Skip to content
BY 4.0 license Open Access Published by De Gruyter March 16, 2018

Regularity for scalar integrals without structure conditions

  • Michela Eleuteri ORCID logo , Paolo Marcellini ORCID logo EMAIL logo and Elvira Mascolo ORCID logo

Abstract

Integrals of the Calculus of Variations with p,q-growth may have not smooth minimizers, not even bounded, for general p,q exponents. In this paper we consider the scalar case, which contrary to the vector-valued one allows us not to impose structure conditions on the integrand f(x,ξ) with dependence on the modulus of the gradient, i.e. f(x,ξ)=g(x,|ξ|). Without imposing structure conditions, we prove that if qp is sufficiently close to 1, then every minimizer is locally Lipschitz-continuous.

1 Introduction

The fundamental classical problem of the Calculus of Variations in the scalar case usually is formulated as finding a function u assuming a given value u0 at the boundary Ω of an open bounded set Ωn which minimizes the integral

(1.1)Ωf(x,Dv)𝑑x

among all functions v:Ω, assuming the same boundary value u0 as u. The precise functional space where to look for solutions depends on the growth conditions of f=f(x,ξ) as ξn grows in modulus to +. Usually this growth is stated in terms of an inequality of the type

(1.2)f(x,ξ)M1|ξ|p

for a.e. xΩ, ξn and for some positive constant M1. Here p=1 is associated to the BV(Ω) space of functions with bounded variation, while p>1 is related to the Sobolev space W1,p(Ω). Usually the condition p>1 and the strict convexity of f(x,ξ) with respect to ξ are sufficient conditions for the existence and uniqueness of minimizers.

A different problem is the regularity of minimizers. A large literature is known about regularity (see for instance [26, 28, 30]) partly based on the nowadays classical well-known Hölder continuity result by De Giorgi [16]. To this aim it seems necessary to impose also a growth condition from above, to be associated to the growth condition from below in (1.2), of the type

(1.3)f(x,ξ)M2(1+|ξ|q)

for a.e. xΩ, ξn, for qp and for some positive constant M2. The so-called “natural growth conditions” appear if q=p, while the more general assumption q>p allows us to consider a much larger class of integrals of the Calculus of Variations, such as for example

(1.4)f(ξ)=|ξ|plog(1+|ξ|)

or

(1.5)f(x,ξ)=|ξ|p(x)orf(x,ξ)=(1+|ξ|2)p(x)2.

We recall also the integrands recently considered in [12, 11, 20, 19], see also [2, 3, 4],

(1.6)f(x,ξ)=a(x)|ξ|p+b(x)|ξ|q,

where a(x),b(x)0 and possibly zero on some part of Ω, being at least one of the two coefficients positive at almost every xΩ. The above examples (1.4), (1.5), (1.6) enter in the theory presented in this paper. However, here we study the more general case with f=f(x,ξ) without a structure, i.e. not necessarily depending on the modulus of ξ of the type f(x,ξ)=g(x,|ξ|).

We assume that f:Ω×n[0,+) is a convex function with respect to the gradient variable and it is strictly convex only at infinity. More precisely, there exists M0>0 such that fξξ, fξx are Carathéodory functions satisfying

(1.7){M1|ξ|p-2|λ|2i,jfξiξj(x,ξ)λiλj,|fξξ(x,ξ)|M2|ξ|q-2,|fξx(x,ξ)|h(x)|ξ|p+q-22

for a.e. xΩ and for all λ,ξn, with |ξ|M0 and for positive constants M1,M2. Here 1<pq and hLr(Ω) for some r>n.

Model integrands satisfying condition (1.7) are, for instance, the function f(x,ξ) in (1.6) and also

(1.8)f(x,ξ)=|ξ|p+c(x)|ξ|s+|ξn|q,

ξn being the last component (or any other component) of the vector ξ=(ξ1,ξ2,,ξn), when

sp+q2.

For instance, when s=p and qp, we are considering energy integrals with integrand of the type (we denote here a(x)=1+c(x) a generic positive coefficient)

(1.9)f(x,ξ)=a(x)|ξ|p+|ξn|q.

Note that the cases (1.8) and (1.9) can be handled with Theorem 1.1. On the other hand example (1.10) below enters in Theorem 1.2:

(1.10)f(x,ξ)=|ξ|p+b(x)|ξ|q.

The main regularity result that we prove here is the following a-priori estimate.

Theorem 1.1 (A-priori estimate).

Let uW1,p(Ω) be a smooth local minimizer of the integral functional (1.1) with exponents p,q fulfilling

(1.11)qp<1+2(1n-1r).

Under the growth assumption (1.7), there exist positive constants C,β,γ depending on n,r,p,q, M0,M1,M2 such that, for every 0<ρ<Rρ+1,

(1.12)DuL(Bρ;n)C(1+hLr(Ω)R-ρ)βγ(BR{1+|Du|p}𝑑x)γp.

Note that to get regularity of solutions it is natural, and also necessary, to assume that the gap q-p is small or that qp is close to 1, because of the known counterexamples [27, 31, 33].

The L-bound of the gradient is obtained through several steps. The first step of the a-priori estimate is Lemma 2.3 below, where on the right-hand side of the a-priori estimate there is the norm of the minimizer u in W1,qm(Ω) (m=rr-2), and the exponents p,q are related by the condition

(1.13)qp<1+2nn-2(1n-1r)

with n3. Note that if r=+ in (1.11) and (1.13), we recover the bounds in [33, Theorems 2.1 and 3.1]. An interpolation method allows us to obtain (1.12).

The mathematical literature on the regularity under p,q-growth is now very large; we refer to [35, 34, 33, 32] and to [36] for a complete survey on the subject. A new impulse to the subject has been given by the recent articles already cited [12, 11, 15, 20] for the case of elliptic equations and by [6, 7, 8] for the case of parabolic equations and systems under p,q-growth. We observe that here the ellipticity and growth assumptions hold only for large values of the gradient variable, i.e. we consider functionals which are uniformly convex only at infinity. In this context see [10, 25, 14] and recently [20, 19, 13]. The Sobolev dependence on x has recently been considered in [37, 1] and for obstacle problems in [21].

The previous a-priori estimate, more precisely Theorem 2.6, under assumptions (1.7) where the last condition is replaced by

(1.14)|fξx(x,ξ)|h(x)|ξ|q-1

for a.e. xΩ, with |ξ|M0, allows us to obtain the following existence and regularity result.

Theorem 1.2 (Existence and regularity).

Assume that f satisfies (1.7) and (1.14) with 1<pq and

qp<1+1n-1r.

The Dirichlet problem min{F(u):uW01,p(Ω)+u0}, with F defined in (1.15) below and u0W1,q(Ω), has at least one locally Lipschitz continuous solution.

Here we emphasize the definition (i.e. the precise meaning) of the integral F(u) to be minimized; in fact, the integral in (1.1) is well defined if uWloc1,q(Ω), due to the growth assumption in (1.3), but a-priori it is not uniquely defined if uW1,p(Ω)Wloc1,q(Ω). In this context of x-dependence, we cannot a-priori exclude the Lavrentiev phenomenon; however, note that in Section 5 we assume a special form of f to rule out this possibility.

For the gap in the Lavrentiev phenomenon we refer to [39, 9] and recently [24, 22, 23] for related results.

For the functional F we adopt the classical definition which refers to the pioneering research by Serrin [38] (see also [29]), which is related to the Γ-convergence theory by De Giorgi [17]. Precisely, for all uW1,p(Ω),

(1.15)F(u)=inf{lim infkΩf(x,Duk)𝑑x:ukW01,q(Ω)+u0,uk𝑤u in W1,p(Ω)}.

We discuss more in details in Section 3 the definition of F in (1.15), while in Section 2 we give the proof of the a-priori estimate. Finally, in Section 4 we give the proof of Theorem 1.2.

2 A-priori estimates

Let us start with two technical lemmas.

Lemma 2.1.

The inequality

(2.1)(1+t)βcβ(1+0t(1+s)β-2s𝑑s)

holds for every t[0,+) and every β(0,+), where

(2.2)cβ=β1-(1+β(β-1))11-β

if β1, while (by continuity)

(2.3)c1=limβ1cβ=ee-1.

Proof.

In order to prove inequality (2.1) we first consider the case β=1.

Step 1 (β=1). We compute the integral on the right-hand side of (2.1):

0t(1+s)-1s𝑑s=0t(1-(1+s)-1)𝑑s=t-log(1+t)

and inequality (2.1) becomes

1+tc1(1+t-log(1+t)),

which is equivalent to

log(1+t)1+tc1-1c1.

A computation shows that g(t)=:log(1+t)1+t is positive for t(0,+) and has a maximum at t=e-1, thus

g(t)=:log(1+t)1+tg(e-1)=1e;

with the position c1-1c1=:1e we find (2.3).

Step 2 (β1). We compute the integral on the right-hand side of (2.1) under the condition β1 and with the notation r=:1+s:

0t(1+s)β-2s𝑑s=1t+1rβ-2(r-1)𝑑r
=1t+1rβ-1𝑑r-1t+1rβ-2𝑑r
=[rββ]r=1r=t+1-[rβ-1β-1]r=1r=t+1
=(t+1)ββ-(t+1)β-1β-1+1β(β-1).

Inequality (2.1) takes then the form

1cβ(1+t)β1+(t+1)ββ-(t+1)β-1β-1+1β(β-1).

We can write it equivalently

(2.4)g(t)1+1β(β-1),

where

g(t)=:(t+1)β-1β-1-(1β-1cβ)(t+1)β.

We can compute the maximum of g(t) when t[0,+). We find that the derivative g(t) is equal to zero if t=βcβ-β and, since cβ>β, we obtain

max{g(t):t[0,+)}=g(βcβ-β)=(cβcβ-β)β-11β(β-1).

Therefore inequality (2.4) holds if we choose cβ to satisfy the condition

(cβcβ-β)β-11β(β-1)=1+1β(β-1).

A further computation gives for cβ the explicit expression in (2.2). Note that cβc1 as β1. ∎

In the sequel we apply the previous lemma to get the a-priori estimates in particular to deal with the left-hand side of (2.26), with β=γ2+p2, for γ0; thus βp2. In the next result in fact we consider β[β0,+) for some fixed β0>0.

Lemma 2.2.

Let β0>0. There exist constants c and c′′, depending on β0 but independent of ββ0 and of t0, such that

(2.5)(1+t)βcβ2log(1+β)(1+0t(1+s)β-2s𝑑s),
(2.6)(1+t)βc′′β2(1+0t(1+s)β-2s𝑑s)

for every β[β0,+) and every t[0,+).

Proof.

First we show that the constant cβ is bounded independently of β1 if β[β0,1] (here we assume that β0(0,1), otherwise nothing to be proved at this step). Precisely, we show that

(2.7)cββ1-e-β0for all β[β0,1].

In fact, by the inequality logtt-1, valid for all t>0, by posing t=1+β(β-1) if β<1, we obtain

(1+β(β-1))11-β=elog(1+β(β-1))1-βeβ(β-1)1-β=e-β

and (2.7) follows if β[β0,1), since

cβ=β1-(1+β(β-1))11-ββ1-e-ββ1-e-β0.

Finally, if β=1, then c1=ee-1<11-e-β0 holds, since it is equivalent to 1<e1-β0.

We now consider the case β>1. By Taylor’s formula we get

(1+β(β-1))11-β=elog(1+β(β-1))1-β=1+log(1+β(β-1))1-β+o(log(1+β(β-1))1-β)

and thus the quantity

cβlog(1+β)β2=log(1+β)β[log(1+β(β-1))β-1+o(log(1+β(β-1))1-β)]

has a finite limit as β+ (equal to 12) and it is a bounded function for β[1,+), let us say bounded by c. This proves (2.5). The other inequality (2.6) is a direct consequence of (2.5). ∎

Let now Ω be an open bounded subset of n for n2 and assume that f satisfies (1.7). We observe that we can transform f(x,ξ) into f(x,M0ξ), which satisfies the same assumptions for |ξ|1 (with different constants depending on M0). Then it is sufficient to obtain the a-priori bound and the regularity results for v=1M0u. Therefore, for clarity of exposition and without loss of generality, we assume M0=1. Throughout the paper we will denote by Bρ and BR balls of radii, respectively, ρ and R (with ρ<R) compactly contained in Ω and with the same center, let us say, x0Ω.

In this section we assume the following supplementary assumptions on f. Assume that f𝒞2(Ω×n) and there exist two positive constants k and K such that for all ξn and all xΩ,

(2.8){k(1+|ξ|2)q-22|λ|2i,jfξiξj(x,ξ)λiλj,|fξξ(x,ξ)|K(1+|ξ|2)q-22,|fξx(x,ξ)|K(1+|ξ|2)q-12.

In the next lemma, we obtain an a-priori estimate for the L-norm of the gradient of u which is independent of k and K.

Lemma 2.3.

Let u be a local minimizer of the integral functional (1.1) with f satisfying (1.7) and (2.8) with

(2.9)qp<1+2αn-2with α=1-nr

if n3 and p<q if n=2. Then there exists a positive constants C depending only on n,r,p,q,M1,M2 (depending also on |Ω| when n=2) such that

(2.10)DuL(Bρ;n)C[1+hLr(Ω)(R-ρ)]θβ~(BR{1+|Du|qm}𝑑x)θqm

for every 0<ρ<Rρ+1, where

(2.11)β~:=2*p2*2-qm,θ:=qm(2*2m-1)p2*2-qm,m:=rr-2.

Remark 2.4.

We observe that

(2.12)1m:=rr-2<nn-2=2*2,since r>n;

the last inequality holds for n>2, while we set 2* equal to any fixed real number greater than 2 if n=2. Moreover, we also have

(2.13)12m-12*=r-22r-n-22n=n(r-2)-r(n-2)2nr=r-nnr=1n-1r=αn,

therefore (2.9) can be equivalently expressed as

(2.14)qp<2*2m

because

1+2αn-2=1+2*αn=(2.13)1+2*(12m-12*)=2*2m.

Therefore, due to (2.14), in (2.11) we have β~>0 and θ>1.

Remark 2.5.

The result obtained is sharp in the sense that if m=1 (r=+), then the relation between p and q reduces to the analogous one in [33, Theorem 2.1], i.e. qp<nn-2.

Proof.

Let uW1,q(Ω) be a local minimizer of (1.1). Then u satisfies the Euler first variation

Ωi=1nfξi(x,Du)φxi(x)dx=0for all φW01,q(Ω).

By (2.8), the technique of the difference quotients (see [30, 18], in particular [28, Chapter 8, Sections 8.1 and 8.2]) gives

(2.15)uWloc1,(Ω)Wloc2,min(2,q)(Ω)and(1+|Du|2)q-22|D2u|2Lloc1(Ω).

Let ηC01(Ω) and for any fixed s{1,,n} define

φ=η2uxsΦ((|Du|-1)+)

for Φ:[0,+)[0,+) increasing, locally Lipschitz continuous function, with Φ and Φ bounded on [0,+), such that Φ(0)=Φ(0)=0 and

(2.16)Φ(s)scΦΦ(s)

for a suitable constant cΦ1. Here (a)+ denotes the positive part of a; in the following we denote

Φ((|Du|-1)+)=Φ(|Du|-1)+.

We have then

(2.17)φxi=2ηηxiuxsΦ(|Du|-1)++η2uxsxiΦ(|Du|-1)++η2uxsΦ(|Du|-1)+[(|Du|-1)+]xi.

Let q2. By (2.15) we have that |D2u|2Lloc1(Ω) . Otherwise if 1<q<2, we use the fact that uWloc1,(Ω) to infer that there exists M=M(suppφ) such that

|Du(x)|Mfor a.e. xsuppφ.

Now since q-2<0, we have

(1+M2)q-22|D2u|2(1+|Du|2)q-22|D2u|2,

and by (2.15) we again get |D2u|2L1(suppφ). Therefore we can insert φxi in the following second variation,

Ω{i,j=1nfξiξj(x,Du)uxjxsφxi+i=1nfξixs(x,Du)φxi}𝑑x=0for all s=1,,n,

and we obtain

0=s[Ω2ηΦ(|Du|-1)+i,jfξiξj(x,Du)ηxiuxsuxsxjdx
+Ωη2Φ(|Du|-1)+i,jfξiξj(x,Du)uxsxiuxsxjdx
+Ωη2Φ(|Du|-1)+i,jfξiξj(x,Du)uxsuxsxj[(|Du|-1)+]xidx
+Ω2ηΦ(|Du|-1)+ifξixs(x,Du)ηxiuxsdx
+Ωη2Φ(|Du|-1)+ifξixs(x,Du)uxsxidx
+Ωη2Φ(|Du|-1)+ifξixs(x,Du)uxs[(|Du|-1)+]xidx]
(2.18)=:s(I1s+I2s+I3s+I4s+I5s+I6s).

In the following, constants will be denoted by C, regardless of their actual value.

Let us start with the estimate of the first integral in (2.18). By the Cauchy–Schwarz inequality, the Young inequality and (1.7), we have

|sI1s|=|Ω2ηΦ(|Du|-1)+i,j,sfξiξj(x,Du)ηxiuxsuxsxjdx|
Ω2ηΦ(|Du|-1)+{i,j,sfξiξj(x,Du)ηxiuxsηxjuxs}12{i,j,sfξiξj(x,Du)uxsxiuxsxj}12𝑑x
CΩ|Dη|2Φ(|Du|-1)+|Du|q𝑑x+12Ωη2Φ(|Du|-1)+i,j,sfξiξj(x,Du)uxsxiuxsxjdx.

Let us consider the third integral in (2.18). First of all we observe that

[(|Du|-1)+]xisuxsuxsxj=[(|Du|-1)+]xi|Du|[(|Du|-1)+]xj.

This entails using (1.7)

sI3s=Ωη2Φ(|Du|-1)+i,j,sfξiξj(x,Du)uxsuxsxj[(|Du-1|+)]xidx
M1Ωη2Φ(|Du|-1)+|Du|p-1|D(|Du|-1)+|2𝑑x0.

We now deal with the fourth integral in (2.18). We have

|sI4s|=|Ω2ηΦ(|Du|-1)+i,sfξixs(x,Du)ηxiuxsdx|
(1.7)Ω2ηΦ(|Du|-1)+h(x)|Du|p+q-22i,s|ηxiuxs|dx
CΩ(η2+|Dη|2)h(x)Φ(|Du|-1)+|Du|q𝑑x.

Consider now the fifth integral in (2.18). We have

|sI5s|=|Ωη2Φ(|Du|-1)+i,sfξixs(x,Du)uxsxjdx|
(1.7)Ωη2Φ(|Du|-1)+h(x)|Du|p+q-22|D2u|𝑑x
Ω[η2Φ(|Du|-1)+|Du|p-2|D2u|2]12[η2Φ(|Du|-1)+|h(x)|2|Du|q]12𝑑x
εΩη2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x+CεΩη2Φ(|Du|-1)+|h(x)|2|Du|q𝑑x,

where in the last line we used the Young inequality. Finally, for any 0<δ<1,

|sI6s|=|Ωη2i,sfξixs(x,Du)uxsΦ(|Du|-1)+[(|Du|-1)+]xidx|
(1.7)Ωη2Φ(|Du|-1)+h(x)|Du|p+q-22|Du||D(|Du|-1)+|𝑑x
Ωη2Φ(|Du|-1)+h(x)|Du|p+q2|D2u|𝑑x
=Ωη2Φ(|Du|-1)+h(x)[(|Du|-1)++δ][(|Du|-1)++δ]-1|Du|p+q2|D2u|𝑑x
Ωη2{1cΦΦ(|Du|-1)+[(|Du|-1)++δ]|Du|p-2|D2u|2}12
×{cΦΦ(|Du|-1)+|h(x)|2|Du|q+2[(|Du|-1)++δ]-1}12dx
CεcΦΩη2Φ(|Du|-1)+|h(x)|2|Du|q+2[(|Du|-1)++δ]-1𝑑x
+εcΦΩη2Φ(|Du|-1)+[(|Du|-1)++δ]|Du|p-2|D2u|2𝑑x.

Since Ω={x:|Du(x)|2}{x:|Du(x)|<2} and (|Du|-1)+1 in {x:|Du(x)|2}, we also have

(2.19)(|Du|-1)++δ2(|Du|-1)+

as long as we have chosen δ<1. Therefore, using (2.16), we can estimate the last integral as

Ωη2Φ(|Du|-1)+[(|Du|-1)++δ]|Du|p-2|D2u|2𝑑x
=|Du|2η2Φ(|Du|-1)+[(|Du|-1)++δ]|Du|p-2|D2u|2𝑑x
+1<|Du|<2η2Φ(|Du|-1)+[(|Du|-1)++δ]|Du|p-2|D2u|2𝑑x
(2.19)2|Du|2η2Φ(|Du|-1)+(|Du|-1)+|Du|p-2|D2u|2𝑑x
+1<|Du|<2η2Φ(|Du|-1)+(|Du|-1)+|Du|p-2|D2u|2𝑑x
+δ1<|Du|<2η2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x
(2.16)2cΦΩη2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x+δ1<|Du|<2η2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x.

Therefore we finally have

|sI6s|CεcΦΩη2Φ(|Du|-1)+|h(x)|2|Du|q+2[(|Du|-1)++δ]-1dx
+2εΩη2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x+δε1<|Du|<2η2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x.

Now, for ε sufficiently small and putting together all the previous estimates, we deduce that there exists a constant C depending on n,r,p,q,M1 such that

Ωη2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x
CcΦΩ(η2+|Dη|2)(1+h(x))2|Du|q[Φ(|Du|-1)++Φ(|Du|-1)+|Du|2[(|Du|-1)++δ]-1]𝑑x
(2.20)+δ1<|Du|<2η2Φ(|Du|-1)+|Du|p-2|D2u|2𝑑x.

Let us now set

Φ(s):=(1+s)γ-2s2,γ0,

with

Φ(s)=(γs+2)s(1+s)γ-3.

It is easy to check that Φ satisfies (2.16) with cΦ=2(1+γ).

We now approximate this function Φ by a sequence of functions Φh, each of them being equal to Φ in the interval [0,h], and then extended to [h,+) with the constant value Φ(h). Since Φh and Φh converge monotonically to Φ and Φ, by inserting Φh in (2.20), it is possible to pass to the limit as h+ by the Monotone Convergence Theorem.

Therefore, for every 0<δ<1, since

(|Du|-1)+(|Du|-1)++δ1for all δ>0

and Φ(t-1)+C(γ) when 1<t<2, we obtain

Ωη2(1+(|Du|-1)+)γ-2(|Du|-1)+2|Du|p-2|D2u|2𝑑x
(2.21)C(1+γ)2Ω(η2+|Dη|2)(1+h(x))2(1+(|Du|-1)+)γ+q𝑑x+δC(γ)1<|Du|<2η2|Du|p-2|D2u|2𝑑x.

Using [20, formula (3.51) of Lemma 3.3], namely the fact that |Du|p-2C(p)(1+|Du|2)p-22 when |Du|>1, we have

1<|Du|<2η2|Du|p-2|D2u|2𝑑xC1<|Du|<2η2(1+|Du|2)p-22|D2u|2𝑑x<+

by (2.15), and for δ0 and the last term in the previous inequality vanishes.

Since hLr(Ω), by the Hölder inequality, since 1m+2r=1, by (2.21) we have

Ωη2(1+(|Du|-1)+)γ-2(|Du|-1)+2|Du|p-2|D((|Du|-1)+)|2𝑑x
(2.22)C(1+γ)21+hLr(Ω)2[Ω(η2+|Dη|2)m(1+(|Du|-1)+)(γ+q)m𝑑x]1m.

Let us introduce

(2.23)G(t)=1+0t(1+s)γ2+p2-2s𝑑s.

We obtain

(2.24)[G(t)]24(1+t)γ+p4(1+t)γ+q,

where we used the fact that pq. On the other hand

(2.25)G(t)=(1+t)γ2+p2-2t,

which in turn allows us to give the following estimate for the gradient of the function w=ηG((|Du|-1)+):

Ω|D(ηG((|Du|-1)+))|2𝑑x
2Ω|Dη|2|G((|Du|-1)+)|2𝑑x+2Ωη2[Gt((|Du|-1)+)]2[D((|Du|-1)+)]2𝑑x
C(1+γ)21+hLr(Ω)2[Ω(η2+|Dη|2)m[1+(|Du|-1)+](γ+q)m𝑑x]1m,

the second inequality by (2.22), (2.24), (2.25). By Sobolev’s inequality there exists a constant C (depending also on |Ω| when n=2) such that

{Ω[ηG((|Du|-1)+)]2𝑑x}22CΩ|D(ηG((|Du|-1)+))|2𝑑x

and by the previous inequality we get (for a different constant)

(2.26){Ω[ηG((|Du|-1)+)]2𝑑x}22C(1+γ)21+hLr(Ω)2[Ω(η2+|Dη|2)m[1+(|Du|-1)+](γ+q)m𝑑x]1m.

We take into account the definition of G(t) in (2.23) and we use Lemma 2.2, and in particular formula (2.6) with β=γ+p2. Being γ0, we have ββ0:=p2>0 and

(1+t)γ+p2c′′(γ+p2)2(1+0t(1+s)γ+p2-2s𝑑s)

for every γ0 and every t[0,+). In terms of G(t)=1+0t(1+s)γ2+p2-2s𝑑s equivalently

(1+t)γ+p2c′′(γ+p2)2G(t)for all γ0 and all t0.

Therefore, if t:=(|Du|-1)+,

(1+(|Du|-1)+)γ+p22(c′′)2(γ+p2)22[G((|Du|-1)+)]2for all γ0,

and by (2.26) we finally get

{Ωη2[1+(|Du|-1)+]γ+p22𝑑x}22(c′′)2(γ+p2)4{Ω[ηG((|Du|-1)+)]2𝑑x}22
C(γ+1)61+hLr(Ω)2[Ω(η2+|Dη|2)m[1+(|Du|-1)+](γ+q)m𝑑x]1m

with a new constant C and for every γ0.

As usual we consider a test function η equal to 1 in a ball Bρ, with suppηBR and such that |Dη|2(R-ρ). We get

(2.27)[Bρ[1+(|Du|-1)+][(γ+p)m]22m𝑑x]2m2C01+hLr(Ω)2m(γ+q)6m(R-ρ)2mBR[1+(|Du|-1)+](γ+q)m𝑑x,

where the constant C0 only depends on n,r,p,q,M1,M2 but is independent of γ.

Fixed 0<ρ0<R0ρ0+1, we define the following decreasing sequence of radii {ρk}k1:

ρk=ρ0+R0-ρ02kfor all k1.

We define recursively a sequence αk in the following way:

(2.28)α1:=0,αk+1:=(αk+pm)2*2m-qm.

Then we have the following representation formula for αk which can easily be proved by induction:

(2.29)αk=(p2*2-qm)[(2*2m)k-1-1]2*2m-1.

We rewrite (2.27) with R=ρk, ρ=ρk+1, γ=αkm and observe that

R-ρ:=ρk-ρk+1=R0-ρ02k+1.

Set, for all k1,

Ak:=(Bρk[1+(|Du|-1)+]αk+qm𝑑x)1αk+qm,
Ck:=C01+hLr(Ω)2m((αk+qm)32k+1R0-ρ0)2m;

we obtain for every k1,

(2.30)Ak+1Ck1αk+pmAkαk+qmαk+pm.

Let θ be defined by

(2.31)θ:=i=1αi+qmαi+pm.

We show that θ is finite and is given by

θ=qm(2*2m-1)p2*2-qm.

Indeed by (2.31), the recursive definition of αk, i.e. (2.28) and the representation formula for αk, namely (2.29), we have

θk:=i=1kαi+qmαi+pm=(2.28)qmαk+pm(2*2m)k-1=(2.29)qm(2*2m)k-1(p2*2-qm)[(2*2m)k-1-1]2*2m-1+pm
=qm(2*2m)k-1(2*2m-1)pm(2*2m-1)+(p2*2-qm)[(2*2m)k-1-1],

which yields (2.31) once we pass to the limit as k as long as 2*2m>1 in view of (2.12). Note that θ makes sense due to (2.12) and the bound (2.14) in Remark 2.4.

Iterating (2.30), we deduce

(2.32)Ak+1C~([1+hLr(Ω)(R0-ρ0)]β~A1)θk,k1,

where

C~:=C0β~θexp[θi=1log[(αi+qm)6m22m(i+1)]αi+pm]<+,

which is finite because the series is convergent (αi from the representation formula (2.29) grows exponentially) and

i=12mαi+pm=(2.29)i=12m(p2*2-qm)[(2*2)i-1-1]2*2m-1+pmi=12m(p2*2-qm)(2*2m)i-12*2m-1
2m(2*2m-1)p2*2-qmi=0(2m2*)i=2m(2*2m-1)p2*2-qm1(1-2m2*)=2*p2*2-qm=:β~,

where in the first inequality we used the fact that

(p2*2-qm)[(2*2m)i-1-1]2*2m-1+pm(p2*2-qm)(2*2m)i-12*2m-1-p2*2-qm2*2m-1+pm0qp.

By letting k+ in (2.32), we have (2.10). Therefore, the proof of Lemma 2.3 is complete. ∎

The a-priori estimate (1.12) in Theorem 1.1 follows by the classical interpolation inequality

(2.33)vLs(Bρ)vLp(Bρ)psvL(Bρ)1-ps

for any sp, which permits to estimate the essential supremum of the gradient of the local minimizer in terms of its Lp-norm.

Proof of Theorem 1.1.

Let us set

V(x):=1+(|Du|(x)-1)+

then estimate (2.10) becomes

(2.34)supxBρ|V(x)|C([1+hLr(Ω)R-ρ]β~VLqm(BR))θ

for every ρ,R such that 0<ρ<Rρ+1 and where C=C(n,r,p,q,M1,M2).

In the following we denote

s:=qm.

At this point, (2.34) and (2.33) give

(2.35)VLs(Bρ)C1-psVLp(Bρ)ps([1+hLr(Ω)R-ρ]β~VLs(BR))θ(1-pqm).

We observe that

(2.36)τ:=θ(1-pqm)<1,

because

θ(1-pqm)<1q2*2-qm-p2*2m+pp2*2-qm<1q<p(1-22*+1m)=(2.13)p(1+2αn).

For 0<ρ<R and for every k0, let us define ρk:=R-(R-ρ)2-k. By inserting in (2.35) ρ=ρk and R=ρk+1, (so that R-ρ=(R-ρ)2-(k+1)) we have, for every k0,

(2.37)VLs(Bρk)C1-psVLp(Bρk)ps(2β~(k+1)[1+hLr(Ω)(R-ρ)]β~VLs(Bρk+1))τ.

By iteration of (2.37), we deduce for k0,

(2.38)VLs(Bρ0)(C1-ps[1+hLr(Ω)(R-ρ)]β~τVLp(Bρk)ps)i=0kτi2β~i=0k+1iτi(VLs(Bρk+1))τk+1.

By (2.36), the series appearing in (2.38) are convergent. Since

VLs(Bρk)VLs(BR),

we can pass to the limit as k+ and we obtain for every 0<ρ<R with a constant C=C(n,r,p,q,M1,M2) independent of k,

(2.39)VLs(Bρ)C([1+hLr(Ω)(R-ρ)]β~τVLp(BR)ps)11-τ.

Combining (2.34) and (2.39), by setting ρ=(R+ρ)2 we have

VL(Bρ)C([1+hLr(Ω)(ρ-ρ)]β~VLs(Bρ))θ
C([1+hLr(Ω)(ρ-ρ)]β~(1-τ)[1+hLr(Ω)(R-ρ)]β~τVLp(BR)ps)θ1-τ;

now, since

(ρ-ρ)=(R-ρ)=R-ρ2,

we get

DuL(Bρ;n)C([1+hLr(Ω)(R-ρ)]β(BR{1+|Du|p}𝑑x)1p)γ,

where

β:=β~qmp=2*qpmp2*2-qm,γ:=θpqm(1-θ(1-pqm)),θ:=qm(2*2m-1)p2*2-qm,

so (1.12) follows. ∎

Let now f satisfy (1.7) and (1.14). Under these assumptions on f, we obtain the following result.

Theorem 2.6.

Let uW1,p(Ω) be a local minimizer of the integral functional (1.1). Assume that f=f(x,ξ) in (1.1) satisfies (1.7), (1.14) and (2.8), with

(2.40)qp<1+1n-1r.

Then there exist positive constants C,β^,γ^ depending on n,r,p,q,M1,M2,ρ,R such that

(2.41)DuL(Bρ;n)C[1+hLr(Ω)]β^γ^(BR{1+f(x,Du)}𝑑x)γ^p

for every 0<ρ<Rρ+1.

Proof.

Let t:=2q-p. Then (1.14) can be written explicitly in the form

|fξx(x,ξ)|h(x)|ξ|t+p-22,|ξ|1.

Moreover, (2.40) in terms of p and t is equivalent to

tp<1+2αnwith α=1-nr.

Thus all the assumptions of Theorem 1.1 are satisfied with q replaced by t. In particular, the second inequality in (1.7) holds with q replaced by t since tq. Then the conclusion of Theorem 1.1 holds with q=t which corresponds to (2.41) with

β^:=2*p(2q-p)mp2*2-(2q-p)m,γ^:=θp(2q-p)m(1-θ)+pθ,

since f(x,ξ)C|ξ|p for every |ξ|1. ∎

3 Extension of the integral energy

Let f:Ω×n[0,+) be a continuous function, convex in ξ such that

(3.1)|ξ|pf(x,ξ)C(1+|ξ|q)for a.e. xΩ and all ξn.

For u0W1,q(Ω), we define the extension to W1,p(Ω) of the integral functional Ωf(x,Du)𝑑x, i.e.

(3.2)F(u)=inf{lim infkΩf(x,Duk)𝑑x:ukW01,q(Ω)+u0,uk𝑤u in W1,p(Ω)}

with

Ωf(x,Du0)𝑑x<+.

It is easy to check that

F(u)=Ωf(x,Du)𝑑xfor uW01,q(Ω)+u0.

In fact, for uk=u for all k,

F(u)Ωf(x,Du)𝑑x.

On the other hand, by the semicontinuity of Ωf(x,Du)𝑑x with respect to the weak topology of W1,p, the inverse inequality also holds.

Lemma 3.1.

For each vW01,p(Ω)+u0, there exists a sequence vkW01,q(Ω)+u0 such that vkv weakly in W1,p(Ω) and

F(v)=limk+Ωf(x,Dvk)𝑑x.

Proof.

The proof follows similarly as in [5]. We give the sketch of the proof.

Let vW01,p(Ω)+u0 such that F(v)<. Then, for all k, there exists vh(k)W01,q(Ω)+u0 such that vh(k)𝑤v, as h+, weakly in W1,p(Ω) and

F(v)limh+Ωf(x,Dvh(k))𝑑xF(v)+1k.

Moreover, by the weak convergence of vh(k) in W1,p(Ω) we get

limh+vh(k)-vLp(Ω)=0

and for h sufficiently large,

Ω|Dvh(k)|p𝑑xΩf(x,Dvh(k))𝑑xF(v)+1.

Then for all k there exists hk such that for all hhk,

vh(k)-vLp(Ω)<1k

and for h=hk, by denoting wk=vhk(k), we have

wk-vLp(Ω)<1kandΩ|Dwk|p𝑑xC;

then wk𝑤v as k+ in the weak topology of W1,p(Ω) and

F(v)Ωf(x,Dwk)𝑑xF(v)+1k,

i.e.

limk+Ωf(x,Dwk)𝑑x=F(v).

4 Existence and regularity

First of all we prove an approximation theorem for f through a suitable sequence of regular functions.

Proposition 4.1.

Let f be satisfying the growth conditions (3.1), fξξ and fξx Carathéodory functions, satisfying (1.7) and (1.14) with M0=1 and f strictly convex at infinity. Then there exists a sequence of C2-functions fk:Ω×Rn[0,+), fk convex in the last variable and strictly convex at infinity, such that fk converges to f as and k for a.e. xΩ, for all ξRn and uniformly in Ω0×K, where Ω0Ω and K being a compact set of Rn. Moreover:

  1. there exists C~, independently of k,, such that

    (4.1)|ξ|pfk(x,ξ)C~(1+|ξ|q)for a.e. xΩ and all ξn,
  2. there exists M~1>0 such that for |ξ|>2 and a.e. xΩ,

    (4.2)M~1|ξ|p-2|λ|2i,jfξiξjk(x,ξ)λiλj,λn,
  3. there exists c(k)>0 such that for all (x,ξ)Ω×n and λn,

    (4.3)c(k)(1+|ξ|2)q-22|λ|2i,jfξiξjk(x,ξ)λiλj,
  4. there exists M~2>0 such that for |ξ|>2 and a.e. xΩ,

    (4.4)|fξξk(x,ξ)|M~|ξ|q-2,
  5. there exists C(k) such that for a.e. xΩ and ξn,

    (4.5)|fξξk(x,ξ)|C(k)(1+|ξ|2)q-22,
  6. there exists a constant C>0 such that for a.e. xΩ and |ξ|>2,

    (4.6)|fξxk(x,ξ)|Ch(x)|ξ|q-1,

    where h𝒞(Ω) is the regularized function of h which converges to h in Lr(Ω),

  7. for Ω0Ω, there exists a constant C such that for xΩ0 and ξn,

    (4.7)|fξxk(x,ξ)|C(k,,Ω0)(1+|ξ|2)q-12.

Proof.

We argue as in the proof of [25, Theorem 2.7 (Step 3)] and [20, Lemma 4.3]. For the sake of completeness, we give a sketch of the arguments of the proof.

Let B be the unit ball of n centered in the origin and consider a positive decreasing sequence ε0. We introduce

f(x,ξ)=B×Bρ(y)ρ(η)f(x+εy,ξ+εη)𝑑η𝑑y,

where ρ is a positive symmetric mollifier, and

(4.8)fk(x,ξ)=f(x,ξ)+1k(1+|ξ|2)q2.

It is easy to check that the sequence fk satisfies conditions (4.1), (4.2), (4.3), (4.4), (4.5), (4.7). Let us verify (4.6). For |ξ|>2 we have

|fξxk(x,ξ)|B×Bρ(y)ρ(η)|ξ+εη|q-1h(x+εy)𝑑y𝑑ηCh(x)|ξ|q-1,

where

h(x)=Bρ(y)h(x+εy)𝑑y,

h is a smooth function and it converges to h in Lr(Ω). Moreover,

|fξxk(x,ξ)|C(k,Ω0)[1+hL(Ω0)](1+|ξ|2)q-12.

This concludes the proof. ∎

Proof of Theorem 1.2.

For u0W1,q(Ω), let us consider the variational problems

(4.9)inf{Ωfk(x,Dv)𝑑x:vW01,q(Ω)+u0},

where fk are defined in (4.8). By semicontinuity arguments, there exists vku0+W01,q(Ω), a solution to (4.9). By the growth conditions and the minimality of vk, we get

Ω|Dvk|p𝑑xΩfk(x,Dvk)𝑑x
Ωfk(x,Du0)𝑑x
=Ωf(x,Du0)𝑑x+1kΩ(1+|Du0|2)q2𝑑x.

Moreover, the properties of the convolutions imply that

f(x,Du0)f(x,Du0)a.e. in Ω,

and since

Ωf(x,Du0)𝑑xCΩ(1+|Du0|2)q2𝑑x,

by the Lebesgue Dominated Convergence Theorem we deduce therefore

limΩ|Dvk|p𝑑xlimΩf(x,Du0)𝑑x+1kΩ(1+|Du0|2)q2𝑑x
=Ωf(x,Du0)𝑑x+1kΩ(1+|Du0|2)q2𝑑x.

By Proposition 4.1, the functions fk satisfy (1.7), (1.14) and (2.8), so we can apply the a-priori estimate (2.41) to vk and obtain, by standard covering arguments for all ΩΩ,

DvkL(Ω;n)C(Ω)[1+hLr(Ω)]β^γ^[Ω(1+fk(x,Dvk))𝑑x]γ^p.

Since 1+hLr(Ω)=(1+h)Lr(Ω)1+hLr(Ω), we obtain

DvkL(Ω;n)C(Ω)[1+hLr(Ω)]β^γ^[Ω(1+fk(x,Dvk))𝑑x]γ^p
C(Ω)[1+hLr(Ω)]β^γ^[Ω1+f(x,Du0)+1k(1+|Du0|2)q2dx]γ^p,

where C,γ^,β^ depend on n,r,p,q,M1,M2,ρ,R but are independent of ,k. Therefore we conclude that

vkvkweakly in W01,p(Ω)+u0,
vkvkweakly star in Wloc1,(Ω),

and by the previous estimates

DvkLp(Ω;n)lim infDvkLp(Ω;n)
Ωf(x,Du0)𝑑x+Ω(1+|Du0|2)q2𝑑x

and

DvkL(Ω;n)lim infDvkL(Ω;n)
C(Ω)[1+hLr(Ω)]β^γ^[Ω1+f(x,Du0)dx+Ω(1+|Du0|2)q2𝑑x]γ^p.

Thus we can deduce that there exists, up to subsequences, u¯u0+W01,p(Ω) such that

vku¯weakly in W01,p(Ω)+u0,
vku¯weakly star in Wloc1,(Ω).

Now, for any fixed k, using the uniform convergence of f to f in Ω0×K (for any K compact subset of n) and the minimality of vk, we get for all vW01,q(Ω)+u0,

Ω0f(x,Dvk)𝑑xlim infΩ0f(x,Dvk)𝑑x
=lim infΩ0f(x,Dvk)𝑑x
lim infΩ0f(x,Dvk)𝑑x+1kΩ(1+|Dvk|2)q2𝑑x
lim infΩf(x,Dvk)𝑑x+1kΩ(1+|Dvk|2)q2𝑑x
lim infΩf(x,Dv)𝑑x+1kΩ(1+|Dv|2)q2𝑑x.

Then, for Ω0Ω,

Ωf(x,Dvk)𝑑xΩf(x,Dv)𝑑x+1kΩ(1+|Dv|2)q2𝑑x.

By definition (3.2), we have

(4.10)F(u¯)lim infkΩf(x,Dvk)𝑑xΩf(x,Dv)𝑑xfor all vW01,q(Ω)+u0.

Let wW01,p(Ω)+u0. By Lemma 3.1, there exists vkW01,q(Ω)+u0 such that vkw weakly in W1,p(Ω) and

limkΩf(x,Dvk)𝑑x=F(w).

By (4.10),

F(u¯)Ωf(x,Dvk)𝑑x,

and we can conclude that

F(u¯)limkΩf(x,Dvk)𝑑x=F(w)for all wW01,p(Ω)+u0.

Then u¯Wloc1,(Ω) is a solution to the problem min{F(u):uW01,p(Ω)+u0}. ∎

5 Regularity of local minimizers in a special case

Let us consider now the case of a special form of integrand

(5.1)f(x,ξ)=i=1Nai(x)gi(ξ)

with ai(x)>0 a.e. in Ω, aiW1,r(Ω), r>n, gi:n[0,+) convex in ξ and strictly convex for ξ such that |ξ|M0. The following regularity result holds.

Theorem 5.1.

Assume that f=f(x,ξ) as in (5.1) satisfies the assumptions of Theorem 2.6. Then every local minimizer uW1,p(Ω) of the integral functional

(5.2)Ωf(x,Dv)𝑑x=Ωi=1Nai(x)gi(Du)dx

is locally Lipschitz continuous in Ω.

Proof.

Let uW1,p(Ω) be a local minimizer of the integral functional (5.2). For a suitable φσ mollifier, consider uσ=uφσWloc1,q(Ω). Consider the following sequence of problems in BRΩ:

(5.3)inf{BRfk(x,Dv)𝑑x:vW01,q(BR)+uσ},

where fk are defined in Proposition 4.1.

For fixed σ,,k, problem (5.3) has a unique solution vσkW01,q(BR)+uσ. By proceeding as in the previous theorem, we have that for each fixed σ, by the minimality of vσk,

vσkvσkweakly in W01,p(BR)+uσ,
vσkvσkweakly star in Wloc1,(BR).

We also have

vσkkvσweakly in W01,p(BR)+uσ,
vσkkvσweakly star in Wloc1,(BR)

and

DvσL(Bρ;n)Clim infk[1+BRf(x,Duσ)𝑑x+1kBR(1+|Duσ|2)q2𝑑x]γ^p
(5.4)=Clim infk[1+BRf(x,Duσ)𝑑x]γ^p

for any 0<ρ<R and where C is independent of k,σ. For fixed k, by proceeding as in the previous theorem we have

BRf(x,Dvσk)𝑑xBRf(x,Duσ)𝑑x+1kBR(1+|Duσ|2)q2𝑑x.

Then, by semicontinuity,

BRf(x,Dvσ)𝑑xlim infkBRf(x,Duσ)𝑑x+1kBR(1+|Duσ|2)q2𝑑x
(5.5)BRf(x,Duσ)𝑑x.

Now we claim that, by the particular form of f, we may deduce

(5.6)lim infσ0BRf(x,Duσ)𝑑xBRf(x,Du)𝑑x.

Since gi is convex, for i=1,,N, Jensen’s inequality (applied to each gi) yields

BRai(x)gi(Duσ)𝑑x=BRai(x)gi(BσDu(y)φσ(x-y)𝑑y)𝑑x
BRai(x)Bσgi(Du(y))φσ(x-y)𝑑y𝑑x
=BRBσai(x)φσ(x-y)𝑑ygi(Du(y))𝑑x
BR+σ(ai)σ(y)gi(Du(y))𝑑y.

Then

i=1NBRai(x)gi(Duσ)𝑑xi=1NBR+σ(ai)σ(x)gi(Du)𝑑x

so that passing to the limit as σ0,

lim infσ0i=1NBRai(x)gi(Duσ)𝑑xi=1NBRai(x)gi(Du)𝑑x

because (ai)σai in L(BR), gi(Du)L1(BR), the Dominated Convergence Theorem may be applied, and (5.6) holds.

By collecting (5.5) and (5.6),

(5.7)lim infσ0BRf(x,Dvσ)𝑑xBRf(x,Du)𝑑x.

On the other hand, the growth assumption on f yields, since u is a local minimizer of (5.2),

lim infσ0BR|Dvσ|p𝑑xlim infσ0BRf(x,Dvσ)𝑑x(5.7)BRf(x,Du)𝑑x<+.

Thus there exists v¯u+W01,p(BR) such that, up to a subsequence,

vσv¯weakly in W1,p(BR).

By the semicontinuity of the functional, using (5.5) and (5.7),

(5.8)BRf(x,Dv¯)𝑑xlim infσ0BRf(x,Dvσ)𝑑xBRf(x,Du)𝑑x.

Moreover, since (5.4) holds, Dvσ converges to Dv¯ as σ0 in the weak star topology of L and there exists a constant C such that, for any 0<ρ<R,

Dv¯L(Bρ;n)C[1+BRf(x,Du)𝑑x]γ^p.

Consider the following problem in BRΩ:

(5.9)inf{BRf(x,Dv)𝑑x:vW01,p(BR)+u}.

Then (5.8) implies that v¯ and u are solutions to (5.9) and v¯Wloc1,(BR).

In the present case the functional is not strictly convex; we proceed as in [20, Theorem 2.1] (see also [25]) and we have that uWloc1,(BR). Indeed, set

E0:={xBR:|Du(x)+Dv¯(x)2|>M0,Du(x)Dv¯(x)}andw:=u+v¯2.

If E0 has positive measure, then from the convexity of f(x,) we have

(5.10)BRE0f(x,Dw)𝑑x12BRE0f(x,Du)𝑑x+12BRE0f(x,Dv¯)𝑑x.

Now, by the strict convexity of f(x,ξ) for ξ such that |ξ|M0 and applying two times the inequality

f(x,η)>f(x,ξ)+fξ(x,ξ),η-ξfor ξ such that |ξ|M0

first with ξ=Dw and η=Du, then for ξ=Dw and η=Dv¯, finally by adding up the two inequalities obtained, we have

(5.11)BRE0f(x,Dw)𝑑x<12BRE0f(x,Du)𝑑x+12BRE0f(x,Dv)𝑑x.

Adding (5.10) and (5.11), we get a contradiction with the minimality of u and v¯. Therefore the set E0 has zero measure, which implies that

supBρ|Du(x)|supBρ|Du(x)+Dv¯(x)|+supBρ|Dv¯(x)|2M0+supBρ|Dv¯(x)|

and this yields the thesis. ∎


Communicated by Juha Kinnunen


Funding statement: The authors are members of GNAMPA (Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni) of INdAM (Istituto Nazionale di Alta Matematica).

Acknowledgements

The authors wish to express their gratitude to the referee for carefully reading the manuscript providing useful comments and remarks. In particular, the referee pointed out a technical problem which has been modified correctly, which however did not influenced the method and the details in the proof of the a-priori estimates.

References

[1] A. L. Baisón, A. Clop, R. Giova, J. Orobitg and A. Passarelli di Napoli, Fractional differentiability for solutions of nonlinear elliptic equations, Potential Anal. 46 (2017), no. 3, 403–430. 10.1007/s11118-016-9585-7Search in Google Scholar

[2] P. Baroni, M. Colombo and G. Mingione, Harnack inequalities for double phase functionals, Nonlinear Anal. 121 (2015), 206–222. 10.1016/j.na.2014.11.001Search in Google Scholar

[3] P. Baroni, M. Colombo and G. Mingione, Nonautonomous functionals, borderline cases and related function classes, Algebra i Analiz 27 (2015), no. 3, 6–50. 10.1090/spmj/1392Search in Google Scholar

[4] P. Baroni, M. Colombo and G. Mingione, Regularity for general functionals with double phase, preprint (2017), https://arxiv.org/abs/1708.09147. 10.1007/s00526-018-1332-zSearch in Google Scholar

[5] L. Boccardo and P. Marcellini, Sulla convergenza delle soluzioni di disequazioni variazionali, Ann. Mat. Pura Appl. (4) 110 (1976), 137–159. 10.1007/BF02418003Search in Google Scholar

[6] V. Bögelein, F. Duzaar and P. Marcellini, Parabolic systems with p,q-growth: A variational approach, Arch. Ration. Mech. Anal. 210 (2013), no. 1, 219–267. 10.1007/s00205-013-0646-4Search in Google Scholar

[7] V. Bögelein, F. Duzaar and P. Marcellini, Existence of evolutionary variational solutions via the calculus of variations, J. Differential Equations 256 (2014), no. 12, 3912–3942. 10.1016/j.jde.2014.03.005Search in Google Scholar

[8] V. Bögelein, F. Duzaar and P. Marcellini, A time dependent variational approach to image restoration, SIAM J. Imaging Sci. 8 (2015), no. 2, 968–1006. 10.1137/140992771Search in Google Scholar

[9] G. Buttazzo and M. Belloni, A survey on old and recent results about the gap phenomenon in the calculus of variations, Recent Developments in Well-Posed Variational Problems, Math. Appl. 331, Kluwer Academic Publisher, Dordrecht (1995), 1–27. 10.1007/978-94-015-8472-2_1Search in Google Scholar

[10] M. Chipot and L. C. Evans, Linearisation at infinity and Lipschitz estimates for certain problems in the calculus of variations, Proc. Roy. Soc. Edinburgh Sect. A 102 (1986), no. 3–4, 291–303. 10.1017/S0308210500026378Search in Google Scholar

[11] M. Colombo and G. Mingione, Bounded minimisers of double phase variational integrals, Arch. Ration. Mech. Anal. 218 (2015), no. 1, 219–273. 10.1007/s00205-015-0859-9Search in Google Scholar

[12] M. Colombo and G. Mingione, Regularity for double phase variational problems, Arch. Ration. Mech. Anal. 215 (2015), no. 2, 443–496. 10.1007/s00205-014-0785-2Search in Google Scholar

[13] G. Cupini, F. Giannetti, R. Giova and A. Passarelli di Napoli, Higher integrability for minimizers of asymptotically convex integrals with discontinuous coefficients, Nonlinear Anal. 154 (2017), 7–24. 10.1016/j.na.2016.02.017Search in Google Scholar

[14] G. Cupini, M. Guidorzi and E. Mascolo, Regularity of minimizers of vectorial integrals with p-q growth, Nonlinear Anal. 54 (2003), no. 4, 591–616. 10.1016/S0362-546X(03)00087-7Search in Google Scholar

[15] G. Cupini, P. Marcellini and E. Mascolo, Existence and regularity for elliptic equations under p,q-growth, Adv. Differential Equations 19 (2014), no. 7–8, 693–724. 10.57262/ade/1399395723Search in Google Scholar

[16] E. De Giorgi, Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari, Mem. Accad. Sci. Torino. Cl. Sci. Fis. Mat. Nat. (3) 3 (1957), 25–43. Search in Google Scholar

[17] E. De Giorgi and T. Franzoni, Su un tipo di convergenza variazionale, Atti Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Natur. (8) 58 (1975), no. 6, 842–850. Search in Google Scholar

[18] E. DiBenedetto, C1+α local regularity of weak solutions of degenerate elliptic equations, Nonlinear Anal. 7 (1983), no. 8, 827–850. 10.1016/0362-546X(83)90061-5Search in Google Scholar

[19] M. Eleuteri, P. Marcellini and E. Mascolo, Lipschitz continuity for energy integrals with variable exponents, Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl. 27 (2016), no. 1, 61–87. 10.4171/RLM/723Search in Google Scholar

[20] M. Eleuteri, P. Marcellini and E. Mascolo, Lipschitz estimates for systems with ellipticity conditions at infinity, Ann. Mat. Pura Appl. (4) 195 (2016), no. 5, 1575–1603. 10.1007/s10231-015-0529-4Search in Google Scholar

[21] M. Eleuteri and A. Passarelli di Napoli, Higher differentiability for solutions to a class of obstacle problems, preprint (2017). 10.1007/s00526-018-1387-xSearch in Google Scholar

[22] A. Esposito, F. Leonetti and P. V. Petricca, Absence of Lavrentiev gap for non-autonomous functionals with (p,q)-growth, Adv. Nonlinear Anal. (2017), 10.1515/anona-2016-0198. 10.1515/anona-2016-0198Search in Google Scholar

[23] L. Esposito, F. Leonetti and G. Mingione, Regularity results for minimizers of irregular integrals with (p,q) growth, Forum Math. 14 (2002), no. 2, 245–272. 10.1515/form.2002.011Search in Google Scholar

[24] L. Esposito, F. Leonetti and G. Mingione, Sharp regularity for functionals with (p,q) growth, J. Differential Equations 204 (2004), no. 1, 5–55. 10.1016/j.jde.2003.11.007Search in Google Scholar

[25] I. Fonseca, N. Fusco and P. Marcellini, An existence result for a nonconvex variational problem via regularity, ESAIM Control Optim. Calc. Var. 7 (2002), 69–95. 10.1051/cocv:2002004Search in Google Scholar

[26] M. Giaquinta, Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems, Ann. of Math. Stud. 105, Princeton University Press, Princeton, 1983. 10.1515/9781400881628Search in Google Scholar

[27] M. Giaquinta, Growth conditions and regularity, a counterexample, Manuscripta Math. 59 (1987), no. 2, 245–248. 10.1007/BF01158049Search in Google Scholar

[28] E. Giusti, Direct Methods in the Calculus of Variations, World Scientific Publishing, River Edge, 2003. 10.1142/5002Search in Google Scholar

[29] A. D. Ioffe and V. M. Tihomirov, Extension of variational problems (in Russian), Trudy Moskov. Mat. Obšč. 18 (1968), 187–246. Search in Google Scholar

[30] O. A. Ladyzhenskaya and N. N. Ural’tseva, Linear and Quasilinear Elliptic Equations, Academic Press, New York, 1968. Search in Google Scholar

[31] P. Marcellini, Un example de solution discontinue d’un problème variationnel dans le cas scalaire, Preprint 11, Istituto Matematico “U.Dini”, Università di Firenze, 1987. Search in Google Scholar

[32] P. Marcellini, Regularity of minimizers of integrals of the calculus of variations with nonstandard growth conditions, Arch. Ration. Mech. Anal. 105 (1989), no. 3, 267–284. 10.1007/BF00251503Search in Google Scholar

[33] P. Marcellini, Regularity and existence of solutions of elliptic equations with p,q-growth conditions, J. Differential Equations 90 (1991), no. 1, 1–30. 10.1016/0022-0396(91)90158-6Search in Google Scholar

[34] P. Marcellini, Regularity for elliptic equations with general growth conditions, J. Differential Equations 105 (1993), no. 2, 296–333. 10.1006/jdeq.1993.1091Search in Google Scholar

[35] P. Marcellini, Regularity for some scalar variational problems under general growth conditions, J. Optim. Theory Appl. 90 (1996), no. 1, 161–181. 10.1007/BF02192251Search in Google Scholar

[36] G. Mingione, Regularity of minima: an invitation to the dark side of the calculus of variations, Appl. Math. 51 (2006), no. 4, 355–426. 10.1007/s10778-006-0110-3Search in Google Scholar

[37] A. Passarelli di Napoli, Higher differentiability of minimizers of variational integrals with Sobolev coefficients, Adv. Calc. Var. 7 (2014), no. 1, 59–89. 10.1515/acv-2012-0006Search in Google Scholar

[38] J. Serrin, A new definition of the integral for nonparametric problems in the calculus of variations, Acta Math. 102 (1959), 23–32. 10.1007/BF02559566Search in Google Scholar

[39] V. V. Zhikov, On Lavrentiev’s phenomenon, Russian J. Math. Phys. 3 (1995), no. 2, 249–269. Search in Google Scholar

Received: 2017-07-02
Revised: 2017-11-01
Accepted: 2018-02-06
Published Online: 2018-03-16
Published in Print: 2020-07-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 11.5.2024 from https://www.degruyter.com/document/doi/10.1515/acv-2017-0037/html
Scroll to top button