Skip to content
BY 4.0 license Open Access Published by De Gruyter September 8, 2021

On the uniqueness for weak solutions of steady double-phase fluids

  • Mohamed Abdelwahed , Luigi C. Berselli EMAIL logo and Nejmeddine Chorfi

Abstract

We consider a double-phase non-Newtonian fluid, described by a stress tensor which is the sum of a p-Stokes and a q-Stokes stress tensor, with 1 < p<2 < q<∞. For a wide range of parameters (p, q), we prove the uniqueness of small solutions. We use the p < 2 features to obtain quadratic-type estimates for the stress-tensor, while we use the improved regularity coming from the term with q > 2 to justify calculations for weak solutions. Results are obtained through a careful use of the symmetries of the convective term and are also valid for rather general (even anisotropic) stress-tensors.

MSC 2010: 76A05; 35J62; 35Q30; 35J25; 35J55

1 Introduction

In this paper we study the uniqueness of “small” (in an appropriate sense) solutions to a family of double-phase steady problems, arising in the analysis of non-Newtonian fluids. The interest in double-phase problems started with the celebrated result by Zhikov [29] concerning the Lavrentiev phenomenon. The problem was set in the framework of functionals with (p, q)-growth, for which we refer also to the pioneering paper by Marcellini [22]; recent results can be found in the works of Esposito, Leonetti, and Mingione [15], Baroni, Colombo, and Mingione [3], and Colombo and Mingione [9], especially in the context of regularity of minimizers. For results regarding the applications to spectral analysis and multiplicity of solutions, see Chorfi and Radulescu [8], Baraket, Chebbi, Chorfi, and Radulescu [2], and the review in Radulescu [25]. In the context of fluid mechanics, problems with more than one phase arise especially in the case of fluids with complex rheologies, as introduced by Růžička [23], where the modeling leads to a problem with variable exponent. See also the review in Rădulescu and Repovš [26] for general problems involving partial differential equations with variable exponents.

Here, instead that regularity or multiplicity of solutions, we focus on some specific problems concerning uniqueness in the class of small weak solutions. In order to introduce the problem, we recall that a basic result for the steady Navier-Stokes equations in a smooth and bounded domain Ω⊂ℝ3

(1.1) ν0Δu+(u)u+π=fin Ω,divu=0in Ω,u=0on Ω,

is that of uniqueness of weak solutions, under the assumption of smallness of the L3(Ω)-norm of u. (In this case the possible non-uniqueness is neither a special feature of the problem in ℝ3, nor coming from the fact that it is a nonlinear system with the divergence-free constraint. Possible multiplicity of solutions is a common result even for semi-linear elliptic scalar equations.) More precisely, for the steady Navier-Stokes equations (1.1), it holds that if u1,u2W01,2(Ω) are weak solutions corresponding to the same external force f ∈ W−1,2(Ω), then there exists ϵ0 > 0 such that if ∥u13ϵ0, then u1=u2; see the review in Galdi [17, Ch. IX.2]. The proof is based on writing the system satisfied by the difference U = u1u2 and testing by U itself (a procedure which is legitimate in the steady case also for weak solutions). Next, one uses the inequality

(1.2) Ω(U)Uu1dxU6U2u13CU22u13,

with a constant C depending only on Ω, obtained by application of the Hölder inequality and of the Sobolev embedding H01(Ω)L6(Ω) .

In this way from (1.1) one easily gets, after integration by parts, that

v0U22Cϵ0U22,

which implies U = 0, provided that ϵ0<ν0/C. Observe that from the energy estimate valid for weak solutions one knows a priori just that

u121v0fW1,2,

hence, uniqueness for small forces/large viscosities follows by using a Sobolev embedding to ensure smallness in L3(Ω), at least when Ω is bounded.

We stress the critical role played by the exponent “two” of ∥∇U2 in the energy estimate (1.2). It appears at the same time as exponent in: i) the lower bound for the dissipative term; and ii) the upper bound for the quantity obtained by estimating of the convective term. If there is a mismatch in the powers, then this easy but powerful argument may fail.

The same argument can be also applied if ∥∇u13/2ϵ1 (for a possibly different small constant ϵ1 > 0), by using a similar estimate for the convective term and observing that W1,3/2(Ω)↪ L3(Ω). Note that the two alternative assumptions have the same scaling.

In the case of steady non-Newtonian fluids, the situation becomes more complex. If one considers for 1 < p<∞ the following system, describing a family of shear dependent fluids,

(1.3) v0Δudiv(δ+|Du|)p2Du+(u)u+π=f in Ω,divu=0 in Ω,u=0 on Ω,

then, for any δ≥0 and for all ν0 > 0, the same argument valid for the Navier-Stokes equations can be applied. In fact, the additional stress tensor is monotone and the following inequality holds:

Ωδ+Du1p2Du1δ+Du2p2Du2:Du1u2dx0.

The above inequality and the fact that weak solutions still belong to W01,2(Ω) will be enough to disregard the effect of the additional non-linear stress tensor. It is possible to treat problem (1.3) in the same way as for the unperturbed Navier-Stokes equations, at least for what concerns uniqueness. The argument applies also to fluids such that the stress tensor has, in addition to the linear part, a nonlinear stress tensor S(Du)=Sp, δ(Du) with a (p, δ)-structure (see Section 2). We recall that the stress tensor

(1.4) Sp,δ(Du):=(δ+|Du|)p2Duδ0,1<p<,

is just the prototypical example of a stress tensor with (p, δ)-structure. These results are reviewed for instance in [16, Sec. 2.2.1(c)].

A similar approach has been recently used by Gasiński and Winkert [18] to treat the following anisotropic double-phase scalar problem

(1.5) div|u|p2u+μ(x)|u|q2u=F(x,u,u) in Ω,u=0 on Ω,

in the case p=2<q < 3, with μ≥0 Lipschitz continuous, under appropriate growth conditions on f. Our results improve also those in [18], since we consider a vector valued problem, with the constraint of divergence-free, and with term F(x, u, ∇u)=f(x)−(u·∇) u.

In addition to the results above, we observe that the problem of uniqueness for non-Newtonian fluids becomes more complex when the linear part is missing (that is system (1.3) with ν0=0 and, in the case p > 2, also δ=0), since one cannot take advantage of the classical estimates. In fact, the convective term is still quadratic as in (1.1); this has to be balanced in some way by a non-quadratic term. Moreover, in the context also of electro-rheological fluids uniqueness for small (and smooth) solutions is proved in Crispo and Grisanti [10, 11], at least in the non-degenerate case.

Note that for a stress tensor with (p, δ)-structure (see precise definition in the Assumption 2:1) as in the example (1.4), it is well-known that the following point-wise estimate holds:

(S(A)S(B)):(AB)c1(δ+|A|+|AB|)p2|AB|2,

for all A, B symmetric matrices. For a review see for instance Růžička [24]. Hence, at least if δ>0 and p > 2, it follows that

(S(A)S(B)):(AB)c1δp2|AB|2,

which allows us to employ the same machinery.

On the other hand, the case 1 < p<2 presents some peculiarities as exploited in Blavier and Mikelić [6], which permits to prove uniqueness for small solutions if 95p<2 . (Note that 95 is the critical value to apply the classical monotonicity argument and to use the solution itself as a test function in the weak formulation.) We will review these results and provide more details in Section 2.1, explaining how we improve them.

Here, we will then consider as basic example the following boundary value problem for a double-phase non-Newtonian fluid

(1.6) div|Du|p2Dudiv|Du|q2Du+(u)u+π=f in Ω,divu=0 in Ω,u=0 on Ω,

with 1 < p<2 < q, proving uniqueness of small solutions (recall again that if at least one among p or q is equal to 2 the result is trivial). Our main aim is to consider the degenerate case and to exploit the interplay between the p-growth and the q-growth in double-phase problems, as a way to enforce uniqueness, which cannot be proved in the single-phase cases. In addition, a non-standard estimation of the convective term using its symmetries is also employed.

The main result we prove, see Theorem 3.1, is the uniqueness of small solutions, for 6 5 p < 2 and q=q(p) large enough. Then, some extensions to an anisotropic problem are proved in Theorem 4.2.

2 Notation and preliminary results

In the sequel, Ω⊂ℝ3 will be a smooth and bounded open set. As usual, we write x=(x1, x2, x3)=(x′, x3), for all x ∈ ℝ3. If the boundary ∂Ω is at least of class C0,1, then the normal unit vector n at the boundary is well defined. We recall that a domain is of class Ck,1, if for each point P ∈ ∂Ω, there are local coordinates such that in these coordinates we have P = 0 and ∂Ω is locally described by a Ck,1-function, i.e., there exist RP,  RP(0,),rP(0,1) and a Ck,1-function aP:BRP2(0)BRP1(0) such that

  1. xΩBRP2(0)×BRP1(0)x3=aPx,

  2. ΩP:=xR3:xBRP2(0),aPx<x3<aPx+RPΩ,

  3. aP(0)=0, and xBRP2(0)aPx<rP,

where Brk(0) denotes the k-dimensional open ball with center 0 and radius r > 0.

For our analysis, we will use the customary Lebesgue (Lp(Ω), ∥ . ∥p) and Sobolev spaces (Wk, p(Ω), ∥ . ∥k, p) of integer index k ∈ ℕ, with 1 ≤ p ≤ ∞. As usual, p=pp1 denotes the conjugate exponent. We do not distinguish between scalar and vector valued function spaces, we just use boldface for vectors and tensors. We recall that L0p(Ω) denotes the subspace with zero mean value, while W01,p(Ω) is the closure of smooth and compactly supported functions with respect to the ∥ . ∥1,p norm. We denote by W1,p(Ω):=W01,p(Ω) its dual space, with norm ∥ . ∥−1,p.

If Ω is bounded and if 1 < p<∞, the following two relevant inequalities hold:

  1. the Poincaré inequality

    (2.1) CP(p,Ω)>0:upCPupuW01,p(Ω);
  2. the Korn inequality

    (2.2) CK(p,Ω)>0:upCKDupuW01,p(Ω),

where Du denotes the symmetric part of the matrix of derivatives ∇u.

For 1 ≤ p<3 we have, as a combination of (2.1)-(2.2), also the Sobolev-type inequality

CS>0:upCSDupuW01,p(Ω),

where p:=3p3p .

When working with incompressible fluids it is natural to incorporate the divergence-free constraint directly in the definition of the function spaces. These spaces are built upon completing the space of solenoidal smooth functions with compact support (here denoted by C0,σ(Ω) in an appropriate topology. For 1 < p<∞, we define

Lσp(Ω):=ϕC0,σ(Ω)ϕp,W0,σ1,p(Ω):=ϕC0,σ(Ω)ϕp.

For the nonlinear stress tensors, we make the following assumption of being with (p, δ)-structure, which is a generalization of the example from (1.4).

Assumption 2.1

We assume that S:R3×3RSym3×3 belongs to C0R3×3,Rsym3×3C1R3×3{0},Rsym3×3, satisfies S(P)=SPsym, and S(0)=0. Moreover, we assume that S has (p, δ)-structure, i.e., there exist p ∈ (1,∞), δ ∈ [0,∞), and constants C0, C1 > 0 such that

(2.3a) i,j,k,l=13klSij(P)QijQklC0δ+Psymp2Qsym2,
(2.3b) klSij(P)C1δ+Psymp2,

are satisfied for all P,Q ∈ ℝ3x3 with psym0 and all i, j, k, l = 1, . . . , 3. The constants C0, C1, and p are called the characteristics of S.

Remark 2.1

We would like to emphasize that, if not otherwise stated, the constants depend only on the characteristics of S, but are independent of δ≥0.

Defining for t=0 a special N-function φ by

(2.4) φ(t):=0tφ(s)ds with φ(t):=(δ+t)p2t,

we can replace Ciδ+Psymp2 on the right-hand side of (2.3a) and (2.3b) by C˜iφPsym, for i=0, 1. Next, the shifted functions are defined for t=0 by

φa(t):=0tφa(s)ds with φa(t):=φ(a+t)ta+t.

In the following proposition, we recall several useful results, which will be frequently used in the paper. The proofs of these results and more details can be found in [4, 13, 14, 27]. Many inequalities can be written in a compact form by means of the following tensor valued function

(2.5) F(A):=δ+Asymp22Asym,

and we write f ∼ g if there exist constants c1, c2 > 0 such that c1g ≤ f ≤ c2f.

Proposition 2.1

Let S satisfy Assumption 2:1, let φ be defined in (2.4), and let F be defined in (2.5).

  1. For all P, Q ∈ ℝ3x3

    (S(P)S(Q)):(PQ)|F(P)F(Q)|2,φ|Pssm|PsymQsym,φPsym+QsymPsymQsym2,S(Q):Q|F(Q)|2φQsym,|S(P)S(Q)|φPssmPsymQsym,

    and the constants depend only on the characteristics of S.

  2. For all ϵ>0, there exists a constant cϵ > 0 (depending only on ϵ>0 and on the characteristics of S) such that for all u, v,w ∈ W1,p(Ω)

    (S(Du)S(Du),DwDu)ϵF(Du)F(Du)22+cϵF(Dw)F(Du)22.

Since the range of the allowed p ∈ (1, 2) will play a relevant role, we first recall that the restriction p > 6/5 is quite natural for the problem, at least for what concerns existence of weak solutions. In the case ν0=0, the weak formulation of (1.3) is in fact: find uW0,σ1,p(Ω) such that

ΩSp,δ(Du):Dϕuu:ϕdx=f,ϕϕC0,σ(Ω).

To properly define the quadratic term, one needs (at least) that uLloc2(Ω), which follows, for instance, if u ∈ W1,p(Ω) for p65 . The basic a priori estimate obtained testing with u itself shows that, if f ∈ W−1,p(Ω), then ∥DupC. Next, Korn inequality and the embedding W01,p(Ω)L2(Ω) (which holds for p=6/5 in three space-dimensions) give that uu ∈ L1(Ω). This restriction on p is intrinsic to the problem, due to the growth of the convective term. The limiting case p=6/5 is excluded from the existence theorem due to some technical compactness arguments which are used in the proof.

The following existence result holds true.

Theorem 2.1

Let Ω⊂ℝ3 be smooth and bounded, let ν0=0, and let the stress tensor Sp,δ satisfy Assumption 2.1 for some δ and for some p > 6/5. Then, for any given f∈ W-1,p'(Ω), there exists at least a weak solution vW0,σ1,p(Ω) of problem (1.3), which satisfies the estimate

(2.6) DuppCf1,pp,

for some constant C depending on the domain and on the characteristics of the stress tensor.

The proof of existence of weak solutions requires a precise use of the monotonicity of the operator, but for p<95 the usual Browder-Minty approach (as employed by Lions [21] and Ladyžhenskaya [20] also for values of p > 3) is not directly applicable and one has to resort to more technical arguments with bounded or even Lipschitz truncation, plus an appropriate divergence correction, see the reviews in Breit [7] and Růžička [24].

Here, we do not discuss the existence of weak solutions, which easily follows also for the problem (1.6), by employing similar arguments without any relevant change. Hence, we can state the following theorem.

Theorem 2.2

Let Ω⊂ℝ3 be smooth and bounded, let the stress tensors Sp,δp(Du),Sq,δq(Du) , satisfy Assumption 2:1 for some δp≥0, p > 6/5 and for some δq≥0, q > p. Then, for all f ∈ W−1,q(Ω), the problem

(2.7) divSp,δp(Du)divSq,δq(Du)+(u)u+π=f in Ω,divu=0 in Ω,u=0 on Ω,

has at least a weak solution uW0,σ1,q(Ω), which satisfies

ΩSp,δp(Du):Dϕ+Sq,δq(Du):DΦuu:ϕdx=f,ϕϕC0,σ(Ω),

and which satisfies the following inequality:

(2.8) Dupp+DuqqCf1,qq,

for some constant C depending only on the characteristics of Sq and Ω.

2.1 On problems without linear part: some known uniqueness results

When considering the problem (1.3) in the case ν0=δ=0 and p > 2, the problem of uniqueness (even of small solutions) is completely open, since the (sharp) estimates from Proposition 2.1 seem to be not suitable to “absorb” the convective term. This is determined by a “gap” between the powers in the lower bound for the stress tensors and those in the upper bound for the convective term.

On the other hand, in the case ν0=0 and for some p < 2, the following result of uniqueness for weak solutions is proved in Blavier and Mikelić [6], under appropriate smallness of both weak solutions.

Theorem 2.3

Let uiW0,σ1,p(Ω) be weak solutions of (1.3) in the case ν0=0 and let p ∈ [9/5, 2[. Then, there exists a constant ϵ0 > 0 (which depends on p and on the data of the problem) such that, if ∥∇uipϵ0, then u1=u2.

The available proofs of uniqueness go through considering the equation satisfied by the difference and testing with the difference of the solutions, U≔ u1u2. To this end, one needs to be able to rigorously write the integrals

Ωu1u1u2u2u1u2dx=Ωu1u2u1u1u2dx,

(where the equality derives from integration by parts and from the divergence-free constraint) and to prove proper upper bounds (similar to (1.2), already used in the Newtonian case). By using the information ui W0,σ1,p(Ω), and by the Hölder inequality, we get that if

2p+1p=2p23+1p1p95,

then the following estimate holds:

(2.9) Ω(U)u1UdxUp2u1p.

Hence, by using Sobolev and Korn inequalities, we get

Ω(U)u1UdxC(p,Ω)DUp2u1p.

The uniqueness result from [6] exploits this estimation, together with the lower bound

(2.10) ΩSp,0Du1Sp,0Du2:Du1u2dxc(p)Du1Du2p2Du1p2p+Du2p2p,

valid for all 1 < p<2, for some c(p) > 0, when ui u i W 0 1 , p ( Ω ) .

If the smallness is assumed on one solution, namely, if

u1pϵ0,

then the estimates for the convective term and the stress tensors imply that (for any 95p<2 ,

c(p)DUp22maxi=1,2Duip2pCϵ0DUp2.

From this inequality, by using (2.6), one gets

(2.11) cp,f1,pDUp2Cϵ0DUp2,

and uniqueness will follow if ϵ0 > 0 is small enough, that is, if ϵ0<cp,f1,pC.

3 On the double-phase problem

In this section, we consider the problem (2.7) with δp=δq=0 and we observe that–on one hand–the “good” estimates (lower bound) for the stress tensor of the difference are valid if 1 < p<2, hence one would like to have one phase with this properties; On the other hand, to handle the convective term larger (than 2) values of the exponent q will provide the estimates needed to rigorously define the integrals involving the convective term.

First, we have the existence Theorem 2.2 for weak solutions. Next, we observe that the same argument as in [6] can be directly adapted to the problem (2.7) (of which system (1.6) represents a particular case), to produce the following elementary–but original–uniqueness result.

Proposition 3.1

Let the same assumption as in Theorem 2.2 be satisfied. Let be given 65<p<95 and let f ∈ W−1,q(Ω), for some q > 2 such that

(3.1) q3p5p6.

Let u1,u2W0,σ1,q(Ω) be weak solutions to (1.6) corresponding to the same f. Then, there exists ϵ0=ϵ0(q, Ω, ∥f−1,q) such that if

(3.2) u1qϵ0,,

then u1=u2.

Remark 3.1

In the range 65<p<95 the inequality 3p5p6>95 holds and, to have results which are not included in Theorem 2.3, we need to require q > 2; next, when p>65 , the following inequality holds:

3p5p62forp1271.714

In the light of the above observations, Proposition 3.1 can be restated as follows: The uniqueness of weak solutions can be proven under condition (3.2) for some q such that

q 3 p 5 p 6 i n t h e c a s e 6 5 < p < 12 7 , q > 2 i n t h e c a s e 12 7 p < 9 5 .

Proof of Proposition 3.1. The proof is a simple adaption of results in [16]. Taking the difference of two solutions ui W0,σ1,q(Ω) corresponding to the same force f ∈ W−1,q(Ω), one gets for the difference U the following estimates:

c(p)DUp2Du1p2p+Du2p2pΩ(U)u1UdxUp2u1qCDUp2u1q,

obtained by using Hölder, Korn, and Sobolev inequalities, and also (2.10). The calculations are justified if

2p+1q1,

which follows if q is as in (3.1). Then, if one assumes the smallness of the Lq-norm of the gradient or u1, the following estimate holds:

c(p)DUp2Du1p2p+Du2p2pCϵ0DUp2,

which allows again to conclude uniqueness (exactly as in the previous case) by the same argument developed in [6].□

In order to improve this result, here we follow a slightly different approach which is based on more precise point-wise estimates for the stress tensor corresponding to the “p-phase.” This will allow us to require less restrictive conditions on q. The main result we prove is the following one:

Theorem 3.1

Let δp≥0 and δq=0. Let p ∈ ]6/5, 12/7[ and q > 2 such that

(3.3)  (i) q>3p(2p)5p6for65<p3332, (ii) q>3p(4p)7p6for3332<p3713, (iii) q>2for3713<p<127. 

Let u1, u2 be weak solutions of (2.7) corresponding to the same f ∈ W−1,q(Ω). Then, there exists a constant ϵ0=ϵ0(p, q, Ω, ∥f−1,q)>0 such that if at least one solution satisfies

u1qϵ0,

then u1=u2.

Before proving the theorem, we observe that the significant case is when the “q-phase” is such that δq=0. Also the possible degeneracy δp=0 of the p-phase improves previously known results. Moreover, we think that an interesting further development would be the study of the uniqueness for a single-phase fluid by using the results of “higher regularity” from Crispo and Maremonti [12]. This will hold provided that: a) the results from [12] can be adapted to the non-modified p-Stokes system; and b) an explicit estimation of some constants related with singular integrals is available.

We make some remarks to explain the improvements in the various ranges, with respect to the results in Proposition 3.1. A first significant difference comes into play if q is smaller or larger than the space dimension, since this will imply estimates involving the Sobolev exponent q=3q3q or, alternatively, estimates in L(Ω).

Remark 3.2

Concerning the case (i) in (3.3)

65<p33321.2<p1.37228,

we can see that the restriction on the range of p is made to enforce q > 3, since the calculations leading to this estimate are valid for q > 3 (cf. the proof of Lemma 3.2).

Moreover, we also have that

3p(2p)5p6<3p5p6withintherange65<p3332,

hence the condition (i) is less restrictive than (3.1), coming from Proposition 3.1 (in the same range of p).

Remark 3.3

Concerning the case (ii) in (3.3)

3332<p37131.37228<p1.69425,

the restrictions are made to impose that 23p(4p)7p6<3 , hence to make possible the choice of some 2 < q<3. The lower bound on q is requested to have a non-trivial result, and the upper bound is requested to use Lemma 3.2.

Moreover, we have again

3p(4p)7p6<3p5p6,withintherange3332<p3713,

and this condition is less restrictive than that from Proposition 3.1 (in the same range of p).

Remark 3.4

The restriction p < 12/7 is already present in Proposition 3.1 since beyond that value the result is included in Proposition 3.1. Again, the condition (iii) is less restrictive than the one previously obtained, since the following inequality holds:

3p5p6>2,withintherange3713<p<127.

The proof of Theorem 3.1 is heavily based on the following inequality coming from an application of Proposition 2.1:

ΩSp,δpDu1Sp,δpDu2:DUdxcΩδp+Du1+|DU|p2|DU|2dx,

and then using the following two lemmas about the convective term.

The estimation of the tri-linear term we use is not the direct one as in (2.9). We bound the convective term exploiting some of its symmetries. This fact is relevant since in the stress tensors present in the equations (1.6) the velocity enters only through the symmetric gradient. We first have a result which follows by integrating by parts:

Lemma 3.1

The following equalities hold:

Ω(u)u1udx=ΩuDu1udx=2ΩuDuu1dxu,u1C0,σ(Ω),

where

vAw:=ij=13viAijwjv,wR3,AR3×3.

Proof. The proof of the first equality is simply obtained by interchanging the dummy variable in the double summation

(u)u1u=i,j=13uiiu1juj,

and using that Du1 is, by definition, a symmetric tensor.

The second equality follows by integrating by parts: due to the vanishing of the boundary trace of u we obtain

Ω(u)u1udx=Ω(u)uu1dx.

Then adding the null term

Ωu[u]Tu1dx=Ωujiuju1idx=12Ωi|u|2u1idx=0,

the second identity follows.

The same result clearly holds also for u,u1 in spaces in which C0,σ(Ω) is dense, provided that the integrals are well-defined (as an example Lemma 3.1 is valid for u,u1W0,σ1,q(Ω) , with q=9/5

The nonlinear term is now estimated with the following inequalities:

Lemma 3.2

LetU, u1W0,σ1,q(Ω) for some q95 and let δ>0 be any positive number. If q< 3, (hence if W1,q(Ω)Lq(Ω)forq=3q3q<, then the following inequality holds:

Ω(U)u1Udx2Upδ+Du1+|DU|p22DU2u1qδ+Du1+|DU|2p22q2p,

with

q3p(4p)7p6;

Observe that 3p(4p)7p6<3,for3332p<2 .

If q > 3, then W1,q(Ω) ⊂ L(Ω) and the following estimate holds:

Ω(U)u1Udx2Upδ+Du1+|DU|p22DU2u1δ+Du1+|DU|2p22q2p,

with

q3p(2p)5p6;

observe that 3p(2p)5p6>3, for 65<p<3332.

Finally, if q=3, then W1,q(Ω)⊂ Ls(Ω) for all s ∞ , and the following estimate holds:

Ω(U)u1Udx2Upδ+Du1+|DU|p22DU2u1sδ+Du1+|DU|2p22q2p,

with

q>3p(2p)5p6.

Proof. We observe that the integral Ω ( U ) u 1 U d x is finite by (2.9) since all functions belong to W0,σ1,q(Ω), hence all calculations we perform are completely justified.

We start from the case q < 3 and we observe that the estimate involves a “natural quantity” related to a stress tensor with (p,δ) -structure. We use directly Lemma 3.1, and since δ>0, we can freely multiply and divide by δ+Du1+|DU|2p20 . We use the Hölder inequality which is valid if

1p+12+1q+2p2q1,

and Sobolev inequalities to get

Ω(U)u1Udx=2ΩUDUu1dx=2ΩUδ+Du1+|DU|p22DUu1δ+Du1+|DU|2p2dx2Upδ+Du1+|DU|p22DU2u1qδ+Du1+|DU|2p22q2p.

The proof in the case ? q > 3 follows more or less the same lines and if the following inequality is satisfied

1p+12+2p2q1,

we can write

Ω(U)u1Udx=2ΩUDUu1dx=2ΩUδ+Du1+|DU|p22DUu1δ+Du1+|DU|2p2dx2Upδ+Du1+|DU|p22DU2u1δ+Du1+|DU|2p22q2p.

We can now give the proof of the main result of this paper.

Proof of Theorem 3.1. We write the difference between the two weak solutions of (2.7) and we use U=u1u2 W0,σ1,q(Ω) as test function to get

ΩSp,δpDu1Sp,δpDu2:Du1u2dx+ΩSq,0Du1Sq,0Du2:Du1u2dx=Ω(U)u1Udx.

Since the operator sq,0() is monotone, we get

(3.4) ΩSp,δpDu1Sp,δpDu2:DUdxΩ(U)u1Udx,

and the right-hand side is finite due to the fact that q > 2

We now estimate the left-hand side by Proposition 2.1 and, for ui at least in W0,σ1,p(Ω) , the following inequality holds:

Ωδp+Du1+|DU|p2|DU|2dxCΩSp,δpDu1Sp,δpDu2:DUdx.

Next, the last term from the right-hand side of the inequalities proved in Lemma 3.2 can be estimated for δ=δp by observing that since

forα(0,1)(x+y)αxα+yαx,y0,

by the Minkowski inequality the following holds:

δp+Du1+|DU|2p22q2pδp2p22q2p+Du12p22q2p+|DU|2p22q2pδp2p2|Ω|2p2q+Du1q2p2+DUq2p2δp2p2|Ω|2p2q+2Du1q2p2+Du2q2p2,

and, if δp0,δ0 for some δ0>0 , then

δ p + D u 1 + | D U | 2 p 2 2 q 2 p C δ 0 , p , q , | Ω | 1 + max i = 1 , 2 D u i q 2 p 2 C δ 0 , p , q , | Ω | 1 + f 1 , q 2 p 1 2 ( q 1 ) .

Hence, in the case of the problem with a non-degenerate stress tensor Sp,δp , that is if δp>0 , by collecting the lower bound for the left-hand side of (3.4) and the upper bound for the right-hand side (using Lemma (3.2) with δ=δp ) and simplifying similar terms, we get

Ωδp+Du1+|DU|p2|DU|2dx1/22CDupu1qf1,q2p2(1).

We use now the hypothesis u1qϵ0 and, after squaring both sides,w we get

Ωδp+Du1+|DU|p2|DU|2dxCϵ02Dup2,

where C depends on the data of the problem, since we used the a priori estimate in W 0 1 , q ( Ω ) for both solutions. The proof follows now as in the previous case. In fact, by using (2.10) and (2.11), we get

(3.5) cp,f1,qDUp2c(p)DUp2DUp2p+Du2p2pCϵ02DUp2,

implying uniqueness, provided that ϵ0 is small enough.

In the degenerate case δp = 0, we observe that from the (p, 0)-structure we get

ΩDu1+|DU|p2|DU|2dxCΩ(U)u1UdxC1,

hence the integral from the left-hand side is well defined. Then, we show that the inequalities from Lemma 3.2 are valid also with δ = 0. This can be seen by writing the estimates in Lemma 3.2 for some 0δ ≤ 1 and taking the inferior limit of the right-hand side as δ → 0. In fact, since p < 2 we have the following monotonic increasing convergence

δ+Du1+|DU|p2|DU|2δ0Du1+|DU|p2|DU|2 a.e in Ω,

which implies convergence of the corresponding integrals. This shows convergence of the second term from the right-hand side of estimates from Lemma 3.2.

Moreover, since p < 2, we have the uniform bound

δ+Du1+|DU|2p21+Du1+|DU|2p2L2q2p(Ω),

which allows us to use the Lebesgue dominated convergence, to handle the last term in the right-hand side of estimates from Lemma 3.2. After having justified the limiting step δ → 0, one obtains again (3.5) and the proof proceeds as in the non-degenerate case.

4 Some remarks for anisotropic (weighted) double-phase problems

In the analysis of double-phase problems, it is also interesting to consider, as done in [9, 18, 25], families of anisotropic double-phase problems. We also adapt our uniqueness results to the following problem

div|Du|p2Dudivμ(x)|Du|r2Du+(u)u+π=f in Ω,divu=0 in Ω,u=0 on Ω,

where μ is a non-negative regular function and r > 2.

Stress-tensors of this type occur in fluid mechanics in certain turbulence models introduced by Baldwin and Lomax [1]. The mathematical analysis in terms of existence of weak solutions has been recently developed by one of the authors and D. Breit in [5]. In this case, it makes sense (from the modeling point of view based on Prandtl mixing length), to use as function μ the distance of x from the boundary of Ω

μ(x)=d(x):=d(x,Ω),

or one of its powers. The original model of Baldwin and Lomax concerns the tensor BL(x,ω)=d2(x)|ω|ω and it enters in the momentum equations through its curl

(4.1) curlBL(x,ω)=curld2(x)|ω|ω with ω=curlv,

being written in rotational form. The system resembles the p-curl system studied in certain problems in electromagnetism, especially in mathematical model for superconductors, see Yin [28]. The value 2 for the power of the distance function is critical in terms of being able to recover estimates on the full gradient from those valid for the weighted curl, by means of the theory of weighted Sobolev spaces and Muckenhoupt weights. The expression (4.1) is the rotational form of the stress tensor

divd2(x)|Du(x)|Du(x),

which is a weighted version of a stress tensor with “(3, 0)-structure,” within the notation of Assumption 2.1.

In this section, we will see how to generalize the uniqueness results when Sq,δ(Du), , the stress tensor in (1.6), is replaced by

divdα(x)|Du(x)|r2Du(x) for some r>2, with 0<α<r1.

To handle this term we first recall a well-known lemma about the distance function d(x) , see for instance Kufner [19].

Lemma 4.1

Let Ω be a domain of class C0,1. There exist constants 0 < c0, c1 ∈ ℝ such that

c0d(x)axx3c1d(x)x=x,x3ΩP.

From the above result we get the following weighted estimate

Lemma 4.2

Let f be measurable such that Ωdα|f|rdx< . Then, for 0 < α < r-1 it follows that fLr1+α(Ω) .

Proof. The proof follows by Hölder inequality and Lemma 4.1. In fact we can write

Ω | f | s   d x = Ω d α s / r d α s / r | f | s   d x Ω d a s r s   d x ( r s ) / r Ω d α | f | r   d x s / r c Ω d α | f | r   d x s / r ,

where the last estimate on the integral of the distance function follows immediately from Lemma 4.1, if αsrs< 1.

We can now state the following existence theorem:

Theorem 4.1

Let Ω⊂ℝ3 be smooth and bounded and let Sp,δp satisfy assumption 2.1 for some δp0 and for some p > 6/5. Let r > 2 be given and α ≥ 0. Then, for all fW-1,p'(Ω), the problem

(4.2) divSp,δp(Du)divdα(x)|Du|r2Du+(u)u+π=f in Ω,divu=0 in Ω,u=0 on Ω,

has at least a weak solution uW0,σ1,p(Ω) , such that

Dupp+Ωdα(x)|Du(x)|rdxCf1,pp,

for some constant depending only on ? p and Ω, but not on the solution. Moreover, if 0 < α < r-1, then

DuqrCf1,ppq<r1+α.

Proof. The proof of existence can be obtained by a standard perturbation argument, as the one employed in [5]. The Lq-estimate for the deformation tensor Du follows from Lemma 4.2.

We can now state the uniqueness result for the anisotropic double-phase problem.%

Theorem 4.2

Let be given p ∈ ]6/5,12/7[, r > 2, and 0 < α < r-1 such that

 (i) r1+α>3p(2p)5p6 for 65<p3332, (ii) r1+α>3p(4p)7p6 for 3332<p3713, (iii) r1+α>2 for 3713<p<127.

Let u1, u2 be a weak solutions of (4.2) with fW1,p(Ω) . Then, there exists a constant ϵ0= ϵ0p,r,α,Ω,f1,p>0 such that if at least one solution has a small enough weighted gradient, that is, if

Ωdα(x)Du1(x)rdx1/rϵ0,

then u1 = u2.

Proof. The proof follows the same lines of that presented in Theorem 3.1, where r1+α plays the same role as q (with the caveat that equality is not valid in the estimates, due to the strict inequality in Lemma 4.2).

  1. Conflict of interest

    Conflict of interest statement: Being one of the authors member of the Editorial Board of ANONA, did not affect the final evaluation of the article.

Acknowledgments

The first and third authors would like to extend their sincere appreciation to the Deanship of Scientific Research at King Saud University for funding this Research group (RG-1440-061).

The second author was partially supported by a grant of the group GNAMPA of INdAM and by the project of the University of Pisa within the grant PRA_2018_52 UNIPI: “Energy and regularity: New techniques for classical PDE problems.”

References

[1] B. Baldwin and H. Lomax, Thin-layer approximation and algebraic model for separated turbulent flows, 16th Aerospace Sciences Meeting (Huntsville, AL), U.S.A, Aerospace Sciences Meetings, (1978).10.2514/6.1978-257Search in Google Scholar

[2] S. Baraket, S. Chebbi, N. Chorfi, and V. D. Radulescu, Non-autonomous eigenvalue problems with variable (p1, p2)- growth, Adv. Nonlinear Stud. 17 (2017), no. 4, 781–792.10.1515/ans-2016-6020Search in Google Scholar

[3] P. Baroni, M. Colombo, and G. Mingione, Regularity for general functionals with double phase, Calc. Var. Partial Differential Equations 57 (2018), no. 2, Art. 62, 48.10.1007/s00526-018-1332-zSearch in Google Scholar

[4] L. Belenki, L. C. Berselli, L. Diening, and M. Ružicka, On the finite element approximation of p-Stokes systems, SIAM J. Numer Anal. 50 (2012), no. 2, 373–397.10.1137/10080436XSearch in Google Scholar

[5] L. C. Berselli and D. Breit, On the existence of weak solutions for the steady baldwin-lomax model and generalizations, J. Math. Anal. Appl. 501 (2021), 124643.10.1016/j.jmaa.2020.124633Search in Google Scholar

[6] E. Blavier and A. Mikelic, On the stationary quasi-Newtonian flow obeying a power-law, Math. Methods Appl. Sci. 18 (1995), no. 12, 927–948.10.1002/mma.1670181202Search in Google Scholar

[7] D. Breit, Existence theory for generalized Newtonian fluids, Mathematics in Science and Engineering, Elsevier/Academic Press, London, 2017.Search in Google Scholar

[8] N. Chorfi and V. D. Radulescu, Standing wave solutions of a quasilinear degenerate Schrödinger equation with unbounded potential, Electron. J. Qual. Theory Differ. Equ. (2016), Paper No. 37, 12.10.14232/ejqtde.2016.1.37Search in Google Scholar

[9] 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

[10] F. Crispo and C. R. Grisanti, On the existence, uniqueness and C1,() \ W2,2() regularity for a class of shear-thinningfluids, J. Math. Fluid Mech. 10 (2008), no. 4, 455–487.Search in Google Scholar

[11] F. Crispo and C. R. Grisanti, On the C1,() \ W2,2() regularity for a class of electro-rheological fluids, J. Math. Anal. Appl. 356 (2009), no. 1,119–132.10.1016/j.jmaa.2009.02.013Search in Google Scholar

[12] F. Crispo and P. Maremonti, A high regularity result of solutions to modified p-Stokes equations, Nonlinear Anal. 118(2015), 97–129.10.1016/j.na.2014.10.017Search in Google Scholar

[13] L. Diening and F. Ettwein, Fractional estimates for non-differentiable elliptic systems with general growth, Forum Math. 20(2008), no. 3, 523–556.10.1515/FORUM.2008.027Search in Google Scholar

[14] L. Diening and Ch. Kreuzer, Linear convergence of an adaptive finite element method for the p–Laplacian equation, SIAM J.Numer. Anal. 46 (2008), 614–638.10.1137/070681508Search in Google Scholar

[15] 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

[16] G. P. Galdi, Mathematical problems in classical and non-Newtonian fluid mechanics, Hemodynamical flows, OberwolfachSemin., vol. 37, Birkhäuser, Basel, 2008, pp. 121–273.10.1007/978-3-7643-7806-6_3Search in Google Scholar

[17] G. P. Galdi, An introduction to the mathematical theory of the Navier-Stokes equations. Steady-state problems, SpringerMonographs in Mathematics, Springer-Verlag, New York, 2011.10.1007/978-0-387-09620-9Search in Google Scholar

[18] L. Gasinski and P. Winkert, Existence and uniqueness results for double phase problems with convection term, J. DifferentialEquations 268 (2020), 4183–4193.10.1016/j.jde.2019.10.022Search in Google Scholar

[19] A. Kufner, Weighted Sobolev spaces, A Wiley-Interscience Publication, John Wiley&Sons, Inc., New York, 1985.Search in Google Scholar

[20] O. A. Ladyžhenskaya, The mathematical theory of viscous incompressible flow, Second English edition, revised and enlarged. Translated from the Russian by Richard A. Silverman and John Chu. Mathematics and its Applications, Vol. 2, Gordon and Breach Science Publishers, New York, 1969.Search in Google Scholar

[21] J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod, Gauthier-Villars, Paris, 1969.Search in Google Scholar

[22] 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

[23] M. Ružicka, Electrorheological fluids: modeling and mathematical theory, Lecture Notes in Mathematics, vol. 1748, Springer-Verlag, Berlin, 2000.Search in Google Scholar

[24] A. Kufner, Analysis of generalized Newtonian fluids, Topics in mathematical fluid mechanics, Lecture Notes in Math., vol. 2073, Springer, Heidelberg, (2013), pp. 199–238.Search in Google Scholar

[25] V. D. Radulescu, Isotropic and anistropic double-phase problems: old and new, Opuscula Math. 39 (2019), no. 2, 259–279.10.7494/OpMath.2019.39.2.259Search in Google Scholar

[26] V. D. Radulescu and D. D. Repovš, Partial differential equations with variable exponents, Monographs and Research Notes in Mathematics, CRC Press, Boca Raton, FL, 2015, Variational methods and qualitative analysis.10.1201/b18601Search in Google Scholar

[27] M. Ružicka and L. Diening, Non–Newtonian fluids and function spaces, Proceedings of NAFSA (Prague 2006), vol. 8 (2007), 95–144.Search in Google Scholar

[28] H.-M. Yin, Regularity of weak solution to a p-curl-system, Differential Integral Equations 19 (2006), no. 4, 361–368.10.57262/die/1356050504Search in Google Scholar

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

Received: 2020-09-20
Accepted: 2021-06-09
Published Online: 2021-09-08

© 2021 Mohamed Abdelwahed et al., published by De Gruyter

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

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0196/html
Scroll to top button