Home Large data existence theory for three-dimensional unsteady flows of rate-type viscoelastic fluids with stress diffusion
Article Open Access

Large data existence theory for three-dimensional unsteady flows of rate-type viscoelastic fluids with stress diffusion

  • Michal Bathory EMAIL logo , Miroslav Bulíček and Josef Málek
Published/Copyright: September 2, 2020

Abstract

We prove that there exists a weak solution to a system governing an unsteady flow of a viscoelastic fluid in three dimensions, for arbitrarily large time interval and data. The fluid is described by the incompressible Navier-Stokes equations for the velocity v, coupled with a diffusive variant of a combination of the Oldroyd-B and the Giesekus models for a tensor 𝔹. By a proper choice of the constitutive relations for the Helmholtz free energy (which, however, is non-standard in the current literature, despite the fact that this choice is well motivated from the point of view of physics) and for the energy dissipation, we are able to prove that 𝔹 enjoys the same regularity as v in the classical three-dimensional Navier-Stokes equations. This enables us to handle any kind of objective derivative of 𝔹, thus obtaining existence results for the class of diffusive Johnson-Segalman models as well. Moreover, using a suitable approximation scheme, we are able to show that 𝔹 remains positive definite if the initial datum was a positive definite matrix (in a pointwise sense). We also show how the model we are considering can be derived from basic balance equations and thermodynamical principles in a natural way.

1 Introduction

We aim to establish a global-in-time and large-data existence theory, within the context of weak solutions, to a class of homogeneous incompressible rate-type viscoelastic fluids flowing in a closed three-dimensional container. The studied class of models can be seen as the Navier-Stokes system (for which a similar existence theory is well known, cf. [27]) coupled with a viscoelastic rate-type fluid model that shares the properties of both Oldroyd-B and Giesekus models and is completed with a diffusion term. Such models are frequently encountered in the theory of non-Newtonian fluid mechanics, see [19, 21] and further references cited in [19].

In order to precisely formulate the problems investigated in this study, we start by introducing the necessary notation. For a bounded domain Ω ⊂ ℝ3 with the Lipschitz boundary Ω and a time interval of the length T > 0, we define the time-space cylinder Q := (0, T) × Ω and we also set Σ := (0, T) × Ω for a part of its boundary. The symbol n denotes the outward unit normal vector on ∂Ω and, for any vector z, the vector zτ denotes the projection of the vector to a tangent plane on Ω, i.e., zτ := z − (zn)n. Then, for a given density of the external body forces f : Q → ℝ3, a given initial velocity v0 : Ω → ℝ3 and a given initial extra stress tensor 𝔹0 : Ω R>03×3 (here R>03×3 denotes the set of symmetric positive definite (3 × 3)-matrices), we look for a vector field v : Q → ℝ3, a scalar field p : Q → ℝ and a positive definite matrix field 𝔹 : Q R>03×3 solving the following system in Q:

divv=0, (1.1)
tv+(v)vνΔv+p=2μadiv((1β)(BI)+β(B2B))+f, (1.2)
tB+(v)B+δ1(BI)+δ2(B2B)λΔB=a+12(vB+(vB)T)+a12(Bv+(Bv)T), (1.3)

and being completed by the following boundary conditions on Σ:

vn=0,σvτ=νv+ν(v)T+2μa(1β)(BI)+2μaβ(B2B)nτ,(n)B=O,(here O stands for zero 3×3-matrix) (1.4)

and by the initial conditions in Ω:

v(0,)=v0, (1.5)
B(0,)=B0 (1.6)

The parameters β ∈ (0, 1), ν, λ,σ > 0, δ1,δ2 ≥ 0 and a ∈ ℝ are given numbers.

The main result of this study can be stated as:

Let v0 and 𝔹0 be such that the initial total energy is bounded. Then, for sufficiently regular f, there exists a global-in-time weak solution to (1.1)(1.6).

Although the above result is stated vaguely, we would like to emphasize that we are going to establish the long-time existence of a weak solution for large data and for three-dimensional flows. A more precise and rigorous version of the above result including the correct function spaces and the properly defined weak formulation is stated in the Theorem below, see Section 2.

We complete the introductory part by providing the physical background relevant to the studied problem and by recalling earlier results relevant to the problem (1.1)(1.6) analyzed here.

1.1 Mathematical and physical background

The system (1.1)(1.4) can be rewritten into a more concise form once one recognizes some physical quantities. First of all, let

Dv=12(v+(v)T)andWv=12(v(v)T)

denote the symmetric and antisymmetric parts of the velocity gradient ∇v, respectively. Then, looking at the equation (1.2), we see that (1.2) is obtained from a general form of the balance of linear momentum, namely

ϱv˙=divT+ϱf, (1.7)

once we set the density ϱ = 1 and require that the Cauchy stress tensor 𝕋 has the form

T=pI+2νDv+2aμ((1β)(BI)+β(B2B)). (1.8)

In (1.7), stands for the material time derivative of v, i.e., = tv + (v ⋅ ∇)v. Defining similarly the material time derivative of a tensor 𝔹 as

B˙=tB+(v)B,

we can recognize the presence of a general objective derivative in (1.3). Namely, defining

B=B˙a(DvB+BDv)(WvBBWv),

we can rewrite the system (1.1)(1.3) into a more familiar form as

divv=0, (1.9)
v˙=divT+f, (1.10)
B+δ1(BI)+δ2(B2B)=λΔB, (1.11)

which is supposed to hold true in Q and which is completed by the initial conditions (1.5), (1.6) fulfilled in Ω and by the boundary conditions (1.4) on Σ that take the form:

vn=0, (1.12)
(Tn)τ=σvτ, (1.13)
(n)B=O. (1.14)

We provide several comments regarding (1.8)(1.11) as well as the boundary conditions (1.12)(1.14). The Navier slip boundary condition (1.13) (and in general all boundary conditions allowing the fluid to slip ever so slightly) has recently attracted lot of attention. It was well documented that in certain situations the Navier slip boundary conditions are more appropriate than no slip boundary conditions, we refer e.g. to [12, 20, 23, 25] or [36] and references therein. In addition, it was shown that the Navier slip boundary condition can be understood as an asymptotic limit of no slip boundary conditions in case we consider rough and highly oscillating boundary, see e.g. [1, 6, 9]. Furthermore, for the classical Navier-Stokes equation or the Stokes equation, we can say that the available mathematical theory for no-slip boundary condition has been already “re-proven” for Navier boundary conditions, see e.g. [3] for the existence analysis, [2, 4, 30] for regularity theory for the Stokes system and [7] for a conditional regularity result for Navier-Stokes system. The key difference and also the main mathematical advantage of the Navier slip boundary conditions is, that for smooth domains, namely if Ω ∈ 𝓒1,1, we can introduce the pressure p as an integrable function, e.g., by using an additional layer of approximation as in [11], see also [15, 16] or [8] which discuss the treatment of the pressure in evolutionary models subject to the Navier boundary condition. Nevertheless, since we shall always deal with formulation without the pressure (see the Definition), we can also treat the Dirichlet boundary condition, as well as very general implicitly specified boundary conditions see e.g. [12, 13, 36] or [8]. The Neumann boundary condition for 𝔹 is considered here only for simplicity and without any specific physical meaning.

A further aspect, which makes the above system more complicated than the Navier-Stokes equation is the form of the Cauchy stress tensor 𝕋 as in (1.8). The term −p𝕀+2ν𝔻v corresponds to the standard Newtonian fluid flow model with a constant kinematic viscosity ν. The next part of the Cauchy stress, which depends linearly on 𝔹, appears in all the viscoelastic rate-type fluid models - see, e.g., [32, (7.20b), (8.20e)], [24, (6.43e)] or [19, (43a)]. On the other hand, the addition of the term 2aμ β(𝔹2 − 𝔹) is, to our best knowledge, considered here for the first time. The fact that we require that β is positive (and strictly less than 1) plays a key role in the analysis of the problem, as will be shown below. Note that the linearization of 𝕋 with respect to 𝔹 when 𝔹 is close to the identity 𝕀 yields

T=pI+2νDv+2aμ(BI)

and we recover the standard form of 𝕋 (after possible redefinition of the pressure).

The quantity 𝔹 takes into account the elastic responses of the fluid and the equation (1.11) describes its evolution in the current configuration (Eulerian coordinates), just as the velocity v. It is frequent to call the tensor μ(𝔹-𝕀) the extra stress or conformation tensor and to denote it by τ. More importantly, since the material derivative of 𝔹 is not objective, it must be “corrected” and this is the reason, why in (1.11) the derivative B appears. The parameter a in the definition of B determines the type of the objective derivative. The case a = 1 leads to the upper convected Oldroyd derivative, that has favourable physical properties and that leads to a clear interpretation of 𝔹 within the thermodynamical framework developed in [38], see also [33, 34, 35, 39]. Next, the case a = 0 leads to the corrotational Jaumann-Zaremba derivative and this is the only case for which the analysis is much simpler than in other cases. Furthermore, if a ∈ [−1, 1], one obtains the entire class of Gordon-Schowalter derivatives. However, it turns out that the physical properties of these derivatives are irrelevant for the analysis presented below (except the case a = 0), therefore we may take any a ∈ ℝ. For a = 1 and λ = 0 we distinguish two cases: if δ1 > 0 and δ2 = 0 we obtain the classical Oldroyd-B model while if δ1 = 0 and δ2 > 0 we get the Giesekus model. Next, by considering a ∈ [−1, 1], we obtain the class of Johnson-Segalman models. If we further let λ > 0, we are introducing diffusive variants of the previous models. It has been observed that including the diffusion term in (1.11) is physically reasonable, see, e.g., [21] or [19] and references therein. However, up to now, it has been unknown what precise form should the diffusion term take and also whether it actually helps in the analysis of the model. Our main result provides a partial answer to this question, namely: for β ∈ (0, 1) and with the diffusion term being of the form Δ 𝔹 (or more generally, a linear second order operator), the global existence of a weak solution is available.

The reader familiar with the equations describing flows of the standard Oldroyd-B viscoleastic rate-type fluid can identify two deviations in the set of equations (1.9)(1.11) studied hereafter. We provide a few comments on these differences.

The first deviation concerns the incorporation of the stress diffusion term, i.e. the term −Δ𝔹, into the equations. Following the pioneering work of [21] it is clear that a quantity related to ∣∇𝔹∣2 has to be added into the list of underlying dissipation mechanisms. On the other hand, the precise form in which stress diffusion should appear depends on the choice of a thermodynamical approach and specific assumptions. In fact, using the thermodynamical concepts as in [32] or [19], one can derive models, where the stress diffusion term takes the form BΔBΔBB,B12ΔBB12 etc., however, we would prefer −Δ𝔹 simply because it coincides with the form proposed by [21], and, from the perspective of PDE analysis and numerical approximation, one prefers to deal with stress diffusion that leads to a linear operator.

The second deviation from usual viscoelastic models consists in the presence of the term (𝔹2-𝔹) in the Cauchy stress tensor, see (1.8). This term arises if we slightly modify energy storage mechanism and apply the thermodynamic approach as developed in [32]. In what follows, we shall give a clear interpretation and a thermodynamic derivation of our model.

1.2 Thermodynamical derivation of the model

Viscoelastic models with (nonlinear) stress diffusion, but without the term 𝔹2 in the stress tensor are derived, e.g., in [32] and [19] even in the temperature-dependent case. Here, we will briefly explain the approach in a simplified isothermal setting (sufficient for the purpose of this study), referring to the cited works for the derivation in a complete thermal setting and for more details.

First, we postulate the constitutive equation for the Helmholtz free energy in the form

ψ(B):=μ((1β)(trB3lndetB)+12β|BI|2), (1.15)

where μ > 0 and β ∈ [0, 1] is a parameter interpolating between two forms of the energy. The choice β = 0 would lead to a standard Oldroyd-B diffusive model. To our best knowledge, the case β > 0 was not considered before in literature. The term 12 β∣𝔹-𝕀∣2, which is newly included in ψ is obviously convex with the minimum at 𝔹 = 𝕀 and depends only on tr 𝔹 and on tr (𝔹𝔹), i.e., on invariants of 𝔹, therefore it does not violate any of the basic principles of continuum physics. Moreover, such an addition does not affect the first three terms in the asymptotic expansion of ψ near 𝕀, on the logarithmic scale. To see this, let ℍ denote the Hencky logarithmic tensor satisfying e = 𝔹 (which exists due to the positive definiteness of 𝔹). Using Jacobi’s identity, we compute that

trB3lndetB=tr(eHIH)=tr(12H2+O(H3)).

On the other hand, we easily get

12|BI|2=12tr(e2H2eH+I)=tr(12H2+O(H3)),

hence we also have

(1β)(trB3lndetB)+12β|BI|2=tr(12H2+O(H3))

and we see that for 𝔹 being close to identity, the form of ψ is almost independent of the choice of parameter β and the second part of ψ in (1.5) can be just understood as a correction for large values of 𝔹.

Next, we show how the constitutive equation for 𝕋 (see (1.8)) appears naturally if we start with the choice of the Helmholtz free energy (1.5) and require that the form of the equation for 𝔹 is given by (1.11). For the derivation, we followed the approach developed in [32] that stems from the balance equations and requires the knowledge of how the material stores the energy, but we simplify the derivation presented there by assuming that the density is constant (in fact we set for simplicity ϱ = 1) and hence divv = 0 and the flow is isothermal, i.e., the temperature θ is constant as well. Under these assumptions the balance equations of continuum physics (for linear and angular momenta, energy and for formulation of the second law of thermodynamics) take the form

v˙=divT,T=TT,e˙=TDvdivje,η˙=ξdivjη with ξ0,

where e is the (specific) internal energy, η is the entropy, ξ is the rate of entropy production, 𝕋 is the Cauchy stress tensor and the quantities je, jη represent the internal and the entropy fluxes, respectively. Since the quantities ψ, e, θ and η are related through the thermodynamical identity

e=ψ+θη,

we can easily deduce from above identities that

θξ=θη˙+div(θjη)=TDvdiv(jeθjη)ψ˙. (1.16)

To evaluate the last term, we rewrite (1.11) as

B˙=λΔBa(DvB+BDv)(WvBBWv)+δ1(BI)+δ2(B2B). (1.17)

Next, it follows from (1.5) that

ψ(B)B=J, (1.18)

where 𝕁 is defined as

J:=μ(1β)(IB1)+μβ(BI).

Consequently, taking the inner product of (1.17) with 𝕁 we observe that (since 𝔹𝕁 = 𝕁 𝔹, the term with 𝕎v vanishes)

ψ˙=λΔBJa(DvB+BDv)J(WvBBWv)J+δ1(BI)J+δ2(B2B)J=λdiv(ψ(B))a(DvB+BDv)J+δ1(BI)J+δ2(B2B)J+λBJ. (1.19)

To evaluate the terms on the last line, we use the symmetry and the positive definiteness of the matrix 𝔹 to obtain

(BI)J=μ(1β)|B12B12|2+μβ|BI|2,(B2B)J=μ(1β)|BI|2+μβ|B32B12|2,BJ=μβ|B|2μ(1β)BB1=μβ|B|2+μ(1β)BB1BB1=μβ|B|2+μ(1β)|B12BB12|2. (1.20)

Similarly, we obtain

a(BDv+DvB)J=2μa((1β)(BI)+β(B2B))Dv. (1.21)

Thus, using (1.19)(1.21) in (1.16), we conclude that

θξ=div(λψ(B)+jeθjη)+T2aμ((1β)(BI)+β(B2B))Dv+μλ(β|B|2+(1β)|B12BB12|2)+μ(1β)δ1|B12B12|2+βδ2|B32B12|2+μ((1β)δ2+βδ1)|BI|2. (1.22)

Hence, assuming that the fluxes fulfil

λψ(B)+jeθjη=0, (1.23)

and setting (compare with (1.8))

T=pI+2νDv+2aμ((1β)(BI)+β(B2B)),

the identity (1.22) reduces to (noticing that −p𝕀 ⋅ 𝔻v = −p div v = 0)

θξ=μλ(β|B|2+(1β)|B12BB12|2)+2ν|Dv|2+μ(1β)δ1|B12B12|2+βδ2|B32B12|2+μ((1β)δ2+βδ1)|BI|2, (1.24)

which gives the nonnegative rate of the entropy production. Moreover, we have seen how the form of the Cauchy stress tensor 𝕋 in (1.8) is dictated by the second line in (1.22). Furthermore, we can also see in (1.24) (and also in the last line of (1.20)) how the choice of the free energy (1.5) affects the entropy production due to the presence of the diffusive term Δ 𝔹 in (1.3).

1.3 The concept of weak solution and energy (in)equality

In order to introduce the proper concept of weak solution, we first derive the basic energy estimates based on the observations from the previous section. First, taking the scalar product of (1.10) and v, we deduce the kinetic energy identity

12t|v|2+12div(|v|2v)div(Tv)+TDv=fv (1.25)

and replacing the term 𝕋 ⋅ 𝔻v from the equation (1.16), and using then also (1.23) and (1.24), we finally obtain

t(ψ+12|v|2)+div((ψ+12|v|2)v)div(Tv+λψ(B))+2ν|Dv|2+μλβ|B|2+(1β)|B12BB12|2+μ(1β)δ1|B12B12|2+βδ2|B32B12|2+((1β)δ2+βδ1)|BI|2=fv. (1.26)

Integrating the above identity over Ω, using integration by parts and the boundary conditions (1.12)(1.14), we obtain

ddtΩ12|v|2+ψ(B)+2νΩ|Dv|2+σΩ|v|2+μλΩβ|B|2+(1β)|B12BB12|2+μΩ((1β)δ1|B12B12|2+βδ2|B32B12|2+((1β)δ2+βδ1)|BI|2)=Ωfv. (1.27)

The identity (1.27) indicates the proper choice of the function spaces for the solution (v, 𝔹) and the form of the (weak) formulation of the solution to (1.1)(1.6).

1.4 Notation

In order to formulate the definition of a weak solution conveniently, let us fix some notation. By Lp(Ω) and Wn,p(Ω), 1 ≤ p ≤ ∞, n ∈ ℕ, we denote the usual Lebesgue and Sobolev space, with their usual norms denoted as ∥⋅∥p and ∥⋅∥n,p, respectively. The trace operator that maps W1,p(Ω) into Lq(∂Ω), for certain q ≥ 1, will be denoted by 𝓣. Further, we set W−1,p(Ω) = (W1,p(Ω))*, where p′ = p/(p − 1). We shall use the same notation for the function spaces of scalar-, vector-, or tensor-valued functions, but we will distinguish the functions themselves using different fonts such as a for scalars, a for vectors and 𝔸 for tensors. Also, we do not specify the meaning of the duality pairing 〈⋅, ⋅〉, assuming that it is clear from the context. Moreover, for certain subspaces of vector valued functions, we shall use the following notation:

Cn={w:ΩR3:w infinitely differentiable,wn=0 on Ω},Cn,div={wCn:divw=0 in Ω},Ln,div2=Cn,div¯2,Wn,div1,2=Cn,div¯1,2,Wn,div3,2=Cn,div¯3,2,Wn,div1,2=(Wn,div1,2),Wn,div3,2=(Wn,div3,2).

Occasionally, we shall denote the standard inner products in L2(Ω) and L2(∂Ω) as (⋅, ⋅) and (⋅, ⋅)∂Ω, respectively. The Bochner spaces of mappings from (0, T) to a Banach space X will be denoted as Lp(0, T; X) with the norm Lp(0,T;X)=(0TXp)1p. If X = Lq(Ω), or X = Wk,q(Ω), we will write just ∥⋅∥LpLq, or ∥⋅∥LpWk,q, respectively. The space 𝓒weak(0, T; X) ⊂ L(0, T; X) denotes a space of weakly continuous functions, i.e., for every f ∈ 𝓒weak(0, T; X), every t0 ∈ [0, T] and every gX* there holds

limtt0f(t),g=f(t0),g.

The symbol Rsym3×3 denotes the set of symmetric 3 × 3 real matrices. Furthermore, by R>03×3 we denote the subset of Rsym3×3 which consists of positive definite matrices, i.e., those which satisfy

Azz>0for all zR3{0}.

2 The definition of a weak solution and its existence

In this section we state and prove the main result.

Definition

Let T > 0 and assume that Ω ⊂ ℝ3 is a Lipschitz domain. Let β ∈ (0, 1), ν,σ,λ > 0, δ1,δ2 ≥ 0, a ∈ ℝ, and fL2(0,T;Wn,div1,2),v0Ln,div2(Ω). Furthermore, let 𝔹0L2(Ω) be such that

ΩlndetB0<. (2.1)

Then, we say that a couple (v, 𝔹) : Q → ℝ3 × R>03×3 is a weak solution to (1.1)(1.6) if the following hold:

vL2(0,T;Wn,div1,2)L(0,T;L2(Ω)),tvL43(0,T;Wn,div1,2),BL2(0,T;W1,2(Ω))L(0,T;L2(Ω)),tBL43(0,T;W1,2(Ω));

For all φL4(0,T;Wn,div1,2) we have

0Ttv,φ+Q(v)vφ+σ0TΩTvTφ=Q(2νDv+2aμ((1β)(BI)+β(B2B)))φ+0Tf,φ; (2.2)

For all 𝔸 ∈ L4(0, T; W1,2(Ω)), 𝔸 = 𝔸T, we have

0TtB,A+Q((v)B+2BWv2aBDv)A+Q(δ1(BI)+δ2(B2B))A+λQBA=0. (2.3)

The initial conditions are satisfied in the following sense

limt0+(v(t)v02+B(t)B02)=0. (2.4)

Moreover, we say that the solution satisfies the energy inequality if, for all t ∈ (0, T):

Ω|v(t)|22+ψ(B(t))+0t2νDv22+σTv2,Ω2+μλ0t((1β)B12BB1222+βB22)+μ0t((1β)δ1B12B1222+βδ2B32B1222+(βδ1+(1β)δ2)BI22)Ω|v0|22+ψ(B0)+0tf,v. (2.5)

The key result of the paper is the following

Theorem

Let T > 0 and assume that Ω ⊂ ℝ3 is a Lipschitz domain. Suppose that β ∈ (0, 1), ν, σ, λ > 0, δ1,δ2 ≥ 0, a ∈ ℝ, and fL2(0,T;Wn,div1,2),v0Ln,div2(Ω). Furthermore, let 𝔹0L2(Ω) be such that (2.1) holds. Then there exists a weak solution to (1.1)(1.6) satisfying the energy inequality.

Let us briefly explain the main difficulties connected with the analysis of the system (1.9)(1.13) and our ideas how to solve them. In the standard models where β = 0, to get an a priori estimate for 𝔹, the appropriate test function to take in (1.11) is 𝕀 − 𝔹−1. Then, using (1.9) and (1.10) tested by v, one can eliminate the problematic terms, such as 𝔹 ⋅ 𝔻v coming from the objective derivative. However, the non-negative quantity to be controlled, which comes from the diffusion term, turns out to be just |B12BB12|2 and this provides little to no information. In particular, the terms ∇v𝔹 appearing in (1.11) are going to be just integrable and it is unclear if one can show strong convergence of 𝔹. Instead, one would like to test also by 𝔹 to achieve control over ∣∇𝔹∣2. But this is not possible, since the resulting term ∇v𝔹 ⋅ 𝔹 cannot be estimated without some serious simplifications (such as boundedness of ∇v, two or one dimensional setting or small data). Quite remarkably, this problem is solved simply by adding 12 β ∣𝔹-𝕀∣2 into the constitutive form for ψ. More precisely, considering β ∈ (0, 1), we observe that the appropriate test function in (1.11) is in fact (1−β)(𝕀 − 𝔹−1) + β(𝔹 − 𝕀). Indeed, the terms from the objective derivative cancel again due to the presence of β(𝔹2 − 𝔹) in 𝕋. But now, we also get β∣∇𝔹∣2 under control, which is much better information than in the case β = 0 and it will imply compactness of all the terms appearing in (1.10) and (1.11). We have seen above that such a modification of ψ, and consequently of 𝕋, is not ad-hoc and that it rests on solid physical grounds.

The second and also the last major difficulty which we will encounter is how one can justify testing of (1.11) by 𝔹−1 on the approximate (discrete level), where 𝔹−1 might not even exist. This we overcome by designing a delicate approximation scheme, which takes into account the smallest eigenvalue of 𝔹, and also by noting that testing (1.11) only by 𝔹 yields sufficiently strong a priori estimates for the initial limit passage (in the Galerkin approximation of 𝔹).

Up to now, there have been no results on global existence of weak solutions to Oldroyd-B models in three dimensions, including either the standard, or diffusive variants. The closest result so far is probably [37, Theorem 4.1], however there it is assumed that δ2 > 0 and λ = 0 (Giesekus model), whereas we treat also the case δ2 = 0, but with λ > 0 (diffusive Oldroyd-B or Giesekus model). Moreover, in [37], only the weak sequential stability of a hypothetical approximation is proved. We, on the other hand, provide the complete existence proof, including the construction of approximate solutions (which, in viscoelasticity, is generally a non-trivial task). In the article [28], Lions and Masmoudi prove the global existence in three dimensions, but only for a = 0 (corrotational case), which is known to be much easier. The local in time existence of regular solutions for the non-diffusive variants of the models above (λ = 0) is proved in the pioneering work [22, Theorem 2.4.]. There, also the global existence for small data is shown. In two dimensions, the problem is solved in [18] in the case λ > 0, δ1 > 0, δ2 = 0 (diffusive Oldroyd-B model). There are also global large data existence results in three dimensions for slightly different classes of diffusive rate-type viscoelastic models, but under some simplifying assumptions. For example, in [14] and [10], the authors consider the case where 𝔹= b𝕀. This assumption, however, turns (1.11) into a much simpler scalar equation. Moreover, note that if 𝔹= b𝕀, then the equations (1.10) and (1.11) decouple (which is not the case in [14] and [10] since there the considered constitutive relation for 𝕋 is more complicated than here). Furthermore, in [29], the authors consider yet another class of Peterlin viscoelastic models with stress diffusion and prove existence of a global two- or three-dimensional solution. However, the free energy associated with these models depends only on the trace of the extra stress tensor. This is a significant simplification, which can even be seen as unphysical. See also [17] for various modifications of Oldroyd-B viscoelastic models, for which an existence theory is available. Finally, in [5] (see also [26]), the global existence of a weak solution is shown for a certain regularized Oldroyd-B model (including a cut-off or nonlinear p−Laplace operator in the diffusive term in 𝔹). Thus, one might argue that since the case β > 0 could be also seen as a regularization of the original model, we are just proving an existence of a solution to another regularization. However, this argument is not, in our opinion, correct for several reasons. First of all, the “regularization” β > 0 does not touch the equation (1.11) at all. Second, it is not obvious why the nonlinear term β(𝔹 − 𝕀)2 should have any regularization effect. And, perhaps most importantly, we already showed in Section 1.2 that the model with β > 0 is physically well founded and worthy of studying in its own right.

Remark

Finally, we close this section with several concluding remarks on possible extensions, but we do not provide their proofs in this paper.

  1. The Theorem holds also in arbitrary dimensions d > 3 (in d ≤ 2, it is known), however with worse function spaces for the time derivatives and better for the test functions. Indeed, the only dimension-specific argument in the proof below is in the derivation of interpolation inequalities, which are then used to estimate tv and t𝔹. Moreover, all of the non-linear terms in (2.2), (2.3) are integrable for arbitrary d if the test functions are smooth. In addition, if d = 2, then we can prove the existence of a weak solution satisfying even the energy equality, i.e., (2.5) holds with the equality sign.

  2. When Ω has C1,1 boundary, then, in addition, there exists a pressure pL53(Q), which appears in (1.2). Then, the test functions in (2.2) need not be divergence-free if we include the term Ω p divφ in (2.2). This follows in a standard way, using the Helmholtz decomposition of v (see, e.g., [8] for details).

  3. It is possible to replace (1.12), (1.13) by the no-slip boundary condition v = 0 on ∂Ω. Then, we only need to change the space Wn1,2 to W01,2, and so on. However, then it seems that the pressure p can be only obtained as a distribution (see [8]).

3 Proof of the Theorem

Throughout the proof, we shall simplify notation by assuming

λ=μ=ν=σ=1

and refer to Section 1.2 for a detailed computation for general parameters. To shorten all formulae, we also denote

S(A):=(1β)(AI)+β(A2A)for AR3×3,R(A):=δ1(AI)+δ2(A2A)for AR3×3.

The general scheme of the proof is the following: In order to invert the matrix 𝔹 and to avoid problems with low integrability in the objective derivative, we introduce the special cut-off function

ρε(A):=max{0,Λ(A)ε}Λ(A)(1+ε|A|3)for ARsym3×3,

where Λ(𝔸) denotes a minimal eigenvalue of 𝔸 (whose spectrum is real due to its symmetry)[1]. Since eigenvalues of a matrix depend continuously on its entries, the function ρε is continuous. Moreover, for any positive definite matrix 𝔸 there holds ρε(𝔸) → 1 as ε → 0+. We construct a solution by an approximation scheme with parameters k, l and ε, where k, l ∈ ℕ correspond to the Galerkin approximation for v and 𝔹, respectively, and ε corresponds to the presence of the cut-off function ρε in certain terms. The first limit we take is l → ∞, which corresponds to the limit in the equation for 𝔹. This way, the limiting object 𝔹 is infinite-dimensional and, using the properties of ρε, we prove that 𝔹−1 exists. With the help of this information, we derive the energy estimates that are uniform with respect to all the parameters. Next, we let ε → 0+ in order to remove the truncation function and finally we take k → ∞, which corresponds to the limiting procedure in the equation for the velocity v.

3.1 Galerkin approximation

Following e.g., [31, Appendix A.4], we know that there exists a basis {wi}i=1 of Wn,div3,2, which is orthonormal in L2(Ω) and orthogonal in Wn,div3,2 . Moreover, the projection Pk:L2(Ω)span{wi}i=1k, defined as[2]

Pkφ=i=1k(φ,wi)wi,φL2(Ω),

is continuous in L2(Ω) and also in Wn,div3,2 independently of k, i.e.,

Pkφ2Cφ2andPkφWn,div3,2CφWn,div3,2

for all φ Wn,div3,2 , where the constant C is independent of k. Furthermore, by the standard embedding, we also have that Wn,div3,2 W2,6(Ω) ↪ W1,∞(Ω). Similarly, we construct the basis {Wj}j=1 of W1,2(Ω), which is L2-orthonormal, W1,2-orthogonal and the projection

QlA=j=1l(A,Wj)Wj,AL2(Ω),

is continuous in L2(Ω) and in W1,2(Ω) independently of l.

Then for fixed k, l ∈ ℕ and ε ∈ (0, 1), we look for the functions vεk,l,Bεk,l of the form

vεk,l(t,x)=i=1kcik,l,ε(t)wi(x)andBεk,l(t,x)=j=1ldjk,l,ε(t)Wj(x),

where cik,l,ε,djk,l,ε,i=1,,k,j=1,,l, are unknown functions of time, and we require that vεk,l,Bεk,l (and consequently the functions cik,l,ε(t) and djk,l,ε(t)) satisfy the following system of (k+l) ordinary differential equations in time interval (0, T):

ddt(vεk,l,wi)+((vεk,l)vεk,l,wi)+2(Dvεk,l,wi)+(Tvεk,l,Twi)Ω=2a(ρε(Bεk,l)S(Bεk,l),wi)+f,wifor i=1,,k, (3.1)
ddt(Bεk,l,Wj)+((vεk,l)Bεk,l,Wj)+(ρε(Bεk,l)R(Bεk,l),Wj)+(Bεk,l,Wj)=2(ρε(Bεk,l)Bεk,l(aDvεk,lWvεk,l),Wj)for j=1,,l. (3.2)

Due to the L2-orthonormality of the bases {wi}i=1 and {Wj}j=1, the system (3.1)(3.2) can be rewritten as a nonlinear system of ordinary differential equations for cik,l,ε and djk,l,ε, where i = 1, …, k and j = 1, …, l, and we equip this system with the initial conditions

cik,l,ε(0)=(v0,wi) and djk,l,ε(0)=(B0ε,Wj). (3.3)

Here, B0ε is defined as

B0ε(x):={B0(x)if Λ(B0(x))>ε,Ielsewhere.

Since 𝔹0(x) ∈ R>03×3 for almost every xΩ, we have that Λmbda(𝔹0(x)) > 0 for almost all xΩ. Consequently, using the fact 𝔹0L2(Ω), we obtain, as ε → 0+, that

B0εB022=Λ(B0)ε|IB0|20

Note also that the initial conditions (3.3) can be rewritten as vεk,l(0)=Pkv0 and Bεk,l(0)=QlB0ε.

For the system (3.1)(3.3), Carathéodory’s theorem can be applied and therefore there exists T* > 0 and absolutely continuous functions cik,l,ε,djk,l,ε satisfying (3.3) and (3.1)(3.2) almost everywhere in (0,T*). If T* is the maximal time, for which the solution exists, and T* < T, then at least one of the functions cik,l,ε,djk,l,ε must blow up as tT. But using the estimate presented below (see (3.8) valid for all t ∈ (0,T*)), this will be seen never to happen. Thus, we can set T* = T.

3.2 Limit l → ∞

In this part, we simplify the notation and denote the approximating solution, constructed in the previous section, by (vl,Bl):=(vεk,l,Bεk,l). We start by proving estimates independent of l. Since 𝔹l(t) and vl(t) belong for almost all t to the linear hull of {Wj}j=1l and {wi}i=1k, respectively, we can use vl instead of wi in (3.1) and 𝔹l instead of 𝕎j in (3.2) to deduce,

12ddtBl22+Bl22=2a(ρε(Bl)BlDvl,Bl)(ρε(Bl)R(Bl),Bl), (3.4)
12ddtvl22+2Dvl22+Tvl2,Ω2=2a(ρε(Bl)S(Bl),Dvl)+f,vl, (3.5)

where we used the integration by parts formula and the facts that div vl = 0 and 𝓣vn = 0. Next, it follows from the definition of ρε, ℝ and 𝕊 that

ρε(Bl)|S(Bl)|+|R(Bl)||Bl|+|Bl|2C1+|Bl|31+ε|Bl|3C(ε). (3.6)

Here, the notation C(ε) emphasizes that the constant C depends on ε; we keep this notation in what follows. Summing (3.4) and (3.5) and using the estimate (3.6) to bound the term on the right-hand side, we obtain with the help of Hölder’s, Young’s and Korn’s inequalities that

ddtvl22+Bl22+Dvl22+Tvl2,Ω2+Bl22C(ε)+CfWn,div1,22.

After integrating over (0, T) with respect to time, we obtain the following bound:

supt(0,T)vl22+Bl22+0TDvl22+Tvl2,Ω2+Bl22C(ε)+Pkv022+QlB0ε22+C0TfWn,div1,22C(ε), (3.7)

where the last inequality follows from the continuity of the projections Pk and Ql and from the assumptions on data, namely that

v022+B022+lndetB01+C0TfWn,div1,22<.

Next, we focus on the estimate for time derivatives. First, it follows from L2-orthonormality of the bases and the estimate (3.7) that

i=1kci(t)2+j=1ldj(t)2C(ε). (3.8)

Then, since vl is a linear combination of {wi}i=1kW1,(Ω), we can estimate

vlLW1,esssupt(0,T)i=1k|ci(t)|wi1,C(ε,k), (3.9)

and we can deduce from (3.1) and fL2(0,T;Wn,div1,2) that

tvlL2W1,C(ε,k). (3.10)

Finally, it follows from (3.2) and (3.7) that

tBlL2W1,2C(ε,k). (3.11)

Using (3.7), (3.9)(3.11) and Banach-Alaoglu’s theorem, we can find subsequences (which we do not relabel) and corresponding weak limits (denoted with the subscript ε), such that, for l → ∞, we get

vlvεweakly in L2(0,T;Wn,div1,2), (3.12)
vlvεweakly in L(0,T;W1,(Ω)), (3.13)
tvltvεweakly in L2(0,T;W1,(Ω)), (3.14)
TvlTvεweakly in L2(0,T;L2(Ω)), (3.15)
BlBεweakly in L2(0,T;W1,2(Ω)), (3.16)
tBltBεweakly in L2(0,T;W1,2(Ω)). (3.17)

Moreover, it follows from (3.12), (3.14), (3.16), (3.17) and from the Aubin-Lions lemma that for some further subsequences, we have

vlvεstrongly in L2(Q), (3.18)
BlBεstrongly in L2(Q) and a.e. in Q (3.19)
ρε(Bl)ρε(Bε)a.e. in Q. (3.20)

Using the convergence results (3.12)(3.20), it is rather standard to let l → ∞ in (3.1)(3.2). This way, for almost all t ∈ (0, T), we obtain

(tvε,wi)+((vε)vε,wi)+2(Dvε,wi)+(Tvε,Twi)Ω=2a(ρε(Bε)S(Bε),wi)+f,wi (3.21)

for i = 1, …, k, and

tBε,A+((vε)Bε,A)+(Bε,A)=2(ρε(Bε)Bε(aDvεWvε),A)(ρε(Bε)R(Bε),A) (3.22)

for all 𝔸 ∈ W1,2(Ω). Moreover, from (3.16) and (3.17), we get 𝔹ε ∈ 𝓒(0, T; L2(Ω)) and it is standard to show that 𝔹ε(0, ⋅) = B0ε and vε(0, ⋅) = Pkv0.

3.3 Limit ε → 0

In this part we consider the solutions (vε, 𝔹ε) constructed in the preceding section for ε ∈ (0, 1) and we study their behaviour as ε → 0+. To do so, we first have to derive estimates that are uniform with respect to ε. Following the ideas used before in the derivation of the model, we wish to test (3.22) by the function

Jε:=(1β)(IBε1)+β(BεI). (3.23)

This test function, however, contains Bε1 and we need to justify that it exists (for any ε ∈ (0, 1)).

3.3.1 Estimates for the inverse matrix, still ε-dependent

First, we prove that Λ(𝔹ε) ≥ ε. For this purpose, let z ∈ ℝ3 be arbitrary and consider[3]

A=(Bεzzε|z|2)(zz),where(zz)ij:=zizj (3.24)

in (3.22). Due to the properties of 𝔹ε (see (3.16)), we know that 𝔸 belongs to L2(0, T;W1,2(Ω)) and we can use it as a test function in (3.22). Upon inserting 𝔸 into (3.22), we integrate the result over (0,τ) with some fixed τ ∈ (0, T). We evaluate all terms in (3.22) separately. For the time derivative, we have

0τtBε,A=0τt(Bεzzε|z|2),(Bεzzε|z|2)=12(Bε(τ)zzε|z|2)2212(B0εzzε|z|2)22=12(Bε(τ)zzε|z|2)22, (3.25)

where, for the last equality, the definition of B0ε was used. Furthermore, we obtain

QBεA=0τΩ(BεεI)((Bεzzε|z|2)(zz))=0τ(Bεzzε|z|2)22 (3.26)

and

Q(vε)BεA=0τΩvε(Bεzzε|z|2)(Bεzzε|z|2)=120τΩvε((Bεzzε|z|2)2=120τΩ((Bεzzε|z|2)2divvε=0, (3.27)

integrating by parts and using the fact that div vε = 0 and 𝓣vε = 0. Since

BεzzΛ(Bε)|z|2a.e. in Q,

we also observe, that

0(Λ(Bε)ε)+(Bεzzε|z|2)(Λ(Bε)ε)+(Λ(Bε)ε)|z|2=0.

Hence, we get

ρε(Bε)A=Oa.e. in Q. (3.28)

Consequently, inserting 𝔸 of the form (3.24) into (3.22), we see that the right-hand side is identically zero. Therefore, relations (3.25), (3.26), (3.27) and (3.28) yield

(Bεzzε|z|2)22(τ)(Bεzzε|z|2)22(τ)+20τ(Bεzzε|z|2)22=0,

which implies

Bεzzε|z|2for every zR3 and a.e. in Q. (3.29)

Thus, we have the following estimate for the minimal eigenvalue of 𝔹ε:

Λ(Bε)inf0zR3Bεzz|z|2ε.

Therefore, the inverse matrix Bε1 is well defined and satisfies

|Bε1|Cεa.e. in Q. (3.30)

Furthermore, since

Bε1=Bε1BεBε1=Bε1(BεBε1)Bε1(Bε)Bε1=Bε1(Bε)Bε1,

we conclude from (3.7) and (3.30), that

Q|Bε1|2Q|Bε1|4|Bε|2C(ε).

Hence, the inverse of 𝔹ε exists and Bε1 L2(0, T; W1,2(Ω)).

3.3.2 Estimates independent of (ε, k)

At this point, we can test (3.22) with 𝕁ε defined in (3.23). This way, we obtain

tBε,Jε+((vε)Bε,Jε)+(Bε,Jε)=2(ρε(Bε)Bε(aDvεWvε),Jε)(ρε(Bε)R(Bε),Jε).

Next, we evaluate all terms. Here, we follow very closely the procedure developed in Section 1.2, see the derivation of (1.19) and consequent identities. Since

Jε=ψ(Bε)Bε,

where ψ is defined in (1.5), it is clear that

tBε,Jε=ddtΩψ(Bε),((vε)Bε,Jε)=Ωvεψ(Bε)=0.

Next, recalling (1.20), we get

(ρε(Bε)R(Bε),Jε)=Ωρε(Bε)δ1(1β)|Bε12Bε12|2+(δ1β+δ2(1β))|BεI|2+δ2β|Bε32Bε12|2,(Bε,Jε)=βBε22+(1β)Bε12BεBε1222

and due to the fact that 𝔹ε𝕁ε = 𝕁ε𝔹ε we also have

(ρε(Bε)(WvεBεBεWvε),Jε)=0,a(ρε(Bε)(DvεBε+BεDvε),Jε)=2a(ρε(Bε)Dvε,BεJε)=2a(ρε(Bε)Dvε,(1β)(BεI)+β(Bε2Bε))=2a(ρε(Bε)S(Bε),Dvε),

where we used the fact that the trace of 𝔻vε is identically zero. Hence, using 𝔸 := 𝕁ε (defined in (3.23)) in (3.22) and taking into account the above identities, we deduce that

ddtΩψ(Bε)+(1β)Bε12BεBε1222+βBε22+(βδ1+(1β)δ2)ρε(Bε)(BεI)22+(1β)δ1ρε(Bε)(Bε12Bε12)22+βδ2ρε(Bε)(Bε32Bε12)22=2a(ρε(Bε)S(Bε),Dvε). (3.31)

Similarly as in previous section, replacing wi in (3.21) by vε, we get

12ddtvε22+2Dvε22+Tvε2,Ω2=f,vε2a(ρε(Bε)S(Bε),Dvε). (3.32)

Thus, summing (3.31) and (3.32) and integrating the result with respect to time t ∈ (0, τ), we deduce the identity

12vε(τ)22+Ωψ(Bε(τ))+0τ(2Dvε22+Tvε2,Ω2+(1β)Bε12BεBε1222+βBε22+(βδ1+(1β)δ2)ρε(Bε)(BεI)22+(1β)δ1ρε(Bε)(Bε12Bε12)22+βδ2ρε(Bε)(Bε32Bε12)22)=12Pkv022+Ωψ(B0ε)+0τf,vε12v022+Ωψ(B0)+0τf,vε, (3.33)

where, for the last inequality we used the continuity of Pk, the definition of B0ε and the fact that ψ(𝕀) = 0.

From (3.33), we get, using Korn’s, Sobolev’s, Hölder’s and Young’s inequalities, that

vεLL2+vεL2L6+vεL2W1,2+BεL2W1,2+BεL2L6C, (3.34)

where the constant C depends only on Ω, v0, 𝔹0 and f. Furthermore, the interpolation inequalities yield

vεL103L103+vεL4L3+BεL103L103+BεL4L3+BεL83L4C. (3.35)

Finally, we focus on the estimate for time derivatives. Let φL4(0,T;Wn,div3,2) be such that ∥φL4W3,2 ≤ 1. Then, since vε is a linear combination of {wi}i=1k, we obtain, using (3.21), Hölder’s inequality, (3.33), (3.35) and W3,2-continuity of Pk, that

0Ttvε,φC,

hence

tvεL43Wn,div3,2C. (3.36)

Similarly, by considering 𝔸 ∈ L4(0, T; W1,2(Ω)) in (3.22), we get

tBεL43W1,2C. (3.37)

3.3.3 Limit ε → 0+

From (3.34), (3.36), (3.37), Banach-Alaoglu’s theorem and the Aubin-Lions lemma, we obtain the existence of a couple (vk, 𝔹k) satisfying the following convergence results[4]

vεvkweakly in L2(0,T;Wn,div1,2),tvεtvkweakly in L43(0,T;Wn,div3,2),TvεTvkweakly in L2(0,T;L2(Ω)),BεBkweakly in L2(0,T;W1,2(Ω)),tBεtBkweakly in L43(0,T;W1,2(Ω)),vεvkstrongly in L3(Q) and a.e. in Q, (3.38)
BεBkstrongly in L3(Q) and a.e. in Q. (3.39)

Using (3.39) and letting ε → 0+ in (3.29), we obtain

Bkzz0a.e. in Q and for all zR3.

Hence Λ(𝔹k) ≥ 0 and det 𝔹k ≥ 0 a.e. in Q. Therefore, using (3.39) again and the continuity of ψ, there exists (still possibly infinite) limit

ψ(Bε)ψ(Bk)a.e. in Q.

However, since ψ ≥ 0, Fatou’s lemma implies that, for almost every t ∈ (0, T), we have

Ωψ(Bk)(t)lim infε0+Ωψ(Bε)(t)C.

Thus, we deduce that

ψ(Bk)LL1C. (3.40)

If there existed a set EQ of a positive measure, where Λ(𝔹k) = 0, then also –ln det 𝔹k = ∞ on that set, which contradicts (3.40). Thus, we have

Λ(Bk)>0 a.e. in Q. (3.41)

Therefore, it directly follows from the continuity of Λ, that ρε(𝔹ε) → 1 a.e. in Q. Then, since ρε(𝔹ε) ≤ 1, we further get, by Vitali’s theorem, that

ρε(Bε)1strongly in Lp(Q) for all p[1,).

Using the established convergence results, it is easy to let ε → 0+ in (3.21) and (3.22) and obtain, for almost all t ∈ (0, T), that

tvk,wi+((vk)vk,wi)+2(Dvk,wi)=(Tvk,Twi)Ω2a(S(Bk),wi)+f,wi

for i = 1, …, k and that

tBk,A+((vk)Bk,A)+(Bk,A)=2(Bk(aDvkWvk),A)(R(Bk),A)

for all AW1,2(Ω). Furthermore, we can take the limit in the estimates (3.33), (3.35), (3.36) and (3.37) using either the weak lower semi-continuity of norms or, in the terms which depend on 𝔹ε, e.g. Qρε(Bε)|Bε32Bε12|2, we apply (3.41) to conclude the pointwise limit and then use Fatou’s lemma. Thus, inequalities (3.33), (3.35), (3.36) and (3.37) continue to hold in the same form, but for (vk, 𝔹k) instead of (vε, 𝔹ε) and with 1 instead of ρε(𝔹ε). In particular, for almost all t ∈ (0, T), we have

12vk(τ)22+Ωψ(Bk(τ))+0τ(2Dvk22+Tvk2,Ω2+(1β)Bk12BkBk1222+βBk22+(βδ1+(1β)δ2)BkI22+(1β)δ1Bk12Bk1222+βδ2Bk32Bk1222)12v022+Ωψ(B0)+0τf,vk.

The attainment of initial conditions is standard (see the last section for details in a more complicated case).

3.4 Limit k → ∞

Since we start from the same a~priori estimates as in the previous section, we follow, step by step, the procedure developed when taking the limit ε → 0+. The only difference is that the term ρε(𝔹ε) is not present. Thus, using the density of {wi}i=1 in Wn,div3,2 , we obtain, after letting k → ∞, for almost all t ∈ (0, T), that

tv,φ+((v)v,φ)+2(Dv,φ)=(Tv,Tφ)Ω2a(S(B),φ)+f,φfor all φWn,div3,2 (3.42)

and that

tB,A+((v)B,A)+(B,A)=2(B(aDvWv),A)(R(B),A)for all AW1,2(Ω).

Moreover, from the weak lower semi-continuity of norms, we obtain the energy inequality (2.5) for almost all t ∈ (0, T). Furthermore, the same argument as above implies that 𝔹 is positive definite a.e. in Q. Now observe that, by Hölder’s inequality and (3.35), all the terms in (3.42) except the first one, are integrable for every φL4(0,T;Wn,div1,2)L4(0,T;L6(Ω)). Indeed, for example for the non-linear terms, we get

Q|(v)vφ|vL4L3vL2L2φL4L6

and

Q|S(B)φ|CBL83L42φL4L2.

Hence, the functional tv can be uniquely extended to tvL43(0,T;Wn,div1,2) and we can use the density argument to conclude (2.2). Analogously, we obtain (2.3). Hence, it remains to show that (2.5) holds for all t ∈ (0, T) and that the initial data fulfil (2.4).

3.4.1 Energy inequality for all t ∈ (0, T)

First, we observe, that due to (3.34), (3.36) and (3.37), we have that

vCweak(0,T;L2(Ω))andBCweak(0,T;L2(Ω)). (3.43)

Next, we notice that the function ψ is convex on the convex set R>03×3 . Indeed, evaluating the second Fréchet derivative of ψ, we get

2ψ(A)A2=(1β)A1A1+βIIfor all AR>03×3,

which is obviously a positive definite operator for any β ∈ [0, 1] and consequently, ψ must be convex on R>03×3 .

Further, we integrate (2.5) over (t1, t1 + δ), where t1 ∈ (0, T), and divide the result by δ. Using also an elementary inequality

0t1g1δt1t1+δ(0tg)dt

valid for every integrable non-negative g, we get

12δt1t1+δv22+1δt1t1+δΩψ(B)+0t1(2Dv22+Tv2,Ω2+(1β)B12BB1222+βB22+(βδ1+(1β)δ2)BI22+(1β)δ1B12B1222+βδ2B32B1222)12v022+Ωψ(B0)+1δt1t1+δ0τf,v.

Finally, we let δ → 0+. The limit on the right hand side is standard and consequently, if we show that

12v(t1)22+Ωψ(B(t1))lim infδ0+1δt1t1+δ(v222+Ωψ(B)), (3.44)

then (2.5) will hold for all t ∈ (0, T). To show it, we notice that due to (3.43)

v(t)v(t1)weakly inL2(Ω)astt1,B(t)B(t1)weakly inL2(Ω)astt1, (3.45)

Consequently, due to the weak lower semicontinuity and the convexity of ψ we also have for all t ∈ (0, T)

Ω|v(t)|2+ψ(B(t))C.

Hence denoting by ΩMΩ the set where |v(t1, ⋅)| + |𝔹(t1, ⋅)| + |𝔹–1(t1, ⋅)| ≤ M, it follows from the previous estimate that |ΩΩM| → 0 as M → ∞. Hence, since ψ is nonnegative and convex, we have for all t ∈ (t1, t1 + δ) that

Ω|v(t)|22+ψ(B(t))ΩM|v(t)|22+ψ(B(t))ΩM|v(t1)|22+ψ(B(t1))+ΩMv(t1)(v(t)v(t1))+ψ(B(t1))B(B(t)B(t1)).

Since, v(t1) and 𝔹ψ(𝔹(t1)) are bounded on ΩM, we can integrate the above estimate over (t1, t1 + δ) and it follows from (3.45) that

lim infδ0+1δt1t1+δΩ|v|22+ψ(B)ΩM|v(t1)|22+ψ(B(t1)).

Hence, letting M → ∞, we deduce (3.44) and the proof of (2.5) is complete.

3.4.2 Attainment of initial conditions

First, it is standard to show from the construction and from the weak continuity (3.45), that for arbitrary φ, 𝔸 ∈ L2(Ω) there holds

limt0+(v(t),φ)=(v0,φ)andlimt0+(B(t),A)=(B0,A). (3.46)

Next, using the convexity of ψ and (3.46) (and consequently weak lower semicontinuity of the corresponding integral) and letting t → 0+ in (2.5), we deduce that

v022+2Ωψ(B0)lim inft0+v(t)22+2Ωψ(B(t))lim supt0+v(t)22+2Ωψ(B(t))v022+2Ωψ(B0). (3.47)

We claim that this implies that

v022=limt0+v(t)22andΩψ(B0)=limt0+Ωψ(B(t)). (3.48)

Indeed, assume for a moment that

v022<lim inft0+v(t)22.

But then it follows from (3.47) that

Ωψ(B0)>lim inft0+Ωψ(B(t)),

which contradicts (3.46) and convexity of ψ. Consequently, (3.48) holds.

It directly follows from (3.46)1 and (3.48)1 that

limt0+v(t)v022=0.

To claim the same result also for 𝔹, we simply split ψ as follows

ψ(A)=β2|AI|2+(1β)(trA3lndetA)=:βψ1(A)+(1β)ψ2(A).

Similarly as above, it is easy to observe that ψ1 as well as ψ2 are convex on the set of positive definite matrices. Therefore, (3.48)2 and (3.46)2 imply

Ω|B0I|2=2Ωψ1(B0)=2limt0+Ωψ1(B(t))=limt0+Ω|B(t)I|2,Ωψ2(B0)=limt0+Ωψ2(B(t)). (3.49)

Finally, (3.46) and (3.49)1 lead to

limt0+B(t)B022=limt0+(B(t)I)+(IB0)22=limt0+B(t)I22+B0I222Ω(B(t)I)(B0I)=0,

which finishes the proof of (2.4) and consequently also the proof of the Theorem.

Acknowledgements

Michal Bathory has been supported by Charles University Research program UNCE/SCI/023 and by the project No. 1652119 financed by the Charles University Grant Agency (GAUK). Also, Michal Bathory appreciates the hospitality provided by the Callery family from Sligo, where the essence of this work was created. Miroslav Bulíček and Josef Málek acknowledge the support of the project No. 18-12719S financed by Czech science foundation (GAČR). Miroslav Bulíček and Josef Málek are members of the Nečas center for mathematical modelling.

References

[1] Y. Amirat, D. Bresch, J. Lemoine, J. Simon, Effect of rugosity on a flow governed by stationary Navier-Stokes equations, Quart. Appl. Math. 59 (2001), no. 4, 769–785.10.1090/qam/1866556Search in Google Scholar

[2] C. Amrouche, A. Rejaiba, Lp-theory for Stokes and Navier–Stokes equations with Navier boundary condition, J. Differential Equations 256 (2014), no. 4, 1515–1547.10.1016/j.jde.2013.11.005Search in Google Scholar

[3] C. Amrouche, N.E.H. Seloula, On the Stokes equations with the Navier-type boundary conditions, Differ. Equ. Appl. 3 (2011), no. 4, 581–607.10.7153/dea-03-36Search in Google Scholar

[4] H. Al Baba, Maximal Lp-Lq regularity to the Stokes problem with Navier boundary conditions, Adv. Nonlinear Anal. 8 (2019), no. 1, 743–761.10.1515/anona-2017-0012Search in Google Scholar

[5] J. Barrett, S. Boyaval, Existence and approximation of a (regularized) Oldroyd-B model, Math. Models Methods Appl. Sci. 21 (2011), no. 09, 1783–1837.10.1142/S0218202511005581Search in Google Scholar

[6] A. Basson, D. Gérard-Varet, Wall laws for fluid flows at a boundary with random roughness, Comm. Pure Appl. Math. 61 (2008), no. 7, 941–987.10.1002/cpa.20237Search in Google Scholar

[7] H. Beirão da Veiga, J. Yang, Regularity criteria for Navier-Stokes equations with slip boundary conditions on non-flat boundaries via two velocity components, Adv. Nonlinear Anal. 9 (2020), no. 1, 633–643.10.1515/anona-2020-0017Search in Google Scholar

[8] J. Blechta, J. Málek, K.R. Rajagopal, On the classifcation of incompressible fluids and a mathematical analysis of the equations that govern their motion, SIAM J. Math. Anal. 52 (2020), no. 2, 1232–1289.10.1137/19M1244895Search in Google Scholar

[9] D. Bucur, E. Feireisl, The incompressible limit of the full Navier-Stokes-Fourier system on domains with rough boundaries, Nonlinear Anal. Real World Appl. 10 (2009), no. 5, 3203–3229.10.1016/j.nonrwa.2008.10.024Search in Google Scholar

[10] M. Bulíček, E. Feireisl, J. Málek, On a class of compressible viscoelastic rate-type fluids with stress-diffusion, Nonlinearity 32 (2019), no. 12, 4665–4681.10.1088/1361-6544/ab3614Search in Google Scholar

[11] M. Bulíček, J. Málek, K.R. Rajagopal, Mathematical analysis of unsteady flows of fluids with pressure, shear-rate, and temperature dependent material moduli that slip at solid boundaries, SIAM J. Math. Anal. 41 (2009), no. 2, 665–707.10.1137/07069540XSearch in Google Scholar

[12] M. Bulíček, J. Málek, Internal flows of incompressible fluids subject to stick-slip boundary conditions, Vietnam J. Math. 45 (2017), no. 1-2, 207–220.10.1007/s10013-016-0221-zSearch in Google Scholar

[13] M. Bulíček, J. Málek, Large data analysis for Kolmogorov’s two-equation model of turbulence, Nonlinear Anal. Real World Appl. 50 (2019), 104–143.10.1016/j.nonrwa.2019.04.008Search in Google Scholar

[14] M. Bulíček, J. Málek, V. Průša, E. Süli, PDE analysis of a class of thermodynamically compatible viscoelastic rate-type fluids with stress-diffusion, Contemp. Math., vol. 710, Amer. Math. Soc., Providence, RI, 2018, pp. 25–51.10.1090/conm/710/14362Search in Google Scholar

[15] M. Bulíček, J. Málek, J. Žabenský, On generalized Stokes’ and Brinkman’s equations with a pressure-and shear-dependent viscosity and drag coefficient, Nonlinear Anal. Real World Appl. 26 (2015), 109–132.10.1016/j.nonrwa.2015.05.004Search in Google Scholar

[16] M. Bulíček, J. Žabenský, Large data existence theory for unsteady flows of fluids with pressure- and shear-dependent viscosities, Nonlinear Anal. 127 (2015), 94–127.10.1016/j.na.2015.07.001Search in Google Scholar

[17] L. Chupin, Global strong solutions for some differential viscoelastic models, SIAM J. Appl. Math. 78 (2018), no. 6, 2919–2949.10.1137/18M1186873Search in Google Scholar

[18] P. Constantin, M. Kliegl, Note on global regularity for two-dimensional Oldroyd-B fluids with diffusive stress, Arch. Ration. Mech. Anal. 206 (2012), no. 3, 725–740.10.1007/s00205-012-0537-0Search in Google Scholar

[19] M. Dostalík, V. Průša, T. Skřivan, On diffusive variants of some classical viscoelastic rate-type models, AIP Conference Proceedings 2107 (2019).10.1063/1.5109493Search in Google Scholar

[20] S.-Q. Wang, P.A. Drda, Stick–slip transition in capillary flow of polyethylene. 2. Molecular weight dependence and low temperature anomaly, Macromolecules 29 (11) (1996) 4115–4119.10.1021/ma951512eSearch in Google Scholar

[21] A.W. El-Kareh, L.G. Leal, Existence of solutions for all Deborah numbers for a non-Newtonian model modified to include diffusion, J. Non-Newton. Fluid Mech. 33 (1989), no. 3, 257–287.10.1016/0377-0257(89)80002-3Search in Google Scholar

[22] C. Guillopé, J.C. Saut, Existence results for the flow of viscoelastic fluids with a differential constitutive law, Nonlinear Anal. 15 (1990), no. 9, 849–869.10.1016/0362-546X(90)90097-ZSearch in Google Scholar

[23] S.G. Hatzikiriakos, Wall slip of molten polymers, Prog. Polym. Sci. 37 (2012) 624–643.10.1016/j.progpolymsci.2011.09.004Search in Google Scholar

[24] J. Hron, V. Miloš, V. Průša, O. Souček, K. Tůma, On thermodynamics of incompressible viscoelastic rate type fluids with temperature dependent material coefficients, Internat. J. Non-Linear Mech. 95 (2017), 193–208.10.1016/j.ijnonlinmec.2017.06.011Search in Google Scholar

[25] W. Jäger, A. Mikelić, On the roughness-induced effective boundary conditions for an incompressible viscous flow, J. Differential Equations 170 (2001), no. 1, 96–122.10.1006/jdeq.2000.3814Search in Google Scholar

[26] O. Kreml, M. Pokorný, P. Šalom, On the global existence for a regularized model of viscoelastic non-Newtonian fluid, Colloq. Math. 139 (2015), no. 2, 149–163.10.4064/cm139-2-1Search in Google Scholar

[27] J. Leray, Sur le mouvement d’un liquide visqueux emplissant l’espace, Acta Math. 63 (1934), no. 1, 193–248.10.1007/BF02547354Search in Google Scholar

[28] P.L. Lions, N. Masmoudi, Global solutions for some Oldroyd models of non-newtonian flows, Chin. Ann. Math. Ser. B 21 (2000), no. 2, 131–146.10.1142/S0252959900000170Search in Google Scholar

[29] M. Lukáčová-Medviďová, H. Mizerová, Š. Nečasová, M. Renardy, Global existence result for the generalized Peterlin viscoelastic model, SIAM J. Math. Anal. 49 (2017), no. 4, 2950–2964.10.1137/16M1068505Search in Google Scholar

[30] V. Mácha, J. Tichý, Higher integrability of solutions to generalized Stokes system under perfect slip boundary conditions, J. Math. Fluid Mech. 16 (2014), no. 4, 823–845.10.1007/s00021-014-0190-5Search in Google Scholar

[31] J. Málek, J. Nečas, M. Rokyta, M. Růžička, Weak and Measure-valued Solutions to Evolutionary PDEs, Chapman & Hall, 1996.10.1007/978-1-4899-6824-1Search in Google Scholar

[32] J. Málek, V. Průša, T. Skřivan, E. Süli, Thermodynamics of viscoelastic rate-type fluids with stress diffusion, Phys. Fluids 30 (2018).10.1063/1.5018172Search in Google Scholar

[33] J. Málek, V. Průša, Derivation of Equations for Continuum Mechanics and Thermodynamics of Fluids, Handbook of mathematical analysis in mechanics of viscous fluids, Springer, Cham, 2018, pp. 3–72.10.1007/978-3-319-13344-7_1Search in Google Scholar

[34] J. Málek, K.R. Rajagopal, K. Tůma, On a variant of the Maxwell and Oldroyd-B models within the context of a thermodynamic basis, Internat. J. Non-Linear Mech. 76 (2015), 42–47.10.1016/j.ijnonlinmec.2015.03.009Search in Google Scholar

[35] J. Málek, K.R. Rajagopal, K. Tůma, Derivation of the variants of the Burgers model using a thermodynamic approach and appealing to the concept of evolving natural configurations, Fluids 3 (2018), no. 4.10.3390/fluids3040069Search in Google Scholar

[36] E. Maringová, J. Žabenský, On a Navier-Stokes-Fourier-like system capturing transitions between viscous and inviscid fluid regimes and between no-slip and perfect-slip boundary conditions, Nonlinear Anal. Real World Appl. 41 (2018), 152–178.10.1016/j.nonrwa.2017.10.008Search in Google Scholar

[37] N. Masmoudi, Global existence of weak solutions to macroscopic models of polymeric flows, J. Math. Pures Appl. (9) 96 (2011), no. 5, 502–520.10.1016/j.matpur.2011.04.008Search in Google Scholar

[38] K.R. Rajagopal, A.R. Srinivasa, A thermodynamic frame work for rate type fluid models, J. Non-Newton. Fluid Mech. 88 (2000), no. 3, 207–227.10.1016/S0377-0257(99)00023-3Search in Google Scholar

[39] K.R. Rajagopal, A.R. Srinivasa, On thermomechanical restrictions of continua, Proc. R. Soc. Lond., Ser. A, Math. Phys. Eng. Sci. 460 (2004), no. 2042, 631–651.10.1098/rspa.2002.1111Search in Google Scholar

Received: 2020-02-27
Accepted: 2020-06-29
Published Online: 2020-09-02

© 2021 Michal Bathory et al., published by De Gruyter

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

Articles in the same Issue

  1. Editorial
  2. Editorial to Volume 10 of ANA
  3. Regular Articles
  4. Convergence Results for Elliptic Variational-Hemivariational Inequalities
  5. Weak and stationary solutions to a Cahn–Hilliard–Brinkman model with singular potentials and source terms
  6. Single peaked traveling wave solutions to a generalized μ-Novikov Equation
  7. Constant sign and nodal solutions for superlinear (p, q)–equations with indefinite potential and a concave boundary term
  8. On isolated singularities of Kirchhoff equations
  9. On the existence of periodic oscillations for pendulum-type equations
  10. Multiplicity of concentrating solutions for a class of magnetic Schrödinger-Poisson type equation
  11. Nehari-type ground state solutions for a Choquard equation with doubly critical exponents
  12. Gradient estimate of a variable power for nonlinear elliptic equations with Orlicz growth
  13. The structure of 𝓐-free measures revisited
  14. Solvability of an infinite system of integral equations on the real half-axis
  15. Positive Solutions for Resonant (p, q)-equations with convection
  16. Concentration behavior of semiclassical solutions for Hamiltonian elliptic system
  17. Global existence and finite time blowup for a nonlocal semilinear pseudo-parabolic equation
  18. On variational nonlinear equations with monotone operators
  19. Existence results for nonlinear degenerate elliptic equations with lower order terms
  20. Blow-up criteria and instability of normalized standing waves for the fractional Schrödinger-Choquard equation
  21. Ground states and multiple solutions for Hamiltonian elliptic system with gradient term
  22. Positivity of solutions to the Cauchy problem for linear and semilinear biharmonic heat equations
  23. Convex solutions of Monge-Ampère equations and systems: Existence, uniqueness and asymptotic behavior
  24. Multiple solutions for critical Choquard-Kirchhoff type equations
  25. Regularity for sub-elliptic systems with VMO-coefficients in the Heisenberg group: the sub-quadratic structure case
  26. Inducing strong convergence of trajectories in dynamical systems associated to monotone inclusions with composite structure
  27. A posteriori analysis of the spectral element discretization of a non linear heat equation
  28. Liouville property of fractional Lane-Emden equation in general unbounded domain
  29. Large data existence theory for three-dimensional unsteady flows of rate-type viscoelastic fluids with stress diffusion
  30. On some classes of generalized Schrödinger equations
  31. Variational formulations of steady rotational equatorial waves
  32. On a class of critical elliptic systems in ℝ4
  33. Exponential stability of the nonlinear Schrödinger equation with locally distributed damping on compact Riemannian manifold
  34. On a degenerate hyperbolic problem for the 3-D steady full Euler equations with axial-symmetry
  35. Existence, multiplicity and nonexistence results for Kirchhoff type equations
  36. Combined effects of Choquard and singular nonlinearities in fractional Kirchhoff problems
  37. Convergence analysis for double phase obstacle problems with multivalued convection term
  38. Multiple solutions for weighted Kirchhoff equations involving critical Hardy-Sobolev exponent
  39. Boundary value problems associated with singular strongly nonlinear equations with functional terms
  40. Global solvability in a three-dimensional Keller-Segel-Stokes system involving arbitrary superlinear logistic degradation
  41. Multiplicity and concentration behaviour of solutions for a fractional Choquard equation with critical growth
  42. Concentration results for a magnetic Schrödinger-Poisson system with critical growth
  43. Periodic solutions for a differential inclusion problem involving the p(t)-Laplacian
  44. The concentration-compactness principles for Ws,p(·,·)(ℝN) and application
  45. Regularity for commutators of the local multilinear fractional maximal operators
  46. An improvement to the John-Nirenberg inequality for functions in critical Sobolev spaces
  47. Local versus nonlocal elliptic equations: short-long range field interactions
  48. Analysis of a diffusive host-pathogen model with standard incidence and distinct dispersal rates
  49. Blowing-up solutions of the time-fractional dispersive equations
  50. Fixed point of some Markov operator of Frobenius-Perron type generated by a random family of point-transformations in ℝd
  51. Non-stationary Navier–Stokes equations in 2D power cusp domain
  52. Non-stationary Navier–Stokes equations in 2D power cusp domain
  53. Nontrivial solutions to the p-harmonic equation with nonlinearity asymptotic to |t|p–2t at infinity
  54. Iterative methods for monotone nonexpansive mappings in uniformly convex spaces
  55. Optimality of Serrin type extension criteria to the Navier-Stokes equations
  56. Fractional Hardy-Sobolev equations with nonhomogeneous terms
  57. New class of sixth-order nonhomogeneous p(x)-Kirchhoff problems with sign-changing weight functions
  58. On the set of positive solutions for resonant Robin (p, q)-equations
  59. Solving Composite Fixed Point Problems with Block Updates
  60. Lions-type theorem of the p-Laplacian and applications
  61. Half-space Gaussian symmetrization: applications to semilinear elliptic problems
  62. Positive radial symmetric solutions for a class of elliptic problems with critical exponent and -1 growth
  63. Global well-posedness of the full compressible Hall-MHD equations
  64. Σ-Shaped Bifurcation Curves
  65. On the critical behavior for inhomogeneous wave inequalities with Hardy potential in an exterior domain
  66. On singular quasilinear elliptic equations with data measures
  67. On the sub–diffusion fractional initial value problem with time variable order
  68. Partial regularity of stable solutions to the fractional Geľfand-Liouville equation
  69. Ground state solutions for a class of fractional Schrodinger-Poisson system with critical growth and vanishing potentials
  70. Initial boundary value problems for the three-dimensional compressible elastic Navier-Stokes-Poisson equations
Downloaded on 4.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2020-0144/html
Scroll to top button