In this paper, we investigate the blow-up criterion of an evolutionary partial differential equation system controlling the flows of incompressible viscoelastic rate-type fluids with stress-diffusion, where the extra stress tensor describing the elastic response of the fluid is purely spherical. By utilizing this criterion, the global well-posedness of the system can be readily obtained. Despite being a physical simplification, this model exhibits features that necessitate novel mathematical approaches to tackle the technically complex structure of the associated internal energy, as well as the more complicated forms of the corresponding entropy and energy fluxes. The paper provides the first rigorous proof of the existence of a global solution to the model under small initial data.
Citation: Xi Wang. Blow-up criterion for the 3-D inhomogeneous incompressible viscoelastic rate-type fluids with stress-diffusion[J]. AIMS Mathematics, 2025, 10(1): 1826-1841. doi: 10.3934/math.2025084
[1] | Xi Wang, Xueli Ke . Global unique solutions for the 2-D inhomogeneous incompressible viscoelastic rate-type fluids with stress-diffusion. AIMS Mathematics, 2024, 9(11): 29806-29819. doi: 10.3934/math.20241443 |
[2] | Jonathan D. Evans, Morgan L. Evans . Stress boundary layers for the Giesekus fluid at the static contact line in extrudate swell. AIMS Mathematics, 2024, 9(11): 32921-32944. doi: 10.3934/math.20241575 |
[3] | Abdelbaki Choucha, Salah Boulaaras, Asma Alharbi . Global existence and asymptotic behavior for a viscoelastic Kirchhoff equation with a logarithmic nonlinearity, distributed delay and Balakrishnan-Taylor damping terms. AIMS Mathematics, 2022, 7(3): 4517-4539. doi: 10.3934/math.2022252 |
[4] | Yaxin Zhao, Xiulan Wu . Asymptotic behavior and blow-up of solutions for a nonlocal parabolic equation with a special diffusion process. AIMS Mathematics, 2024, 9(8): 22883-22909. doi: 10.3934/math.20241113 |
[5] | Mingyu Zhang . On the Cauchy problem of 3D nonhomogeneous micropolar fluids with density-dependent viscosity. AIMS Mathematics, 2024, 9(9): 23313-23330. doi: 10.3934/math.20241133 |
[6] | Salah Boulaaras, Abdelbaki Choucha, Bahri Cherif, Asma Alharbi, Mohamed Abdalla . Blow up of solutions for a system of two singular nonlocal viscoelastic equations with damping, general source terms and a wide class of relaxation functions. AIMS Mathematics, 2021, 6(5): 4664-4676. doi: 10.3934/math.2021274 |
[7] | Yina Lin, Qian Zhang, Meng Zhou . Global existence of classical solutions for the 2D chemotaxis-fluid system with logistic source. AIMS Mathematics, 2022, 7(4): 7212-7233. doi: 10.3934/math.2022403 |
[8] | Keltoum Bouhali, Sulima Ahmed Zubair, Wiem Abedelmonem Salah Ben Khalifa, Najla ELzein AbuKaswi Osman, Khaled Zennir . A new strict decay rate for systems of longitudinal $ m $-nonlinear viscoelastic wave equations. AIMS Mathematics, 2023, 8(1): 962-976. doi: 10.3934/math.2023046 |
[9] | Ahmed Himadan . Well defined extinction time of solutions for a class of weak-viscoelastic parabolic equation with positive initial energy. AIMS Mathematics, 2021, 6(5): 4331-4344. doi: 10.3934/math.2021257 |
[10] | Qian Li . General and optimal decay rates for a viscoelastic wave equation with strong damping. AIMS Mathematics, 2022, 7(10): 18282-18296. doi: 10.3934/math.20221006 |
In this paper, we investigate the blow-up criterion of an evolutionary partial differential equation system controlling the flows of incompressible viscoelastic rate-type fluids with stress-diffusion, where the extra stress tensor describing the elastic response of the fluid is purely spherical. By utilizing this criterion, the global well-posedness of the system can be readily obtained. Despite being a physical simplification, this model exhibits features that necessitate novel mathematical approaches to tackle the technically complex structure of the associated internal energy, as well as the more complicated forms of the corresponding entropy and energy fluxes. The paper provides the first rigorous proof of the existence of a global solution to the model under small initial data.
In this paper, we investigate the global well-posedness of the following system:
{ut+u⋅∇u−Δu+∇Π+σdiv(∇b⊗∇b−12|∇b|2I)=0,bt+u⋅∇b+1ν(e′(b)−σΔb)=0,divu=0,(u,b)(t,x)|t=0=(u0,b0)(x), | (1.1) |
where u=(u1,u2,u3) and b=b(x,t) represent the velocity field of the fluid and the spherical component of the elastic strain tensor, respectively. The scalar pressure function is denoted by Π=Π(x,t). Generally, the viscosity coefficients ν and σ are positive constants, and e(⋅) denotes a smooth convex function of b. Furthermore, it is assumed that the second derivative of e with respect to b, which is denoted as e″(b), is bounded by a constant C.
The system we study in this paper controls the motion of non-Newtonian fluids, which is described by a simplified viscoelastic rate-type model incorporating a stress-diffusion component. It retains many qualitative characteristics of more complex viscoelastic rate-type models, which makes it an important tool for studying the mathematical properties of non-Newtonian fluids. In this paper, we present a proof regarding the blow-up criterion applicable to this model. In the context of differential equations, when we say that the solution of an equation containing the "time" variable undergoes "blow-up", it typically means that the domain of definition of the solution is finite, and some kind of "undesirable" situation occurs at the endpoint of the time interval: The solution may tend towards infinity or it may lose smoothness (which can cause the differential equation to lose physical meaning), among other possibilities. This is an important and challenging phenomenon. As time progresses, whether it is displacement, velocity, heat, field intensity, or any other form of response, they will all tend towards infinity.
On the other hand, the presence of the stress-diffusion term in the governing equations is significant for several reasons. First, it enhances the qualitative mathematical properties of the governing equations to a certain degree. Second, the existence of the diffusion term has a profound impact on the dynamical behavior predicted by the system of governing equations, as exemplified by its utilization in modeling the shear banding phenomenon. There are two interpretations of the stress-diffusion term: One views it as a consequence of a nonlocal energy storage mechanism, while the other sees it as a result of a nonlocal entropy production mechanism. These different interpretations of the stress-diffusion mechanism lead to distinct evolution equations for temperature. Hence, understanding the thermodynamic foundation of the derived models is crucial when examining the nonlinear stability of equilibrium resting states. Moreover, in the current model, the presence of the additional stress tensor disrupts the original structure outlined in [1]. Growth strain induces stress, which, in turn, affects the diffusion process through chemical potential and diffusivity. Consequently, substantial modifications must be made to this part of the theory.
This model combines features of the Oldroyd-B and Giesekus models, and incorporates diffusion terms. Moreover, the fluid type studied in this investigation is characterized by its elastic response, which can be represented by spherical strain (a scalar multiplier of the identity tensor), specifically a scalar multiplication of the identity tensor. A detailed derivation process of the mathematical model can be found in the appendix of [2] (carrying on the developments outlined in [3,4,5]). Although mathematically complex, the model is capable of capturing the complex microstructure of fluids, which is essential for predicting and simulating the flow of non-Newtonian fluids. The rate-type fluid models with stress-diffusion are popular in the modeling of shear and vorticity banding phenomena; see, for example, the reviews in [6,7,8]. The results of this model provide profound insights into the behavior of viscoelastic fluids and offer mathematical tools and theoretical foundations for fluid dynamics problems in engineering and scientific applications. This task is particularly interesting if one considers fluid models that include terms that are associated with the elastic properties of the material.
When b=0, the system reduces to the incompressible Navier-Stokes equations. The theoretical investigation of these equations was initiated by the foundational works of Leray [9] and Hopf [10] on the global existence of a weak solution. Regarding the local well-posedness of strong solutions, Ciga and Miyakawa [11], Kato [12], and others researched it using semigroup theory. Abidi, Gui, and Zhang [13] explored the global well-posedness of solutions in Besov spaces for systems with a bounded density. More recently, the authors of [14] also investigated the global existence and uniqueness of solution to two-dimensional (2-D) inhomogeneous incompressible Navier-Stokes equations in critical spaces.
Moreover, in the case where σ=0, our Eq (1.1) can be conceptualized as the coupling of two systems: The inhomogeneous Navier-Stokes equation governing a fluid with constant density, and a transport equation with the damped term e′(b). As we mentioned above, research on the three-dimensional (3-D) incompressible Navier-Stokes equation is indeed extensive, with numerous studies, such as [1,15,16] and others. On the other hand, for when σ>0, Wang and Zhang [17] investigated the global well-posedness of the 2-D Boussinesq system incorporating the temperature-dependent viscosity and thermal diffusivity. Abidi and Zhang [18] explored the global well-posedness of a 3-D Boussinesq system with variable viscosity.
For the system we study in this paper, M. Bulíček, J. Málek, Vit Pr˚usǎ, and E. Süli have conducted some related research in recent years. They [3] established the long-time and large-data existence of a weak solution to the evolutionary spatially periodic problem associated with the set of governing equations for compressible fluids with variable density in 3-D space. M. Bathory, M. Bulíček, and J. Málek [19] explored 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 confined in a closed 3-D container.
Compared with these reference papers, our primary focus is on the simplified model of viscoelastic fluids with stress-diffusion, which was derived in [2]. In this model, we specifically consider the scenario where the extra stress tensor, which describes the elastic response of the fluid, is purely spherical. Furthermore, we have incorporated the divergence-free condition to investigate the incompressible situation. In this paper, we have rigorously proven, for the first time, the existence and uniqueness of the local solution to the system using the energy method. Additionally, we have obtained global well-posedness under the assumption of small initial data through the application of a blow-up criterion.
Our main result reads as follows:
Theorem 1.1. Assume that when s>32 and the initial data (u0,b0) satisfy u0∈Hs(R3), b0∈Hs+1(R3) and divu0=0, the following hold:
(1) (Local existence) There exists a small time T∗>0 such that (1.1) has a unique strong solution (u,b) on (0,T∗) satisfying
u∈C(0,T∗;Hs)∩L2(0,T∗;Hs+1),b∈C(0,T∗;Hs+1)∩L2(0,T∗;Hs+2). |
(2) (Blow-up criterion) T∗ is the finite blow-up time of (u,b) if and only if
∫T∗0(‖∇u‖4L2+‖∇2b‖4L2)dτ=+∞. |
(3) (Global existence) If there exists a small constant ε>0 such that
‖u0‖Hs+‖b0‖Hs+1<ε, |
then T∗=∞.
The remainder of this paper is structured as follows. In Section 2, we introduce preliminary knowledge that will be subsequently used throughout the text. Next, the primary focus of the subsequent sections is on proving Theorem 1.1. In Section 3, we utilize Friedrich's method to construct approximate solutions for the system in (1.1). By employing a priori estimates to the approximate system, we establish the existence and uniqueness of the local strong solution for (1.1). In Section 4, we derive a blow-up criterion by a contradiction argument to determine whether the local solution exhibits blow-up behavior. Finally, we demonstrate the global existence of the solution under the assumption that the initial data are sufficiently small, thereby completing the proof of (1.1).
Let us complete this section with the notations we are going to use in the following.
(1) If u, v are two functions, let us set u:=v to define u as equal to v.
(2) For X being a Banach space and I being an interval of R, 1≤q≤∞, we denote by Lq(I;X) the set of measurable function on I with values in X. Sometimes, for convenience in this paper, we write it as Lq(X).
(3) For m∈N, fi∈Lp(i=1,2,...,m), we write ‖(f1,f2,...,fm)‖Lp:=m∑i=1‖fi‖Lp.
(4) For a≲b, we mean that there exists a uniform constant C, which may be different in different lines, such that a≤Cb.
(5) For m>1, x∈Rn, and a vector-valued function u(x)=(u1,u2,...,um), we define
∇u:=(u1x1⋯u1xn⋮⋱⋮u1x1⋯umxn), |
Δu:=(n∑i=1u1xixi,n∑i=1u2xixi,...,n∑i=1umxixi). |
First, in order to facilitate the construction of an approximate equation for the system in (1.1), we introduce two projection operators Pn and P here.
Definition 2.1. (see [20]) Let us define the Helmholtz projection operator P with a divergence-free condition by
P=(δij+RiRj)1≤i,j≤3, |
where Rj denotes the Riesz transform and satisfies F(Rjf)(ξ)=−iξj|ξ|Ff(ξ).
Definition 2.2. (see [21]) The projection operator Pn is defined as follows:
Pnf(x)=F−1(1Bn)(ξ)Ff(ξ)(x), |
where the Fourier operator Ff(ξ):=∫R3f(x)e−ixξdx, and the characteristic function 1Bn is defined on the surface of a sphere Bn centered at the origin with a radius n.
Clearly, based on the definitions of Pn and P provided above, it is straightforward to deduce that PnP=PPn and P2n=Pn. The proofs of these statements are omitted here for brevity.
Next, we recall some basic definitions and properties related to the Littlewood-Paley theory (see [21] for more details). Assume that ψ and θ are two functions in C∞(R3) with
Suppˆψ⊆{34≤|ξ|≤83},Suppˆθ⊆{|ξ|≤43}, |
and ˆθ(ξ)+∑j≥0ˆψ(2−jξ)=1(∀ξ∈R3).
Definition 2.3. (see [21]) We define the Littlewood-Paley operator Δj in R3 as follows:
Δjf(x):=ψ∗f(x)=∫R3ψj(x−y)f(y)dy;ψj(x):=23jψ(2jx), j≥0;Δ−1f(x):=θ∗f(x);Sjf(x):=j−1∑k=−1Δkf(x). |
In particular, utilizing the definition of the Littlewood-Paley operator provided above, we can re-describe the Sobolev space Hs(R3).
Definition 2.4. (see [22]) Let s∈R; the Sobolev space Hs(R3) can then be defined as
Hs(R3):={u∈D′(R3):∑j≥−122js‖Δju‖2L2<∞}. |
We set the norm as
‖u‖2Hs:=∑j≥−122js‖Δju‖2L2. |
Definition 2.5. (see [22]) We define the space ˜L∞(0,T;Hs) by the norm:
‖u‖2˜L∞(0,T;Hs):=∑j≥−122js‖Δju‖2L∞(0,T;L2)<∞. |
Next, we introduce some lemmas that will be used in the following sections.
Lemma 2.1. (see [23]) Assume that k∈N, 1≤p≤q≤∞, and j≥0, α is a multi-index (i.e., an element of N3), ∀f∈Lp(R3), and there exists a positive constant C which is independent of j such that
‖∂αΔjf‖Lq+‖∂αSjf‖Lq≤C2j|α|+3j(1p−1q)‖f‖Lp. |
Lemma 2.2. (see [24]) Let s>0 and f∈Hs(R3)∩L∞(R3); F(⋅) is a smooth function on R and F(0)=0. The following thus holds
‖F(f)‖Hs≲(1+‖f‖L∞)⌊s⌋+1‖f‖Hs. |
Lemma 2.3. (see [25]) Let s>0 and set Λs:=√−Δ, then
‖u(x)‖2Hs=∫R3(Λsu)2(x)dx, |
and for ∀f,g∈Hs(R3)∩L∞(R3), we have
‖Λs(fg)‖L2≲‖f‖L∞‖Λsg‖L2+‖g‖L∞‖Λsf‖L2. |
Lemma 2.4. (see [26]) (Lions-Aubin lemma) Assume that X↪↪Y↪Z, where X, Z are reflexible spaces, and X is dense in Z, then let W:={u∈Lp0(0,T;X),ut∈Lp1(0,T;X),and1<p0,p1≤∞}. It then holds that W↪↪Lp0(0,T;Y).
This section is devoted to the proof of the local well-posedness of Eq (1.1). To clearly illustrate the process, we will divide it into five steps.
Step 1. Construct approximate equations.
We use the Friedrich's method to construct the approximate solutions. Recall the projector operator Pn and the Helmholtz projection operator P stated in Definitions 2.1 and 2.2, respectively, and then utilize them to introduce the following approximate system for (1.1):
{∂tun+PnP(Pnun⋅∇Pnun)−PnΔPnun+σPnP(∇PnbnΔPnbn)=0,∂tbn+Pn(Pnun⋅∇Pnbn)+1νPn(e′(Pnbn)−σΔPnbn)=0,(un,bn)(t,x)|t=0=(Pnu0,Pnb0)(x), | (3.1) |
where we used the fact that
div(∇b⊗∇b−12|∇b|2I)=∇bΔb. | (3.2) |
Step 2. Energy estimates to approximate equations.
First, we take the scalar product of (3.1)1 with un and integrate the resulting expression over R3, which gives
12ddt‖un‖2L2+‖∇Pnun‖2L2+σ∫R3∇Pnbn⋅ΔPnbn⋅Pnundx=0. | (3.3) |
Next, by multiplying (3.1)2 by −σΔbn and integrating in space, we have
σ2ddt‖∇bn‖2L2+σ2ν‖∇2Pnbn‖2L2+σν∫R3e″(Pnbn)|∇Pnbn|2dx=−σ∫R3Pnun⋅∇Pnbn⋅ΔPnbn. | (3.4) |
Hence, by summing up (3.3) and (3.4), we get
ddt(12‖un‖2L2+σ2‖∇bn‖2L2)+‖∇Pnun‖2L2+σ2ν‖∇2Pnbn‖2L2≤0. | (3.5) |
By then integrating it over [0,Tn], one eventually gets
un∈L∞(0,Tn;L2),∇Pnun∈L2(0,Tn;L2),∇bn∈L∞(0,Tn;L2),∇2Pnbn∈L2(0,Tn;L2). | (3.6) |
In addition, taking the L2 inner product of the bn equation with Pn(e′(Pnbn)) gives
ddt‖Pn(e(bn))‖L1+1ν‖Pn(e′(Pnbn))‖2L2≤12ν‖Pn(e′(Pnbn))‖2L2+C‖ΔPn(bn)‖2L2. |
Next, doing integration by parts and integrating it over [0,Tn], which, along with (3.6) shows that
Pn(e(Pnbn))∈L∞(0,Tn;L1),Pn(e′(Pnbn))∈L2(0,Tn;L2). | (3.7) |
Therefore, from (3.6) and (3.7), we observe that each term in (3.1)1 and (3.1)2 is bounded in L2(R3). Consequently, it follows from the Cauchy-Lipschitz theorem that there exists a unique solution (un,bn)∈C(0,Tn;L2) to the system in (3.1). Due to the uniqueness of the solution, one has Pnun=un and Pnbn=bn. Hence the approximate system in (3.1) simplifies to the following form:
{∂tun+PnP(un⋅∇un)−Δun+σPn(∇bnΔbn)=0,∂tbn+Pn(un⋅∇bn)+1νPn(e′(bn)−σΔbn)=0,(un,bn)(t,x)|t=0=(Pnu0,Pnb0)(x). | (3.8) |
Moreover, the inequalities in (3.6) and (3.7) can also be reformulated as follows:
un∈L∞(0,Tn;L2)∩L2(0,Tn;˙H1),bn∈L∞(0,Tn;H1)∩L2(0,Tn;H2),e(bn)∈L∞(0,Tn;L1),e′(bn)∈L2(0,Tn;L2). | (3.9) |
In what follows, we focus on estimating ‖un‖Hs and ‖bn‖Hs+1.
Applying the operator Λs to (3.8)1 and taking the L2 inner product of the resulting equation with Λsun, followed by integrating it over R3, one obtains
12ddt‖Λsun‖2L2+∫R3(Λs(un⋅∇un)⋅Λsun+|Λs∇un|2+σΛs(∇bnΔbn)⋅Λsun)dx=0. | (3.10) |
Similarly, we perform the same operation on the bn equation. Applying the Λs operator to (3.8)2 and multiplying the resulting expression by Λsbn, −ΛsΔbn, by integrating them over R3, we then have
12ddt‖Λsbn‖2L2+∫R3(Λs(un⋅∇bn)Λsbn+1νΛse′(bn)Λsbn+σν|Λs∇bn|2)dx=0 | (3.11) |
and
12ddt‖Λs∇bn‖2L2+∫R3(−Λs(un⋅∇bn)ΛsΔbn−1νΛse′(bn)ΛsΔbn+σν|ΛsΔbn|2)dx=0. | (3.12) |
Summing Eqs (3.10)–(3.12) gives
12ddt‖(Λsun,Λsbn,Λs∇bn)‖2L2+(‖Λs∇un‖2L2+σν‖Λs∇bn‖2L2+σν‖ΛsΔbn‖2L2)=∫R3[−Λs(un⋅∇un)⋅Λsun−σΛs(∇bnΔbn)⋅Λsun−Λs(un⋅∇bn)Λsbn−1νΛse′(bn)Λsbn+Λs(un⋅∇bn)ΛsΔbn+1νΛse′(bn)ΛsΔbn]dx. |
According to Hölder's inequality, Young's inequality, and Lemmas 2.2 and 2.3, we can estimate each term on the right side of the above inequality:
∫R3−Λs(un⋅∇un)⋅Λsundx≤|∫R3Λs(un⊗un)Λs∇undx|≤‖Λs(un⊗un)‖L2‖Λs∇un‖L2≤2‖un‖L∞‖Λsun‖L2‖Λs∇un‖L2≤14‖Λs∇un‖2L2+C‖un‖2L∞‖Λsun‖2L2,∫R3−σΛs(∇bnΔbn)⋅Λsundx≤σ|∫R3Λs(∇bn⊗∇bn)Λs∇undx|≤14‖Λs∇un‖2L2+C‖∇bn‖2L∞‖Λs∇bn‖2L2,∫R3−Λs(un⋅∇bn)Λsbndx≤|∫R3Λs(unbn)Λs∇bndx|≤σ2ν‖Λs∇bn‖2L2+C‖un‖2L∞‖Λsbn‖2L2+C‖bn‖2L∞‖Λsun‖2L2,∫R3−1νΛse′(bn)Λsbndx≤|∫R3Λse′(bn)Λsbndx|≤CJ(‖bn‖L∞)‖Λsbn‖2L2,∫R3Λs(un⋅∇bn)ΛsΔbndx≤|∫R3Λs(un⋅∇bn)ΛsΔbndx|≤σ2ν‖ΛsΔbn‖2L2+C‖un‖2L∞‖Λs∇bn‖2L2+C‖∇bn‖2L∞‖Λsun‖2L2,∫R31νΛse′(bn)ΛsΔbndx≤1ν|∫R3Λse′(bn)ΛsΔbndx|≤1ν|∫R3Λs+1e′(bn)Λs∇bndx|≤CJ(‖bn‖L∞)‖Λs∇bn‖2L2, |
where J(‖bn‖L∞)=(1+‖bn‖L∞)⌊s⌋+2.
Hence, plugging these estimates into the above inequality yields
ddt(‖un‖2Hs+‖bn‖2Hs+‖∇bn‖2Hs)+(‖∇un‖2Hs+‖∇bn‖2Hs+‖∇2bn‖2Hs)≲(J(‖bn‖L∞)+‖un‖2L∞+‖bn‖2L∞+‖∇bn‖2L∞)(‖un‖2Hs+‖bn‖2Hs+‖∇bn‖2Hs), |
which, together with Grönwall's inequality, gives rise to
En(t)≲En(0)exp∫t0Hn(τ)dτ, | (3.13) |
where we defined
En(t):=‖un(t)‖2Hs+‖bn(t)‖2Hs+1+∫t0(‖un‖2Hs+1+‖bn‖2Hs+2)(τ)dτ,En(0):=‖Pnu0‖2Hs+‖Pnb0‖2Hs+1,Hn(t):=J(‖bn‖L∞)+‖un‖2L∞+‖bn‖2L∞+‖∇bn‖2L∞. |
Subsequently, assume that Tn is the maximal existence time of the approximate system in (3.8). Furthermore, there exists a uniform lower bound for Tn, which ensures that the limit of time is in a certain interval. We omit the detailed proof here; the readers can refer to [27]. In addition, we set
~Tn:=sup |
Since J(\cdot) is an increasing function and s > \frac{3}{2} , it follows from the Sobolev embedding theorem that when t\in[0, \widetilde{T_n}) ,
\begin{align*} E_n(t)&\leq E_n(0)\exp\{\int_{0}^{t} [J(\|b_n\|_{H^s})+\|u_n\|_{H^s}^2+\|b_n\|_{H^s}^2+\|\nabla b_n\|_{H^s}^2]d\tau\}\\ &\leq E_n(0)\exp\{\int_{0}^{t}[J(E_n)+E_n]d\tau\}\\ &\leq E_n(0)\exp\{J(4E_n(0))t+4E_n(0)t\}. \end{align*} |
Clearly, \exp\{J(4E_n(0))t+4E_n(0)t\} is an increasing and continuous function with respect to t . Therefore, we can choose T^*\in[0, \widetilde{T_n}) to be sufficiently small such that \exp\{J(4E_n(0))t+4E_n(0)t\} < 2 . Consequently, for any t\in[0, T^*] , one has
\begin{align} E_n(t)\leq 2E_n(0), \end{align} | (3.14) |
which implies that
\begin{align} u_n\in L^\infty(0,T^*;H^s)\cap L^2(0,T^*;H^{s+1}),\quad b_n\in L^\infty(0,T^*;H^{s+1})\cap L^2(0,T^*;H^{s+2}). \end{align} | (3.15) |
Finally, with the help of Lemma 2.3 and Sobolev's embedding theorem, it follows from (3.8)_1 that
\begin{align*} &\|\partial_tu_n\|_{H^{s-1}}^2\le\|u_n\cdot\nabla u_n\|_{H^{s-1}}^2+\|\nabla^2 u_n\|_{H^{s-1}}^2+\|\nabla b_n\Delta b_n\|_{H^{s-1}}^2\\ &\le \|u_n\|_{L^\infty}^2\|\nabla u_n\|_{H^{s-1}}^2+\|\nabla u_n\|_{L^\infty}^2\| u_n\|_{H^{s-1}}^2+\|u_n\|_{H^{s+1}}^2+\|\nabla b_n\|_{L^\infty}^2\|\Delta b_n\|_{H^{s-1}}^2\\ &\quad+\|\Delta b_n\|_{L^\infty}^2\|\nabla b_n\|_{H^{s-1}}^2\\ &\le \|u_n\|_{H^s}^2\|u_n\|_{H^{s+1}}^2+\|u_n\|_{H^{s+1}}^2+\|\nabla b_n\|_{H^s}^2\|b_n\|_{H^{s+1}}^2+\| b_n\|_{H^{s+2}}^2\| b_n\|_{H^{s}}^2, \end{align*} |
which, combined with (3.15) , shows that
\begin{align} \partial_tu_n\in L^2(0,T^*;H^{s-1}). \end{align} | (3.16) |
Similarly, performing the same operation on (3.8)_2 yields
\begin{align} \partial_tb_n\in L^2(0,T^*;H^s). \end{align} | (3.17) |
Thus, we have completed the desired estimates for the approximate solutions.
Step 3. The existence of a local solution for the system in (1.1).
Based on the estimates of (3.15) to (3.17) derived in Step 2, the Lions-Aubin compactness theorem guarantees the existence of the subsequence (u_{n_k}, b_{n_k}) of (u_n, b_n) and a function (u, b)\in L^\infty(0, T^*; H^s)\cap L^2(0, T^*; H^{s+1})\times L^\infty(0, T^*; H^{s+1})\cap L^2(0, T^*; H^{s+2}) such that for any s' < s , when k\rightarrow \infty ,
\begin{align*} u_{n_k}\rightarrow u \qquad in \quad L^2(0,T^*;H_{loc}^{s'+1}),\\ b_{n_k}\rightarrow b \qquad in \quad L^2(0,T^*;H_{loc}^{s'+2}). \end{align*} |
Additionally, it is straightforward to observe that (u, b) satisfies Eq (1.1) by taking the limit of the approximate equations; thus we have proved the uniqueness of the system in (1.1).
Step 4. Continuity of the solution.
To establish the continuity of the solution, we apply the operator \Delta_j to both sides of (3.5) and then integrate it on [0, T^*] . Subsequently, multiplying the integrated expression by 2^{2js} and ultilizing Definition 2.5 gives rise to
\begin{align*} \|u_n\|_{\tilde{L}^{\infty}(0,T^*;H^{s})}+\|b_n\|_{\tilde{L}(0,T^*;H^{s+1})}\le C. \end{align*} |
Therefore, we easily infer the improved estimates for (u, b) that
\begin{align} \|u\|_{\tilde{L}^{\infty}(0,T^*;H^{s})}+\|b\|_{\tilde{L}^{\infty}(0,T^*;H^{s+1})}\le C. \end{align} | (3.18) |
Meanwhile, Definition 2.4 indicates that there exists a positive integer N such that
\begin{align*} \sum\limits_{j = N}^{\infty}2^{2js}\|\Delta_j u\|_{L^\infty(0,T^*;L^2)}^2\le\frac{\varepsilon}{4}. \end{align*} |
Hence, for any t\in (0, T^*) and any \delta such that t+\delta\in (0, T^*) , one obtains
\begin{align*} &\|u(t+\delta)-u(t)\|_{H^s}^2\\ = &\sum\limits_{j = -1}^{N}2^{2js}\|\Delta_ju(t+\delta)-\Delta_ju(t)\|_{L^2}^2+\sum\limits_{j = N}^{\infty}2^{2js}\|\Delta_j u(t+\delta)-\Delta_ju(t)\|_{L^2}^2\\ \le&\sum\limits_{j = -1}^{N}2^{2js}\|\int_{t}^{t+\delta}\partial_\tau\Delta_ju(\tau)d\tau\|_{L^2}^2+\frac{\varepsilon}{2}\\ \le&\sum\limits_{j = -1}^{N}2^{2js}|\delta|\int_{t}^{t+\delta}\|\partial_\tau u(\tau)\|_{L^2}^2d\tau+\frac{\varepsilon}{2}\\ \le&2N2^{2sN}\|\partial_t u\|_{L^2(0,T^*;L^2)}^2|\delta|+\frac{\varepsilon}{2}, \end{align*} |
where we have used Hölder's inequality and Lemma 2.1.
Taking \delta which is small enough that 2N2^{2sN}|\delta|\|\partial_tu(t)\|_{L^2(0, T^*; L^2)}^2\le\frac{\varepsilon}{2} , one can easily check that for any \varepsilon > 0 , a \delta > 0 exists such that
\|u(t+\delta)-u(t)\|_{H^s}^2\le\varepsilon, |
which yields
\begin{align} u\in C(0,T^*;H^s). \end{align} | (3.19) |
On the other hand, by the same argument, the fact b\in \tilde{L}^\infty(0, T^*; H^{s+1}) ensures that b\in C(0, T^*; H^{s+1}) . Therefore, the continuity of the solution (u, b) to Eq (1.1) has been proved.
Step 5. Uniqueness of the solution.
Let (u^{(1)}, b^{(1)}, \Pi^{(1)}) and (u^{(2)}, b^{(2)}, \Pi^{(2)}) be two solutions of Eq (1.1) with the same initial data. We denote \tilde{u}: = u^{(2)}-u^{(1)} , \tilde{b}: = b^{(2)}-b^{(1)} , and \tilde{\Pi}: = \Pi^{(2)}-\Pi^{(1)} ; (\tilde{u}, \tilde{b}, \tilde{\Pi}) then solves the following equation:
\begin{equation} \begin{cases} \tilde{u}_t+\tilde{u}\cdot\nabla u^{(2)}+u^{(1)}\cdot\nabla\tilde{u}-\Delta\tilde{u}+\nabla \tilde{\Pi}+\sigma \nabla b^{(2)}\Delta\tilde{b}+\sigma \nabla\tilde{b}\Delta b^{(1)} = 0,\\ \tilde{b}_t+\tilde{u}\cdot\nabla b^{(2)}+u^{(1)}\cdot\nabla\tilde{b}+\frac{1}{\nu}(e'(b^{(2)})-e'(b^{(1)}))-\frac{\sigma}{\nu}\Delta\tilde{b} = 0,\\ \mathrm{div}\,\tilde{u} = 0,\\ \tilde{u} = 0,\quad \tilde{b} = 0. \end{cases} \end{equation} | (3.20) |
As we proved in Step 2, we can apply the standard L^2 energy estimate to \tilde{u} and \tilde{b} , which gives
\begin{align*} &\frac{1}{2}\frac{d}{dt}(\|\tilde{u}\|_{L^2}^2+\| \tilde{b}\|_{L^2}^2+\sigma\|\nabla \tilde{b}\|_{L^2}^2)+(\|\nabla \tilde{u}\|_{L^2}^2+\frac{\sigma}{\nu}\|\nabla\tilde{b}\|_{L^2}^2+\frac{\sigma^2}{\nu}\|\Delta\tilde{b}\|_{L^2}^2)\\ \le &\int_{\mathbb{R}^3}\Big[\tilde{u}\cdot\nabla u^{(2)}\cdot\tilde{u} +\sigma\nabla\tilde{b}\Delta b^{(1)}\cdot\tilde{u}+ \tilde{u}\cdot\nabla b^{(2)}\cdot\tilde{b}+u^{(1)}\cdot\nabla\tilde{b}\tilde{b}+\frac{1}{\nu}(e'(b^{(2)})-e'(b^{(1)}))\tilde{b}\\ &\quad+ \sigma u^{(1)}\cdot\nabla \tilde{b}\Delta \tilde{b}+\frac{\sigma}{\nu}\big(e'(b^{(2)})-e'(b^{(1)})\big)\Delta\tilde{b}\Big]dx. \end{align*} |
Combining integration by parts with Gagliardo-Nirenberg inequality, H \ddot{o} lder's inequality, and Young's inequality, we can estimate the right side of the inequality above as follows:
\begin{align*} &\int_{\mathbb{R}^3}\tilde{u}\cdot\nabla u^{(2)}\cdot\tilde{u}dx\le \|\nabla u^{(2)}\|_{L^\infty}\|\tilde{u}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}\sigma\nabla\tilde{b}\Delta b^{(1)}\cdot\tilde{u}dx\le \|\nabla\tilde{b}\|_{L^2}\|\Delta b^{(1)}\|_{L^\infty}\| \tilde{u}\|_{L^2}\le\frac{\sigma}{6\nu}\|\nabla\tilde{b}\|_{L^2}^2+\|\Delta b^{(1)}\|_{L^\infty}^2\| \tilde{u}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}\tilde{u}\cdot\nabla b^{(2)}\cdot\tilde{b}dx\le \|\tilde{u}\|_{L^2}^{\frac{1}{2}}\|\nabla\tilde{u}\|_{L^2}^{\frac{1}{2}}\|\nabla b^{(2)}\|_{L^{2}}\|\nabla\tilde{b}\|_{L^2}\\ &\qquad\qquad\qquad\qquad\le \frac{1}{2}\|\nabla\tilde{u}\|_{L^2}^2+\frac{\sigma}{6\nu}\|\nabla\tilde{b}\|_{L^2}^2+C\|\nabla b^{(2)}\|_{L^{2}}^4\|\tilde{u}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}u^{(1)}\cdot\nabla\tilde{b}\tilde{b}dx\le\| u^{(1)}\|_{L^\infty}\|\nabla\tilde{b}\|_{L^2}\|\tilde{b}\|_{L^2}\le \frac{\sigma}{6\nu}\|\nabla\tilde{b}\|_{L^2}^2+C\| u^{(1)}\|_{L^\infty}^2\|\tilde{b}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}\frac{1}{\nu}(e'(b^{(2)})-e'(b^{(1)}))\tilde{b}dx\le e''(\xi)\|\tilde{b}\|_{L^2}^2\le C\|\tilde{b}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}\sigma u^{(1)}\cdot\nabla \tilde{b}\Delta \tilde{b}dx\le \| u^{(1)}\|_{L^\infty}\|\nabla\tilde{b}\|_{L^2}\|\Delta\tilde{b}\|_{L^2}\le\| u^{(1)}\|_{L^\infty}^2\|\nabla\tilde{b}\|_{L^2}^2+\frac{\sigma^2}{4\nu}\|\Delta\tilde{b}\|_{L^2}^2,\\ &\int_{\mathbb{R}^3}\frac{\sigma}{\nu}\big(e'(b^{(2)})-e'(b^{(1)})\big)\Delta\tilde{b}dx\le\|e''(\xi)\tilde{b}\|_{L^2}\|\Delta\tilde{b}\|_{L^2}\le \frac{\sigma^2}{4\nu}\|\Delta\tilde{b}\|_{L^2}^2+C\| \tilde{b}\|_{L^2}^2, \end{align*} |
where \xi is a function between b^{(1)} and b^{(2)} . Putting all these estimates into the former inequality and utilizing the Sobolev embedding theorem, we can then rewrite the inequality as follows:
\begin{align*} &\frac{d}{dt}(\|\tilde{u}\|_{L^2}^2+\|\tilde{b}\|_{L^2}^2+\|\nabla\tilde{b}\|_{L^2}^2)+\|\nabla \tilde{u}\|_{L^2}^2+\|\nabla\tilde{b}\|_{L^2}^2+\|\Delta\tilde{b}\|_{L^2}^2\\ &\le C(\|u^{(2)}\|_{H^{s+1}}+\| b^{(1)}\|_{H^{s+2}}^2+\| u^{(1)}\|_{H^{s}}^2+\| b^{(2)}\|_{H^1}^4+1)(\|\tilde{u}\|_{L^2}^2+\| \tilde{b}\|_{L^2}^2+\sigma\|\nabla \tilde{b}\|_{L^2}^2). \end{align*} |
Note that
\begin{align*} \int_{0}^{t}(\|u^{(2)}\|_{H^{s+1}}+\| b^{(1)}\|_{H^{s+2}}^2+\| u^{(1)}\|_{H^{s}}^2+\| b^{(2)}\|_{H^1}^4+1)(\tau)d\tau\le C(t). \end{align*} |
Grönwall's inequality then yields
\begin{align*} \sup\limits_{t > 0}(\|\tilde{u}(t)\|_{L^2}^2+\|\tilde{b}(t)\|_{L^2}^2+\|\nabla\tilde{b}(t)\|_{L^2}^2)\le \exp (C(t))\Big(\|\tilde{u}_0\|_{L^2}^2+\|\tilde{b}_0\|_{L^2}^2+\|\nabla\tilde{b}_0\|_{L^2}^2\Big). \end{align*} |
Therefore, one can easily conclude that
\begin{align*} \tilde{u} = 0,\quad \tilde{b} = 0,\quad \nabla\tilde{b} = 0. \end{align*} |
Hence the uniqueness of the solution has been proved.
In this section, we prove the blow-up criterion of the system in (1.1) by a contradiction argument.
First of all, let T^* < \infty be the maximal existence time of the solution and assume that
\begin{align} \int_{0}^{T^*}({\|\nabla u\|}_{L^{2}}^{4}+{\|\nabla^2 b\|}_{L^{2}}^{4})d\tau < +\infty. \end{align} | (4.1) |
Taking the L^2 inner product of (1.1)_1 with -\Delta u , which, along with Hölder's and the Gagliardo-Nirenberg inequality, yields
\begin{align*} &\frac{1}{2}\frac{d}{dt}\|\nabla u\|_{L^2}^2+\|\Delta u\|_{L^2}^2\le|\int_{\mathbb{R}^3}\big(u\cdot\nabla u\cdot\Delta u+ \sigma\nabla b\Delta b\cdot\Delta u\big)dx|\\ & \le\|\nabla u\|_{L^{2}}^{\frac{3}{2}}\|\nabla^2 u\|_{L^{2}}^{\frac{3}{2}}+\sigma\|\nabla^2 b\|_{L^{2}}^{\frac{3}{2}}\|\nabla^3 b\|_{L^{2}}^{\frac{1}{2}}\|\nabla^2 u\|_{L^{2}}\\ &\le\frac{1}{4}\|\nabla ^2 u\|_{L^2}^2+\frac{\sigma}{4\nu}\|\nabla ^3 b\|_{L^2}^2+C\|\nabla u\|_{L^2}^{6}+C\|\nabla^2 b\|_{L^2}^6, \end{align*} |
Next, by applying the Laplace operator \Delta to (1.1)_2 and multiplying the resulting expression by \Delta b , and then integrating the resulting expression over \mathbb{R}^3 , we obtain
\begin{align*} &\frac{1}{2}\frac{d}{dt}\|\Delta b\|_{L^2}^2+\frac{\sigma}{\nu}\|\nabla^3 b\|_{L^2}^2 = -\int_{\mathbb{R}^3} [\Delta(u\cdot\nabla b)\cdot\Delta b+\frac{1}{\nu}\Delta e'\cdot\Delta b]dx\\ & = -\int_{\mathbb{R}^3}(\Delta u\cdot\nabla b\cdot\Delta b+ u\cdot\nabla \Delta b \cdot\Delta b+ 2\nabla u\cdot\nabla^2 b\cdot\Delta b-\frac{1}{\nu}\nabla e'\nabla^3 b)dx. \end{align*} |
As we computed in the previous example, each term on the right-hand side of the inequality can also be estimated as follows:
\begin{align*} &-\int_{\mathbb{R}^3}\Delta u\cdot\nabla b\cdot\Delta bdx\le 2 \|\nabla^2 u\|_{L^2}\|\nabla^2 b\|_{L^{2}}^{\frac{3}{2}}\|\nabla^3 b\|_{L^{2}}^{\frac{1}{2}}\le \frac{\sigma}{8\nu}\|\nabla^3 b\|_{L^2}^2+\frac{1}{4}\|\nabla^2 u\|_{L^2}^{2}+C\|\nabla^2 b\|_{L^{2}}^{6},\\ &-\int_{\mathbb{R}^3}u\cdot\nabla \Delta b \cdot\Delta bdx\le \|\nabla u\|_{L^{2}}\|\nabla^3 b \|_{L^{2}}^{\frac{3}{2}}\|\nabla ^2 b\|_{L^{2}}^{\frac{1}{2}}\le \frac{\sigma}{8\nu}\|\nabla^3 b\|_{L^2}^2+C\|\nabla u\|_{L^2}^{4}\|\nabla^2 b\|_{L^{2}}^{2},\\ &-2\int_{\mathbb{R}^3}\nabla u\cdot\nabla^2 b\cdot\Delta bdx\le\|\nabla u\|_{L^{2}}\|\nabla^3 b \|_{L^{2}}^{\frac{3}{2}}\|\nabla ^2 b\|_{L^{2}}^{\frac{1}{2}}\le \frac{\sigma}{8\nu}\|\nabla^3 b\|_{L^2}^2+C\|\nabla u\|_{L^2}^{4}\|\nabla^2 b\|_{L^{2}}^{2},\\ &\frac{1}{\nu}\int_{\mathbb{R}^3}\nabla e'\nabla^3 bdx\le C\|\nabla b\|_{L^2}\|\nabla^3 b\|_{L^2}\le \frac{\sigma}{8\nu}\|\nabla^3 b\|_{L^2}^2+C\|\nabla b\|_{L^{2}}^{2}. \end{align*} |
Hence, it follows from (3.9) and a simple simplification that
\begin{align*} \frac{d}{dt}\|(\nabla u,\Delta b)\|_{L^2}^2+\|(\Delta u,\nabla^3 b)\|_{L^2}^2 \lesssim(\|\nabla u\|_{L^{2}}^4+\|\nabla^2 b\|_{L^2}^4)(\|\nabla u\|_{L^2}^2+\|\nabla^2 b\|_{L^2}^2)+1. \end{align*} |
Then, by virtue of Grönwall's inequality and (4.1), one deduces that
\begin{align} u\in L^\infty(0,T^*;H^1)\cap L^2(0,T^*;H^2),\quad b\in L^\infty(0,T^*;H^2)\cap L^2(0,T^*;H^3). \end{align} | (4.2) |
Furthermore, the Sobolev embedding theorem gives
\begin{align} \|u\|_{L^2(L^\infty)}\le\|u\|_{L^2(H^2)},\quad\| b\|_{L^{\infty}(L^\infty)}\le \|b\|_{L^{\infty}(H^2)},\quad\|\nabla b\|_{L^2(L^\infty)}\le \|\nabla b\|_{L^2(H^2)}. \end{align} | (4.3) |
Thus, with (4.2) and (4.3), by using the same method as employed in proving Eq (3.13), one eventually finds that when s > \frac{3}{2} ,
\begin{align} \|(u,b,\nabla b)\|_{H^s}^2(T^*)& \le\|(u_0,b_0,\nabla b_0)\|_{H^s}^2\exp\{\int_{0}^{T^*}[J(\|b\|_{H^2})+\|u\|_{H^2}^2+\|b\|_{H^2}^2+\|\nabla b\|_{H^2}^2]d\tau\}\\ &\le C. \end{align} | (4.4) |
The result implies that the solution can be extended even when t = T^* , which contradicts the definition of T^* . Therefore, we have
\begin{align} \int_{0}^{T^*}({\|\nabla u\|}_{L^{2}}^{4}+{\|\nabla^2 b\|}_{L^{2}}^{4})d\tau = +\infty. \end{align} | (4.5) |
Finally, the converse is readily obvious from (4.4).
First, for simplicity, we define
\begin{align*} E(t): = \|u(t)\|_{H^s}^2+\|b(t)\|_{H^{s+1}}^2+\int_{0}^{t}(\|\nabla u\|_{H^s}^2+\| b\|_{H^{s+2}}^2(\tau)d\tau, \end{align*} |
\begin{align*} E(0): = \|u_0\|_{H^s}^2+\|b_0\|_{H^{s+1}}^2. \end{align*} |
Assuming that the solution (u, b) blows up at T^* and T^* < \infty , by utilizing the same proof of (3.14), one can find that when t\in[0, T^*) ,
\begin{align*} E(t)\le 2E(0)\le 2\varepsilon, \end{align*} |
then
\begin{align*} \int_{0}^{T^*}({\|\nabla u\|}_{L^{2}}^{4}+{\|\nabla^2 b\|}_{L^{2}}^{4})d\tau\le 4\varepsilon^2 < +\infty. \end{align*} |
On the basis of the blow-up criterion we proved in Section 4, we conclude that T^* = \infty . Therefore, the proof of Theorem 1.1 is completed.
In this paper, we studied a streamlined model that shares numerous similarities with the more complicated viscoelastic rate-type models commonly used in modeling fluids with complex microstructures. We have proven the blow-up result for this fluid, which implies the global well-posedness of the 3-D incompressible viscoelastic rate-type fluid system. In other words, this model has been mathematically proven to be thermodynamically consistent, and the blow-up criteria implies that these models are capable of describing the long-term behavior of fluids under small initial conditions and external forces. There are still many issues to explore concerning this equation, such as studying the stability of the solution or the existence of the smooth solution in the case of variable density, among others. These results will provide important preliminary insights into the mathematical properties of those more complex and practically significant models of non-Newtonian fluids used in applications.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The author would like to thank the anonymous referees for their suggestions which made the paper more readable.
The author has no relevant financial or non-financial interests to disclose. The author has no competing interests to declare that are relevant to the content of this article.
[1] | P. L. Lions, Mathematical topics in fluids mechanics. Vol.1 incompressible models, New York: Oxford Univeristy Press, 1996. https://doi.org/10.1080/00107514.2013.869264 |
[2] |
M. Bulíček, E. Feireisl, J. Málek, On a class of compressible viscoelastic rate-type fluids with stress-diffusion, Nonlinearity, 32 (2019), 4665–4681. https://doi.org/10.1088/1361-6544/ab3614 doi: 10.1088/1361-6544/ab3614
![]() |
[3] |
M. Bulíček, J. Málek, V. Pruša, E. A. S\ddot{u}li, PDE analysis of a class of thermodynamically compatible viscoelastic rate-type fluids with stress-diffusion, Amer. Math. Soc., 710 (2018), 25–53. https://doi.org/10.48550/arXiv.1707.02350 doi: 10.48550/arXiv.1707.02350
![]() |
[4] |
J. Málek, V. Pruša, T. Skřivan, E. A. S\ddot{u}li, Thermodynamics of viscoelastic rate-type fuids with stress diffusion, Phys. Fluids, 30 (2018), 023101. https://doi.org/10.1063/1.5018172 doi: 10.1063/1.5018172
![]() |
[5] |
J. Málek, V. Pruša, Derivation of equations for continuum mechanics and thermodynamics of fluids, Handbook Math. Anal. Mech. Viscous Flu., 2016, 3–72. https://doi.org/10.1007/978-3-319-10151-4_1-1 doi: 10.1007/978-3-319-10151-4_1-1
![]() |
[6] |
T. Divoux, M. A. Fardin, S. Manneville, S. Lerouge, Shear banding of complex fluids, Annu. Rev. Fluid Mech, 48 (2016), 81–103. https://doi.org/10.1146/annurev-fluid-122414-034416 doi: 10.1146/annurev-fluid-122414-034416
![]() |
[7] |
J. K. G. Dhont, W. J. Briels, Gradient and vorticity banding, Rheol. Acta, 47 (2008), 257–281. https://doi.org/10.1007/s00397-007-0245-0 doi: 10.1007/s00397-007-0245-0
![]() |
[8] |
M. A. Fardin, T. J. Ober, C. Gay, G. Gregoire, G. H. McKinley, S. Lerouge, Potential "ways of thinking" about the shear-banding phenomenon, Soft Matt., 8 (2012), 910–922. https://doi.org/10.1039/C1SM06165H doi: 10.1039/C1SM06165H
![]() |
[9] |
J. Leray, Sur le mouvement d'un liquide visqueux emplissant l'espace, Acta Math., 63 (1934), 193–248. https://doi.org/10.1007/BF02547354 doi: 10.1007/BF02547354
![]() |
[10] |
E. Hopf, über die Anfangswertaufgabe für die hydrodynamischen Grundgleichungen. Erhard Schmidt zu seinem 75. Geburtstag gewidmet, Math. Nachr., 4 (1950), 213–231. https://doi.org/10.1002/mana.3210040121 doi: 10.1002/mana.3210040121
![]() |
[11] |
Y. Giga, T. Miyawa, Solutions in L^r of the Navier-Stokes initial value problem, Arch. Ration. Mech. Anal., 85 (1985), 267–281. https://doi.org/10.1007/BF00276875 doi: 10.1007/BF00276875
![]() |
[12] |
T. Kato, Strong L^p-solutions of the Navier-Stokes equation in \mathbb{R}^m, with applications to weak solutions, Math. Z., 187 (1984), 471–480. https://doi.org/10.1007/BF01174182 doi: 10.1007/BF01174182
![]() |
[13] |
H. Abidi, G. Gui, P. Zhang, On the decay and stability of global solutions to the 3-D inhomogeneous Navier-Stokes equations, Commun. Pure. Appl. Math., 64 (2011), 832–881. https://doi.org/10.1002/cpa.20351 doi: 10.1002/cpa.20351
![]() |
[14] |
H. Abidi, G. Gui, P. Zhang, On the global existence and uniqueness of solution to 2-D inhomogeneous incompressible Navier-Stokes equations in critical spaces, J. Diff. Equ., 406 (2024), 126–173. https://doi.org/10.1016/j.jde.2024.06.004 doi: 10.1016/j.jde.2024.06.004
![]() |
[15] |
R. Danchin, Local and global well-posedness reuslts for flows of inhomogeneous viscous fluids, Adv. Differ. Equ., 9 (2004), 353–386. https://doi.org/10.57262/ade/1355867948 doi: 10.57262/ade/1355867948
![]() |
[16] |
B. Desjardins, Global existence results for the incompressible density-dependent Navier-Stokes equations in the whole space, Differ. Int. Equ., 10 (1997), 587–598. http://dx.doi.org/10.1016/j.jmaa.2016.09.050 doi: 10.1016/j.jmaa.2016.09.050
![]() |
[17] |
C. Wang, Z. Zhang, Global well-posedness for the 2-D Boussinesq system with the temperature-dependent viscosity and thermal diffusivity, Adv. Math., 228 (2011), 43–62. https://doi.org/10.1016/j.aim.2011.05.008 doi: 10.1016/j.aim.2011.05.008
![]() |
[18] |
H. Abidi, P. Zhang, On the global well-posedness of 3-D Boussinesq system with variable viscosity, Chin. Ann. Math. Ser. B, 40 (2019), 643–688. https://doi.org/10.1007/s11401-019-0156-2 doi: 10.1007/s11401-019-0156-2
![]() |
[19] |
M. Bathory, M. Bulíček, J. Málek, Large data existence theory for three-dimensional unsteady flows of rate-type viscoelastic fluids with stress diffusion, Adv. Nonlinear Anal., 10 (2020), 501–521. https://doi.org/10.1515/anona-2020-0144 doi: 10.1515/anona-2020-0144
![]() |
[20] | J. Bedrossian, V. Vicol, The mathematical analysis of the incompressible Euler and Navier-Stokes equations: An introduction, Rhode Island: American Mathematical Society, 2022. https://doi.org/10.1090/gsm/225 |
[21] | H. Bahouri, J. Chemin, R. Danchin, Fourier analysis and nonlinear partial differential equations, New York: Berlin Springer, 2011. |
[22] |
J. Chemin, Théorémes d'unicité pour le systéme de Navier-Stokes tridimensionnel, J. Anal. Math., 77 (1999), 27–50. https://doi.org/10.1007/BF02791256 doi: 10.1007/BF02791256
![]() |
[23] | J. Chemin, Perfect incompressible fluids, New York: Oxford University Press, 1998. |
[24] | H. Triebel, Theory of function spaces, Basel, Boston, Berlin: Birkhäuser Verlag, 1983. https://doi.org/10.1007/978-3-0346-0416-1 |
[25] |
T. Kato, G. Ponce, Commutator estimates and the Euler and Navier-Stokes equations, Commun, Pure Appl. Math, 41 (1988), 891–907. https://doi.org/10.1002/cpa.3160410704 doi: 10.1002/cpa.3160410704
![]() |
[26] | J. L. Lions, Some methods of solution of nonlinear boundary-value problems, Moscow: Mir, 1972. |
[27] | A. J. Majda, A. L. Bertozzi, A. Ogawa, Vorticity and incompressible flow, NewYork: Cambridge University Press, 2002. |