• Keine Ergebnisse gefunden

A pointwise differential inequality and second-order regularity for nonlinear elliptic systems

N/A
N/A
Protected

Academic year: 2022

Aktie "A pointwise differential inequality and second-order regularity for nonlinear elliptic systems"

Copied!
50
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

https://doi.org/10.1007/s00208-021-02249-9

Mathematische Annalen

A pointwise differential inequality and second-order regularity for nonlinear elliptic systems

Anna Kh. Balci1·Andrea Cianchi2 ·Lars Diening1·Vladimir Maz’ya3,4

Received: 23 March 2021 / Revised: 22 July 2021 / Accepted: 25 July 2021

© The Author(s) 2021

Abstract

A sharp pointwise differential inequality for vectorial second-order partial differential operators, with Uhlenbeck structure, is offered. As a consequence, optimal second- order regularity properties of solutions to nonlinear elliptic systems in domains inRn are derived. Both local and global estimates are established. Minimal assumptions on the boundary of the domain are required for the latter. In the special case of the p-Laplace system, our conclusions broaden the range of the admissible values of the exponentppreviously known.

Mathematics Subject Classification 35J25·35J60·35B65

Communicated by Y. Giga.

B

Andrea Cianchi andrea.cianchi@unifi.it Anna Kh. Balci

akhripun@math.uni-bielefeld.de Lars Diening

lars.diening@uni-bielefeld.de Vladimir Maz’ya

vladimir.mazya@liu.se

1 Fakultät für Mathematik, University Bielefeld, Universitätsstrasse 25, 33615 Bielefeld, Germany 2 Dipartimento di Matematica e Informatica “U. Dini”, Università di Firenze, Viale Morgagni

67/A, 50134 Firenze, Italy

3 Department of Mathematics, Linköping University, 581 83 Linköping, Sweden

4 Peoples Friendship University of Russia (RUDN University), 6 Miklukho-Maklay St, Moscow 117198, Russian Federation

(2)

1 Introduction

A classical identity, which links the Laplacianuof a vector-valued functionuC3(,RN)to its Hessian∇2u, tells us that

|u|2=div

(u)T∇u−12∇|∇u|2

+ |∇2u|2 in , (1.1) where is an open set in Rn. Here, and in what follows,n ≥ 2, N ≥ 1, and the gradient∇uof a functionu:→RNis regarded as the matrix inRN×nwhose rows are the gradients inR1×nof the componentsu1, . . . ,uN ofu. Moreover, the suffix

“T” stands for transpose.

Identity (1.1) can be found as early as more than one century ago in [11] forn =2

—see also [43,58]. It has applications, for instance, in the second-orderL2-regularity theory for solutions to the Poisson system for the Laplace operator

u=f in , (1.2)

wheref:→RN. Indeed, identity (1.1) enables one to bound the integral of|∇2u|2 over some set inby the integral of|u|2over the same set, plus a boundary integral involving the expression under the divergence operator. Of course, since the equations in the linear system (1.2) are uncoupled, its theory is reduced to that of its single equations.

The second-order regularity theory of nonlinear equations and systems is much less developed, yet for the basic p-Laplace equation or system

div(|∇u|p2∇u)=f in , (1.3) where p>1 and “div” denotes theRN-valued divergence operator. Standard results concern weak differentiability properties of the expression|∇u|p22u. They trace back to [61] for p >2, and to [1,22] for every p >1. The case of a single equation was earlier considered in [62]. Further developments are in [8,20,33].

As demonstrated by several more recent contributions, the regularity of solutions to p-Laplacian type equations and systems is often most neatly described in terms of the expression|∇u|p2∇uappearing under the divergence operator in (1.3). This surfaces, for instance, from BMO and Hölder bounds of [39], potential estimates of [47], rearrangement inequalities of [28], pointwise oscillation estimates of [13], regularity results up to the boundary of [14]. More results in this connection can be found e.g. in [3,30,48].

Differentiability properties of|∇u|p2∇u have customarily been detected under strong regularity assumptions on the right-hand sidef. This is the case of [50], where local solutions are considered. High regularity of the right-hand side is also assumed in [35], where results for boundary value problems can be found under smoothness assumptions on. Both papers [35,50] deal with scalar problems, i.e. with the case when N = 1. Fractional-order regularity of the gradient of solutions to quasilinear equations of p-Laplacian type has been studied in [57], and in the more recent con-

(3)

tributions [3,18,21,55,56]. The question of fractional-order regularity of the quantity

|∇u|p2u, whenN =1 and the right-hand side of Eq. (1.3) is in divergence form, is addressed in [5], where, in particular, sharp results are obtained forn =2.

Optimal second-orderL2-estimates for solutions to a class of problems, including (1.3) for everyp>1, in the scalar case, have recently been established in [31]. Loosely speaking, these estimates tell us that|∇u|p2uW1,2if and only iffL2. Such a property is shown to hold both locally, and, under minimal regularity assumptions on the boundary, also globally. Parallel results are derived in [32] for vectorial problems, namely for N ≥ 2, but for the restricted range of powers p > 32. The results of [31,32] rely upon the idea that, in the nonlinear case, the role of the pointwise identity (1.1) can be performed by a pointwise inequality. The latter amounts to a bound from below for the square of the right-hand side of (1.3) by the square of the derivatives of

|∇u|p2u, plus an expression in divergence form. The restriction for the admissible values of pin the vectorial case stems from this pointwise inequality.

In the present paper we offer an enhanced pointwise inequality in the same spirit, with best possible constant, for a class of nonlinear differential operators of the form

div(a(|∇u|)∇u). The relevant inequality holds under general assumptions on the functiona, which also allow growths that are not necessarily of power type. Impor- tantly, our inequality improves the available results even in the case when the operator is thep-Laplacian, namely whena(t)=tp2. In particular, for this special choice, it entails the existence of a constantc>0 such that

div(|∇u|p2∇u)2

≥div

|∇u|2(p2)

(u)Tu12∇|∇u|2

+c|∇u|2(p2)|∇2u|2 (1.4) in{∇u=0}if and only eitherN =1 and p >1, orN ≥2 and p >2(2−√

2)≈ 1.1715.

The differential inequality to be presented, in its general version, is the crucial point of departure in our proof of the local and globalW1,2-regularity for the expression a(|∇u|)∇ufor systems of the form

div(a(|∇u|)∇u)=f in . (1.5) Regularity issues for equations and systems driven by non standard nonlinearities, encompassing (1.5), are nowadays the subject of a rich literature. A non exhaustive sample of contributions along this direction of research includes [2,4,6,7,16,19,23,26, 27,36,39,40,44,45,49,51,60].

Let us incidentally note that system (1.5) is the Euler equation of the functional J(u)=

B(|∇u|)f·ud x. (1.6)

Here, the dot “·” stands for scalar product, andB: [0,∞)→ [0,∞)is the function defined as

B(t)= t

0

b(s)ds for t ≥0, (1.7)

(4)

where the functionb: [0,∞)→ [0,∞)is given by

b(t)=a(t)t for t>0, (1.8)

andb(0)=0.

Under the assumptions to be imposed ona, the function B and the functional J turn out to be strictly convex. In particular, ifa(t)=tp2, thenB(t)= 1ptp, andJ agrees with the usual energy functional associated with thep-Laplace system (1.3).

We shall focus on the case when N ≥ 2, the case of equations being already fully covered by the results of [31]. In particular, our regularity results apply to the

p-Laplacian system (1.3) for every p>2(2−√

2)≈1.1715. (1.9)

Hence, we extend the range of the admissible exponents pknown until now, which was limited top> 32.

In the light of the pointwise inequality (1.4), the lower bound (1.9) forpis optimal for our approach to the second-order regularity of solutions to thep-Laplace system (1.3). The question of whether such a restriction is really indispensable for this regu- larity, or it can be dropped as in the case whenN =1, where everyp>1 is admitted, is an open challenging problem.

2 Main results

The statement of the general differential inequality requires a few notations. Given a positive functionaC1((0,∞)), we define the indices

ia=inf

t>0

ta(t)

a(t) and sa=sup

t>0

ta(t)

a(t) , (2.1)

whereastands for the derivative ofa. Plainly, ifa(t)=tp2, thenia=sa=p−2.

Moreover, we denote, for N ≥ 1, the continuously increasing function κN : [1,∞)→Ras

κ1(p)=

(p−1)2 if p∈ [1,2)

1 if p∈ [2,∞), (2.2)

ifN =1, and

κN(p)=

⎧⎪

⎪⎩

1−18(4p)2 if p∈ [1,43) (p−1)2 if p∈ [43,2) 1 if p∈ [2,∞),

(2.3)

ifN ≥2.

Theorem 2.1 (General pointwise inequality).Let n≥2and N ≥1. Letbe an open set inRnand letuC3(,RN).

(5)

(i) [Nonsingular case]Assume that the function aC0([0,∞))is such that:

a(t) >0 for t >0, (2.4)

ia≥ −1, (2.5)

and

bC1([0,∞)), (2.6)

where b is the function defined by(1.8). Then div

a(|∇u|)∇u2

≥div

a(|∇u|)2

(u)T∇u−12∇|∇u|2

+κN(ia+2)a(|∇u|)2|∇2u|2 (2.7) in, whereκNis defined as in(2.2)–(2.3). Moreover, the constantκN(ia+2)is sharp.

(ii) [General case]If a is just defined in(0,∞), a∈ C1((0,∞)), and conditions (2.4)and(2.5)are fulfilled, then inequality(2.7)continues to hold in the set{∇u=0}.

Remark 2.2 Observe that the assumption (2.6) need not be fulfilled by the functionsa appearing in the elliptic systems (1.5) to be considered. Such an assumpton fails, for instance, whena(t)=tp2with 1<p<2. This calls for a regularization argument forain our applications of inequality (2.7) to the solutions to the systems in question.

The solutions to the regularized systems will also enjoy the smoothness properties required on the functionuin Part (i) of Theorem2.1. On the other hand, the functions ain the original systems do satisfy the conditions required in Part (ii) of Theorem2.1 for the validity of inequality (2.7) outside the set{∇u=0}of critical points of the functionu.

In the special case whena(t)=tp2, Theorem2.1yields the following inequality for the p-Laplace operator we alluded to in Sect.1.

Corollary 2.3 (Pointwise inequality for the p-Laplacian).Let n ≥2and N ≥ 1. Let be an open set inRnand letuC3(,RN). Assume that p≥1. Then

div(|∇u|p2∇u)2

≥div

|∇u|2(p2)

(u)Tu12∇|∇u|2

+κN(p)|∇u|2(p2)|∇2u|2 (2.8) in{∇u=0}. Moreover, the constantκN(p)is sharp.

Notice that, ifN =1, then

κ1(p) >0 if p>1, (2.9) whereas, ifN ≥2,

κN(p) >0 if p>2(2−√

2). (2.10)

(6)

The gap between (2.9) and (2.10) is responsible for the different implications of inequality (2.7) in view of second-orderL2-estimates for solutions to

div(a(|∇u|)∇u)=f in , (2.11) according to whether N = 1 or N ≥ 2. Indeed, inequality (2.7) is of use for this purpose only ifκN(ia+2) >0.

Since we are concerned withL2-estimates, the datumfin (2.11) is assumed to be merely square integrable. Solutions in a suitably generalized sense have thus to be considered. For instance, the existence of standard weak solutions to the p-Laplace system (1.3) is only guaranteed if pn2n+2. In the scalar case, various definitions of solutions—entropy solutions, renormalized solutions, SOLA—that allow for right- hand sides that are just integrable functions, or even finite measures, are available in the literature, and turn out to be a posteriori equivalent. Note that these solutions need not be even weakly differentiable. The case of systems is more delicate and has been less investigated. A notion of solution, which is well tailored for our purposes and will be adopted, is patterned on the approach of [41]. Loosely speaking, the solutions in question are only approximately differentiable, and are pointwise limits of solutions to approximating problems with smooth right-hand sides.

The outline of the derivation of the second-orderL2-bounds for these solutions to system (2.11) via Theorem2.1is analogous to the one of [31]. However, new technical obstacles have to be faced, due to the non-polynomial growth of the coefficientain the differential operator. In particular, anL1-estimate, of independent interest, for the expressiona(|∇u|)∇ufor merely integrable datafis established. Such an estimate is already available in the literature for equations, but seems to be new for systems, and its proof requires an ad hoc Sobolev type inequality in Orlicz spaces.

Our local estimate for system (2.11) reads as follows. In the statement,BRandB2R

denote concentric balls, with radiusRand 2R, respectively.

Theorem 2.4 (Local estimates).Let be an open set in Rn, with n ≥ 2, and let N ≥2. Assume that the function a:(0,∞)→(0,∞)is continuously differentiable, and satisfies

ia>2(1−√

2) , (2.12)

and

sa<. (2.13)

LetfL2loc(,RN)and letu be an approximable local solution to system(2.11).

Then

a(|∇u|)∇uWloc1,2(,RN×n), (2.14) and there exists a constant C =C(n,N,ia,sa)such that

R1a(|∇u|)∇u

L2(BR,RN×n)+ ∇

a(|∇u|)∇u

L2(BR,Rn×n)

C

fL2(B2R,RN)+Rn21a(|∇u|)∇uL1(B2R,RN×n)

(2.15) for any ball B2R ⊂⊂.

(7)

Remark 2.5 In particular, if=Rnand, for instance,a(|∇u|)∇uL1(Rn,RN×n), then passing to the limit in inequality (2.15) asR→ ∞tells us that

a(|∇u|)∇u

L2(Rn,Rn×n)CfL2(Rn,RN). (2.16) We next deal with global estimates for solutions to system (2.11), subject to Dirichlet homogeneous boundary conditions. Namely, we consider solutions to problems of the

form

−div(a(|∇u|)∇u)=f in

u=0 on ∂ . (2.17)

As shown by classical counterexamples, yet in the linear case, global estimates involving second-order derivatives of solutions can only hold under suitable regu- larity assumptions on∂. Specifically, information on the (weak) curvatures of∂ is relevant in this connection. Convexity of the domain, which results in a posi- tive semidefinite second fundamental form of∂, is well known to ensure bounds in W2,2(,RN×n)for the solutionuto the homogeneous Dirichlet problem associated with the linear system (1.2) in terms of the L2(,RN)norm of f – see [43]. The following result provides us with an analogue for problem (2.17), for the same class of nonlinearitiesaas in Theorem2.4.

Theorem 2.6 (Global estimates in convex domains)Let be any bounded convex open set inRn, with n≥ 2, and let N ≥ 2. Assume that the function a :(0,∞)→ (0,∞)is continuously differentiable and fulfills conditions (2.12) and (2.13). Let fL2(,RN)and let u be an approximable solution to the Dirichlet problem (2.17). Then

a(|∇u|)∇uW1,2(,RN×n), (2.18) and

C1fL2(,RN)≤ a(|∇u|)∇uW1,2(,RN×n)C2fL2(,RN) (2.19) for some positive constants C1=C1(n,N,ia,sa)and C2=C2(N,ia,sa, ).

The global assumption on the signature of the second fundamental form of entailed by the convexity ofcan be replaced by local conditions on the relevant fundamental form. This is the subject of Theorem2.7.

The finest assumption on∂, that we are able to allow for, amounts to a decay estimate of the integral of its weak curvatures over subsets ofwhose diameter approaches zero, in terms of their capacity. Specifically, suppose thatis a bounded Lipschitz domain such thatW2,1. This means that the domainis locally the subgraph of a Lipschitz continuous function of(n−1)variables, which is also twice weakly differentiable. Denote byBthe weak second fundamental form on∂, by|B|

its norm, and set

K(r)= sup

E⊂∂∩Br(x) x∈∂

E|B|dHn1

capB1(x)(E) for r(0,1) . (2.20)

(8)

Here,Br(x)stands for the ball centered atx, with radiusr, the notation capB1(x)(E) is adopted for the capacity of the set E relative to the ball B1(x), andHn1is the (n−1)-dimensional Hausdorff measure. The decay we hinted to above consists in a smallness condition on the limit at asr →0+of the functionK(r). The smallness depends onthrough its diameterdand its Lipschitz characteristicL. The latter quantity is defined as the maximum among the Lipschitz constants of the functions that locally describe the intersection ofwith balls centered on, and the reciprocals of their radii. Here, and in similar occurrences in what follows, the dependence of a constant ondandLis understood just via an upper bound for them.

Theorem2.7also provides us with an ensuing alternate assumption on∂, which only depends on integrability properties of the weak curvatures of∂. Precisely, it requires the membership of|B|in a suitable function spaceX(∂)overdefined in terms of weak type norms, and a smallness condition on the decay of these norms of|B|over balls centered on∂. This membership will be denoted by∂W2X.

The relevant weak space is defined as X(∂)=

Ln1,∞(∂) if n≥3,

L1,∞logL(∂) if n =2. (2.21)

Here,Ln1,∞(∂)denotes the weak-Ln1(∂)space, andL1,∞logL(∂)denotes the weak-LlogL(∂)space (also called Marcinkiewicz spaces), with respect to the (n−1)-dimensional Hausdorff measure.

Theorem 2.7 (Global estimates under minimal boundary regularity). Let be a bounded Lipschitz domain inRn, n≥2, such thatW2,1, and let N ≥2. Assume that the function a :(0,∞)→(0,∞)is continuously differentiable and fulfills con- ditions(2.12)and(2.13). LetfL2(,RN)and letube an approximable solution to the Dirichlet problem(2.17). Then there exists a constant c=c(n,N,ia,sa,L,d) such that, if

lim

r0+K(r) <c, (2.22) then a(|∇u|)∇uW1,2(,RN×n), and inequality(2.19)holds.

In particular, if∂W2X , where X(∂)is the space defined by(2.21), then there exists a constant c=c(n,N,ia,sa,L,d)such that, if

rlim0+

sup

x∈∂BX(∂∩Br(x))

<c, (2.23)

then a(|∇u|)∇uW1,2(,RN×n), and inequality(2.19)holds.

Remark 2.8 We emphasize that the assumptions onin Theorem2.7are essentially sharp. For instance, the mere finiteness of the limit in (2.22) is not sufficient for the conclusion to hold. As shown in [53,54], there exists a one-parameter family of domainssuch thatK(r) <∞forr(0,1)and the solution to the homogeneous Dirichlet problem for (1.2), with a smooth right-hand sidef, belongs toW2,2(,RN)

(9)

only for those values of the parameter which make the limit in (2.22) smaller than a critical (explicit) value.

A similar phenomenon occurs in connection with assumption (2.23). An example from [46] applies to demonstrate its optimality yet for the scalarp-Laplace equation.

Actually, open sets⊂R3, withW2L2,∞, are displayed where the solutionu to the homogeneous Dirichlet problem for (1.3), withN =1,p(32,2]and a smooth right-hand sidef, is such that|∇u|p2∇u ∈/ W1,2(,Rn). This lack of regularity is due to the fact that the limit in (2.23), though finite, is not small enough. Similarly, ifn = 2 there exist open sets, with∂W2L1,∞logL, for which the limit in (2.23) exceeds some threshold, and where the solution to the homogeneous Dirichlet problem for (1.2), with a smooth right-hand side, does not belong toW2,2()—see [53].

Remark 2.9 The one-parameter family of domainsmentioned in the first part of Remark2.8with regard to condition (2.22) is such that∂ /W2Ln1,∞ifn ≥ 3.

Hence, assumption (2.23) is not fulfilled even for those values of the parameter which render (2.22) true. This shows that the latter assumption is indeed weaker than (2.23).

Remark 2.10 Condition (2.23) certainly holds if n ≥ 3 and W2,n1, and if n =2 andW2LlogL(and hence, ifW2,qfor someq >1). This is due to the fact that, under these assumptions, the limit in (2.23) vanishes. In particular, assumption (2.23) is satisfied ifC2.

3 The pointwise inequality

This section is devoted to the proof of Theorem2.1, which is split in several lemmas.

The point of departure is a pointwise identity, of possible independent use, stated in Lemma3.1.

Given a positive functionaC1(0,∞), we define the functionQa : [0,∞)→R as

Qa(t)=ta(t)

a(t) for t>0. (3.1)

Hence,

ia=inf

t>0Qa(t) and sa=sup

t>0

Qa(t), (3.2)

whereiaandsaare the indices given by (2.1).

Lemma 3.1 Let n, N ,andube as in Theorem2.1.

(i)Assume that the function aC0([0,∞))and satisfies conditions(2.4)–(2.6).

Then

(10)

div(a(|∇u|)∇u)2=div

a(|∇u|)2

(u)T∇u−12∇|∇u|2 +a(|∇u|)2

|∇2u|2+2Qa(|∇u|)|∇|∇u||2 +Qa(|∇u|)2

∇u

|∇u|(∇|∇u|)T 2

in , (3.3)

where the last two addends in square brackets on the right-hand side of equation(3.3) have to interpreted as0ifu=0.

(ii)If a is just defined in(0,∞), aC1((0,∞)), and conditions(2.4)and(2.5) are fulfilled, then equation(3.3)continues to hold in the set{∇u=0}.

The next corollary follows from Lemma3.1. applied witha(t)=tp2. Corollary 3.2 Let n, N ,andube as in Theorem2.1. Assume that p≥1. Then

div(|∇u|p2u)2=div

|∇u|2(p2)

(u)T∇u−12∇|∇u|2 + |∇u|2(p2)

|∇2u|2+2(p−2)|∇|∇u||2 +(p−2)2

u

|∇u|(∇|∇u|)T 2

(3.4) in{∇u=0}.

Proof of Lemma3.1 Part (i). The following chain can be deduced via straightforward computations:

div

a(|∇u|)∇u2

=a(|∇u|)u+a(|∇u|)∇u(∇|∇u|)T2

= a(|∇u|)2

|u|2− |∇2u|2

+a(|∇u|)2|∇2u|2+

+a(|∇u|)2|∇u(∇|∇u|)T2+2a(|∇u|)a(|∇u|)u· ∇u(∇|∇u|)T

=a(|∇u|)2

div((u)Tu)12div(∇|∇u|2)

+a(|∇u|)2|∇2u|2+

+a(|∇u|)2∇u(∇|∇u|)T2+2a(|∇u|)a(|∇u|)u· ∇u(∇|∇u|)T. (3.5) Notice that equation (3.5) also holds at the points where|∇u| =0, provided that the terms involving the factora(|∇u|)are intepreted as 0. This is due to the fact that all the terms in question also contain the factor∇uand, by assumption (2.6),

tlim0+a(t)t =0.

(11)

Moreover,

a(|∇u|)2div((u)Tu)

=div

a(|∇u|)2(u)T∇u

−2a(|∇u|)a(|∇u|)u· ∇u(∇|∇u|)T, (3.6) and

1

2a(|∇u|)2div

∇|∇u|2

= 12div

a(|∇u|)2∇|∇u|2

−2a(|∇u|)a(|∇u|)|∇u||∇|∇u||2. (3.7) From equations (3.5)–(3.7) one deduces that

div(a(|∇u|)∇u)2=div

a(|∇u|)2(u)T∇u

12div

a(|∇u|)2∇|∇u|2 +a(|∇u|)2|∇2u|2+a(|∇u|)2u(∇|∇u|)T2

+2a(|∇u|)a(|∇u|)|∇u||∇|∇u||2. (3.8) If∇u = 0, then the last two addends on the right-hand side of Eq. (3.8) vanish.

Hence, Eq. (3.3) follows. Assume next that∇u=0. Then, from Eq. (3.8) we obtain that

div(a(|∇u|)∇u)2=div

a(|∇u|)2(u)Tu

12div

a(|∇u|)2∇|∇u|2 +a(|∇u|)2

|∇2u|2+

a(|∇u|)|∇u|

a(|∇u|) 2

∇u

|∇u|(∇|∇u|)T 2 +2a(|∇u|)|∇u|

a(|∇u|) |∇|∇u||2

. The proof of Eq. (3.3) is complete.

Part (ii). The conclusion follows from the above computations, on disregarding the

comments on the points where∇u=0.

Having identity (3.3) at our disposal, the point is now to derive a sharp lower bound for the second addend on its right-hand side. This will be accomplished via Lemma3.6 below. Its proof requires a delicate analysis of the quadratic form, depending on the entries of the Hessian matrix∇2u, which appears in square brackets in the expression to be bounded. This analysis relies upon some critical linear-algebraic steps that are presented in the next three lemmas.

In what follows,Rnsym×ndenotes the space of symmetric matrices inRn×n. The dot

“·” is employed to denote scalar product of vectors or matrices, and the symbol “⊗”

for tensor product of vectors. Also,I stands for the identity matrix inRn×n. Lemma 3.3 Letω∈Rnbe such that|ω| =1. Then

|Hω|212|ω·Hω|212|H|2= −12|Hω|2 (3.9) for every H ∈Rnsym×n, where Hω=(Iωω)H(Iωω).

(12)

Proof Let{e1, . . . ,en}denote the canonical basis inRn and let{θ1, . . . , θn}be an orthonormal basis of Rn such thatθ1 = ω. Let Q ∈ Rn×n be the matrix whose columns areθ1, . . . , θn. Hence,ω=Qe1. Next, letR=QTH Q. Clearly,R∈Rnsym×n. Denote byri j the entries ofR. Computations show that

|Hω|212|ω·Hω|212|H|2= |Re1|212|e1·Re1|212|R|2

= n i=1

|ri1|212|r11|212 n i,j=1

|ri j|2

= 12 n

j=1

|r1j|2+12 n i=1

|ri1|212|r11|212 n i,j=1

|ri j|2

= −12

i,j2

|ri j|2

= −12|(I−e1e1)R(Ie1e1)|2

= −12|(I−ωω)H(Iωω)|2.

Hence, Eq. (3.9) follows.

Given a vectorω∈Rn, define the set E(ω)=

: H∈Rnsym×n,|H| ≤1 .

It is easily verified that E(ω)is a convex set inRn for everyω ∈ Rn. Lemma3.4 below tells us that, in fact,E(ω)is an ellipsoid, centered at 0 (which reduces to{0}if ω=0). This assertion will be verified by showing that, for eachω∈Rn, there exists a positive definite matrixW ∈Rnsym×nsuch thatE(ω)agrees with the ellipsoid

F(W)=

x∈Rn: x·W1x≤1

, (3.10)

whereW1stands for the inverse ofW. This is the content of Lemma3.4below. In its proof, we shall make use of the alternative representation

F(W)=

x∈Rn: y·x

y·W y for every y∈Rn

, (3.11)

which follows, for instance, via a maximization argument for the ratio of the two sides of the inequality in (3.11) for each givenx∈Rn.

Also, observe that, as a consequence of Eq. (3.11),

|x| =x·x≤√

x·Wx for every xF(W)\ {0}. (3.12) Here, and in what follows, we adopt the notation

x= x

|x| for x ∈Rn\ {0}.

(13)

Lemma 3.4 Givenω∈Rn, let W(ω)∈Rnsym×nbe defined as W(ω)= 12

|ω|2I+ωω

. (3.13)

Then W(ω)is positive definite, and

E(ω)=F(W(ω)). (3.14)

In particular,

∈ |H|F W(ω)

for every ω∈Rnand H ∈Rnsym×n. (3.15) Proof Equation (3.14) trivially holds ifω=0. Thus, by a scaling argument, it suffices to consider the case when|ω| =1. We begin showing that E(ω)F(W(ω)). One can verify that, since|ω| =1,

W(ω)1=2Iωω. (3.16)

LetH∈Rnsym×nbe such that|H| ≤1. Owing to equation (3.16) and to Lemma3.3, ·W(ω)1=2|Hω|2ω·2≤ |H|2≤1. (3.17) This shows thatF(W(ω)). The inclusionE(ω)F(W(ω))is thus established.

Let us next prove that F(W(ω))E(ω). LetxF(W(ω)). We have to detect a matrix H ∈ Rnsym×n such that|H| ≤ 1 andx = Hω. To this purpose, consider the decompositionx =+, for suitables,t ∈R, whereωωand|ω| =1.

SincexF(W(ω)), one has thatx·W(ω)1x≤1. Furthermore,

x·W(ω)1x=(tω+sω)·(2Iωω)(tω+sω)=2(t2+s2)t2=t2+2s2. Hence,t2+2s2≤1. We claim that the matrix H, defined as

H=ω+s(ωω+ωω), has the desired properties. Indeed,H ∈Rnsym×n,

|H|2=tr(HTH)=t2+2s2≤1, =+=x.

This proves thatxE(ω). The inclusionF(W(ω))E(ω)hence follows.

In view of the statement of the next lemma, we introduce the following notation. Given N vectorsωα ∈RnandNmatricesHα ∈Rnsym×n, withα=1, . . .N, we set

(14)

J = N

α=1

Hαωα

2

, J0= N α=1

ωα· N β=1

Hβωβ

2

, J1= N α=1

|Hα|2. (3.18)

Lemma 3.5 Let N ≥2,0≤δ12andδ+σ ≥1. Assume that the vectorsωα ∈Rn and the matrices Hα ∈Rnsym×n, withα=1, . . .N , satisfy the following constraints:

N α=1

α|2≤1, (3.19)

N α=1

|Hα|2≤1. (3.20)

Then

JδJ0σJ1

0 if δ∈ [0,13], max

0,(δ+8δ1)2σ

if δ(13,12]. (3.21) Proof Givenδandσas in the statement, set

Dδ,σ =JδJ0σJ1.

The quantities J0, J and J1 are 1-homogeneous with respect to the quantity N

j=1|Hj|2. Moreover, inequality (3.21) trivially holds if the latter quantity vanishes.

Thereby, it suffices to prove this inequality under the assumption thatN

j=1|Hj|2=1, namely that

J1=1. (3.22)

On settingζ =N

α=1Hαωα, one has that J= |ζ|2 and J0=

N α=1

α·ζ|2.

Therefore,

J0≤ |ζ|2 N α=1

α|2≤ |ζ|2=J. (3.23)

Owing to Lemma3.4,

Hαωα ∈ |Hα|F(Wα)

forα=1, . . . ,N, whereWα = |ωα|2 12(Id+ωαωα). Thus, by equations (3.15) and (3.12),

(15)

Hαωα·ζ ≤ |Hα|

ζ ·Wαζ = |Hα||ωα|

1

2+12|ωα·ζ|2 (3.24) forα=1, . . . ,N. Since

ζ =·ζ)ζ = N α=1

(Hαωα·ζ )ζ ,

equation (3.24) implies that

|ζ| ≤ N α=1

Hαωα·ζN α=1

|Hα||ωα|

1

2+12|ωα·ζ|2.

Hence,

|ζ|212 N

α=1

|Hα||ωα|

1

2+12|ωα·ζ|2 2

. (3.25)

On settingJ0=N

α=1α·ζ|2, we obtain that J0=

N α=1

α|2|ωα·ζ|2 and J0= |ζ|2J0.

Note thatJ0≤1, inasmuch asJ0J = |ζ|2. Moreover, by equation (3.22), Dδ,σ = JδJ0σ = |ζ|2

1−δJ0

σ. (3.26)

From inequalities (3.25) and (3.26) we deduce that Dδ,σ12

N α=1

|Hα||ωα|

1

2+12|ωα·ζ|2 2

1−δ N α=1

α|2|ωα·ζ|2

σ.

(3.27) Next, define the function withg : [0,1]N× [0,1]N× [0,1]N →Ras

g(h,s,t)= 12N

α=1

hαtα

1+sα2 2

1−δN

α=1

tα2sα2σ

for (h,s,t)∈ [0,1]N× [0,1]N× [0,1]N,

(3.28)

whereh =(h1, . . . ,hN),s = (s1, . . . ,sN)andt =(t1, . . . ,tN). Inequality (3.27) then takes the form

Dδ,σg((|H1|, . . . ,|HN|), (|ω1|, . . . ,|ωN|), (|ω1·ζ|, . . . ,|ωN·ζ|)).

(16)

Our purpose is now to maximize the functiongunder the constraints N

α=1

tα2≤1, N α=1

h2α=1. (3.29)

We claim that the maximum ofgcan only be attained ifN

α=1tα2=1. To verify this claim, it suffices to show that

g(h,s, τt)g(h,s,t)

for every (h,s,t)∈ [0,1]N× [0,1]N× [0,1]Nandτ ∈ [0,1]. (3.30) Plainly,

g(h,s, τt)= 12τ2N

α=1

hαtα

1+sα2 2

1−τ2δN

α=1

tα2sα2σ

for(h,s,t)∈ [0,1]N× [0,1]N× [0,1]Nandτ ∈ [0,1]. Note that

0≤δN

α=1

tα2sα2

δN

α=1

tα2

=δ12. (3.31)

Thus, for each fixed(h,s,t)∈ [0,1]n× [0,1]n× [0,1]n, we have that

g(h,s, τt)=c1τ(1−c2τ)β for τ ∈ [0,1], (3.32) for suitable constants c1 ≥ 0 and 0 ≤ c212, depending on(h,s,t). Since the polynomial on the right-hand side of Eq. (3.32) is increasing forτ ∈ [0,1], inequality (3.30) follows. As a consequence, constraints (3.29) can be equivalently replaced by

N α=1

tα2=1 and N α=1

h2α=1. (3.33)

Let us maximize the function g(h,s,t) with respect to h, under the constraint N

α=1h2α =1. Let(h1, . . . ,hN)be any point where the maximum is attained. Then, there exists a Langrange multiplierλ∈Rsuch that

tα

1+sα2 N

γ=1

hγtγ

1+sγ2

1−δN

γ=1

tγ2sγ2

=2λhα for α=1, . . . ,N.

(3.34)

Referenzen

ÄHNLICHE DOKUMENTE

In Section 2, we define Bessel potential and Besov spaces on euclidean (half) space and manifolds with (boundary and) bounded geometry and collect the relevant results for these

This analog in some sense was suggested by the deep connection of mixed order problems with large parameter to the Luysternik–Vishik theory of boundary layers, developed for

Vector-v alued elliptic and parabolic boundary value problems subject to general boundary con-.. ditions have been investigated recently in [DHP01] in

Herein, we follow an approach based on recent advances in Calculus of Variations in the space L ∞ (see e.g. [26] and references therein) developed for functionals involving higher

Elliptic-parabolic systems, semilinear equations, Newton method, proper orthogonal decomposition, empirical interpolation, error estimates.. The authors gratefully acknowledges

In particular, we consider the realization of the boundary value problem as an unbounded oper- ator with the ground space being a closed subspace of a Sobolev space and give

This paper is concerned with a boundary value problem defined over a bounded region of Euclidean space, and in particular it is devoted to the establishment of a priori estimates

In particular, we consider the realization of the boundary value problem as an unbounded operator with the ground space being a closed subspace of a Sobolev space and give necessary