This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the existence of weak solutions for a family of unsteady rotational Smagorinsky models

Luigi C. Berselli Dipartimento di Matematica, Università di Pisa, Via F. Buonarroti 1/c, I-56127 Pisa, Italy luigi.carlo.berselli@unipi.it Alex Kaltenbach Institute of Applied Mathematics, Albert-Ludwigs-University Freiburg, Ernst-Zermelo-Straß 1, 79104 Freiburg, alex.kaltenbach@mathematik.uni-freiburg.de rose@mathematik.uni-freiburg.de Roger Lewandowski IRMAR, UMR 6625, Université Rennes 1, Campus Beaulieu, 35042 Rennes cedex FRANCE Roger.Lewandowski@univ-rennes1.fr  and  Michael Ružička
Abstract.

In this paper we show that the rotational Smagorinsky model for turbulent flows, can be put, for a wide range of parameters in the setting of Bochner pseudo-monotone evolution equations. This allows to prove existence of weak solutions a) identifying a proper functional setting in weighted spaces and b) checking some easily verifiable assumptions, at fixed time. We also will discuss the critical role of the exponents present in the model (power of the distance function and power of the curl) for what concerns the application of the theory of pseudo-monotone operators.

Key words and phrases:
Bochner pseudo-monotone operators, Rotational turbulence models, evolution equations
2010 Mathematics Subject Classification:
35Q35, 76F02, 47H05, 47J35

1. Introduction

In this paper we introduce the unsteady general rotational Smagorinsky model for incompressible turbulence

t𝐯¯+𝝎¯×𝐯¯+curl(Cαα|𝝎¯|𝝎¯)+q¯\displaystyle\partial_{t}\overline{\bf v}+\overline{\boldsymbol{\omega}}\times\overline{\bf v}+\operatorname{curl}\big{(}C_{\alpha}\ell^{\alpha}|\overline{\boldsymbol{\omega}}|\overline{\boldsymbol{\omega}}\big{)}+\nabla\overline{q} =𝐟\displaystyle={\bf f}    in (0,T)×Ω(0,T)\times\Omega, (1.1)
𝝎¯\displaystyle\overline{\boldsymbol{\omega}} =curl𝐯¯\displaystyle=\operatorname{curl}\overline{\bf v}    in (0,T)×Ω(0,T)\times\Omega,
div𝐯¯\displaystyle\mathrm{div}\,\overline{\bf v} =0\displaystyle=0    in (0,T)×Ω(0,T)\times\Omega,
𝐯¯\displaystyle\overline{\bf v} =𝟎\displaystyle=\mathbf{0}    on (0,T)×Ω(0,T)\times\partial\Omega,
𝐯¯(0)\displaystyle\overline{\bf v}(0) =𝐯0¯\displaystyle=\overline{{\bf v}_{0}}    in Ω\Omega,

where Ω\Omega is a smooth bounded domain in 3\mathbb{R}^{3}, \ell is the Prandtl mixing length, α>0\alpha>0 is a given exponent, Cα>0C_{\alpha}>0 is a calibration constant, 𝐯¯\overline{\bf v} is the mean velocity, 𝝎¯\overline{\boldsymbol{\omega}} is the mean vorticity, and q¯\overline{q} is the sum of the Bernoulli pressure of the fluid and certain potentials such as the turbulent kinetic energy and others. Here and in the sequel, for each smooth vector 𝐮:33{\bf u}:\mathbb{R}^{3}\to\mathbb{R}^{3} we define as curl the vector

(curl𝐮)i:=j,k=13ϵijkujxk,(\operatorname{curl}{\bf u})_{i}:=\sum_{j,k=1}^{3}\epsilon_{ijk}\frac{\partial u_{j}}{\partial x_{k}},

where ϵijk\epsilon_{ijk} is the Levi–Civita totally anti-symmetric tensor.

Note that in the equations (1.1) the linear dissipative term νΔ𝐯¯-\nu\Delta\overline{\bf v} is not present, since we are considering flows at very high Reynolds number, and then viscous effects are negligible compared to the Reynolds stresses. Note that the presence of the linear dissipative term, for any ν>0\nu>0 will allow for some simplifications of the proofs. Nevertheless, to obtain result and estimates independent of ν>0\nu>0, a treatment as the one we provide is requested.

According to standard assumptions (see (2.7) and (2.8) in Section 2.3), we will assume that \ell behaves as the distance to the boundary. This means that (𝐱)d(𝐱)\ell({\bf x})\approx d({\bf x}) when 𝐱Ω{\bf x}\in\Omega and 𝐱{\bf x} is close to the boundary Ω\partial\Omega (see (2.9) and (2.10) in what follows). As it is common in turbulence modeling, we assume that the flow fields are stochastic processes, and the bar operator stands for the expectation in the Reynolds decomposition 𝐯=𝐯¯+𝐯{\bf v}=\overline{\bf v}+{\bf v}^{\prime}, π=π¯+π\pi=\overline{\pi}+\pi^{\prime}, where π\pi denotes the pressure and π¯\overline{\pi} the mean pressure (see Section 2.1 below, even if other choices are possible, as for instance denoting by the bar operator the long time-averaging).

The natural value of the parameter α\alpha is equal to 2 and this model is similar to the widely used Smagorinsky model, but the term div(Cs2|D𝐯¯|D𝐯¯)\hbox{div}(C_{s}\ell^{2}|D\overline{\bf v}|D\overline{\bf v}) is replaced here by curl(C22|𝝎¯|𝝎¯)\operatorname{curl}(C_{2}\ell^{2}|\overline{\boldsymbol{\omega}}|\overline{\boldsymbol{\omega}}). The equivalence between both models can be understood for homogeneous isotropic turbulence, by the equality of the enstrophy |𝝎|2¯\overline{|\boldsymbol{\omega}|^{2}} to the total mean deformation 2|D𝐯|2¯2\overline{|D{\bf v}|^{2}}. The equivalence can be obtained by a straightforward generalization of [9, Lemma 4.7]. Then, according to this equality, in [9, Section 5.5.1] it is proved that the 5/3-5/3 Kolmogorov law yields to express the eddy viscosity as νT=C22|𝝎¯|\nu_{T}=C_{2}\ell^{2}|\overline{\boldsymbol{\omega}}|. The rotational structure of the eddy diffusion is a peculiarity of the model which is suitable for high-speed flows with thin attached boundary-layers. The mathematical treatment of rotational models is one of the main theoretical contribution of this paper.

The numerical performance of this model in the steady state case has been initially tested by Baldwin and Lomax [2], so that this model is also known as the Baldwin–Lomax model. Numerical analysis foundations also in the statistical non-equilibrium setting can be found in [23].

The analytical properties of a steady version of this model have been recently studied in [3] in the setting of weighted Sobolev spaces. Some unsteady versions, with the presence of a dispersive term –which allows for a more classical treatment– have been recently studied in [21, 5].

The steady version can be treated within the standard theory of monotone operators, plus a localization argument, while the unsteady one requires a more delicate argument to deal with the precise choice of spaces and formulation of the problem. As we will prove, a proper definition of the functional setting will make system (1.1) to fit into the framework of evolution problems with Bochner pseudo-monotone operators, for which the theory have been recently developed by two of the authors in [13]. The theory developed in [13] represents an extension and an adaption to unsteady problems of the classical theory of pseudo-monotone operators from Brézis [6], [7], already described in the classical monograph of Lions [19]. Our main result is the following, which covers all possible positive powers of the distance function which are strictly smaller than the critical value α=2\alpha=2.

Theorem 1.1.

Let us suppose that (𝐱)=d(𝐱,Ω)\ell({\bf x})=d({\bf x},\partial\Omega) and let α[0,2)\alpha\in[0,2), 0<T<0<T<\infty, 𝐯0¯Lσ2(Ω)\overline{{\bf v}_{0}}\in L^{2}_{\sigma}(\Omega), and 𝐟L3/2(0,T;(W01,3(Ω,dα)){\bf f}\in L^{3/2}(0,T;(W^{1,3}_{0}(\Omega,d^{\alpha})^{*}). Then, there exists a weak solution to the initial boundary value problem (1.1) such that

𝐯¯C([0,T];Lσ2(Ω))L3(0,T;W0,σ1,3(Ω,dα)),\displaystyle\overline{\bf v}\in C([0,T];L^{2}_{\sigma}(\Omega))\cap L^{3}(0,T;W^{1,3}_{0,\sigma}(\Omega,d^{\alpha})),
and for all t[0,T]\displaystyle\text{and for all }t\in[0,T]
12𝐯¯(t)2+0tΩCαdα(𝐱)|𝝎¯(s,𝐱)|3𝑑𝐱𝑑s=12𝐯0¯2+0t𝐟,𝐯¯W01,3(Ω,dα)𝑑s.\displaystyle\frac{1}{2}\|\overline{\bf v}(t)\|^{2}+\int_{0}^{t}\int_{\Omega}C_{\alpha}d^{\alpha}({\bf x})|{\overline{\boldsymbol{\omega}}}(s,{\bf x})|^{3}\,d{\bf x}\,ds=\frac{1}{2}\|\overline{{\bf v}_{0}}\|^{2}+\int_{0}^{t}\langle{\bf f},\overline{\bf v}\rangle_{W^{1,3}_{0}(\Omega,d^{\alpha})}ds.

The limitation α<2\alpha<2 seems to be intrinsic to the problem due to the fact that dαd^{\alpha} is not anymore a Muckenhoupt weight for α2\alpha\geq 2 (cf. Definition 4.4). Hence, for α2\alpha\geq 2 most of the analytical properties may fail, since we cannot ensure that the quantity from the energy estimates controls the (weighted) full gradient of the solution. For values of α\alpha larger or equal than 22, even the weak formulation, the density of smooth functions, and the meaning of the boundary conditions may fail; the solution of the problem, if possible, would pass through the introduction of a more general setting, of very weak solutions.

In the last section we will also consider the existence for a family of problems with different powers of the vorticity in the turbulent stress tensor, still with the distance function raised to any exponent smaller than the critical one, cf. Thm. 5.5.


Plan of the paper. In Section 2 we derive the Rotational Smagorinsky model from a classical turbulence modeling process, in Section 3 we define the notion of Bochner pseudo-monotone operators and we recall the main result for general evolutionary problems, in Section 4 we recall the main results on weighted spaces, which will be used to properly formulate the problem. Next in the final Section 5 we show how the hypotheses apply to problem (1.1), for relevant choices of the weight functions and discuss generalization and critical values of the parameters.

2. Modeling

2.1. Reynolds decomposition

Let us consider the Navier–Stokes equations (NSE in the sequel) written with the convective term in the rotational formulation:

𝐯t+𝝎×𝐯νΔ𝐯+(π+|𝐯|22)\displaystyle{\bf v}_{t}+\boldsymbol{\omega}\times{\bf v}-\nu\Delta{\bf v}+\nabla\left(\pi+\dfrac{|{\bf v}|^{2}}{2}\right) =𝐟\displaystyle={\bf f}    in (0,T)×Ω(0,T)\times\Omega, (2.1)
𝝎\displaystyle\boldsymbol{\omega} =curl𝐯\displaystyle=\operatorname{curl}{\bf v}    in (0,T)×Ω(0,T)\times\Omega,
div𝐯\displaystyle\mathrm{div}\,{\bf v} =0\displaystyle=0    in (0,T)×Ω(0,T)\times\Omega,
𝐯\displaystyle{\bf v} =𝟎\displaystyle={\bf 0}    on (0,T)×Ω(0,T)\times\partial\Omega,
𝐯(0)\displaystyle{\bf v}(0) =𝐯0\displaystyle={\bf v}_{0}    in Ω\Omega,

where 𝐯=𝐯(t,𝐱,ω){\bf v}={\bf v}(t,{\bf x},\omega) is the velocity field, π=π(t,𝐱,ω)\pi=\pi(t,{\bf x},\omega) the pressure, 𝝎=curl𝐯\boldsymbol{\omega}=\operatorname{curl}{\bf v} the vorticity, (t,𝐱)+×Ω(t,{\bf x})\in\mathbb{R}_{+}\times\Omega, ωX(,P)\omega\in X({\mathcal{B}},P), where X(,P)X({\mathcal{B}},P) is a given probability space on the space of initial data.

For instance, if 𝐟=𝟎{\bf f}=\mathbf{0} (the argument can be adapted also to include a smooth enough external force) it holds that for each divergence-free element of 𝐯0H1/2(Ω){\bf v}_{0}\in H^{1/2}(\Omega) there exists a lower bound T=T(𝐯01/2)>0T=T(\|{\bf v}_{0}\|_{1/2})>0 for the life-span of the unique Fujita–Kato mild solution. Since the life-span can be estimated with the norm of the initial datum, by fixing X=B(0,R)¯H1/2(Ω){𝐯=0}X=\overline{B(0,R)}\subseteq H^{1/2}(\Omega)\cap\{\nabla\cdot{\bf v}=0\} for some R>0R>0, then the life-span is bounded from below by some TX>0T_{X}>0. This means that for each 𝐯0X{\bf v}_{0}\in X, there exists a unique 𝐯C(0,TX;H1/2(Ω)){\bf v}\in C(0,T_{X};H^{1/2}(\Omega)) solution of the NSE.

We then introduce PP, which is a probability measure on the Borel sets of XX. More specifically, PP can be constructed as limit of averages of Dirac measures as in [9] or the renormalized Lebesgue measure constructed from the Borel sets of XX. The final result does not depend on the choice of PP. Let us denote the expectation with a bar, hence

𝐯0¯=X𝐯0𝑑P(𝐯0),\overline{{\bf v}_{0}}=\int_{X}{\bf v}_{0}\,dP({\bf v}_{0}),

and

𝐯¯(t,𝐱)=X𝐯(t,𝐱,𝐯0)𝑑P(𝐯0)π¯(t,𝐱)=Xπ(t,𝐱,𝐯0)𝑑P(𝐯0).\overline{\bf v}(t,{\bf x})=\int_{X}{\bf v}(t,{\bf x},{\bf v}_{0})\,dP({\bf v}_{0})\qquad\overline{\pi}(t,{\bf x})=\int_{X}\pi(t,{\bf x},{\bf v}_{0})\,dP({\bf v}_{0}).

More generally, for any field Ψ=t𝐯,𝝎,𝐯,Δ𝐯,|𝐯|22\Psi=\partial_{t}{\bf v},\boldsymbol{\omega},\nabla{\bf v},\Delta{\bf v},{|{\bf v}|^{2}\over 2}..., we can define the statistical mean as

Ψ¯(t,𝐱)=XΨ(t,𝐱,𝐯0)𝑑P(𝐯0),\overline{\Psi}(t,{\bf x})=\int_{X}\Psi(t,{\bf x},{\bf v}_{0})\,dP({\bf v}_{0}),

and consequently we can perform the usual decomposition of Ψ\Psi as

Ψ=Ψ¯+Ψ,\Psi=\overline{\Psi}+\Psi^{\prime},

which is known as the Reynolds decomposition. The properties of the statistical averaging process imply (Reynolds rules) that for all Ψ,ΘX\Psi,\Theta\in X

tΨ¯=tΨ¯,Ψ¯=Ψ¯,Ψ¯=0,Ψ¯Θ¯=Ψ¯Θ¯,\partial_{t}\overline{\Psi}=\overline{\partial_{t}\Psi},\qquad\nabla\overline{\Psi}=\overline{\nabla\Psi},\qquad\overline{\Psi^{\prime}}=0,\qquad\overline{\overline{\Psi}\,\Theta}=\overline{\Psi}\,\overline{\Theta},

hence, taking the expectation of the NSE (2.1) yields

𝐯¯t+𝝎¯×𝐯¯+𝝎×𝐯¯νΔ𝐯¯+(π¯+|𝐯¯|22+|𝐯|22¯)\displaystyle\overline{{\bf v}}_{t}+\overline{\boldsymbol{\omega}}\times\overline{\bf v}+\overline{\boldsymbol{\omega}^{\prime}\times{\bf v}^{\prime}}-\nu\Delta\overline{\bf v}+\nabla\left(\overline{\pi}+{|\overline{\bf v}|^{2}\over 2}+\overline{|{\bf v}^{\prime}|^{2}\over 2}\right) =𝐟¯,\displaystyle=\overline{{\bf f}}, (2.2)
𝝎¯\displaystyle\overline{\boldsymbol{\omega}} =curl𝐯¯,\displaystyle=\operatorname{curl}\overline{\bf v},
div𝐯¯\displaystyle\mathrm{div}\,\overline{{\bf v}} =0,\displaystyle=0,
𝐯¯|Ω\displaystyle\overline{\bf v}|_{\partial\Omega} =𝟎,\displaystyle={\bf 0},
𝐯¯|t=0\displaystyle\overline{\bf v}|_{t=0} =𝐯0¯.\displaystyle=\overline{{\bf v}_{0}}.

The basic closure and modeling problems concern expressing 𝝎×𝐯¯\overline{\boldsymbol{\omega}^{\prime}\times{\bf v}^{\prime}} in terms of averaged variables.

2.2. Rotational Reynolds stress

When taking the expectation of the NSE with the convective term written in the usual form, we get the term div(𝐯𝐯¯)\hbox{div}(\overline{{\bf v}^{\prime}\otimes{\bf v}^{\prime}}). The quantity 𝝈(r)=𝐯𝐯¯\boldsymbol{\sigma}^{(\textsc{r})}=\overline{{\bf v}^{\prime}\otimes{\bf v}^{\prime}} is called the Reynolds stress and the Boussinesq assumption consists in assuming that

𝝈(r)=νTD𝐯¯,\boldsymbol{\sigma}^{(\textsc{r})}=-\nu_{T}D\overline{\bf v},

where νT0\nu_{T}\geq 0 is an eddy viscosity which remains to be determined and modeled in terms of 𝐯¯\overline{\bf v}. If we want to use such a Boussinesq assumption, we must express the turbulent stress (which is a vector in the rotational formulation)

𝐬:=𝝎×𝐯¯{\bf s}:=\overline{\boldsymbol{\omega}^{\prime}\times{\bf v}^{\prime}}

in terms of derivatives of mean quantities. This is similar to the approach used when modeling the more standard Reynolds stress tensor. We prove in what follows the following theorem

Theorem 2.1.

Assume that Ω\Omega is connected and of class C1C^{1}. Then, there exists a vector 𝐚(R)=𝐚(R)(t,𝐱){\bf a}^{\text{(R)}}={\bf a}^{(\text{R})}(t,{\bf x}) and a scalar potential Φ=Φ(t,𝐱)\Phi=\Phi(t,{\bf x}) such that

𝐯¯t+𝝎¯×𝐯¯+curl𝐚(R)νΔ𝐯¯+(π¯+12|𝐯¯|2+kΦ)\displaystyle\overline{{\bf v}}_{t}+\overline{\boldsymbol{\omega}}\times\overline{\bf v}+\operatorname{curl}{\bf a}^{(\hbox{\tiny R})}-\nu\Delta\overline{\bf v}+\nabla(\overline{\pi}+\tfrac{1}{2}|\overline{\bf v}|^{2}+k-\Phi) =𝐟¯,\displaystyle=\overline{{\bf f}}, (2.3)
div𝐯¯\displaystyle\mathrm{div}\,\overline{{\bf v}} =0,\displaystyle=0,
𝐯¯|Ω\displaystyle\overline{\bf v}|_{\partial\Omega} =𝟎,\displaystyle={\bf 0},
𝐯¯|t=0\displaystyle\overline{\bf v}|_{t=0} =𝐯0¯,\displaystyle=\overline{{\bf v}_{0}},

where k=12|𝐯|2¯k={1\over 2}\overline{|{\bf v}^{\prime}|^{2}} is the turbulent kinetic energy.

Proof.

Let 𝐚(R){\bf a}^{(\hbox{\tiny R})} and Φ\Phi be given by:

𝐚(R)(t,𝐱)\displaystyle{\bf a}^{(\hbox{\tiny R})}(t,{\bf x}) =14πΩcurl𝐬(t,𝐱)|𝐱𝐱|𝑑𝐱+14πΩ𝐬(t,𝐱)|𝐱𝐱|×𝑑σ(𝐱),\displaystyle={1\over 4\pi}\int_{\Omega}{\operatorname{curl}{\bf s}(t,{\bf x}^{\prime})\over|{\bf x}-{\bf x}^{\prime}|}d{\bf x}^{\prime}+{1\over 4\pi}\int_{\partial\Omega}{{\bf s}(t,{\bf x}^{\prime})\over|{\bf x}-{\bf x}^{\prime}|}\times d\sigma({\bf x}^{\prime}), (2.4)
Φ(t,𝐱)\displaystyle\Phi(t,{\bf x}) =14πΩdiv𝐬(t,𝐱)|𝐱𝐱|𝑑𝐱14πΩ𝐬(t,𝐱)|𝐱𝐱|𝑑σ(𝐱).\displaystyle={1\over 4\pi}\int_{\Omega}{\hbox{div}\,{\bf s}(t,{\bf x}^{\prime})\over|{\bf x}-{\bf x}^{\prime}|}d{\bf x}^{\prime}-{1\over 4\pi}\int_{\partial\Omega}{{\bf s}(t,{\bf x}^{\prime})\over|{\bf x}-{\bf x}^{\prime}|}\cdot d\sigma({\bf x}^{\prime}).

Therefore, by the Helmholtz–Hodge theorem, we have the relation

𝝎×𝐯¯=curl𝐚(R)Φ.\overline{\boldsymbol{\omega}^{\prime}\times{\bf v}^{\prime}}=\operatorname{curl}{\bf a}^{(\hbox{\tiny R})}-\nabla\Phi. (2.5)

Inserting (2.5) into (2.2) gives (2.3). ∎

The vector 𝐚(R){\bf a}^{(\hbox{\tiny R})}, which is continuously and uniquely determined by formula (2.4), is called the rotational Reynolds stress tensor. From now on we write q¯\bar{q} instead of π¯+12|𝐯¯|2+kΦ\overline{\pi}+\tfrac{1}{2}|\overline{\bf v}|^{2}+k-\Phi.

2.3. Closure assumption: Rotational Smagorinsky model

In order to finish the modeling of turbulent quantities, it remains to link 𝐚(R){\bf a}^{(\hbox{\tiny R})} to the mean vorticity 𝝎¯\overline{\boldsymbol{\omega}}. Notice that 𝐚(R){\bf a}^{(\hbox{\tiny R})} has the dimension of a squared velocity, while 𝝎¯\overline{\boldsymbol{\omega}} those of a frequency. Therefore, adapting the Boussinesq assumption to this case yields to assume

𝐚(R)=νT𝝎¯,{\bf a}^{(\hbox{\tiny R})}=\nu_{T}\,\overline{\boldsymbol{\omega}},

in which νT0\nu_{T}\geq{0} is a quantity with the dimensions of a viscosity. According to the 5/3-5/3 Kolmogorov law and following [9, Section 5.5.1], we can assume (for an homogeneous and isotropic flow, in the limit ν0\nu\to 0)

νT=νT(,|𝝎¯|),\nu_{T}=\nu_{T}(\ell,|\overline{\boldsymbol{\omega}}|),

where \ell is the Prandtl mixing length. The dimensional analysis of the expression shows that a consistent expression is

νT=Cω2|𝝎¯|,\nu_{T}=C_{\omega}\ell^{2}|\overline{\boldsymbol{\omega}}|, (2.6)

with CωC_{\omega} a dimensionless constant. This raises the question of the determination of \ell. In the case of a flow over a plate, one finds in Obukhov [22] the following classical law:

=(z)=κz,\ell=\ell(z)=\kappa z, (2.7)

where z0z\geq 0 is the distance from the plate and κ\kappa the von Kármán constant. The Van Driest formula [26] defines \ell by:

(z):=κz(1ez/A);\ell(z):=\kappa\,z\,(1-\mathrm{e}^{-z/A}); (2.8)

here AA depends on the oscillations of the plate and on the kinematic viscosity ν\nu, while z0z\geq 0 is again the distance from the plate.

According to these formula, we shall assume throughout the rest of the paper that the function :Ω¯+\ell:\,\overline{\Omega}\to\mathbb{R}^{+} is of class C2C^{2} and satisfies the two following properties:

a)\displaystyle a)\ (𝐱)d(𝐱,Ω)for 𝐱 close to Ω;\displaystyle\ell({\bf x})\approx d({\bf x},\partial\Omega)\qquad\text{for }{\bf x}\hbox{ close to }\,\partial\Omega; (2.9)
b)\displaystyle b)\ KΩ,K>0s.t.(𝐱)K>0𝐱K,\displaystyle\forall\ K\subset\subset\Omega,\,\,\exists\,\ell_{K}>0\quad\hbox{s.t.}\quad\ell({\bf x})\geq\ell_{K}>0\quad\forall\,{\bf x}\in K, (2.10)

where d(𝐱,Ω)d({\bf x},\partial\Omega) denotes the distance from the boundary. In practice, we could have directly assumed (𝐱)=d(𝐱)\ell({\bf x})=d({\bf x}), i.e.,

νT=Cωd2|𝝎¯|.\nu_{T}=C_{\omega}d^{2}|\overline{\boldsymbol{\omega}}|. (2.11)

2.4. Generalised Rotational Smagorinsky models by dimensional analysis

The analysis of the previous section can be put also in a more general framework of Large Eddy Simulation (LES) models, looking also at possible modifications of the parameters present in the expression of the turbulent (rotational) stress vector. Let 0>0\ell_{0}>0 be a typical length scale of the motion. For instance, in the case of a flow over a plate, one can take

0=νv,\ell_{0}={\nu\over v_{*}},

where ν\nu is the kinematic viscosity and vv_{*} is the so-called friction velocity (cf. [1]).

We consider (modulo introducing an appropriate non-dimensionalization of the equations) the following operator

curl(02αα|𝝎¯|𝝎¯)\operatorname{curl}\big{(}\ell_{0}^{2-\alpha}\ell^{\alpha}|\overline{\boldsymbol{\omega}}|\overline{\boldsymbol{\omega}}\big{)} (2.12)

with α[0,2]\alpha\in[0,2], which is degenerate at the boundary and for which the natural treatment is through scales of weighted Banach spaces.

We report some discussion about the relationships between the scaling of the weight and that of the power of the curl. In the framework of LES methods we show that even starting with

νT=02αα|𝝎¯|p2\nu_{T}=\ell_{0}^{2-\alpha}\ell^{\alpha}|\overline{\boldsymbol{\omega}}|^{p-2} (2.13)

this determines a link between powers α\alpha and pp. Nevertheless, in the last section we will also point out the limiting behavior of the exponent p=3p=3 present in model (1.1), when p=α+1p=\alpha+1.

If one thinks of a flow as composed of eddies of different sizes in different places, then in a region of large eddies the changes of velocity and its curl are both 𝒪(1)\mathcal{O}(1) of the typical distance. In a region of smaller eddies the velocity changes over a distance of 𝒪\mathcal{O}(eddy length scale), so the local deformation is 𝒪\mathcal{O}(1/eddy length scale), cf. [4, § 3.3.2]. Hence, the rotational Smagorinsky model introduces a turbulent viscosity νT=(Cδ)2|𝝎¯|\nu_{T}=(C\delta)^{2}|\overline{\boldsymbol{\omega}}|, where δ\delta is the (local) smallest resolved scale, such that

νT={𝒪(δ2)in regions where |𝝎¯|=𝒪(1),𝒪(δ)in the smallest resolved scale where |𝝎¯|=𝒪(δ1).\nu_{T}=\left\{\begin{aligned} &\mathcal{O}(\delta^{2})&&\qquad\text{in regions where }|\overline{\boldsymbol{\omega}}|=\mathcal{O}(1),\\ &\mathcal{O}(\delta)&&\qquad\text{in the smallest resolved scale where }|\overline{\boldsymbol{\omega}}|=\mathcal{O}(\delta^{-1}).\end{aligned}\right.

By extrapolation, motivated by experiments with central difference approximations to linear convection diffusion problems, the following alternate scaling has also been proposed (cf. [4] and Layton [17]) νT=(Cδ)p1|𝐃𝐯¯|p2\nu_{T}=(C\delta)^{p-1}|{\bf D}\overline{\bf v}|^{p-2}, and we consider here the rotational counterpart

νT=(Cδ)p1|𝝎¯|p2,1<p<,\nu_{T}=(C\delta)^{p-1}|\overline{\boldsymbol{\omega}}|^{p-2},\qquad 1<p<\infty,

which resembles general power laws for non-Newtonian fluids. The above choice of νT\nu_{T} satisfies

νT={𝒪(δp)in regions where |𝝎¯|=𝒪(1),𝒪(δ)in the smallest resolved scale where |𝝎¯|=𝒪(δ1),\nu_{T}=\left\{\begin{aligned} &\mathcal{O}(\delta^{p})&&\qquad\text{in regions where }|\overline{\boldsymbol{\omega}}|=\mathcal{O}(1),\\ &\mathcal{O}(\delta)&&\qquad\text{in the smallest resolved scale where }|\overline{\boldsymbol{\omega}}|=\mathcal{O}(\delta^{-1}),\end{aligned}\right.

The justification of the presence of the critical value p1p-1 as power of the distance function can be done directly by dimensional arguments as in [3]. In fact, recall that both 𝐯¯\nabla\overline{\bf v} and 𝝎¯\overline{\boldsymbol{\omega}} have dimensions T1T^{-1}, where TT is a time, and in (2.6) the turbulent viscosity νT=d2|𝝎¯|L2T1\nu_{T}=d^{2}|\overline{\boldsymbol{\omega}}|\sim L^{2}T^{-1} (where LL is a length) has the dimensions of a viscosity. This is the only way to identify (by using just a typical length and the vorticity) a quantity with the dimensions of a viscosity. Introducing as third parameter as the friction velocity vLT1v_{*}\sim LT^{-1}, one can consider more general combinations. The outcome is to find a turbulent eddy viscosity of the following form

νT=vθdα|𝝎¯|p2,\nu_{T}=v_{*}^{\theta}d^{\alpha}|\overline{\boldsymbol{\omega}}|^{p-2},

for some constants θ,α,p\theta,\,\alpha,\,p. It turns out (cf. [3]) that the dimensions of this quantity are νTLθ+αT2θp\nu_{T}\sim L^{\theta+\alpha}T^{2-\theta-p}, and to respect dimensions of the viscosity one has to fix

θ=3pandα=p1.\theta=3-p\qquad\text{and}\qquad\alpha=p-1.

A sound generalization of the rotational Smagorinsky model is then the one with rotational stress

𝐒(v,d,𝝎¯)=Cv3pdp1|𝝎¯|p2𝝎¯,\mathbf{S}(v_{*},d,\overline{\boldsymbol{\omega}})=Cv_{*}^{3-p}d^{p-1}|\overline{\boldsymbol{\omega}}|^{p-2}\overline{\boldsymbol{\omega}},

and, after re-scaling, one can assume Cv3p=1Cv_{*}^{3-p}=1. Note that, even for different values of pp, the power of the distance is always the critical one (in terms of analytical properties of the weight functions), since dp1Apd^{p-1}\not\in{A}_{p}, cf. Lemma 4.5.

In the last section we will show that from the point of view of mathematical properties, the turbulent eddy viscosity

νT=dp1|𝝎¯|p2,\nu_{T}=d^{p-1}|\overline{\boldsymbol{\omega}}|^{p-2},

can be handled in terms of an existence theory by (pseudo)monotone operators only for p3p\geq 3. Hence, the exponent p=3p=3 plays for the weighted rotational operators, the same role that the exponent p=11/5p=11/5 plays for the usual pp-NSE with stress tensor S(𝐃𝐯¯)=c|𝐃𝐯¯|p2S({\bf D}\overline{\bf v})=c|{\bf D}\overline{\bf v}|^{p-2}.

From now and so far no risk of confusion occurs, we do not write the bar anymore.

3. Evolution equations in an abstract setting

As already claimed in the introduction, a proper setting to the rotational Smagorinsky model is that of pseudo-monotone evolution problems so we briefly recall the abstract existence result we will use on the sequel.

For the convenience of the reader, we recall the following known definition.

Definition 3.1.

Let XX, YY be Banach spaces. An operator A:XY{A:X\rightarrow Y} is called

  1. (i)

    bounded, if for all bounded MXM\subseteq X, the image A(M)YA(M)\subseteq Y is bounded.

  2. (ii)

    coercive, if Y=XY=X^{*} and limxXAx,xXxX=\lim\limits_{\|x\|_{X}\to\infty}\frac{\langle Ax,x\rangle_{X}}{\|x\|_{X}}=\infty.

  3. (iii)

    pseudo-monotone, if Y=XY=X^{*} and for any sequence (xn)nX{(x_{n})_{n\in\mathbb{N}}\subseteq X} from

    xnnx in X,\displaystyle x_{n}\overset{n\rightarrow\infty}{\rightharpoonup}x\text{ in }X,
    lim supnAxn,xnxX0,\displaystyle\limsup_{n\rightarrow\infty}{\langle Ax_{n},x_{n}-x\rangle_{X}}\leq 0,

    it follows that Ax,xyXlim infnAxn,xnyX\langle Ax,x-y\rangle_{X}\leq\liminf_{n\rightarrow\infty}{\langle Ax_{n},x_{n}-y\rangle_{X}} for all yX{y\in X}.

It is well-known that for each fXf\in X^{*}, the steady problem Ax=fAx=f admits a solution if AA is bounded, coercive and pseudo-monotone, see  [6], [7]. A typical example of a pseudo-monotone operator is the sum of a hemi-continuous monotone and a compact operator. Recently, two of the authors in [13] developed an abstract framework for evolution problems, by using the concepts of Bochner pseudo-monotone and Bochner coercive operators to generalize the ideas of [19, Sec. 2.5][16][11][12] and [24]. We want to access this theory for our concrete example. Therefore, for the remainder of this section, we assume that (V,H,id)(V,H,\textrm{id}) is an evolution triple, i.e., VV is a separable, reflexive Banach space, HH a separable Hilbert space and VV embeds densely into HH. For I:=(0,T){I:=\left(0,T\right)}, T(0,)T\in\left(0,\infty\right), and p(1,){p\in\left(1,\infty\right)}, we set

𝒳¬𝒱and𝒴¬\mathbfcal{X}:=L^{p}(I,V)\qquad\text{and}\qquad\mathbfcal{Y}:=L^{\infty}(I,H).

In this framework we have the following notion of a time derivative.

Definition 3.2.

A function 𝐮𝒳\mathbf{u}\in\mathbfcal{X} has a generalized time derivative if there exists a function 𝐰Lp(I,V){\mathbf{w}\in L^{p^{\prime}}(I,V^{*})} such that

I(𝐮(s),v)Hφ(s)𝑑s=I𝐰(s),vVφ(s)𝑑s\displaystyle-\int_{I}{(\mathbf{u}(s),v)_{H}\varphi^{\prime}(s)\,ds}=\int_{I}{\langle\mathbf{w}(s),v\rangle_{V}\varphi(s)\,ds}

for every vVv\in V and φC0(I)\varphi\in C_{0}^{\infty}(I). Since such a function is unique, d𝐮dt:=𝐰{\frac{d\mathbf{u}}{dt}:=\mathbf{w}} is well-defined. By

𝒲¬𝒲𝒱𝒱¬{𝒳𝒱}\displaystyle\mathbfcal{W}:=W^{1,p,p^{\prime}}(I,V,V^{*}):=\big{\{}\mathbf{u}\in\mathbfcal{X}\mid\exists\,\tfrac{d\mathbf{u}}{dt}\in L^{p^{\prime}}(I,V^{*})\big{\}},

we denote the Bochner–Sobolev space with respect to the evolution triple (V,H,id)(V,H,\textrm{id}).

In the context of evolutionary problems, the following generalized notions of pseudo-monotonicity and coercivity (cf. Definition 3.1) are particularly relevant and useful.

Definition 3.3 (Bochner pseudo-monotonicity).

An operator 𝒜¬𝒳𝒴𝒳{\mathbfcal{A}:\mathbfcal{X}\cap\mathbfcal{Y}\rightarrow\mathbfcal{X}^{*}} is said to be Bochner pseudo-monotone if for a sequence (𝐮n)n𝒳𝒴{(\mathbf{u}_{n})_{n\in\mathbb{N}}\subseteq\mathbfcal{X}\cap\mathbfcal{Y}} from

𝐮nn\displaystyle\mathbf{u}_{n}\overset{n\rightarrow\infty}{\rightharpoonup} 𝐮\displaystyle\mathbf{u}\quad in 𝒳\displaystyle\text{ in }\mathbfcal{X}.
𝐮n\displaystyle\mathbf{u}_{n}\;\;\overset{\ast}{\rightharpoondown}\;\; 𝐮\displaystyle\mathbf{u} in 𝒴\\displaystyle\text{ in }\mathbfcal{Y}\quad(n\rightarrow\infty),
𝐮n(t)n\displaystyle\mathbf{u}_{n}(t)\overset{n\rightarrow\infty}{\rightharpoonup} 𝐮(t)\displaystyle\mathbf{u}(t)\quad in Hfor a.e. tI,\displaystyle\text{ in }H\quad\text{for a.e. }t\in I,

and

lim supn𝒜\\𝒳0,\displaystyle\limsup_{n\rightarrow\infty}{\langle\mathbfcal{A}\mathbf{u}_{n},\mathbf{u}_{n}-\mathbf{u}\rangle_{\mathbfcal{X}}}\leq 0,

it follows that 𝒜𝒳lim inf\𝒜\\𝒳{\langle\mathbfcal{A}\mathbf{u},\mathbf{u}-\mathbf{v}\rangle_{\mathbfcal{X}}\leq\liminf_{n\rightarrow\infty}{\langle\mathbfcal{A}\mathbf{u}_{n},\mathbf{u}_{n}-\mathbf{v}\rangle_{\mathbfcal{X}}}} for every 𝐯𝒳{\mathbf{v}\in\mathbfcal{X}}.

Definition 3.4 (Bochner coercivity).

An operator 𝒜¬𝒳𝒴𝒳{\mathbfcal{A}:\mathbfcal{X}\cap\mathbfcal{Y}\rightarrow\mathbfcal{X}^{*}} is called:

  1. (i)

    Bochner coercive with respect to 𝐟𝒳\mathbf{f}\in\mathbfcal{X}^{*} and u0Hu_{0}\in H if there is a constant M:=M(𝐟,u0,𝒜M:=M(\mathbf{f},u_{0},\mathbfcal{A})>0 such that for every 𝐮𝒳𝒴{\mathbf{u}\in\mathbfcal{X}\cap\mathbfcal{Y}} from

    12𝐮(t)H2+𝒜{χ𝒳 for a.e. \displaystyle\tfrac{1}{2}\|\mathbf{u}(t)\|_{H}^{2}+\langle\mathbfcal{A}\mathbf{u}-\mathbf{f},\mathbf{u}\chi_{\left[0,t\right]}\rangle_{\mathbfcal{X}}\leq\tfrac{1}{2}\|u_{0}\|_{H}^{2}\quad\text{ for a.e. }t\in I,

    it follows that 𝐮𝒳𝒴=𝐮𝒳+𝐮𝒴M\|\mathbf{u}\|_{\mathbfcal{X}\cap\mathbfcal{Y}}=\|\mathbf{u}\|_{\mathbfcal{X}}+\|\mathbf{u}\|_{\mathbfcal{Y}}\leq M.

  2. (ii)

    Bochner coercive if it is Bochner coercive with respect to 𝐟\mathbf{f} and u0u_{0}, for every 𝐟𝒳\mathbf{f}\in\mathbfcal{X}^{*} and u0Hu_{0}\in H.

The critical role of the above definitions is that they identify a vast class of problems for which existence can be established. In fact, if 𝒜¬𝒳𝒴𝒳{\mathbfcal{A}:\mathbfcal{X}\cap\mathbfcal{Y}\to\mathbfcal{X}^{*}} is bounded, Bochner pseudo-monotone, and Bochner coercive, then the corresponding evolution problem d𝐮dt+𝒜{\frac{d{\bf u}}{dt}+\mathbfcal{A}{\bf u}={\bf f} is solvable for any initial datum u0Hu_{0}\in H. This result was recently obtained in [13, Thm. 4.1].

This result is particularly relevant since the difficulty is then shifted to the verification of the properties of induced operators, which can be performed time-by-time in the known steady setting. We will not describe the full result, but we propose a particular, simplified setting enough to solve (1.1).

The existence result is mainly based on the following proposition giving sufficient conditions which have to be checked at any fixed time slice tI{t\in I} and which is a particular case of [13, Prop. 3.13].

Proposition 3.5.

Let A:VVA:V\to V^{*} be an operator. Assume that there exists a number p(1,)p\in(1,\infty) and constants c0,c1>0c_{0},c_{1}>0 such that111 For a pseudo-monotone operator A:XXA:X\to X^{*} (local) boundedness implies demi-continuity, i.e., xnxx_{n}\to x in XX (n)(n\to\infty) implies AxnAxAx_{n}\rightharpoonup Ax in XX^{*} (n)(n\to\infty), hence we do not need here to make any further assumptions of demi-continuity.:

(C.1):

For every vVv\in V there holds

AvVc0vVp1.\displaystyle\left\|Av\right\|_{V^{*}}\leq c_{0}\|v\|_{V}^{p-1}.
(C.2):

A:VVA:V\rightarrow V^{*} is pseudo-monotone.

(C.3):

For every vVv\in V there holds

Av,vVc1vVp.\displaystyle\langle Av,v\rangle_{V}\geq c_{1}\|v\|_{V}^{p}.

Then, the induced operator 𝒜¬𝒳𝒴𝒳\mathbfcal{A}:\mathbfcal{X}\cap\mathbfcal{Y}\rightarrow\mathbfcal{X}^{*}, for all 𝐮𝒳𝒴\mathbf{u}\in\mathbfcal{X}\cap\mathbfcal{Y} and 𝐯𝒳\mathbf{v}\in\mathbfcal{X} defined by

𝒜𝒳¬𝒜𝒱\displaystyle\langle\mathbfcal{A}\mathbf{u},\mathbf{v}\rangle_{\mathbfcal{X}}:=\int_{I}{\langle A(\mathbf{u}(t)),\mathbf{v}(t)\rangle_{V}\,dt},

is well-defined, bounded, Bochner pseudo-monotone, and Bochner coercive.

On the basis of Proposition 3.5, we immediately obtain the following existence result, which will be used to study the families of rotational models just checking that the conditions (C.1)–(C.3) are satisfied, after a proper choice of the functional setting.

Theorem 3.6.

Let A:VVA:V\to V^{*} be an operator satisfying (C.1)–(C.3). Then, for arbitrary u0Hu_{0}\in H and 𝐟Lp(I,V){\mathbf{f}\in L^{p^{\prime}}(I,V^{*})}, there exists a solution 𝐮𝒲\mathbf{u}\in\mathbfcal{W} of the evolution equation

Id𝐮dt(t)+A(𝐮(t)),𝐯(t)V\displaystyle\int_{I}\Big{\langle}\frac{d\mathbf{u}}{dt}(t)+A(\mathbf{u}(t)),\mathbf{v}(t)\Big{\rangle}_{V} =I𝐟(t),𝐯(t)V𝑑t𝐯𝒳\displaystyle=\int_{I}{\langle\mathbf{f}(t),\mathbf{v}(t)\rangle_{V}\,dt}\quad\forall\mathbf{v}\in\mathbfcal{X},
𝐮c(0)\displaystyle\mathbf{u}_{c}(0) =u0 in H.\displaystyle=u_{0}\quad\text{ in }H.

Here, the initial condition has to be understood in the sense of the unique continuous representation 𝐮cC0(I¯,H)\mathbf{u}_{c}\in C^{0}(\overline{I},H) of 𝐮𝒲\mathbf{u}\in\mathbfcal{W} (cf. [27, Prop. 23.23]).

4. Weighted spaces

Since (1.1) is a boundary value problem with the principal part given by a space dependent (and degenerate at the boundary) operator, a natural functional setting would be that of weighted Sobolev spaces. Apart from classical Lebesgue and Sobolev spaces, we will use their weighted counterparts. We follow the notation from the classical book of Kufner et al. [14].

A weight ϱ\varrho on n\mathbb{R}^{n} is a locally integrable function satisfying almost everywhere 0<ϱ(𝐱)<0<\varrho({\bf x})<\infty. The weighted space Lp(Ω,ϱ)L^{p}(\Omega,\varrho), 1<p<1<p<\infty, is defined as follows

Lp(Ω,ϱ):={𝐟:Ωn measurable |Ω|𝐟(𝐱)|pϱ(𝐱)d𝐱<}.L^{p}(\Omega,\varrho):=\Big{\{}{\bf f}:\ \Omega\to\mathbb{R}^{n}\text{ measurable }{\,\big{|}\,}\int_{\Omega}|{\bf f}({\bf x})|^{p}\,\varrho({\bf x})\,\mathrm{d}{\bf x}<\infty\Big{\}}.

For p>1p>1 we have by using Hölder’s inequality that

ϱ1/(p1)Lloc1(Ω)Lp(Ω,ϱ)Lloc1(Ω)𝒟(Ω),\varrho^{-1/(p-1)}\in L^{1}_{\textup{loc}}(\Omega)\quad\Rightarrow\quad L^{p}(\Omega,\varrho)\subset L^{1}_{\textup{loc}}(\Omega)\subset\mathcal{D}^{\prime}(\Omega),

allowing to work in the standard setting of distributions. It turns out that C0(Ω)C^{\infty}_{0}(\Omega) is dense in Lp(Ω,ϱ)L^{p}(\Omega,\varrho) if the weight satisfies ϱ1p1Lloc1(n)\varrho^{\frac{-1}{p-1}}\in L^{1}_{\textup{loc}}(\mathbb{R}^{n}), see [14]. In addition, Lp(Ω,ϱ)L^{p}(\Omega,\varrho) is a Banach space when equipped with the norm

𝐟p,ϱ:=(Ω|𝐟(𝐱)|pϱ(𝐱)d𝐱)1/p.\|{\bf f}\|_{p,\varrho}:=\bigg{(}\int_{\Omega}|{\bf f}({\bf x})|^{p}\varrho({\bf x})\,\mathrm{d}{\bf x}\bigg{)}^{1/p}.

Next, we define weighted Sobolev spaces

Wk,p(Ω,ϱ):={𝐟:Ωn|Dα𝐟Lp(Ω,ϱ) for all α s.t. |α|k},W^{k,p}(\Omega,\varrho):=\left\{{\bf f}:\ \Omega\to\mathbb{R}^{n}{\,\big{|}\,}D^{\alpha}{\bf f}\in L^{p}(\Omega,\varrho)\text{ for all }\alpha\text{ s.t. }|\alpha|\leq k\right\},

equipped with the norm

𝐟k,p,ϱ:=(|α|kDα𝐟p,ϱp)1/p,\|{\bf f}\|_{k,p,\varrho}:=\Bigg{(}\sum_{|\alpha|\leq k}\|D^{\alpha}{\bf f}\|_{p,\varrho}^{p}\Bigg{)}^{1/p},

and, as usual, we define W0k,p(Ω,ϱ)W^{k,p}_{0}(\Omega,\varrho) as follows

W0k,p(Ω,ϱ):={ϕC0(Ω)}¯.k,p,ϱ.W^{k,p}_{0}(\Omega,\varrho):=\overline{\left\{{\boldsymbol{\phi}}\in C^{\infty}_{0}(\Omega)\right\}}^{\|\,.\,\|_{k,p,\varrho}}.

In our application the weight ϱ(𝐱)\varrho({\bf x}) will be a power of the distance d(𝐱)0d({\bf x})\geq 0 of the point 𝐱Ω{\bf x}\in\Omega from the boundary Ω\partial\Omega. Consequently, we specialize to this setting and give specific notions regarding these so-called power-type weights, see Kufner [14]. First, it turns out that Wk,p(Ω,dα)W^{k,p}(\Omega,d^{\alpha}) is a separable Banach space provided α\alpha\in\mathbb{R}, kk\in\mathbb{N} and 1p<1\leq p<\infty. In this special setting, since d(𝐱)CK>0d({\bf x})\geq C_{K}>0 for each compact KΩK\subset\subset\Omega, several results are stronger or more precise due to the inclusion Lp(Ω,dα)Llocp(Ω)L^{p}(\Omega,d^{\alpha})\subset L^{p}_{\textup{loc}}(\Omega), valid for all α\alpha\in\mathbb{R}.

We recall the following classical result about the distance function (cf. [14]).

Lemma 4.1.

Let Ω\Omega be a domain of class C0,1C^{0,1}, which means that in a small enough neighborhood ΩP\Omega_{P}, for PΩP\in\partial\Omega, the boundary ΩΩP\partial\Omega\cap\Omega_{P} can be expressed (after a rigid rotation) as x3=a(x1,x2)x_{3}=a(x_{1},x_{2}) for Lipschitz continuous aa. Then, there exist constants 0<c0,c10<c_{0},c_{1}\in\mathbb{R} such that

c0d(𝐱)|a(x)x3|c1d(𝐱)𝐱=(x,x3)ΩP.c_{0}\,d({\bf x})\leq|a(x^{\prime})-x_{3}|\leq c_{1}\,d({\bf x})\qquad\forall\,{\bf x}=(x^{\prime},x_{3})\in\Omega_{P}.

One of the most relevant properties of the distance function is that the following embedding holds true

Lp(Ω,dα)L1(Ω)ifα<p1.L^{p}(\Omega,d^{\alpha})\subset L^{1}(\Omega)\quad\text{if}\quad\alpha<p-1. (4.1)

It follows directly from Hölder’s inequality

Ω|f|d𝐱=Ωdα/p|f|dα/p𝑑𝐱(Ωdα|f|p𝑑𝐱)1/p(Ωdαp/p𝑑𝐱)1/p,\int_{\Omega}|f|\,\mathrm{d}{\bf x}=\int_{\Omega}d^{\alpha/p}|f|d^{-\alpha/p}d{\bf x}\leq\Big{(}\int_{\Omega}d^{\alpha}|f|^{p}d{\bf x}\Big{)}^{1/p}\Big{(}\int_{\Omega}d^{-\alpha p^{\prime}/p}d{\bf x}\Big{)}^{1/p^{\prime}},

and using Lemma 4.1 the latter integral is finite if and only if

αpp=αp1<1.\frac{\alpha\,p^{\prime}}{p}=\frac{\alpha}{p-1}<1.

In the same way we have also that

α[0,p1[Lp(Ω,dα)Lq(Ω)q[1,p1+α[.\forall\,\alpha\in[0,p-1[\qquad L^{p}(\Omega,d^{\alpha})\subset L^{q}(\Omega)\qquad\forall\,q\in\big{[}1,\frac{p}{1+\alpha}\big{[}. (4.2)

As in [14, Prop. 9.10] it can be shown that:

Lemma 4.2.

The quantity (Ωdα|𝐟|pd𝐱)1p\Big{(}\int_{\Omega}d^{\alpha}|\nabla{\bf f}|^{p}\,\mathrm{d}{\bf x}\Big{)}^{\frac{1}{p}} is an equivalent norm in W01,p(Ω,dα)W^{1,p}_{0}(\Omega,d^{\alpha}), provided that 0α<p10\leq\alpha<p-1.

In this case functions from W01,p(Ω,dα)W^{1,p}_{0}(\Omega,d^{\alpha}) are zero on Ω\partial\Omega in the sense that they can be approximated by smooth functions with compact support. In the sequel we will use certain Hardy–Sobolev inequalities. Note that inequalities of this kind, when dd is replaced by |𝐱|=d(𝐱,0)|{\bf x}|=d({\bf x},0) are known as Caffarelli–Kohn–Nirenberg inequalities [8].

Lemma 4.3.

Let Ωn\Omega\subseteq\mathbb{R}^{n} be a bounded Lipschitz domain. For p[1,n)p\in[1,n), αp1\alpha\neq p-1 and q[p,npnp]q\in[p,\frac{np}{n-p}\big{]} there exists a constant c>0c>0 such that for all fW01,p(Ω,dα)f\in W^{1,p}_{0}(\Omega,d^{\alpha}) there holds

(Ωdqp(np+α)n|f|q𝑑𝐱)1qc(Ωdα|f|p𝑑𝐱)1p.\left(\int_{\Omega}d^{\frac{q}{p}(n-p+\alpha)-n}|f|^{q}\,d{\bf x}\right)^{\frac{1}{q}}\leq c\,\left(\int_{\Omega}d^{\alpha}|\nabla f|^{p}\,d{\bf x}\right)^{\frac{1}{p}}. (4.3)
Proof.

This follows from the definition of the space W01,p(Ω,dα)W^{1,p}_{0}(\Omega,d^{\alpha}), [18, Theorem 2.1] and the classical (p,α)(p,\alpha) Hardy inequality

(Ωdαp|f|p𝑑𝐱)1pc(Ωdα|f|p𝑑𝐱)1p,\left(\int_{\Omega}d^{\alpha-p}|f|^{p}\,d{\bf x}\right)^{\frac{1}{p}}\leq c\,\left(\int_{\Omega}d^{\alpha}|\nabla f|^{p}\,d{\bf x}\right)^{\frac{1}{p}}, (4.4)

which is valid for all p(1,)p\in(1,\infty) and αp1\alpha\not=p-1, for functions in W01,p(Ω,dα)W^{1,p}_{0}(\Omega,d^{\alpha}) (cf. [20], [15, Theorem 8.10.14]). ∎

In addition to (4.1) and its role in Hardy-type inequalities, the critical nature of the power α=p1\alpha=p-1 also occurs in the notion of Muckenhoupt weights and their relation with the maximal function.

Definition 4.4.

We say that a weight ϱLloc1(3)\varrho\in L^{1}_{\textup{loc}}(\mathbb{R}^{3}) belongs to the Muckenhoupt class ApA_{p}, for 1<p<1<p<\infty, if there exists CC such that

supQn(Qϱ(𝐱)d𝐱)(Qϱ(𝐱)1/(1p)d𝐱)p1C,\sup_{Q\subset\mathbb{R}^{n}}\Bigg{(}\fint_{Q}\varrho({\bf x})\,\mathrm{d}{\bf x}\Bigg{)}\Bigg{(}\fint_{Q}\varrho({\bf x})^{1/(1-p)}\,\mathrm{d}{\bf x}\Bigg{)}^{p-1}\leq C,

where QQ denotes a cube in 3\mathbb{R}^{3}.

The powers of the distance function belong to the class ApA_{p} according to the following well-known result for general domains (say it is enough that Ω\partial\Omega is a n1n-1-dimensional closed set, see [10]). Here and in the sequel the boundary will be at least locally Lipschitz to have the outward unit vector properly defined.

Lemma 4.5.

The function ϱ(𝐱)=(d(𝐱))α\varrho({\bf x})=\big{(}d({\bf x})\big{)}^{\alpha} is a Muckenhoupt weight of class ApA_{p} if and only if 1<α<p1-1<\alpha<p-1.

4.1. Solenoidal spaces

A standard approach in fluid mechanics, is to incorporate the divergence-free constraint directly in the function spaces. These spaces are built upon completing the space of solenoidal smooth vector fields with compact support, denoted as ϕC0,σ(Ω){\boldsymbol{\phi}}\in C^{\infty}_{0,\sigma}(\Omega). For α\alpha\in\mathbb{R} define

Lσp(Ω,dα):={ϕC0,σ(Ω)}¯.p,dα,\displaystyle L^{p}_{\sigma}(\Omega,d^{\alpha}):=\overline{\left\{{\boldsymbol{\phi}}\in C^{\infty}_{0,\sigma}(\Omega)\right\}}^{\|\,.\,\|_{p,d^{\alpha}}},
W0,σ1,p(Ω,dα):={ϕC0,σ(Ω)}¯.1,p,dα.\displaystyle W^{1,p}_{0,\sigma}(\Omega,d^{\alpha}):=\overline{\left\{{\boldsymbol{\phi}}\in C^{\infty}_{0,\sigma}(\Omega)\right\}}^{\|\,.\,\|_{1,p,d^{\alpha}}}\,.

For α=0\alpha=0 they reduce to the classical spaces Lσp(Ω)L^{p}_{\sigma}(\Omega) and W0,σ1,p(Ω)W^{1,p}_{0,\sigma}(\Omega). Next, we will extensively use the following extension of classical inequalities linking curl/divergence and full gradient estimates (cf. [3]).

Lemma 4.6.

Let 1<p<1<p<\infty and assume that the weight ϱ\varrho belongs to the class ApA_{p}. Then, there exists a constant CC, depending on the domain Ω\Omega and on the weight ϱAp\varrho\in A_{p}, such that

𝐮p,ϱC(div𝐮p,ϱ+curl𝐮p,ϱ)𝐮W01,p(Ω,ϱ).\|\nabla{\bf u}\|_{p,\varrho}\leq C(\|\mathrm{div}\,{\bf u}\|_{p,\varrho}+\|\operatorname{curl}{\bf u}\|_{p,\varrho})\qquad\forall\,{\bf u}\in W^{1,p}_{0}(\Omega,\varrho).

In particular, we will use the latter result in the following special form

Corollary 4.7.

For 1<α<p1-1<\alpha<p-1 there exists a constant C=C(Ω,α,p)C=C(\Omega,\alpha,p) such that

Ωdα|𝐯|pd𝐱CΩdα|curl𝐯|pd𝐱𝐯W0,σ1,p(Ω,dα).\int_{\Omega}d^{\alpha}|\nabla{\bf v}|^{p}\,\mathrm{d}{\bf x}\leq C\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|^{p}\,\mathrm{d}{\bf x}\qquad\forall\,{\bf v}\in W^{1,p}_{0,\sigma}(\Omega,d^{\alpha}). (4.5)

5. Application to the rotational turbulence models: the proof of Theorem 1.1

In this section we verify that the initial boundary value problem (1.1), after a proper selection of parameters, and definition of both the operators and functional spaces, can be put in the framework of the abstract Theorem 3.6. This will be enough to give a proof of the main result of this paper, that is the existence of weak solutions in Theorem 1.1.

In our setting the choice of the natural spaces is determined by the problem itself which yields, by the a priori estimate obtained by testing with the velocity 𝐯¯\overline{\bf v}, that the integral

0TΩdα|curl𝐯¯|3𝑑𝐱𝑑t\int_{0}^{T}\int_{\Omega}d^{\alpha}|\operatorname{curl}\overline{\bf v}|^{3}\,d{\bf x}dt

is finite. Hence, for almost all t[0,T]t\in[0,T] the integral Ωdα|curl𝐯¯|3𝑑𝐱\int_{\Omega}d^{\alpha}|\operatorname{curl}\overline{\bf v}|^{3}\,d{\bf x} will be finite, determines the choice for the Banach space VV.

In order to identify the evolution triple to be used for the proper formulation, we need to clarify the relationship with the L2(Ω)L^{2}(\Omega) norm. We have the following result which immediately derives from the basic results on weighted spaces of the previous section.

Lemma 5.1.

Let 𝐮C0,σ(Ω){\bf u}\in C^{\infty}_{0,\sigma}(\Omega) and α[0,2)\alpha\in[0,2). Then, there exists C=C(α,Ω)C=C(\alpha,\Omega) such that

(Ω|𝐮|2𝑑𝐱)1/2C(Ωdα|curl𝐮|3𝑑𝐱)1/3.\left(\int_{\Omega}|{\bf u}|^{2}\,d{\bf x}\right)^{\smash{1/2}}\leq C\left(\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf u}|^{3}\,d{\bf x}\right)^{\smash{1/3}}. (5.1)
Proof.

For α<2\alpha<2, combining (4.3) with q=3p3p+αq=\frac{3p}{3-p+\alpha} and (4.5), it follows for every p(α+1,3)p\in(\alpha+1,3)

Ω|𝐮|3p3p+αd𝐱c(Ωdα|𝐮|pd𝐱)qpc(Ωdα|curl𝐮|pd𝐱)qp,\int_{\Omega}{|{\bf u}|^{\frac{3p}{3-p+\alpha}}\,\mathrm{d}{\bf x}}\leq c\,\left(\int_{\Omega}d^{\alpha}|\nabla{\bf u}|^{p}\,\mathrm{d}{\bf x}\right)^{\frac{q}{p}}\leq c\left(\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf u}|^{p}\,\mathrm{d}{\bf x}\right)^{\frac{q}{p}},

for all 𝐯C0,σ(Ω){\bf v}\in C^{\infty}_{0,\sigma}(\Omega). Since 23p3p+α2\leq\frac{3p}{3-p+\alpha} the assertion follows from Hölder’s inequality as Ω\Omega is bounded. ∎

Lemma 5.1 shows that one can work with the following evolution triple for all α[0,2)\alpha\in[0,2)

(V,H,id):=(W0,σ1,3(Ω,dα),Lσ2(Ω),id).(V,H,\textrm{id}):=\big{(}W^{1,3}_{0,\sigma}(\Omega,d^{\alpha}),L^{2}_{\sigma}(\Omega),\textrm{id}\big{)}.

and as functional setting for (1.1) we use the following spaces and operators, where 0α<20\leq\alpha<2

V\displaystyle V :=W0,σ1,3(Ω,dα)𝐯V:=(Ωdα|curl𝐯|3𝑑𝐱)1/3\displaystyle:=W^{1,3}_{0,\sigma}(\Omega,d^{\alpha})\qquad\|{\bf v}\|_{V}:=\left(\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|^{3}\,d{\bf x}\right)^{1/3}
H\displaystyle H :=Lσ2(Ω)𝐯H:=(Ω|𝐯|2𝑑𝐱)1/2\displaystyle:=L^{2}_{\sigma}(\Omega)\qquad\|{\bf v}\|_{H}:=\left(\int_{\Omega}|{\bf v}|^{2}\,d{\bf x}\right)^{1/2}
𝒳\displaystyle\mathbfcal{X} :=L3(I,V),𝒴¬\displaystyle:=L^{3}(I,V),\qquad\mathbfcal{Y}:=L^{\infty}(I,H)
𝒲\displaystyle\mathbfcal{W} :={𝐮L3(I,V)|d𝐮dtL3/2(I,V)},\displaystyle:=\Big{\{}\mathbf{u}\in L^{3}(I,V){\,\big{|}\,}\exists\,\frac{d\mathbf{u}}{dt}\in L^{3/2}(I,V^{*})\Big{\}},

and define the operator A:=S+B:VVA:=S+B:V\to V^{*} via

S𝐯,𝐰V\displaystyle\langle S{\bf v},{\bf w}\rangle_{V} :=Ωdα|curl𝐯|curl𝐯curl𝐰d𝐱,\displaystyle:=\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|\operatorname{curl}{\bf v}\cdot\operatorname{curl}{\bf w}\,d{\bf x},
B𝐯,𝐰V\displaystyle\langle B{\bf v},{\bf w}\rangle_{V} :=Ω(curl𝐯×𝐯)𝐰𝑑𝐱.\displaystyle:=\int_{\Omega}(\operatorname{curl}{\bf v}\times{\bf v})\cdot{\bf w}\,d{\bf x}.

The induced operator 𝒮¬𝒳𝒴𝒳\mathbfcal{S}:\ \mathbfcal{X}\cap\mathbfcal{Y}\rightarrow\mathbfcal{X}^{*} inherits the properties of the operator SS (cf. [28, Chapter 30]). Note that SS is a strictly monotone, bounded, coercive, and continuous operator. These properties are practically the same known for the pp-Laplace operator. In fact, from the definition, one obtains directly the following two inequalities:

S𝐯V\displaystyle\left\|S{\bf v}\right\|_{V^{*}} 𝐯V2𝐯V,\displaystyle\leq\|{\bf v}\|_{V}^{2}\qquad\forall\,{\bf v}\in V,
S𝐯,𝐯V\displaystyle\langle S{\bf v},{\bf v}\rangle_{V} =𝐯V3𝐯V.\displaystyle=\|{\bf v}\|_{V}^{3}\qquad\forall\,{\bf v}\in V.

The monotonicity of SS derives from the following lemma (cf. [3, Lemma 3.3]).

Lemma 5.2.

For smooth enough vector field 𝛚i{\boldsymbol{\omega}}_{i} (it is actually enough that dαp𝛚iLp(Ω)d^{\frac{\alpha}{p}}{\boldsymbol{\omega}}_{i}\in L^{p}(\Omega), with 1<p<1<p<\infty) and for α+\alpha\in\mathbb{R}^{+} it holds that

Ω(dα|𝝎1|p2𝝎1dα|𝝎2|p2𝝎2)(𝝎1𝝎2)d𝐱0,\int_{\Omega}(d^{\alpha}|{\boldsymbol{\omega}}_{1}|^{p-2}{\boldsymbol{\omega}}_{1}-d^{\alpha}|{\boldsymbol{\omega}}_{2}|^{p-2}{\boldsymbol{\omega}}_{2})\cdot({\boldsymbol{\omega}}_{1}-{\boldsymbol{\omega}}_{2})\,\mathrm{d}{\bf x}\geq 0,

for any (not necessarily the distance) bounded function such that d:Ω+d:\Omega\to\mathbb{R}^{+} for a.e. 𝐱Ω{\bf x}\in\Omega.

The proof of the above lemma is based on the observation that it can be proved that dα(|𝝎1|p2𝝎1|𝝎2|p2𝝎2)(𝝎1𝝎2)0d^{\alpha}(|{\boldsymbol{\omega}}_{1}|^{p-2}{\boldsymbol{\omega}}_{1}-|{\boldsymbol{\omega}}_{2}|^{p-2}{\boldsymbol{\omega}}_{2})\cdot({\boldsymbol{\omega}}_{1}-{\boldsymbol{\omega}}_{2})\geq 0 point-wise. Then weighted integrability of the functions this used to prove that the integral is finite.

To treat the operator BB, and the induced one ¬𝒳𝒴𝒳\mathbfcal{B}:\,\mathbfcal{X}\cap\mathbfcal{Y}\rightarrow\mathbfcal{X}^{*}, we need to properly adapt the estimates on the convective term in weighted spaces and this is mainly based on the previously Hardy-type inequalities (4.3).

Lemma 5.3 (Boundedness of BB).

For all α[0,2)\alpha\in[0,2) the operator B:VV{B:V\to V^{*}} is bounded. It satisfies B𝐮,𝐯Vc𝐮V2𝐯V\langle B{\bf u},{\bf v}\rangle_{V}\leq c\|{\bf u}\|_{V}^{2}\|{\bf v}\|_{V} and B𝐮,𝐮V=0\langle B{\bf u},{\bf u}\rangle_{V}=0, for all 𝐮,𝐯V{\bf u},{\bf v}\in V.

Proof.

The proof is based on the estimation of the space integral, by using appropriate weighted version of classical Sobolev spaces tools. We have in fact, for all smooth functions with compact support the following inequality (obtained multiplying and dividing a.e. 𝐱Ω{\bf x}\in\Omega by the positive function dα/3d^{\alpha/3})

|Ω(curl𝐯×𝐮)𝐰𝑑𝐱|Ωdα/6|𝐮|dα/3|curl𝐯|dα/6|𝐰|𝑑𝐱\displaystyle\left|\int_{\Omega}(\operatorname{curl}{\bf v}\times{\bf u})\cdot{\bf w}\,d{\bf x}\right|\leq\int_{\Omega}d^{-\alpha/6}|{\bf u}|\,d^{\alpha/3}|\operatorname{curl}{\bf v}|\,d^{-\alpha/6}|{\bf w}|\,d{\bf x}
(Ωdα/2|𝐮|3𝑑𝐱)1/3(Ωdα|curl𝐯|3𝑑𝐱)1/3(Ωdα/2|𝐰|3𝑑𝐱)1/3.\displaystyle\quad\leq\left(\int_{\Omega}d^{-\alpha/2}|{\bf u}|^{3}\,d{\bf x}\right)^{1/3}\left(\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|^{3}\,d{\bf x}\right)^{1/3}\left(\int_{\Omega}d^{-\alpha/2}|{\bf w}|^{3}\,d{\bf x}\right)^{1/3}.

It remains to show that all 𝐮V{\bf u}\in V also belong to the weighted space L3(Ω,dα2)L^{3}(\Omega,d^{-\frac{\alpha}{2}}), with a continuous embedding. From (4.3) it follows for all p[1,3)p\in[1,3), α[0,2)\alpha\in[0,2) and 𝐮W0,σ1,p(Ω,dα){\bf u}\in W^{1,p}_{0,\sigma}(\Omega,d^{\alpha}) that

(Ωdα2|𝐮|p(6α)2(3p+α)𝑑𝐱)1qc(Ωdα|𝐮|p𝑑𝐱)1p,\left(\int_{\Omega}d^{-\frac{\alpha}{2}}|{\bf u}|^{\frac{p(6-\alpha)}{2(3-p+\alpha)}}\,d{\bf x}\right)^{\frac{1}{q}}\leq c\,\left(\int_{\Omega}d^{\alpha}|\nabla{\bf u}|^{p}\,d{\bf x}\right)^{\frac{1}{p}},

with

q:=p(6α)2(3p+α)<p.q:=\frac{p(6-\alpha)}{2(3-p+\alpha)}<p^{*}.

One easily checks that for all α[0,2)\alpha\in[0,2) there exists a p(1+α,3)p\in(1+\alpha,3) such that 3<q<p3<q<p^{*}. Since Ω\Omega is bounded we deduce from this VL3(Ω,dα2)V\hookrightarrow L^{3}(\Omega,d^{-\frac{\alpha}{2}}) by using Hölder’s inequality.

Once the integral Ω(𝐮×curl𝐯)𝐰𝑑𝐱\int_{\Omega}({\bf u}\times\operatorname{curl}{\bf v})\cdot{\bf w}\,d{\bf x} is well-defined for 𝐮,𝐯,𝐰V{\bf u},{\bf v},{\bf w}\in V, it immediately follows that B𝐮,𝐮V=0\langle B{\bf u},{\bf u}\rangle_{V}=0 for all 𝐮V{\bf u}\in V, since a.e. in Ω\Omega it holds (𝐯×curl𝐯)𝐯=0({\bf v}\times\operatorname{curl}{\bf v})\cdot{\bf v}=0. ∎

This is enough for what concerns the growth and coercivity. We need now to show compactness for BB in order to prove pseudo-monotonicity.

Lemma 5.4 (Compactness of BB).

Let α[0,2)\alpha\in[0,2). Then, the weak convergence 𝐮n𝐮{\bf u}_{n}\rightharpoonup{\bf u} in VV implies (up to a sub-sequence) that

B𝐮nB𝐮in V,B{\bf u}_{n}\to B{\bf u}\qquad\text{in }V^{*},

i.e., the operator BB is compact.

Proof.

By the boundedness of the weakly converging sequence (𝐮n)nV({\bf u}_{n})_{n\in\mathbb{N}}\subseteq V and by (4.2) we get that

𝐮nW1,r(Ω)Cr[1,31+α[.\|{\bf u}_{n}\|_{W^{1,r}(\Omega)}\leq C\qquad\forall\,r\in\big{[}1,\frac{3}{1+\alpha}\big{[}.

Hence, by the usual (unweighted) compact Sobolev embedding W1,r(Ω)Lr~(Ω)W^{1,r}(\Omega)\hookrightarrow\hookrightarrow L^{\tilde{r}}(\Omega), valid for all r~<(31+α)=3α\tilde{r}<(\frac{3}{1+\alpha})^{*}=\frac{3}{\alpha} we get also that (up to a sub-sequence)

𝐮n𝐮 a.e. and in Lr~(Ω).{\bf u}_{n}\to{\bf u}\qquad\text{ a.e. and in }L^{\tilde{r}}(\Omega).

By using the definition of BB, the properties of the curl\operatorname{curl} (with summation over repeated indices), and integration by parts, we have that for all 𝐮,𝐯V{\bf u},{\bf v}\in V

B𝐮,𝐯\displaystyle\langle B{\bf u},{\bf v}\rangle =Ωϵjklϵjlmukviumxl𝑑𝐱=Ω(δklδimδkmδil)ukviumxl𝑑𝐱\displaystyle=\int_{\Omega}\epsilon_{jkl}\epsilon_{jlm}u_{k}v_{i}\frac{\partial u_{m}}{\partial x_{l}}\,d{\bf x}=\int_{\Omega}(\delta_{kl}\delta_{im}-\delta_{km}\delta_{il})u_{k}v_{i}\frac{\partial u_{m}}{\partial x_{l}}\,d{\bf x}
=Ωukuivixk𝑑𝐱=Ω(𝐮𝐮):𝐯d𝐱.\displaystyle=-\int_{\Omega}u_{k}u_{i}\frac{\partial v_{i}}{\partial x_{k}}\,d{\bf x}=-\int_{\Omega}({\bf u}\otimes{\bf u}):\nabla{\bf v}\,d{\bf x}.

Hence, we have

B𝐮n,𝐯B𝐮,𝐯\displaystyle\langle B{\bf u}_{n},{\bf v}\rangle-\langle B{\bf u},{\bf v}\rangle =Ω(𝐮n𝐮n):𝐯(𝐮𝐮):𝐯d𝐱\displaystyle=-\int_{\Omega}({\bf u}_{n}\otimes{\bf u}_{n}):\nabla{\bf v}-({\bf u}\otimes{\bf u}):\nabla{\bf v}\,d{\bf x}
=Ω((𝐮n𝐮)𝐮n):𝐯+(𝐮(𝐮n𝐮)):𝐯d𝐱.\displaystyle=-\int_{\Omega}\big{(}({\bf u}_{n}-{\bf u})\otimes{\bf u}_{n}\big{)}:\nabla{\bf v}+\big{(}{\bf u}\otimes({\bf u}_{n}-{\bf u})\big{)}:\nabla{\bf v}\,d{\bf x}.

By Hölder inequality we get, as in the proof of Lemma 5.3,

|Ω((𝐮n𝐮)𝐮n):𝐯d𝐱|\displaystyle\left|\int_{\Omega}\big{(}({\bf u}_{n}-{\bf u})\otimes{\bf u}_{n}\big{)}:\nabla{\bf v}\,d{\bf x}\right|
(Ωdα/2|𝐮n𝐮|3𝑑𝐱)1/3(Ωdα|𝐯|3𝑑𝐱)1/3(Ωdα/2|𝐮n|3𝑑𝐱)1/3\displaystyle\leq\left(\int_{\Omega}d^{-\alpha/2}|{\bf u}_{n}-{\bf u}|^{3}\,d{\bf x}\right)^{1/3}\left(\int_{\Omega}d^{\alpha}|\nabla{\bf v}|^{3}\,d{\bf x}\right)^{1/3}\left(\int_{\Omega}d^{-\alpha/2}|{\bf u}_{n}|^{3}\,d{\bf x}\right)^{1/3}
(Ωdα/2|𝐮n𝐮|3𝑑𝐱)1/3𝐯V𝐮nV.\displaystyle\leq\left(\int_{\Omega}d^{-\alpha/2}|{\bf u}_{n}-{\bf u}|^{3}\,d{\bf x}\right)^{1/3}\|{\bf v}\|_{V}\|{\bf u}_{n}\|_{V}.

We now observe that the last two terms are uniformly bounded, while

dα/2|𝐮n𝐮|30a.e. 𝐱Ω.d^{-\alpha/2}|{\bf u}_{n}-{\bf u}|^{3}\to 0\qquad\text{a.e. }{\bf x}\in\Omega.

Consequently, to show that the integral vanishes it is enough to prove that for some q>3q>3 there holds

𝐮n𝐮Lq(Ω,dα/2)C\|{\bf u}_{n}-{\bf u}\|_{L^{q}(\Omega,d^{-\alpha/2})}\leq C

uniformly in nn\in\mathbb{N}, which permits to apply the Vitali theorem in the weighted space L3(Ω,dα/2)L^{3}(\Omega,d^{-\alpha/2}). However, this was already obtained in the proof of Lemma 5.3. The other term in the decomposition of B𝐮n,𝐯B𝐮,𝐯\langle B{\bf u}_{n},{\bf v}\rangle-\langle B{\bf u},{\bf v}\rangle can be treated in the same way. ∎

Proof of Theorem 1.1.

In the previous lemmas we have proved that A=S+BA=S+B is continuous and pseudo-monotone since it is the sum of a monotone continuous and a compact one. Collecting the estimates we have that in particular that the boundedness and coercivity are as follows

A𝐯V\displaystyle\left\|A{\bf v}\right\|_{V^{*}} c0𝐯V2,\displaystyle\leq c_{0}\|{\bf v}\|_{V}^{2},
A𝐯,𝐯V\displaystyle\langle A{\bf v},{\bf v}\rangle_{V} 𝐯V3,\displaystyle\geq\|{\bf v}\|_{V}^{3},

since B𝐯,𝐯V=0\langle B{\bf v},{\bf v}\rangle_{V}=0. Hence all hypotheses from (𝐂.1)\mathbf{(C.1)} to (𝐂.3)\mathbf{(C.3)} are satisfied.

This shows that the induced operator 𝒜\mathbfcal{A} is Bochner pseudo-monotone and coercive, hence all the hypotheses of the abstract existence Theorem 3.6 are satisfied. This proves the main result of this paper, that is the existence of weak solution in Theorem 1.1. ∎

5.1. The case p>3p>3

In this section we show that most of the results of the previous section can be extended (even with easier proofs) to the system with the following operator

Sp𝐯,𝐰V:=Ωdα|curl𝐯|p2curl𝐯curl𝐰d𝐱 with p>3, 0α<p1,\langle S_{p}{\bf v},{\bf w}\rangle_{V}:=\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|^{p-2}\operatorname{curl}{\bf v}\cdot\operatorname{curl}{\bf w}\,d{\bf x}\quad\text{ with }p>3,\ 0\leq\alpha<p-1,

while the use of the tools typical of pseudo-monotone operators fails for p<3p<3. We can then prove the following result

Theorem 5.5.

Let p>3p>3, α[0,p1)\alpha\in[0,p-1), 0<T<0<T<\infty, 𝐯0¯Lσ2(Ω)\overline{{\bf v}_{0}}\in L^{2}_{\sigma}(\Omega), and 𝐟Lp(0,T;(W01,p(Ω,dα)){\bf f}\in L^{p^{\prime}}(0,T;(W^{1,p}_{0}(\Omega,d^{\alpha})^{*}). Then, there exists a weak solution to the initial boundary value problem

t𝐯¯+𝝎¯×𝐯¯+curl(dα|𝝎¯|p2𝝎¯)+q¯\displaystyle\partial_{t}\overline{\bf v}+\overline{\boldsymbol{\omega}}\times\overline{\bf v}+\operatorname{curl}\big{(}d^{\alpha}|\overline{\boldsymbol{\omega}}|^{p-2}\overline{\boldsymbol{\omega}}\big{)}+\nabla\overline{q} =𝐟\displaystyle={\bf f}    in (0,T)×Ω(0,T)\times\Omega,
𝝎¯\displaystyle\overline{\boldsymbol{\omega}} =curl𝐯¯\displaystyle=\operatorname{curl}\overline{\bf v}    in (0,T)×Ω(0,T)\times\Omega,
div𝐯¯\displaystyle\mathrm{div}\,\overline{\bf v} =0\displaystyle=0    in (0,T)×Ω(0,T)\times\Omega,
𝐯¯\displaystyle\overline{\bf v} =𝟎\displaystyle=\mathbf{0}    on (0,T)×Ω(0,T)\times\partial\Omega,
𝐯¯(0)\displaystyle\overline{\bf v}(0) =𝐯0¯\displaystyle=\overline{{\bf v}_{0}}    in Ω\Omega,

such that

𝐯¯C([0,T];Lσ2(Ω))Lp(0,T;W0,σ1,p(Ω,dα))\overline{\bf v}\in C([0,T];L^{2}_{\sigma}(\Omega))\cap L^{p}(0,T;W^{1,p}_{0,\sigma}(\Omega,d^{\alpha}))

and for all t[0,T]t\in[0,T]

12𝐯¯(t)2+0tΩCαdα(𝐱)|𝝎¯(s,𝐱)|p𝑑𝐱𝑑s\displaystyle\frac{1}{2}\|\overline{\bf v}(t)\|^{2}+\int_{0}^{t}\int_{\Omega}C_{\alpha}d^{\alpha}({\bf x})|{\overline{\boldsymbol{\omega}}}(s,{\bf x})|^{p}\,d{\bf x}\,ds
=12𝐯0¯2+0t𝐟(s),𝐯¯(s)W01,p(Ω,dα)𝑑s.\displaystyle=\frac{1}{2}\|\overline{{\bf v}_{0}}\|^{2}+\int_{0}^{t}\langle{\bf f}(s),\overline{\bf v}(s)\rangle_{W^{1,p}_{0}(\Omega,d^{\alpha})}ds.

The proof of this result is again just a verification that the hypotheses of the abstract theorem are satisfied, but we will highlight the critical role of the parameters.

First note that for p>3p>3 and 0α<p10\leq\alpha<p-1 the inclusion W01,p(Ω,dα)L2(Ω)W^{1,p}_{0}(\Omega,d^{\alpha})\subset L^{2}(\Omega) holds true. Directly by Hölder’s inequality with δ=p/2\delta=p/2, δ=p/(p2)\delta^{\prime}=p/(p-2), and Hardy inequality (4.4) we get

Ω|𝐮|2𝑑𝐱\displaystyle\int_{\Omega}|{\bf u}|^{2}\,d{\bf x} =Ωd2p(αp)|𝐮|2d2p(pα)𝑑𝐱\displaystyle=\int_{\Omega}d^{\frac{2}{p}(\alpha-p)}|{\bf u}|^{2}d^{\frac{2}{p}(p-\alpha)}\,d{\bf x}
(Ωdαp|𝐮|p𝑑𝐱)2p(Ωd2pαp2𝑑𝐱)p2p\displaystyle\leq\left(\int_{\Omega}d^{\alpha-p}|{\bf u}|^{p}\,d{\bf x}\right)^{\frac{2}{p}}\left(\int_{\Omega}d^{2\frac{p-\alpha}{p-2}}\,d{\bf x}\right)^{\frac{p-2}{p}}
C(p,α)(Ωdα|𝐮|p𝑑𝐱)2p,\displaystyle\leq C(p,\alpha)\left(\int_{\Omega}d^{\alpha}|\nabla{\bf u}|^{p}\,d{\bf x}\right)^{\frac{2}{p}},

since pα>0p-\alpha>0. This shows that one can work with the evolution triple

(V,H,id):=(W0,σ1,p(Ω,dα),Lσ2(Ω),id)withp>3, 0α<p1.(V,H,\textrm{id}):=\big{(}W^{1,p}_{0,\sigma}(\Omega,d^{\alpha}),L^{2}_{\sigma}(\Omega),\textrm{id}\big{)}\qquad\text{with}\quad p>3,\ 0\leq\alpha<p-1.

The properties of the operator SpS_{p} are practically the same as those of the operator SS, hence one can directly show

Sp𝐯V\displaystyle\left\|S_{p}{\bf v}\right\|_{V^{*}} 𝐯Vp1𝐯V,\displaystyle\leq\|{\bf v}\|_{V}^{p-1}\qquad\forall\,{\bf v}\in V,
Sp𝐯,𝐯V\displaystyle\langle S_{p}{\bf v},{\bf v}\rangle_{V} =𝐯Vp𝐯V.\displaystyle=\|{\bf v}\|_{V}^{p}\qquad\forall\,{\bf v}\in V.

On the other hand, the properties of BB are to be checked. The operator is the same as before, but the functional setting is different.

To show the boundedness of B:VVB:\,V\to V^{*} for 0α<p10\leq\alpha<p-1, we proceed as in Lemma 5.3 and we get

|Ω(curl𝐯×𝐮)𝐰𝑑𝐱|Ωdα/2p|𝐮|dα/p|curl𝐯|dα/2p|𝐰|𝑑𝐱\displaystyle\left|\int_{\Omega}(\operatorname{curl}{\bf v}\times{\bf u})\cdot{\bf w}\,d{\bf x}\right|\leq\int_{\Omega}d^{-\alpha/2p}|{\bf u}|\,d^{\alpha/p}|\operatorname{curl}{\bf v}|\,d^{-\alpha/2p}|{\bf w}|\,d{\bf x}
(Ωdαp/p|𝐮|2p𝑑𝐱)12p(Ωdα|𝐯|p𝑑𝐱)1p(Ωdαp/p|𝐰|2p𝑑𝐱)12p.\displaystyle\leq\left(\int_{\Omega}d^{-\alpha p^{\prime}/p}|{\bf u}|^{2p^{\prime}}\,d{\bf x}\right)^{\frac{1}{2p^{\prime}}}\left(\int_{\Omega}d^{\alpha}|\nabla{\bf v}|^{p}\,d{\bf x}\right)^{\frac{1}{p}}\left(\int_{\Omega}d^{-\alpha p^{\prime}/p}|{\bf w}|^{2p^{\prime}}\,d{\bf x}\right)^{\frac{1}{2p^{\prime}}}.

Next, observe that 2p=2p/(p1)<p2p^{\prime}=2p/(p-1)<p is satisfied for p>3p>3. Consequently, in this case we can directly apply Hölder inequality with exponents δ=(p1)/2\delta=(p-1)/2 and δ=(p1)/(p3)\delta^{\prime}=(p-1)/(p-3) to bound the first and third integrals as follows:

Ωdα/(p1)|𝐮|2p𝑑𝐱=Ωd(αp)2/(p1)|𝐮|2pd(2p3α)/(p1)𝑑𝐱\displaystyle\int_{\Omega}d^{-\alpha/(p-1)}|{\bf u}|^{2p^{\prime}}\,d{\bf x}=\int_{\Omega}d^{(\alpha-p)2/(p-1)}|{\bf u}|^{2p^{\prime}}d^{(2p-3\alpha)/(p-1)}\,d{\bf x}
(Ωdαp|𝐮|p𝑑𝐱)2p1(Ωd(2p3α)/(p3)𝑑𝐱)p3p1.\displaystyle\leq\left(\int_{\Omega}d^{\alpha-p}|{\bf u}|^{p}\,d{\bf x}\right)^{\frac{2}{p-1}}\left(\int_{\Omega}d^{(2p-3\alpha)/(p-3)}\,d{\bf x}\right)^{\frac{p-3}{p-1}}.

The first term from the right-hand side is bounded with (Ωdα|𝐮|p𝑑𝐱)2p1\left(\int_{\Omega}d^{\alpha}|\nabla{\bf u}|^{p}\,d{\bf x}\right)^{\frac{2}{p-1}} by using (4.4), while the second is finite if

2p3αp3>1α<p1.\frac{2p-3\alpha}{p-3}>-1\quad\Longleftrightarrow\quad\alpha<p-1.

This shows that

B𝐮,𝐰C(Ω,α,p)𝐮V2𝐰V,\langle B{\bf u},{\bf w}\rangle\leq C(\Omega,\alpha,p)\|{\bf u}\|_{V}^{2}\|{\bf w}\|_{V},

and the compactness of BB follows with the same arguments used in the previous section (almost everywhere convergence and Vitali theorem).

Remark 5.6.

The case 1<p<31<p<3 does not fit with the theory for the reasons we now explain. The argument with Hardy inequality as in the previous lemma requires p>3p>3. If we try to apply the same argument used for p=3p=3 with Hardy–Sobolev inequality (4.3), we can write

|Ω(curl𝐯×𝐮)𝐰𝑑𝐱|\displaystyle\left|\int_{\Omega}(\operatorname{curl}{\bf v}\times{\bf u})\cdot{\bf w}\,d{\bf x}\right|
(Ωdαp/p|𝐮|2p𝑑𝐱)12p(Ωdα|curl𝐯|p𝑑𝐱)1p(Ωdαp/p|𝐰|2p𝑑𝐱)12p,\displaystyle\leq\left(\int_{\Omega}d^{-\alpha p^{\prime}/p}|{\bf u}|^{2p^{\prime}}\,d{\bf x}\right)^{\frac{1}{2p^{\prime}}}\left(\int_{\Omega}d^{\alpha}|\operatorname{curl}{\bf v}|^{p}\,d{\bf x}\right)^{\frac{1}{p}}\left(\int_{\Omega}d^{-\alpha p^{\prime}/p}|{\bf w}|^{2p^{\prime}}\,d{\bf x}\right)^{\frac{1}{2p^{\prime}}},

and then estimate the first and third integral with (4.3) for q=2p<pq=2p^{\prime}<p^{*}, which holds for p>95p>\frac{9}{5}. Hence, to apply (4.3) the precise exponent will be q=pp13p3α3p+αq=\frac{p}{p-1}\frac{3p-3-\alpha}{3-p+\alpha}, and since q2pq\geq 2p^{\prime} this implies that we have to request for

α5p93.\alpha\leq\frac{5p-9}{3}.

Since we would like to treat cases with α\alpha smaller but “arbitrarily close” to p1p-1, the inequality

p15p93,p-1\leq\frac{5p-9}{3},

should be correct. On the other hand the latter can be satisfied only for p3p\geq 3. Since we are out of the range of permitted pp this shows that the estimate can not be used. Being the inequalities Hardy–Sobolev inequalities sharp, this proves that operator BB is not bounded for 95<p<3\frac{9}{5}<p<3, when α\alpha is close to p1p-1, hence the basic assumptions to use the pseudo-monotone methods are not satisfied. The existence of weak solutions, if possible, should be obtained with different methods and possibly considering different weak formulations of the problem.

Acknowledgments

Luigi C. Berselli was partially supported by a grant of the group GNAMPA of INdAM and by the University of Pisa within the grant PRA_2018_52\_{}2018\_{}52 UNIPI: “Energy and regularity: New techniques for classical PDE problems.

References

  • [1] C. Amrouche, L. C. Berselli, R. Lewandowski, and D. D. Nguyen, Turbulent flows as generalized Kelvin–Voigt materials: Modeling and analysis, Nonlinear Anal. 196 (2020), 111790. MR 4065455
  • [2] 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.
  • [3] L. C. Berselli and D. Breit, On the existence of weak solutions for the steady Baldwin–Lomax model and generalizations, J. Math. Anal. Appl. (2021), 124633.
  • [4] L. C. Berselli, T. Iliescu, and W. J. Layton, Mathematics of Large Eddy Simulation of turbulent flows, Scientific Computation, Springer-Verlag, Berlin, 2006. MR MR2185509 (2006h:76071)
  • [5] L. C. Berselli, R. Lewandowski, and D. D. Nguyen, Rotational Forms of Large Eddy Simulation Turbulence Models: Modeling and Mathematical Theory, Chin. Ann. Math. Ser. B 42 (2021), no. 1, 17–40. MR 4209051
  • [6] H. Brezis, Les opérateurs monotones, Séminaire Choquet: 1965/66, Initiation à l’ Analyse, Fasc. 2, Exp. 10, Secrétariat mathématique, Paris, 1968, p. 33. MR 0236777
  • [7] H. Brezis, Équations et inéquations non linéaires dans les espaces vectoriels en dualité, Ann. Inst. Fourier (Grenoble) 18 (1968), no. fasc. 1, 115–175. MR 270222
  • [8] L. Caffarelli, R. Kohn and L. Nirenberg, First order interpolation inequalities with weights, Compositio Math. 53 (1984), no. 3, 259–275. MR 768824
  • [9] T. Chacón Rebollo and R. Lewandowski, Mathematical and numerical foundations of turbulence models and applications, Modeling and Simulation in Science, Engineering and Technology, Birkhäuser/Springer, New York, 2014. MR 3288092
  • [10] B. Dyda, L. Ihnatsyeva, H. Lehrbäck, J. Tuominen, and A. V. Vähäkangas, Muckenhoupt ApA_{p}-properties of distance functions and applications to Hardy-Sobolev–type inequalities, Potential Anal. 50 (2019), no. 1, 83–105. MR 3900847
  • [11] N. Hirano, Nonlinear Volterra equations with positive kernels, Nonlinear and convex analysis (Santa Barbara, Calif., 1985), Lecture Notes in Pure and Appl. Math., vol. 107, Dekker, New York, 1987, pp. 83–98.
  • [12] N. Hirano, Nonlinear evolution equations with nonmonotonic perturbations, Nonlinear Anal. 13 (1989), no. 6, 599–609.
  • [13] A. Kaltenbach and M. Ružička, Note on the existence theory for pseudo-monotone evolution problems, J. Evol. Equ. 21 (2021), no. 1, 247–276. MR 4238205
  • [14] A. Kufner, Weighted Sobolev spaces, A Wiley-Interscience Publication, John Wiley & Sons, Inc., New York, 1985, Translated from the Czech. MR 802206
  • [15] A. Kufner, O. John, and S. Fučík, Function Spaces, Academia, Praha, 1977.
  • [16] R. Landes and V. Mustonen, A strongly nonlinear parabolic initial-boundary value problem, Ark. Mat. 25 (1987), no. 1, 29–40. MR 918378
  • [17] W. J. Layton, A nonlinear, subgridscale model for incompressible viscous flow problems, SIAM J. Sci. Comput. 17 (1996), no. 2, 347–357. MR MR1374284 (96k:76073)
  • [18] J. Lehrbäck and A. V. Vähäkangas, In between the inequalities of Sobolev and Hardy, J. Funct. Anal. 271 (2016), no. 2, 330–364.
  • [19] J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod, Gauthier-Villars, Paris, 1969. MR MR0259693 (41 #4326)
  • [20] J. Nečas, Sur une méthode pour résoudre les équations aux dérivées partielles du type elliptique, voisine de la variationnelle, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3) 16 (1962), 305–326.
  • [21] D. D. Nguyen, Some results on turbulent models, Ph.D. thesis, Université Rennes I, 2020.
  • [22] A. M. Obuhov, Turbulence in an atmosphere with inhomogeneous temperature, Akad. Nauk SSSR. Trudy Inst. Teoret. Geofiz. 1 (1946), 95–115. MR 0023169
  • [23] Y. Rong, W. Layton, and H. Zhao, Extension of a simplified Baldwin-Lomax model to nonequilibrium turbulence: model, analysis and algorithms, Numer. Methods Partial Differential Equations 35 (2019), no. 5, 1821–1846. MR 3985938
  • [24] N. Shioji, Existence of periodic solutions for nonlinear evolution equations with pseudo-monotone operators, Proc. Amer. Math. Soc. 125 (1997), no. 10, 2921–2929.
  • [25] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton Mathematical Series, No. 30, Princeton University Press, Princeton, N.J., 1970. MR 0290095
  • [26] E. R. van Driest, On turbulent flow near a wall, J. Aerospace Sci. 23 (1956), 1007–1011.
  • [27] E. Zeidler, Nonlinear functional analysis and its applications. II/A, Springer, New York, 1990, Linear monotone operators.
  • [28] E. Zeidler, Nonlinear functional analysis and its applications. II/B, Springer-Verlag, New York, 1990, Nonlinear monotone operators, Translated from the German by the author and Leo F. Boron. MR 1033498