whenever . Notice that by (5.5) every point is a Lebesgue point for both and . By Lemma 10, and (10) used on , we can apply Lemma 5 on verifying (2.14) with
(5.53) |
As has been chosen arbitrarily, (5) with the choices in (5.53) implies
(5.54) |
for . Recalling the definitions in (3.12)-(3.13), and the choice of in (5.53), after a few elementary manipulations in (5.7) we arrive at
(5.55) |
with . We now take to be such that
(5.56) |
Recalling (5.51), the quantity is now determined as a function of the parameters . This determines a first restriction on the size of via (5.38) and the choice of . Notice that at this stage the number , determining an upper bound on , depends on ; notice also that all the above computation remain valid provided . Finally, by (5.56) and a standard continuity argument we can further restrict the value of finding such that
(5.57) |
Notice also that the functions are increasing. This finally fixes the value of from the statement of Theorem 1.2, with the asserted dependence on the constants. We now want to the estimate the potential terms appearing in the right-hand side of (5.55) by means of Lemma 4. In this respect, notice that allows to verify (2.12), while we can take in (2.13) because
Applying (2.13) gives
(5.58) |
with . Using (5.58) in (5.55), and recalling (5.57), we can use Young’s inequality to finally get
with and . This allows to apply Lemma 2 that provides
from which (5.50) follows using (3.17) and a few elementary manipulations.
5.8 Proof of (1.14)
Keeping in mind the notation of Proposition 2, (5.8) and (5.50) give
so that, recalling (3.13) we find
(5.59) |
It follows that, up to not relabelled subsequences, the convergence in (5.10) can be upgraded to in for every . Letting in (5.59) yields
(5.60) |
that again holds for every , with and . Similarly, as after (5.59), by (5.60) the convergence in (5.12) can be upgraded to in , so that (1.14) (for ) follows letting in (5.60), taking , recalling that and renaming into .
5.9 Local gradient Hölder continuity
Since the result is local, up passing to smaller open subsets we can assume with no loss of generality that is finite (see the comment after (1.15)). We fix an open subset and the radius . We take centred in and recall that the quantities and are defined in (5.59) and (5.60), respectively. With and being a ball concentric to , we define as
so that
(5.61) |
holds with . The derivation of (5.61) follows using first (5.59) (with replaced by and ) and combining it with (4.5) and easy manipulations as in Proposition 1 and Lemma 8. By Proposition 3 in the subsequent step and (5.61), we find that
(5.62) |
holds for every . Here, in the notation of Proposition 3 below, it is and ; these constants are non-decreasing and non-increasing functions of , respectively. Letting first and then , we have that
(see (5.60)). In particular
(5.63) |
holds with and so that only depends on and . Again letting and then , and passing to not relabelled subsequences, we can assume that . Notice that by (5.63) we have and this last quantity depends only on and but it is otherwise independent of the chosen ball . In the following, with some abuse of notation, we keep on denoting by a double-sequence of constants with the above property, typically being itself released via a non-decreasing function of ; the exact value of the numbers may vary from line to line. A similar reasoning can be done for the exponents , that is, we have . Similarly, depends only and and is independent of the ball considered. Proceeding as for (9), and also using (5.61) repeatedly, it follows that
(5.64) |
where . Using (2.4) with and then (5.61) yields
By again using (5.64) in the above display we conclude with
(5.65) |
We can now complete the proof with more standard arguments (see for instance gg2 ; manth1 ; manth2 ), that we recall for completeness. Combining (5.62) and (5.65) yields
for every , and taking finally yields
for every . In the above display we first let and then , and finally conclude with
(5.66) |
where, by the discussion made after (5.63), both and depend only on data, and , but are otherwise independent of the ball considered . Since is arbitrary, (5.66) and a standard covering argument and Campanato-Meyers integral characterization of Hölder continuity yield that for every open subset it holds that , where . The proof of Theorem 1.2 in the case is therefore complete up to (5.62), whose proof will be given in the subsequent section. We now briefly comment on how to obtain the local gradient Hölder continuity in the nonsingular case , which is simpler since it requires no approximation via the additional parameter . For the proof of (1.14) it is sufficient to use functionals as in (5.4), and all the subsequent estimates remain independent of s and ; the approximation and the convergence then only take place with respect to the parameter and leads to (1.14). The same reasoning applies to the proof in this section by taking . It remains to deal with the last issue, that is the local Hölder continuity of with explicit exponent when . This will be done in Section 5.11.
5.10 A technical decay estimate.
Here we prove (5.62); this relies on certain hidden facts from regularity theory of singular parabolic equations that we take as a starting point to treat the nonuniformly elliptic case considered here. Relevant related methods are also in dibe . Consistently with the notation established in Sections 3 and 4, with being a fixed number, in the following we denote
(5.67) |
for every , where and . Here we permanently assume (1.11)-(1.13) and therefore is an integrand of the type in (3.2)2 with replaced by . Indeed, conditions (3.8) apply here when accordingly recasted (see also (5.78) below). Moreover, let us define
(5.68) |
Then (1.13) and Lemma 1 imply that
(5.69) |
holds for every , where .
Proposition 3
The proof of (5.71) is achieved via two preliminary lemmas. In the rest of section, we denote .
Lemma 11
With being a ball, assume that satisfies (5.70)1 and
(5.72) |
There exists , which is independent of and is a non-increasing function of , such that if either
(5.73) |
holds for some , then
(5.74) |
In this case
(5.75) |
where and both depend on , but are otherwise independent of . The constants and are non-decreasing and non-increasing functions of , respectively. On the other hand there exists such that, if fails for every , then
(5.76) |
The constant is a non-decreasing function of .
Proof
For the proof of the first assertion we use and extend some of the arguments in (KuM, , Proposition 3.7); see also dibe for the case of the classical -Laplacian operator. We can of course confine ourselves to the case the first inequality in (5.73) occurs (for a to be determined in the course of the proof), the other being similar. By scaling, as in (5.14), we can assume that . Since conditions (3.10) are satisfied by , it is standard to prove that . We can then differentiate the equation in the direction, thereby obtaining , that is an analogue of (4.8). Recalling that (as in (4.4) and by (5.72)), we then use as test function , where is non-negative and ; this is still possible by (3.10)). Integrating by parts the resulting equation we find
(5.77) |
We define the functions
(5.78) |
for , so that
(5.79) |
holds for . By also using (3.8), (5.69) and (5.72), and the monotonicity features of and in (5.77), we easily find
By then using Young’s inequality and (5.79), and yet recalling that , we find
(5.80) |
For integers we choose levels , radii and cut-off functions , with on , . With
(5.81) |
estimate (5.80) becomes
We then find, via Sobolev embedding
Observing that and that (5.72) implies , the above display gives
(5.82) | |||||
with which is independent of . Inequality (5.82) allows to perform a standard geometric iteration (see for instance (giu, , Lemma 7.1)) leading to once the first inequality in (5.73) is verified with small enough (that corresponds to require that is small, see (5.81)). This allows to prove (5.75) by considering
(5.83) |
this time for every . Recalling (3.8)-(3.9) it follows that satisfies
(5.84) |
a.e. in and for every , where . Therefore a standard application of De Giorgi-Nash-Moser theory to and (5.72) imply the validity of (5.75) (note that for this we obtain (5.75) for , and then the case follows trivially by (5.72)). We now turn to (5.76); the proof is a variant of the one in (KuM, , Proposition 3.11) apart from the parabolic case considered there. The only remark we need to do is that the analogue of equation in (KuM, , (3.53)) is here given by , where if belongs to the support of , and (the identity matrix) otherwise. Then turns out to be a weak subsolution. Moreover, as on the support of , it follows that also satisfies (5.84). We can now use the arguments in (KuM, , pp. 784-786) to conclude with (5.76). Finally, as for the monotone dependence on of the constants , this is classical, and it is a consequence of the fact that all the estimates above feature constants, usually denoted by , that are non-decreasing functions of and ; in turn, is a non-decreasing function too.
Lemma 12
Proof
Proof (of Proposition 3)
We take from Lemma 11 and determine an integer such that ; notice that can be determined as a non-decreasing function of . All in all, once is fixed so is and therefore . We set , , for every , and define as the smallest non-negative integer for which . This implies that
(5.86) |
Note that (5.86) is empty when . We further define
and as if , and if . Using (5.76) and (5.86) iteratively, we find
(5.87) |
On the other hand, by the very definition of and Lemmas 11-12 (the latter is only needed when and therefore when ), we find that
(5.88) |
At this stage (5.71) follows via a standard interpolation argument combining (5.87)-(5.88).
Remark 4
As a consequence of the local gradient boundedness obtained in Section 5.8, the arguments developed for Proposition 3 do not require any upper bound on . Moreover, the specific structure in (5.67) is not indispensable and the methods used here can be used as a starting point to treat general nonautonomous functionals with -growth of the type considered in M2 .
5.11 Improved Hölder exponent when
The -regularity proof follows almost verbatim the one in (piovra, , Section 10) since now we already know that is locally Hölder continuous in with exponent . We confine ourselves to give a few remarks on the main modifications and to facilitate the adaptation we use the same notation introduced in piovra . We set (we can assume this is finite since our result is local). This time we take and recall that holds whenever . With to solve in , as in (5.61) we gain with . With these informations estimate follows as in (5.65) (in turn as in (9)), with . This is the analogue of (piovra, , (10.3)), but for the fact that here we find rather than and the integral is supported in instead of . Next, by defining the matrix , to use piovra we need to prove that there exists , independent of , such that holds for a.e. . As a consequence of (1.11)2 and of in , we have
for a.e. and , where . This is the crucial point where we use After this, the proof follows exactly as (piovra, , (10.6)) and the proof of the entire Theorem 1.2 is finally complete.
6 Theorem 1.1, Corollary 1 and model examples (1.10)
For every integer , we consider the integrand
with as in (1.7)2 and as in (1.10)2. Direct computations show that satisfies (1.11) with , with when , and with any choice of when ; moreover, it also satisfies (1.13). The constants and depend on and . Theorem 1.2 therefore applies to local minimizers of
On the other hand the functional in the above display and
share the same local minimizers, and therefore the regularity results stated in Theorem 1.2 hold for local minimizers of the last functional too. In this way the model case in (1.10) is covered. Theorem 1.1 and Corollary 1 then follow as special cases taking and and , respectively. Note that, in the spirit of Corollary 1, local minimizers of the functional
are locally -regular in , for every , provided (1.7)2 is assumed.
7 Vectorial cases and further generalizations
When functionals as in (1.1) and (1.9) are considered in the vectorial case, i.e. minima are vector valued and , we can still obtain Lipschitz continuity results. In this situation it is unavoidable to impose a so-called Uhlenbeck structure uh , that is
(7.1) |
where is a continuous function such that is -regular for every choice of . This is obviously satisfied by the models in (1.1) and (1.10).
Theorem 7.1
The proof of Theorem 7.1 is essentially the same of the one given for Theorem 1.2, once the content of Proposition 1 is available. This is in fact the only point where (7.1) enters the game. Inspecting the proof of Proposition 1 shows that this works in the vectorial case provided (4.9) holds. In turn, this inequality follows along the lines of the estimates in (BM, , Lemma 5.6). Notice that, without an additional structure assumption as in (7.1), Theorem 7.1 cannot hold and counterexamples to Lipschitz regularity emerge already when considering uniformly elliptic systems SY . We also remark that we expect gradient Hölder continuity in the vectorial case as well; the proof must be different from the one given in Section 5.10, and based on linearization methods as those originally introduced in uh and developed in large parts of the subsequent literature. Another direct generalization, this time in the scalar case, occurs for functionals of the type
(7.2) |
where is as Theorem 1.2 and belongs to and satisfies
for every , . Theorem 1.2 continuous to hold in this last case, with essentially the same proof. The only difference worth pointing out is the new shape of the integrands in (3.1) and (and the related minimal one ) in (3.2), accordingly defined as
Here , denotes a family of radially symmetric mollifiers, defined as , where , , . The definition of , instead, remains the same. Following for instance (dm, , Section 4.5) and (piovra, , Section 6.7) it is possible to show that the newly defined integrands still have the properties described in Section 3. In particular (3.5) and (3.6) still hold for a suitable constant with the same dependence upon the various parameters described there. Finally, when again considering the vectorial case for functionals as in (7.2) we can still obtain local Lipschitz regularity of minima provided (7.1) holds together with and is -regular outside the origin (see for instance BM ; ciccio for precise assumptions).
Acknowledgments. The first author is supported by INdAM-GNAMPA via the project “Problemi non locali: teoria cinetica e non uniforme ellitticità", and by the University of Parma via the project “Local vs nonlocal: mixed type operators and nonuniform ellipticity". Both the authors are grateful to the referees for the careful reading of the original version of the manuscript and for the many suggestions and comments that eventually led to a better presentation.
Funding Open access funding provided by Università degli Studi di Parma within the CRUI-CARE Agreement.
Data Availibility No data is attached to this paper.
Declarations
Conflict of interests and data. The authors declare to have no conflict of interests. No data are attached to this paper.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit creativecommons.org/ licenses/by/4.0/.
References
- (1) E. Acerbi, G. Bouchitté, I. Fonseca, Relaxation of convex functionals: the gap problem, Ann. Inst. Henri Poincaré, Anal. Non Linéaire 20, 359–390 (2003).
- (2) D.R. Adams, L.I. Hedberg, Function spaces and potential theory, Grundlehren der mathematischen Wissenschaften 314. Springer-Verlag, Berlin, 1996.
- (3) A. Balci, L. Diening, M. Surnachev, New examples on Lavrentiev gap using fractals, Calc. Var. & PDE 59:180 (2020).
- (4) A. Balci, C. Ortner, J. Storn, Crouzeix-Raviart finite element method for non-autonomous variational problems with Lavrentiev gap, Numer. Math. 151, 779–805 (2022).
- (5) P. Baroni, M. Colombo, G. Mingione, Regularity for general functionals with double phase, Calc. Var. & PDE 57:62 (2018).
- (6) S. Baasandorj, S.S. Byun, Regularity for Orlicz phase problems, Memoirs Amer. Math. Soc., to appear.
- (7) L. Beck, M. Bulíček, F. Gmeineder, On a Neumann problem for variational functionals of linear growth, Ann. Sc. Norm. Super. Pisa Cl. Sci. (V) 21, 695–737 (2020).
- (8) L. Beck, G. Mingione, Lipschitz bounds and nonuniform ellipticity, Comm. Pure Appl. Math. 73, 944-1034 (2020).
- (9) P. Bella, M. Schäffner, Local boundedness and Harnack inequality for solutions of linear nonuniformly elliptic equations, Comm. Pure Appl. Math. 74, 453-477 (2021).
- (10) P. Bella, M. Schäffner, On the regularity of minimizers for scalar integral functionals with -growth, Anal. & PDE 13, 2241–2257 (2020).
- (11) P. Bella, M. Schäffner, Lipschitz bounds for integral functionals with -growth conditions, Adv. Calc. Var. 10.1515/acv-2022-0016
- (12) M. Bildhauer, Convex variational problems. Linear, nearly linear and anisotropic growth conditions. Lecture Notes in Mathematics, 1818. Springer-Verlag, Berlin, 2003. x+217.
- (13) M. Bildhauer, M. Fuchs, Partial regularity for variational integrals with -growth, Calc. Var. & PDE 13, 537–560 (2001).
- (14) M. Bildhauer, M. Fuchs, -solutions to non-autonomous anisotropic variational problems, Calc. Var. & PDE 24, 309–340 (2005).
- (15) G. Bouchitté, I. Fonseca, J. Malý, The effective bulk energy of the relaxed energy of multiple integrals below the growth exponent, Proc. R. Soc. Edinb., Sect. A, Math. 128, 463-479 (1998).
- (16) M. Bulíček, P. Gwiazda, J. Skrzeczkowski, On a range of exponents for absence of Lavrentiev phenomenon for double phase functionals, Arch. Ration. Mech. Anal. 246, 209–240 (2022).
- (17) S.S. Byun, J. Oh, Regularity results for generalized double phase functionals, Anal. PDE 13, 1269–1300 (2020).
- (18) M. Colombo, G. Mingione, Regularity for double phase variational problems, Arch. Ration. Mech. Anal. 215, 443–496 (2015).
- (19) C. De Filippis, Quasiconvexity and partial regularity via nonlinear potentials, J. Math. Pures Appl. 163, 11–82 (2022).
- (20) C. De Filippis, G. Mingione, A borderline case of Calderón-Zygmund estimates for nonuniformly elliptic problems, St. Petersburg Math. J. 31, 455–477 (2020).
- (21) C. De Filippis, G. Mingione, Manifold constrained nonuniformly elliptic problems, J. Geom. Anal. 30:1661-1723 (2020).
- (22) C. De Filippis, G. Mingione, On the regularity of minima of non-autonomous functionals, J. Geom. Anal. 30:1584-1626 (2020).
- (23) C. De Filippis, G. Mingione, Lipschitz bounds and nonautonomous integrals, Arch. Ration. Mech. Anal. 242, 973-1057 (2021).
- (24) C. De Filippis, G. Mingione, Nonuniformly elliptic Schauder theory, arXiv:2201.07369v1.
- (25) C. De Filippis, B. Stroffolini, Singular multiple integrals and nonlinear potentials, J. Funct. Anal. 285, paper 109952 (2023).
- (26) E. DiBenedetto, -local regularity of weak solutions of degenerate elliptic equations, Nonlinear Anal. 7, 827-850 (1983).
- (27) E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math. 136, 521-573 (2012).
- (28) L. Esposito, F. Leonetti, G. Mingione, Sharp regularity for functionals with growth, J. Diff. Equ. 204, 5-55 (2004).
- (29) I. Fonseca, J. Malý, Relaxation of multiple integrals below the growth exponent, Ann. Inst. H. Poincaré Anal. Non Linéaire 14, 309-338 (1997)
- (30) I. Fonseca, J. Malý, G. Mingione, Scalar minimizers with fractal singular sets, Arch. Ration. Mech. Anal. 172, 295-307 (2004).
- (31) I. Fonseca, P. Marcellini, Relaxation of multiple integral in subcritical Sobolev spaces, J. Geom. Anal. 7, 57-81 (1997).
- (32) J. Frehse, G. Seregin, Regularity of solutions to variational problems of the deformation theory of plasticity with logarithmic hardening, Proc. St. Petersburg Math. Soc. V, 127–152, Amer. Math. Soc. Transl. Ser. 2, 193, Amer. Math. Soc., Providence, RI, 1999.
- (33) M. Fuchs, G. Mingione, Full -regularity for free and constrained local minimizers of elliptic variational integrals with nearly linear growth, manuscripta math. 102 (2000), 227–250.
- (34) M. Fuchs, G. Seregin, Variational methods for problems from plasticity theory and for generalized Newtonian fluids, Lecture Notes in Mathematics, 1749. Springer-Verlag, Berlin, 2000. vi+269 pp.
- (35) M. Giaquinta, E. Giusti, Differentiability of minima of nondifferentiable functionals, Invent. Math. 72, 285-298 (1983).
- (36) M. Giaquinta, G. Modica, Remarks on the regularity of the minimizers of certain degenerate functionals, manuscripta math. 57, 55-99 (1986).
- (37) E. Giusti, Direct Methods in the calculus of variations, World Scientific Publishing Co., Inc., River Edge (2003).
- (38) F. Gmeineder, Partial regularity for symmetric quasiconvex functionals on BD, J. Math. Pures Appl. (IX) 145 83–129 (2021).
- (39) F. Gmeineder, The regularity of minima for the Dirichlet problem on BD, Arch. Ration. Mech. Anal. 237, 1099–1171 (2020).
- (40) F. Gmeineder, J. Kristensen, Partial regularity for BV minimizers, Arch. Ration. Mech. Anal. 232, 1429–1473 (2019).
- (41) F. Gmeineder, J. Kristensen, Sobolev regularity for convex functionals on BD, Calc. Var. & PDE 58:56 (2019).
- (42) C. Hamburger, Regularity of differential forms minimizing degenerate elliptic functionals, J. reine angew. Math. (Crelle J.) 431, 7-64 (1992).
- (43) P. Hästö, J. Ok, Maximal regularity for local minimizers of non-autonomous functionals, J. Eur. Math. Soc. 24, 1285–1334 (2022).
- (44) P. Hästö, J. Ok, Regularity theory for non-autonomous partial differential equations without Uhlenbeck structure, Arch. Ration. Mech. Anal. 245, 1401–1436 (2022).
- (45) A.V. Ivanov, The Dirichlet problem for second order quasilinear nonuniformly elliptic equations, Trudy Mat. Inst. Steklov. 116 (1971) 34–54.
- (46) A.V. Ivanov, Local estimates of the maximum modulus of the first derivatives of the solutions of quasilinear nonuniformly elliptic and nonuniformly parabolic equations and of systems of general form, Proc. Steklov Inst. Math. 110, 48–71 (1970).
- (47) A.V. Ivanov, Quasilinear degenerate and nonuniformly elliptic and parabolic equations of second order, Proc. Steklov Inst. Math. 1984, no. 1 (160), xi+287 pp.
- (48) N.M. Ivočkina, A.P. Oskolkov, Nonlocal estimates of the first derivatives of solutions of the Dirichlet problem for nonuniformly elliptic quasilinear equations, Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 5, 37–109 (1967).
- (49) T. Kilpeläinen, J. Malý, The Wiener test and potential estimates for quasilinear elliptic equations, Acta Math. 172, 137–161 (1994).
- (50) L. Koch, Global higher integrability for minimisers of convex functionals with -growth, Calc. Var. & PDE 60:63 (2021).
- (51) L. Koch, On global absence of Lavrentiev gap for functionals with -growth, arXiv:2210.15454
- (52) J. Kristensen, G. Mingione, The singular set of -minima, Arch. Ration. Mech. Anal. 177, 93–114 (2005).
- (53) T. Kuusi, G. Mingione, Gradient regularity for nonlinear parabolic equations, Ann. Scu. Norm. Sup. Pisa Cl. Sci. (V) 12, 755–822 (2013).
- (54) O.A. Ladyzhenskaya, N.N. Ural’tseva, Local estimates for gradients of solutions of nonuniformly elliptic and parabolic equations, Comm. Pure Appl. Math. 23, 677–703 (1970).
- (55) J. J. Manfredi, Regularity of the gradient for a class of nonlinear possibly degenerate elliptic equations, Ph.D. Thesis, University of Washington, St. Louis.
- (56) J. J. Manfredi, Regularity for minima of functionals with -growth, J. Diff. Equ. 76, 203–212 (1988).
- (57) P. Marcellini, On the definition and the lower semicontinuity of certain quasiconvex integrals, Ann. Inst. H. Poincaré Anal. Non Linéaire 3, 391–409 (1986).
- (58) P. Marcellini, The stored-energy for some discontinuous deformations in nonlinear elasticity, Partial Differential Equations and the Calculus of Variations vol. II, Birkhäuser Boston Inc. (1989).
- (59) P. Marcellini, Regularity of minimizers of integrals of the calculus of variations with non standard growth conditions, Arch. Rat. Mech. Anal. 105, 267–284 (1989).
- (60) P. Marcellini, Regularity and existence of solutions of elliptic equations with -growth conditions, J. Diff. Equ. 90, 1–30 (1991).
- (61) P. Marcellini, Regularity for elliptic equations with general growth conditions, J. Diff. Equ. 105, 296–333 (1993).
- (62) P. Marcellini, G. Papi, Nonlinear elliptic systems with general growth, J. Diff. Equ. 221 412–443 (2006).
- (63) V. Maz’ya, M. Havin, A nonlinear potential theory, Russ. Math. Surveys, 27, 71–148 (1972).
- (64) J. Serrin, The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables, Philos. Trans. R. Soc. Lond., Ser. A 264, 413-496 (1969).
- (65) L. Simon, Interior gradient bounds for nonuniformly elliptic equations, Indiana Univ. Math. J. 25, 821–855(1976).
- (66) V. Sverák, X. Yan, Non-Lipschitz minimizers of smooth uniformly convex functionals, Proc. Natl. Acad. Sci. USA 99, 15269–15276 (2002).
- (67) H. Triebel, Theory of function spaces. III. Monographs in Mathematics, 100. Birkhäuser Verlag, Basel, 2006.
- (68) N. Trudinger, The Dirichlet problem for nonuniformly elliptic equations, Bull. Am. Math. Soc. 73, 410–413 (1967).
- (69) K. Uhlenbeck, K, Regularity for a class of non-linear elliptic systems, Acta Math. 138, 219–240 (1977).
- (70) V. V. Zhikov, On Lavrentiev’s Phenomenon, Russian J. Math. Phys. 3, 249–269 (1995).
- (71) V. V. Zhikov, On some variational problems, Russian J. Math. Phys. 5, 105–116 (1997).
- (72) V. V. Zhikov, S.M. Kozlov, O. A. Oleĭnik, Homogenization of differential operators and integral functionals. Springer-Verlag, Berlin, 1994. xii+570 pp. ISBN: 3-540-54809-2
Dipartimento SMFI
Università di Parma
Parco Area delle Scienze 53/a, I-43124, Parma, Italy
Dipartimento SMFI
Università di Parma
Parco Area delle Scienze 53/a, I-43124, Parma, Italy
email:[email protected]