Processing math: 100%
Research article Topical Sections

Constructing boundary layer approximations in rotating magnetohydrodynamic fluids within cylindrical domains

  • This paper aims to investigate the effects of the Ekman-Hartmann boundary layer on rotating magnetohydrodynamics (MHD) within cylindrical domains, focusing on constructing approximate solutions within the boundary layer. We employed the multiscale analysis method to derive the approximate solutions, emphasizing the solutions at the cylinder's corners and lateral boundaries. Furthermore, we rigorously examined the asymptotic behavior of the rotating MHD flow in the limit case, proving its convergence to a two-dimensional damped and rotating dynamical system. These findings revealed the significant impact of high-speed rotation and strong magnetic fields on the structure and flow characteristics of the boundary layer, providing new insights into the dynamics of rotating MHD flows.

    Citation: Guanglei Zhang, Kexue Chen, Yifei Jia. Constructing boundary layer approximations in rotating magnetohydrodynamic fluids within cylindrical domains[J]. AIMS Mathematics, 2025, 10(2): 2724-2749. doi: 10.3934/math.2025128

    Related Papers:

    [1] Afraz Hussain Majeed, Sadia Irshad, Bagh Ali, Ahmed Kadhim Hussein, Nehad Ali Shah, Thongchai Botmart . Numerical investigations of nonlinear Maxwell fluid flow in the presence of non-Fourier heat flux theory: Keller box-based simulations. AIMS Mathematics, 2023, 8(5): 12559-12575. doi: 10.3934/math.2023631
    [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] Nadeem Abbas, Wasfi Shatanawi, Taqi A. M. Shatnawi, Fady Hasan . Theoretical analysis of induced MHD Sutterby fluid flow with variable thermal conductivity and thermal slip over a stretching cylinder. AIMS Mathematics, 2023, 8(5): 10146-10159. doi: 10.3934/math.2023513
    [4] Jishan Fan, Tohru Ozawa . Magnetohydrodynamics approximation of the compressible full magneto- micropolar system. AIMS Mathematics, 2022, 7(9): 16037-16053. doi: 10.3934/math.2022878
    [5] Kunquan Li . Analytical solutions and asymptotic behaviors to the vacuum free boundary problem for 2D Navier-Stokes equations with degenerate viscosity. AIMS Mathematics, 2024, 9(5): 12412-12432. doi: 10.3934/math.2024607
    [6] Umair Khan, Aurang Zaib, Sakhinah Abu Bakar, Anuar Ishak, Dumitru Baleanu, El-Sayed M Sherif . Computational simulation of cross-flow of Williamson fluid over a porous shrinking/stretching surface comprising hybrid nanofluid and thermal radiation. AIMS Mathematics, 2022, 7(4): 6489-6515. doi: 10.3934/math.2022362
    [7] Yellamma, N. Manjunatha, Umair Khan, Samia Elattar, Sayed M. Eldin, Jasgurpreet Singh Chohan, R. Sumithra, K. Sarada . Onset of triple-diffusive convective stability in the presence of a heat source and temperature gradients: An exact method. AIMS Mathematics, 2023, 8(6): 13432-13453. doi: 10.3934/math.2023681
    [8] Lazhar Bougoffa, Ammar Khanfer, Smail Bougouffa . Qualitative analysis on the electrohydrodynamic flow equation. AIMS Mathematics, 2024, 9(1): 775-791. doi: 10.3934/math.2024040
    [9] Essam R. El-Zahar, Ghaliah F. Al-Boqami, Haifa S. Al-Juaydi . Piecewise approximate analytical solutions of high-order reaction-diffusion singular perturbation problems with boundary and interior layers. AIMS Mathematics, 2024, 9(6): 15671-15698. doi: 10.3934/math.2024756
    [10] Kuiyu Cheng, Abdelraheem M. Aly, Nghia Nguyen Ho, Sang-Wook Lee, Andaç Batur Çolak, Weaam Alhejaili . Exothermic thermosolutal convection in a nanofluid-filled square cavity with a rotating Z-Fin: ISPH and AI integration. AIMS Mathematics, 2025, 10(3): 5830-5858. doi: 10.3934/math.2025268
  • This paper aims to investigate the effects of the Ekman-Hartmann boundary layer on rotating magnetohydrodynamics (MHD) within cylindrical domains, focusing on constructing approximate solutions within the boundary layer. We employed the multiscale analysis method to derive the approximate solutions, emphasizing the solutions at the cylinder's corners and lateral boundaries. Furthermore, we rigorously examined the asymptotic behavior of the rotating MHD flow in the limit case, proving its convergence to a two-dimensional damped and rotating dynamical system. These findings revealed the significant impact of high-speed rotation and strong magnetic fields on the structure and flow characteristics of the boundary layer, providing new insights into the dynamics of rotating MHD flows.



    Magnetohydrodynamics (MHD) investigates the interplay between conducting fluids and electromagnetic fields. It has extensive potential applications across diverse domains, encompassing energy, materials science, astrophysics, and engineering technologies [19]. In particular, the impacts of rotational and boundary layer effects on MHD are of significant research importance, as detailed in prior studies [1,2,4,9,10,11,15,20].

    The classical incompressible MHD equations constitute a set of coupled partial differential equations. Grounded in the fundamental principles of physics, such as the conservation of mass, momentum, and energy, as well as Maxwell's electromagnetic equations, they are employed to depict the behavior of conducting fluids under the influence of electromagnetic fields. The incompressible MHD equations can be summarized as follows:

    {tu+uu=pB×j+νΔu,tB=×(u×B)+ηΔB,j=σ(EB×u),E=φ,u=B=0,

    where u,p,j,B,E,φ correspond to the fluid velocity, the pressure, the current density, the magnetic field, the electric field, and the electric potential, respectively. The coefficients ν,η,σ are the kinematic viscosity, the magnetic diffusivity, and the electrical conductivity, respectively. B×j represents the Lorentz force.

    In this paper, we consider the MHD equations under the influence of the Coriolis force, with the magnetic field B=βε(0,0,1)T.

    In this scenario, the MHD equations are simplified to include only the momentum equation with the Coriolis force and the current density equation. Specifically,

    {tuε+(uε)uεεΔuε+αεe3uε+βεe3jε+1εpε=0,jεφε+e3uε=0,uε=jε=0, (1.1)

    where (t,x)R+×Ω, Ω=S×[0,1], S is smooth bounded domain of R2, αεe3uε is the Coriolis force term, and the charge conservation principle requires jε=0. We also consider Equation (1.1) under the following initial and boundary conditions:

    uε(t,x)|t=0=uε0(x), (1.2)
    uε(t,x)|Ω=0,jεn|Ω=0, (1.3)

    where n is the normal vector of Ω. Since we are considering a system (1.1) in the region Ω=S×[0,1], the boundary condition of jε is equivalent to

    jε3(t,x)|z=0,1=0,jεhns|S=0, (1.4)

    where ns is the normal vector of S.

    It is crucial to acknowledge that boundary layer effects must be considered when examining rotating fluids within bounded regions. The boundary layer concept, originally introduced by the German physicist Ludwig Prandtl, is of paramount importance in fluid dynamics. It delineates the transitional zone wherein the fluid velocity shifts from zero near the solid surface to free-flow velocity due to viscous influences. Extensive experimental and theoretical investigations have established that the flow region adjacent to the solid wall can be bifurcated into two distinct zones: one is a skinny layer near the object, called the boundary layer, where the coefficient of viscosity plays a significant role. The other is the region outside the boundary layer, which has a negligible viscosity coefficient.

    In Model (1.1), the parameter ε>0 is very small (107), with 1/ε used to describe the strength of the magnetic field and the rotation rate of the fluid. Therefore, the system in (1.1) describes the dynamic behavior of incompressible fluids with low viscosity and large force terms. Furthermore, the ratio β/α>0 represents the Elsasser number utilized to describe the relative strength between the magnetic field and fluid flow in MHD. As the Elsasser number increases, the boundary layer transitions from the Ekman type to the Hartmann type. For example, when the external force term is of the Coriolis type (β/α=0), it can simulate rotating fluids in oceans, atmospheres, or containers (see [14,15]). When magnetic effects are considered (β/α1), e3jε represents the Lorentz force. It is linked to uε through Ohm's law, as shown in (1.1)2.

    This paper considers a three-dimensional model subject to high-speed rotation and the effect of a high-intensity magnetic field (β/α=O(1)) within a bounded domain Ω. It is assumed that the direction of the rotation axis aligns with that of the mean magnetic field, both being e3=(0,0,1)T. The hydrodynamic behavior within this region is profoundly influenced by the magnetic field and rotational effects, displaying characteristics that markedly deviate from those of the interior region. Furthermore, the structure of the boundary layer exerts a substantial impact on the stability and performance of the overall flow system. For further details on MHD layers, please refer to [6,7,8,13,16,17].

    The Ekman-Hartmann layer is crucial in MHD systems with strong magnetic fields and rapid rotation. It impacts ocean currents and wind patterns in geophysics, heat management in engineering, and plasma confinement in fusion reactors. Extensive research has been devoted to the mathematical analysis of the Ekman-Hartmann boundary layer. For instance, in [6], the authors employed a matched asymptotic expansion technique to investigate the boundary layer for the half-space domain and the region between two parallel plates. Their findings revealed that the boundary layer displays nonlinear stability when the characteristic Reynolds number, defined within the boundary layer, falls below a critical threshold. This conclusion was corroborated in [16] under more generalized spectral assumptions. It is noteworthy that the models discussed in [6] and [16] represent generalizations of the system in (1.1), wherein Eq (1.1)2 is replaced by an equation governing the evolution of the magnetic field. For the simplified Model (1.1), [13] introduced a unified approach for boundary layer analysis, with special attention given to the derivation of approximate solutions in scenarios involving rotation (the Ekman layer) or magnetic fields (the Hartmann layer). Furthermore, in the intricate setting characterized by concurrent high-speed rotation and intense magnetic fields, [17] undertook a comprehensive investigation of Model (1.1) under Dirichlet boundary conditions applied to the region bounded by two parallel planes. This study effectively extended the nonlinear stability conclusion established for the Ekman-Hartmann boundary layer in [16] to encompass a broader range of initial value conditions. Subsequently, Rousset [18] proved the nonlinear stability of Ekman-Hartmann boundary layers in a spherical geometry for well-prepared initial data.

    Furthermore, investigating rotating fluids within cylindrical domains presents numerous challenges, primarily arising from the intricate interplay among hydrodynamics, rotation, and the container's geometry, particularly in the vicinity of corners and edges. Bresch, Desjardins, and Gérard-Varet [3] addressed these challenges by developing correction terms near the lateral edges while preserving the integrity of the upper and lower boundary terms as well as the interior terms.

    Before presenting the results, we provide the following definitions for convenience.

    Let =(x,y,z)T, h=(x,y)T, and h=(y,x)T. We also write Δ=2x+2y+2z and Δh=2x+2y, f=(fh,f3)T, fh=(f1,f2)T, fh,=(f2,f1)T, and

    A=(βααβ).

    In this paper, we occasionally employ the notation AB to denote the equivalence ACB, where C is a uniform constant.

    For a fixed ε>0, the mathematical behavior of Systems (1.1)–(1.4) closely resembles that of the incompressible Navier–Stokes equations. By the theory of global weak solutions, which is analogous to the Leray solutions of the incompressible Navier–Stokes equations (for further details, see [19]), this paper aims to investigate the asymptotic behavior of the weak solutions as ε approaches zero. The details are as follows.

    Theorem 1.1. Let (uε,jε)L(R+;L2(Ω)) be a family of weak solutions of Systems (1.1)(1.4) associated with the initial data uε0(x)L2(Ω). Under the following well-prepared initial data conditions: uε0=(uε0,h,uε0,3) and ˉu0,h=10uε0,hdz, satisfy

    limε0+uε0=(ˉu0,h,0)=:ˉu0,in L2(S). (1.5)

    For an universal constant C0,

    ˉu0,hL(S)<C0, (1.6)

    then ˉu(t,x,y)=(ˉuh,0) satisfies the following two-dimensional (2D) primitive type equations with the initial data ˉu0:

    {tˉuh+(ˉuhh)ˉuh+γˉuh+ηˉuh,+hˉp+βhˉφ=0,hˉuh=0,Δhˉφ=2cosτ2(α2+β2)14hˉuh,ˉuhS=0,hˉφnsS=0, (1.7)

    where γ and η are defined by γ=2(α2+β2)14(αsin(τ2)+βcos(τ2)), η=2(α2+β2)14(αcos(τ2)βsin(τ2)), sin(τ)=αα2+β2, and cos(τ)=βα2+β2, such that

    limε0+uεˉuL(R+;L2(Ω))=0.

    Remark 1.1. We employ strict asymptotic analysis to demonstrate that System (1.1) converges to the limiting system (1.7) under high rotational conditions, which is a 2D system incorporating both damping and rotational effects, where the term γˉuh represents the damping, and the term ηˉuh, signifies the rotation. The structure of the damping term γˉuh specifically recovers the results obtained by [13], confirming the accuracy and consistency of our analysis. Meanwhile, the term ηˉuh, indicates that the derived limiting state still exhibits rotational effects, aligning with physical expectations and highlighting the enduring impact of rotation on magnetohydrodynamic fluids.

    Remark 1.2. Addressing the challenges posed by the corners and lateral boundaries of cylindrical domains, we adopt an approach inspired by the work of [3]. We refine the boundary conditions by constructing correction terms in a thin layer near the lateral boundaries, ensuring a more accurate representation of the physical system.

    Remark 1.3. In Section 2, we derived the structure of the approximate solution for the MHD fluid within a cylindrical domain. This structure provides a more accurate representation of the fluid characteristics in such geometries. By capturing the essential features of the fluid's behavior under the influence of magnetic fields and rotation, our solution offers a robust framework for constructing numerical models in geophysics and related fields. This approach facilitates more precise simulations and predictions.

    This paper is organized as follows: Section 2 constructs approximate solutions order by order through asymptotic expansion and introduces correction terms to satisfy incompressibility and boundary conditions. Section 3 investigates the properties of the 2D limiting equation. Section 4 proves the convergence results for rotating magnetohydrodynamics in the limiting state.

    This section constructs a linear approximate solution (uappL,pappL,jappL,φappL) of the following form:

    {uappL=Σi=0εi[ui,int(t,x,y,z)+ui,T(t,x,y,λ)+ui,B(t,x,y,θ)],pappL=Σi=0εi[pi,int(t,x,y,z)+pi,T(t,x,y,λ)+pi,B(t,x,y,θ)],jappL=Σi=0εi[ji,int(t,x,y,z)+ji,T(t,x,y,λ)+ji,B(t,x,y,θ)],φappL=Σi=0εi[φi,int(t,x,y,z)+φi,T(t,x,y,λ)+φi,B(t,x,y,θ)], (2.1)

    where θ=zε, λ=1zε, and the superscripts int,T,B, and c represent the interior, top boundary, bottom boundary terms, and correction terms, respectively. We also put a nature boundary condition as follows:

    limλui,T(t,x,y,λ)=limθui,B(t,x,y,θ)=0. (2.2)

    Furthermore, the approximate solution satisfies the following linear approximate equations:

    {tuappLεΔuappL+αεe3uappL+βεe3jappL+1εpappL=RappL,jappLφappL+e3uappL=0,uappL=jappL=0, (2.3)

    where RappL represents the residual term obtained by substituting the linear approximate solution (uappL,pappL,jappL,φappL) into the original linear system, with the boundary conditions

    uappL|Ω=0,jappL,3|z=0,1=0,jappL,hns|S=0. (2.4)

    Next, we decide the precise forms of (2.1) by analyzing the order of ε. Moreover, we substitute the approximate forms of (2.1) for the top and bottom boundaries into System (2.3) and analyze its εi-order parts (i=2,1,0,).

    In this subsection, we analyze the part of the linear approximation system of order εi and determine the specific form of the linear approximated solution by combining the top and bottom boundary conditions and the incompressibility conditions. We mainly construct the bottom boundary and interior terms in the following section. The construction process of the top boundary layer is similar to that of the bottom boundary.

    Through simple computation

    θp0,B=2θφ0,B=0,

    we obtain the highest order term as ε2. Clearly, getting p0,B and θφ0,B is independent of θ. It is natural to take p0,B=0, implying that the boundary layer's highest order pressure term is vanishing.

    Similarly, it can be obtained that p0,T=0, and that λφ0,T is independent of λ.

    From the ε1-order bottom boundary term, we get

    {2θu0,Bh+Au0,Bhβhφ0,B=0,2θu0,B3θp1,B=0,θφ0,B=θu0,B3=2θφ1,B=0. (2.5)

    First, from (2.5)3, we know that u0,B3 is independent of θ. Combining u0,B3 then satisfies the Dirichlet boundary condition, and the Taylor–Proudman theorem leads to the conclusion that u0,B3=0. Next, due to u0,Bh satisfying the boundary condition, take the limit ε0 for (2.5)1, which gives hφ0,B=0. Combined with θφ0,B=0 from (2.5)3, this gives φ0,B=0. Moreover, j0,B3=0 can be obtained from (2.3)2. On this basis, in combination with (2.5)2, we can see that p1,B is also independent of θ.

    Similarly, we take the ε1-order internal terms from the equations as

    {xp0,int=αu0,int2+βj0,int2,yp0,int=αu0,int1βj0,int1,zp0,int=0. (2.6)

    It is natural to show that p0,int(t,x,y) is independent of z. Combining the incompressible conditions of u0,int and j0,int and (2.6)1,2, we get

    yxp0,intxyp0,int=α(hu0,inth)+β(hj0,inth)=z(αu0,int3+βj0,int3)=0.

    Due to u0,B3=j0,B3=0 and their boundary conditions in (2.4), we obtain

    u0,int3|z=0,1=j0,int3|z=0,1=(αu0,int3+βj0,int3)|z=0,1=0. (2.7)

    According to the Taylor–Proudman theorem, z(αu0,int3+βj0,int3)=0 and (2.7); it follows that u0,int3=j0,int3=0. Hence, zφ0,int=0 and φ0,int(t,x,y) is independent of z. Since j0,int satisfies

    j0,int=φ0,inte3u0,int,

    Eq (2.6) can be changed to

    {xp0,int=αu0,int2+βyφ0,intβu0,int1,yp0,int=αu0,int1βxφ0,intβu0,int2,zp0,int=0.

    Since p0,int and φ0,int are independent of z, it follows from expression above that u0,inth(t,x,y) is also independent of z.

    The following inner product of the system in Eq (2.6)1,2 and u0,inth, combined with the incompressibility condition for u0,inth, gives

    Sj0,int2u0,int1+j0,int1u0,int2=0. (2.8)

    Note that u0,inth and j0,inth satisfy the equations and the boundary condition

    {j0,int=φ0,inte3u0,int,j0,inthns|S=0. (2.9)

    Then, by combining (2.8) and (2.9), it can be deduced that

    S|j0,inth|2+j0,inthhφ0,int=S|j0,inth|2+Sj0,inthnsφ0,int=0.

    Thus we obtain j0,inth=0 from the boundary condition in (2.4) for j0,inth and have

    φ0,int(t,x,y)=Δ1hhu0,inth. (2.10)

    On the basis of this analysis, it can be seen that the internal terms in (2.6) can be reduced to

    {xp0,int=αu0,int2,yp0,int=αu0,int1.

    By the incompressibility condition of u0,inth, p0,int can be expressed as

    p0,int(t,x,y)=αΔ1hhu0,inth. (2.11)

    Furthermore, the boundary terms (2.5) can be rewritten as

    {2θu0,Bh+Au0,Bh=0,u0,Bh|θ=0=u0,inth,limθu0,Bh=0. (2.12)

    Equation (2.12) is a fourth-order ordinary differential system in u0,Bh. Solving this differential equation is straightforward, and we can solve it for

    u0,Bh(t,x,y,θ)=eaθ(cos(bθ)u0,inth+sin(bθ)u0,inth,), (2.13)

    where

    a=(α2+β2)14cos(τ2),b=(α2+β2)14sin(τ2).

    Furthermore, from (2.3)2 and φ0,B=0, we have

    j0,Bh(t,x,y,θ)=u0,Bh,. (2.14)

    Similar to the analysis above, we can also obtain the expressions for the top boundary terms as φ0,T=u0,T3=0 and

    u0,Th(t,x,y,λ)=eaλ(cos(bλ)u0,inth+sin(bλ)u0,inth,). (2.15)

    From the O(1)-order bottom boundary term, we get

    {2θu1,Bh+Au1,Bh=hp1,B+βhφ1,B,θp2,B=2θu1,B3,θu1,B3=hu0,Bh,2θφ2,B=hu0,Bh. (2.16)

    Firstly, from (2.16)3 and the expression of (2.13) for u0,Bh, we have

    θu1,B3=h(eaθ(cos(bθ)u0,inth+sin(bθ)u0,inth,)). (2.17)

    If we integrate Equation (2.17) concerning θ, we get

    u1,B3(t,x,y,θ)=(a2+b2)1eaθ(asin(bθ)+bcos(bθ))hu0,inth(a2+b2)1eaθ(acos(bθ)bsin(bθ))hu0,inth. (2.18)

    From the boundary condition in (2.4), we can deduce that

    u1,int3|z=0=u1,B3|θ=0=(a2+b2)1(bhu0,inth+ahu0,inth). (2.19)

    According to the boundary expression in (2.19), we take u1,int3 to be

    u1,int3(t,x,y,z)=(12z)(a2+b2)1(bhu0,inth+ahu0,inth). (2.20)

    We then combine this with the incompressible condition of u1,int that

    hu1,inth=zu1,int3=2(a2+b2)1(bhu0,inth+ahu0,inth). (2.21)

    In this case, u1,inth can be expressed as

    u1,inth=2(a2+b2)1(au0,inthbu0,inth,)+g1(t,z),

    where the expression for g1(t,z) is determined below.

    Remark 2.1. It is worth noting that hu0,inth in (2.18)(2.21) practically vanishes. Since this term affects the construction of u1,inth and hence the limit equations, we keep it in this form.

    Below, we analyze the forms of u1,inth and g1,int. First of all, we know that the O(1)-order interior part in the approximate system is:

    {tu0,inthAu1,inth+hp1,int+βhφ1,int=0,tu0,int3+zp1,int=0,hu0,inth=0,Δφ1,int=hu1,inth. (2.22)

    Given u0,int3=0 and (2.22)2, it follows that p1,int(t,x,y) is independent of z. At this point, the expression for hφ1,int is not determined, so we can assume that hφ1,int=g2(t,x,y)+g3(t,z). Consequently, Eq (2.22)1 can be decomposed into the parts related to (x,y) and the parts related to z, i.e.,

    tu0,inth2(a2+b2)1A(au0,inthbu0,inth,)+hp1,int+g2(t,x,y)=0,

    and

    g1(t,z)+g3(t,z)=0.

    Furthermore, we can set g1(t,z)=g3(t,z)=0, as this assumption does not affect the subsequent analysis. Therefore, both u1,inth and hφ1,int are independent of z, and u1,inth can be expressed as

    u1,inth(t,x,y)=2(a2+b2)1(au0,inthbu0,inth,). (2.23)

    Next, we analyze φ1,int. Assuming φ1,int=g4(x,y)+g5(z), then with the boundary condition in (2.4), we have

    zφ1,int|z=0,1=zg5(z)|z=0,1=0,

    which gives

    g5(z)=n=0ancos(nz),

    where an is a family of constants. Thus

    Δφ1,int=Δhg4(x,y)n=0n2ancos(nz); (2.24)

    however, by (2.22)3 and because u1,inth is independent of z, it follows that Δφ1,int is independent of z. This contradicts (2.24), and thus an=0, i.e., φ1,int(t,x,y) is independent of z.

    With the above analysis and the expression in (2.23) for u1,inth, (2.22) can be rewritten as

    {tu0,inth+γu0,inth+ηu0,inth,+hp1,int+βhφ1,int=0,hu0,inth=0,Δhφ1,int=2cosτ2(α2+β2)14hu0,inth, (2.25)

    where

    γ=2(α2+β2)14(αsinτ2+βcosτ2),η=2(α2+β2)14(αcosτ2βsinτ2).

    On the basis of the expressions for u1,B3 and u0,Bh, we integrate (2.16)2,4 to get

    p2,B(t,x,y,θ)=eaθsin(bθ)hu0,inth, (2.26)

    and

    θφ2,B=(a2+b2)1eaθ(bsin(bθ)acos(bθ))hu0,inth+g6(t,x,y), (2.27)

    where the form of g6(t,x,y) is determined subsequently.

    On the basis of the facts that the O(1)-order term θφ1,B|θ=0=zφ0,int|z=0=0 in the boundary conditions in (2.4) and that θφ1,B is independent of θ, we can determine that φ1,B(t,x,y) is also independent of θ. Combining the boundary condition (2.2) with the boundary terms p1,B and φ1,B(t,x,y), independent of θ, and taking the limit ε to zero at both ends of (2.16)1, we get

    hp1,B=βhφ1,B. (2.28)

    Thus u1,Bh satisfies the following equations and boundary conditions, and the right-hand side of the system are all known terms:

    {2θu1,Bh+Au1,Bh=0,u1,Bh|θ=0=u1,inth,limεu1,Bh=0.

    Duhamel's principle leads to

    u1,Bh(t,x,y,θ)=eaθ(cos(bθ)u1,inthsin(bθ)u1,inth,). (2.29)

    Remark 2.2. Notably, the coefficient γ of the damping term of the linear limit system remains consistent with the results in [13]. Meanwhile, ηu0,inth, is due to the retention of hu0,inth in (2.18)(2.21), reacting to the continuous effect of rotation on the fluid.

    Similarly, on the basis of the analysis above, we can get

    u1,T(t,x,y,λ)=(eaλ(cos(bλ)u1,inthsin(bλ)u1,inth,)(a2+b2)1eaλ(asin(bλ)+bcos(bλ))hu0,inth), (2.30)
    [4pt]p2,T(t,x,y,λ)=eaλsin(bλ)hu0,inth, (2.31)
    [4pt]λφ2,T(t,x,y,λ)=(a2+b2)1eaλ(bsin(bλ)acos(bλ))hu0,inth+g7(x,y), (2.32)

    and hp1,T=βhφ1,T and g7(x,y) are determined subsequently.

    The boundary O(ε)-order term in the incompressibility condition is θu2,B3=hu1,Bh. It can then be found in the case where u1,Bh is known that

    u2,B3(t,x,y,θ)=+θhu1,Bh(t,x,y,s)ds. (2.33)

    Similarly, according to the incompressibility condition, the upper boundary term u2,T3 is

    u2,T3(t,x,y,λ)=+λhu1,Th(t,x,y,s)ds. (2.34)

    Since the internal higher-order terms do not introduce singularities, they do not affect the subsequent analysis. Therefore, we take u2,int=0, then u2,Bh=u2,Th=0. We will correct the boundary conditions for u2,B3 and u2,T3 subsequently.

    On the basis of the facts that the O(ε)-order terms θφ2,B|θ=0=zφ1,int|z=0 and λφ2,T|λ=0=zφ1,int|z=1 in the boundary conditions in (2.4) and that φ1,int is independent of z, we can get λφ2,T|λ=0=θφ2,B|θ=0=0. Thus there is

    θφ2,B=(a2+b2)1eaθ(bsin(bθ)acos(bθ))hu0,inth+(a2+b2)1ahu0,inth,

    and

    λφ2,T=(a2+b2)1eaλ(bsin(bλ)acos(bλ))hu0,inth(a2+b2)1ahu0,inth.

    In this subsection, we construct the top and bottom boundaries as well as the internal terms (see Figure 1), with the approximate solution (u1,appL,p1,appL,φ1,appL,j1,appL) given by

    {u1,appL=(u0,inth+u0,Bh+u0,Th0)+ε(u1,inth+u1,Bh+u1,Thu1,int3+u1,B3+u1,T3)+ε2(0u2,B3+u2,T3),p1,appL=p0,int+ε(p1,int+p1,B+p1,T)+ε2(p2,B+p2,T),φ1,appL=φ0,int+ε(φ1,int+φ1,B+φ1,T)+ε2(φ2,B+φ2,T),j1,appL=φ1,appLe3u1,appL, (2.35)

    where the approximate solution (u1,appL,j1,appL) satisfies

    u1,appL=0, (2.36)
    j1,appL=ε(Δhφ1,B+hu1,Bh+Δhφ1,T+hu1,Th)+ε2(Δhφ2,B+Δhφ2,T), (2.37)

    and

    u1,appLz=0,1=(u0,Bhθ=1ε0)+ε(u1,Bhθ=1ε(1)1zz=0,1u1,B3θ=1ε)+ε2(0(1)zz=0,1u2,B3θ=1ε), (2.38)
    [5pt]j1,appL,3z=0,1=0, (2.39)
    j1,appL,hnsS0,u1,appLS0. (2.40)

    The next goal is to correct these incompressibility conditions and boundary conditions one by one.

    Figure 1.  Schematic of the top and bottom boundary layers in a cylindrical domain Ω.

    This subsection aims to correct the top and bottom boundary conditions in (2.40), and we establish the correction term uc, namely

    uc=u0,c+εu1,c+ε2u2,c.

    Note that we can now construct the correction term uc in such a way as to ensure that uc satisfies the incompressibility condition. We therefore make ui,c(i=0,1,2) satisfy

    u0,c=(cos(2πz)u0,Bhθ=1ε,sin(2πz)2πhu0,Bhθ=1ε), (2.41)
    u1,c=(cos(2πz)u1,Bhθ=1ε,sin(2πz)2πhu1,Bhθ=1ε)+(πsin(πz)+1εu0,Bhdθ,cos(πz)u1,B3θ=1ε), (2.42)
    u2,c=(πsin(πz)+1εu1,Bhdθ,cos(πz)u2,B3θ=1ε). (2.43)

    It is clear from the expression (2.41)–(2.43) above that

    ucW1,(0,T;H1(Ω))=O(ε12). (2.44)

    At this point, an approximate solution (u2,appL,j2,appL) is obtained, i.e.

    {u2,appL=u1,appL+uc,j2,appL=φ1,appLe3(u1,appL+uc),

    and (u2,appL,j2,appL) satisfies

    u2,appL=0, (2.45)
    j2,appL=ε(Δhφ1,B+hu1,Bh+Δhφ1,T+hu1,Th)+ε2(Δhφ2,B+Δhφ2,T)+hu0,ch+εhu1,ch+ε2hu2,ch, (2.46)

    and

    j2,appL,3z=0,1=0,u2,appLz=0,1=0, (2.47)
    j2,appL,hnsS0,u2,appLS0. (2.48)

    In the analysis above, we corrected the top and bottom boundary conditions for the approximate solution of the velocity field. Below, we correct the lateral boundary conditions.

    The purpose of this subsection is to correct the lateral boundary conditions in (2.48) for u2,appL. The horizontal component of the approximate solution u2,app consists of u0,inth. It is therefore natural to impose a Dirichlet boundary condition on the velocity field u0,inth(t,x,y) in the bounded domain S in R2:

    u0,inthS=0. (2.49)

    Thus, we have

    u2,apphS=0. (2.50)

    Below, we correct the vertical component of the approximate solution u2,appL. Referring to [3], we introduce d:SR as a distance to the side S, and construct the lateral correction terms in the region of size εσ(12<σ<1) near the lateral boundary (see Figure 2). The value of σ here will be determined later.

    Figure 2.  Schematic of the boundary layers in a cylindrical domain Ω.

    First, using w0,c=(w0,ch,w0,c3) to correct the ε0-order term, we write

    w0,c3=ed(x,y)εσu0,c3=ed(x,y)εσ(sin(2πz)2πhu0,Bh|θ=1ε), (2.51)

    and w0,ch=0. w0,c3 vanishes at the top and bottom boundaries. Furthermore, the presence of ed(x,y)εσ in w0,c3 causes w0,c3 to vanish when ε is sufficiently small, as well as away from the region where the size of the side edges is εσ. At the same time, w0,c does not satisfy the incompressibility condition and has

    w0,c=ed(x,y)εσcos(2πz)hu0,Bhθ=1ε.

    Concerning w0,c, we have the following estimates:

    {w0,cW1,(R+,L2(Ω))=O(εσ+12),w0,cW1,(R+,L2(Ω))=O(εσ+12),w0,cW1,(R+,H1(Ω))=O(ε12).

    Secondly, using w1,c=(w1,ch,w1,c3) to correct for the ε1-order side boundary term, we write

    w1,c3=ed(x,y)εσε(u1,int3+u1,B3+u1,T3+u1,c3), (2.52)

    and

    w1,ch=εσ(1+ed(x,y)εσ)d(x,y)|d(x,y)|2(θu1,B3λu1,T3)cos(2πz)εσ(1+ed(x,y)εσ)d(x,y)|d(x,y)|2θu1,B3θ=1ε. (2.53)

    Clearly, from the definitions of u1,int3, u1,B3, u1,T3, and u1,c3, as well as the analysis above, it follows that w1,c satisfies all boundary conditions. Nevertheless, it does not satisfy the incompressibility condition:

    w1,c=εσ(1+ed(x,y)εσ)h(d(x,y)|d(x,y)|2(θu1,B3λu1,T3))cos(2πz)ed(x,y)εσθu1,B3θ=1εed(x,y)εσε(zu1,int3+zu1,c3)cos(2πz)εσ(1+ed(x,y)εσ)h(d(x,y)|d(x,y)|2θu1,B3|θ=1ε). (2.54)

    We have the following estimates for w1,c:

    {w1,cW1,(R+,L2(Ω))=O(εσ),w1,cW1,(R+,L2(Ω))=O(εσ),w1,cW1,(R+,L2(Ω))=O(εσ12).

    Finally, utilizing w2,c=(w2,ch,w2,c3) to correct the ε2-order side boundary term, we write

    w2,c3=ed(x,y)εσε2(u2,int3+u2,B3+u2,T3+u2,c3).

    Since the higher-order correction term does not affect the subsequent analysis, we take w2,ch=0. Then w2,c satisfies the boundary conditions, and

    w2,c=ed(x,y)εσε(θu2,B3λu2,T3)ed(x,y)εσε2(zu2,int3+zu2,c3),

    as well as

    {w2,cW1,(R+,L2(Ω))=O(εσ+12),w2,cW1,(R+,L2(Ω))=O(εσ+12),w2,cW1,(R+,H1(Ω))=O(εσ+12).

    Next, we take σ=34 and

    wc=w0,c+w1,c+w2,c.

    Moreover, wc satisfies

    {wcW1,(R+,L2(Ω))=O(ε34),wcW1,(R+,L2(Ω))=O(ε34),wcW1,(R+,H1(Ω))=O(ε14). (2.55)

    It is worth noting that while constructing wc, we need it to satisfy the incompressibility condition. According to [12,21], uwW1,(R+,H1(Ω)) exists such that the following equations hold:

    {uw=wc,uw|Ω=0,

    and

    uwW1,(0,T;H1(Ω))wcW1,(0,T;L2(Ω))=O(ε34). (2.56)

    In the analysis above, we corrected the boundary and incompressibility conditions for the approximate solution of the velocity field. Moreover, we denote this new approximate solution u3,app as

    u3,appL=u1,appL+uc+wc+uw.

    Moreover, let

    j3,appL=φ1,appLe3(u1,appL+uc+wc+uw).

    Due to we construct the correction term uc+wc+uw, relative to which we also correct the magnetic potential.

    In this subsection, we correct the incompressibility condition of j3,appL by constructing a correction term φc for the magnetic potential. By the order of ε in (2.46), we write φc as

    φc=φ0,c+φ1,c+φ2,c.

    Next, according to [12,21], φ0,c,φ1,c,φ2,cL(R+,H1(Ω)) exists such that the following equations hold:

    {(φ0,c)=hu0,ch,φ0,cΩ=0, (2.57)
    {(φ1,c)=h(εu1,ch+εu1,Bh+εu1,Th+wch+uwh)εΔh(φ1,B+φ1,T),φ1,cΩ=0, (2.58)
    {(φ2,c)=ε2hu2,chε2Δh(φ2,B+φ2,T),φ2,cΩ=0. (2.59)

    Thus, we obtain a new approximate solution for the magnetic potential j4,appL, i.e.,

    j4,appL=(φ1,appL+φc)e3(u1,appL+uc+wc+uw), (2.60)

    which satisfies

    j4,appL=0,j4,appL,3z=0,1=0. (2.61)

    In the following, correcting only the lateral boundary conditions of j4,app is necessary. Through the analysis and construction process above, we can get

    j4,appL,HnsS=(εhφ1,int+εhφ1,B+εhφ1,T+ε2hφ2,B+ε2hφ2,T)nsS.

    First, we take

    hφ1,intnsS=0. (2.62)

    Secondly, according to [12,21], hφwL(R+,H1(Ω)) exists such that the following equations hold:

    {h(hφw)=0,hφwΩ=(εhφ1,B+εhφ1,T+ε2hφ2,B+ε2hφ2,T)Ω.

    In summary, we have completed the construction of the approximate solution and satisfied all its incompressibility and boundary conditions.

    The previous subsections considered the approximate system under the linear system in (2.3). On the basis of the analysis above, we construct the approximate solution to the following system:

    {tuappεΔuapp+(uapp)uapp+αεe3uapp+βεe3japp+1εpapp=Rapp,jappφapp+e3uapp=0,uapp=japp=0, (2.63)

    where Rapp represents the residual term obtained by substituting the corrected approximate solution into the original system, with the boundary conditions

    uapp|Ω=0,japp3|z=0,1=0,japphns|S=0. (2.64)

    First, we consider the principal part of the approximate solution uapp and let it be ˉu(t,x,y). By analyzing the linear part above and combining (2.25), (2.49), and (2.62), it is natural to set ˉu(t,x,y)=(ˉuh(t,x,y),0) as

    {tˉuh+(ˉuhh)ˉuh+γˉuh+ηˉuh,+hˉp+βhˉφ=0,hˉuh=0,Δhˉφ=2cosτ2(α2+β2)14hˉuh, (2.65)

    with the boundary conditions

    ˉuhS=0,hˉφnsS=0. (2.66)

    The remaining terms all consist of the central part ˉuh(t,x,y). It may be helpful to use the original notation so that the approximate solution (uapp,papp,φapp,japp) is

    {uapp=ˉu+u0,B+u0,T+ε(u1,int+u1,B+u1,T)+ε2(u2,B+u2,T)+uc+wc+uw,papp=p0,int+ε(ˉp+p1,B+p1,T)+ε2(p2,B+p2,T),φapp=φ0,int+ε(ˉφ+φ1,B+φ1,T)+ε2(φ2,B+φ2,T)+φc+φw,japp=φappe3uapp, (2.67)

    where u0,inth in the original forms is substituted for ˉuh in all but the main part (ˉu,ˉp,ˉφ).

    According to the construction, the following asymptotic behavior holds.

    Proposition 2.1. For the approximate solution uapp given above, if ˉuhL2(R2), it satisfies

    limε0+uappˉuL2(Ω)=0.

    Proof. With the expression (2.67)1 for uapp, it can be shown that

    uappˉuL2(Ω)2i=0εi(ui,BL2(Ω)+ui,TL2(Ω))+εu1,intL2(Ω)+ucL2(Ω)+wcL2(Ω)+uwL2(Ω).

    The presence of the exponential factors eaθand eaλ in the boundary layer terms results in the subsequent estimates being small. As an illustration, consider the example of the bottom boundary term u0,BL2(Ω). From (2.13), we have

    u0,B2L2(Ω)=10S|eaθ(cos(bθ)ˉuh+sin(bθ)ˉuh,)|2dxdydz10S|eazε(ˉuh+ˉuh,)|2dxdydz=aε0Sεa|eazε(ˉuh+ˉuh,)|2dxdydazεεˉuh2L2(S). (2.68)

    Similarly, to estimate other top and bottom boundary terms, by combining (2.68) with the expression (2.23) for the interior term u1,int, we get

    2i=0εi(ui,BL2(Ω)+ui,TL2(Ω))+εu1,intL2(Ω)ε12ˉuhL2(S). (2.69)

    Recalling (2.44), (2.55), and (2.56), one has

    ucL2(Ω)+wcL2(Ω)+uwL2(Ω)ε12ˉuhL2(S). (2.70)

    Combining (2.68)–(2.70) gives

    limε0+uappˉuL2(Ω)=0.

    This section investigates the following properties of 2D limit system:

    {tˉuh+(ˉuhh)ˉuh+γˉuh+ηˉuh,+hˉp+βhˉφ=0,hˉuh=0,Δhˉφ=2cosτ2(α2+β2)14hˉuh, (3.1)

    with the boundary conditions

    ˉuhS=0,hˉφnsS=0. (3.2)

    By applying h to (3.1), and writing ˉω=hˉuh, we can obtain the vorticity system:

    tˉω+(ˉuhh)ˉω+γˉω+βΔhˉφ=0, (3.3)

    where

    Δhˉφ=2cosτ2(α2+β2)14ˉω.

    Therefore, combined with the definition of γ, (3.3) can be rewritten as

    tˉω+(ˉuhh)ˉω+2αsinτ2(α2+β2)14ˉω=0. (3.4)

    As the flow is divergence-free, with hˉuh=0, we have

    ˉω=hˉuh,ˉuh=h(Δh)1ˉω. (3.5)

    Proposition 3.1. Let ˉu0,h(x,y)H1(S) be a divergence-free vector field, ˉω0=hˉu0,h be the initial vorticity, and (ˉu,ˉp,ˉφ) be a pair of solution to the systems in (3.1) and (3.2) with the initial data ˉu0=(ˉu0,h,0). Then the following estimations are valid:

    ˉuh2L2(S),ˉω2L2(S),hˉuh2L2(S)eνt, (3.6)

    where ν=2αsinτ2(α2+β2)14.

    Proof. Given the divergence-free condition, we derive the L2 estimate for ˉuh as follows:

    12ddtˉuh2L2(S)+γˉuh2L2(S)+βhˉφ,ˉuh=0.

    From (2.67)4, it follows that ˉφ=j1,int+e3u1,int, and from ˉu=(ˉuh,0), we have

    hˉφ,ˉuh=e3j1,int,ˉu+e3(e3u1,int),ˉu=Se3j1,intˉuSu1,intˉu=Sj1,int(e3ˉu)+S(2sinτ2(α2+β2)14ˉuh,2cosτ2(α2+β2)14ˉuh)ˉuh=Sj1,intφ0,intS2cosτ2(α2+β2)14|ˉuh|2=2cosτ2(α2+β2)14ˉuh2L2(S).

    A simple derivation gives

    12ddtˉuh2L2(S)+2αsinτ2(α2+β2)14ˉuh2L2(S)=0.

    We write ν=2αsinτ2(α2+β2)14, and thus

    ˉuh2L2(S)eνtˉu0,h2L2(S).

    From (3.5), it is clear that

    12ddtˉω2L2(S)+νˉω2L2(S)=0,

    and thus

    ˉω2L2(S)eνtˉω02L2(S).

    as well as

    hˉuh2L2(S)ˉω2L2(S)eνtˉω02L2(S).

    Remark 3.1. Furthermore, combining this with (3.6) yields an estimate of (3.6) for ˉuL(S).

    In our proof, we require the bound for hˉuhL(S). This necessity arises from the Calderón-Zygmund theory of singular integral operators, which asserts that the mapping ˉωhˉuh is continuous within the space Ls(S) for 1<s<. However, the case when s= presents additional complexities. To address this, we will establish the desired bound by employing the Littlewood-Paley decomposition in the subsequent steps.

    First, let C={ξR2|34|ξ|43}. The radial functions ψ1 and ψ take values in [0,1] and have support in B(0,43) and C, respectively, such that

    ξR2,ψ1(ξ)+j0ψ(2jξ)=1.

    We then take ψj(ξ)=ψ(2jξ). Obviously, ψj(j>1) is supported in 2j1<|ξ|<2j+2. We write

    fj(x)=F1[ψj(ξ)F(f)],jZ, (3.7)

    where F and F1 are the Fourier and inverse Fourier transforms, respectively. Recalling (3.1), we see that any function fL1(S) holds:

    f=j1fj(x). (3.8)

    Proposition 3.2. Let ˉu0,h(x,y)Ha+1(S)(a>1) be a divergence-free vector field, ˉω0=hˉu0,h be the initial vorticity, and (ˉu,ˉp,ˉφ) be a pair of solutions to the system in (3.1) with the initial data ˉu0=(ˉu0,h,0). Then there holds

    hˉuhL(S)eνt. (3.9)

    Proof. From (3.8), one has

    ˉuh=j1F1(ψjF(ˉuh))=j1ˉuh,j,

    and

    hˉuhL(S)j1hˉuh,jL(S)hˉuh,1L(S)+j>1hˉuh,jL(S). (3.10)

    The first term on the right-hand side can easily be bounded using the Bernstein inequality (for a more specific elaboration of the inequality, see [5])

    hˉuh,1L(S)ˉuhL(S). (3.11)

    Combining this with (3.5), we have

    j>1hˉuh,jL(S)=j>1hh(Δh)1ˉωjL(S)j>1ˉωjL(S). (3.12)

    Thus, from (3.10)–(3.12) and the results of Proposition 3.1, we have

    hˉuhL(S)ˉuhL(S)+j>1ˉωjL(S)ˉu0,hL(S)eνt+j>1ˉωjL(S). (3.13)

    Now, we turn to the term j>1ˉωjL(S). Applying δj to (3.4), we have

    {tˉωj+(ˉuhh)ˉωj+νˉωj=[δj,(ˉuhh)]ˉω,ˉωjt=0=ˉωj0, (3.14)

    where [,] stands for the commutator. Let

    ˉω=j1ˉωj,

    with N to be determined later. One then has

    j>1ˉωjL(S)=1<j<NˉωjL(S)+jNˉωjL(S). (3.15)

    From (3.6)–(3.8) and (3.14), for j<N, we get

    ˉωjL(S)eνtˉωj0L(S)+t0eν(ts)[δj,(ˉuhh)]ˉωL(S)dseνtˉωj0L(S)+t0eν(ts)ψjF((ˉuhh)ˉω)L1(S)dseνtˉωj0L(S)+eνt(ˉu0,hL2(S)+ˉω0H1(S)), (3.16)

    where we used the results of Proposition 3.2. Thus we get

    jNˉωjL(S)eνtNjNˉωj0L(S)+eνtN(ˉu0,hL2(S)+ˉω0H1(S)). (3.17)

    Furthermore, to deal with the case j>N, similar to (3.16), we get

    j>NˉωjL(S)eνt2(a1)N2j>Na12hˉωj0L(S)+eνt2(a1)N(ˉu0,hL2(S)+ˉω0H1(S)). (3.18)

    If we combine (3.17) and (3.18), taking N=[log22a1(1+1+j1a12hˉωj0L(S)1+j1ˉωj0L(S))], the following holds:

    j1ˉωjL(S)eνt(ln(1+j1a12hˉωj0L(S))+ˉu0,hL2(S)+ˉω0H1(S))(j1ˉωj0L(S)+ˉu0,hL2(S)+ˉω0H1(S))eνt(ln(1+ˉω0Ha(S))+ˉω0Ha(S))ˉω0Ha(S). (3.19)

    Therefore, the result is derived from from (3.13) and (3.19).

    In this section, we aim to demonstrate that as ε0+, the weak solution uε of the system given by (1.1)–(1.4) converges in the L2(Ω) norm to ˉu. Specifically, we show that uεˉuL2(Ω) tends to zero. Given Proposition 2.1, it suffices to establish that uεuappL2(Ω) also approaches zero.

    Note that uε and uapp satisfy the following systems:

    {tuε+(uε)uεεΔuε+αεe3uε+βεe3jε+1εpε=0,jεφε+e3uε=0,uε=jε=0,uε(t,x)|t=0=uε0(x),uε(t,x)|Ω=0,jε3(t,x)|z=0,1=0,jεhns|S=0,

    and

    {tuapp+(uapp)uappεΔuapp+αεe3uapp+βεe3japp+1εpapp=Rapp,jappφapp+e3uapp=0,uapp=japp=0,uapp|t=0=(ˉu+u0,B+u0,T+ε(u1,int+u1,B+u1,T)+ε2(u2,B+u2,T)+uc+wc+uw)|t=0,uapp(t,x)|Ω=0,japp3(t,x)|z=0,1=0,japphns|S=0,

    where

    Rapp=t(u0,B+u0,T+ε(u1,int+u1,B+u1,T)+ε2(u2,B+u2,T)+uc+wc+uw)εΔ(ˉu+εu1,int+ε2(u2,B+u2,T)+uc+wc+uw)εΔh(u0,B+u0,T+ε(u1,B+u1,T))+(uapp)(εu1,int+uc+wc+uw)+αεe3(uc+wc+uw)+ε(hp2,B+hp2,T0)+βε(hφ2,B+hφ2,T0)+2i=0εi((u0,Bh+u0,Th+ε(u1,inth+u1,Bh+u1,Th))h(ui,B+ui,T)+(ˉu+uc+wc+uw)(ui,B+ui,T)+(ε(u1,int3+u1,B3+u1,T3)+ε2(u2,B3+u2,T3))z(ui,B+ui,T)). (4.1)

    Below, we compute the error estimate between uε and uapp. Let v=uεuapp, jv=jεjapp, φv=φεφapp, and pv=pεpapp. We then have

    {tv+(uε)vεΔv+αεe3v+βεe3jv+1εpv+(v)uapp+Rapp=0,jvφv+e3v=0,v=jv=0,v(t,x)|Ω=0,jv3(t,x)|z=0,1=0,jvhns|S=0. (4.2)

    Estimating v2L2 using Eq (4.2) naturally yields

    12ddtv2L2+εv2L2+αεe3v+1εpv+(uε)v,v+βεe3jv,v=(v)uapp,vRapp,v. (4.3)

    Using the incompressibility condition for v and the structure of e3v, the third term on the left-hand side of (4.3) is

    αεe3v+1εpv+(uε)v,v=0.

    By definition and the boundary conditions in (4.2)2(4.2)4 of jv, the fourth term reduces to

    βεe3jv,v=βεjv,e3×v=βεjv,jvφv=βεΩ|jv|2dxβεΩjvφvdx=βεjv2L20.

    Next, we estimate the right-hand side of (4.3). The first term can be expanded to

    (v)uapp,v=v(ˉu+u0,B+u0,T+ε(u1,int+u1,B+u1,T)+ε2(u2,B+u2,T)+uc+wc+uw),v.

    First, by Hölder's inequality and Proposition 3.2, one obtains

    |vˉu,v|=|vhhˉuh,vh|hˉuhL(S)vh2L2(Ω)hˉu0,hL(S)v2L2(Ω)eνt. (4.4)

    Second, for the ε0-order boundary term, in the case of u0,B, we utilize the integration by parts, which is computed as

    |v(u0,B+u0,T),v|=|vv,u0,B+u0,T|S×[0,12]|v||v||u0,B+u0,T|dx+S×[12,1]|v||v||u0,B+u0,T|dx.

    Due to the boundary conditions on v, we can deduce that

    |v|=|z0zvdz|d(z)12zvL2(0,1),

    where d(z) is the distance to the bottom boundary. Then

    S×[0,12]|v||v||u0,B+u0,T|dxΩzvL2(0,1)|v|d(z)12|u0,B+u0,T|dxv2L2(Ω)d(z)12|u0,B+u0,T|L2(0,1;L(S)),

    where

    d(z)12|u0,B+u0,T|L2(0,1;L(S))εaˉuhL(S)[0,aε]azεeazεdazε+εaˉuhL(S)[0,aε]a(1z)εea(1z)εda(1z)εεˉuhL(S).

    In summary, this gives

    |v(u0,B+u0,T),v|εˉu0,hL(S)v2L2(Ω)eνt. (4.5)

    Next, for higher-order terms, it is easy to obtain

    |v(ε(u1,int+u1,B+u1,T)+ε2(u2,B+u2,T)),v|=ε|vv,(u1,int+u1,B+u1,T)+ε(u2,B+u2,T)|εvL2(Ω)vL2(Ω)ˉu0,hL(S)eνtε4v2L2(Ω)+4εˉu0,hL(S)v2L2(Ω)eνt. (4.6)

    Finally, for the correction term, according to (2.44), (2.55), and (2.56), it follows that

    |v(uc+wc+uw),v|ε14v2L2(Ω)hˉu0,hL(S)eνt. (4.7)

    Combining (4.4)–(4.7), we get

    (v)uapp,vε(14+ˉu0,hL(S)eνt)v2L2(Ω)+(hˉu0,hL(S)+4εˉu0,hL(S)+ε14hˉu0,hL(S))v2L2(Ω)eνt. (4.8)

    The second term on the right-hand side of (4.3), from the expression for Rapp, can be easily obtained as

    |v,Rapp|=RappL2(Ω)vL2(Ω)εˉu0,hL2(S)vL2(Ω)eνtεˉu0,h2L2(S)eνt+εv2L2(Ω)eνt. (4.9)

    Thus, on the basis of (4.3), (4.8) and (4.9), we have

    12ddtv2L2(Ω)+3ε4v2L2(Ω)εˉu0,hL(S)eνtv2L2(Ω)+εˉu0,h2L2(S)eνt+(hˉu0,hL(S)+4εˉu0,hL(S)+ε14hˉu0,hL(S)+ε)v2L2(Ω)eνt. (4.10)

    Due to the initial conditions in (1.5) and (1.6), and by integrating the inequality (4.10) with respect to the variable t, we can complete the proof of the theorem.

    This paper employs a multiscale analysis approach to investigate the impact of the Ekman-Hartmann boundary layer within rotating MHD flows confined to cylindrical domains and develops the corresponding approximate solutions. These solutions are valuable for numerical computations in geophysics and metal engineering industries, aiding in more accurate simulations of fluid dynamic behaviors. Although our model has achieved innovation in handling constant magnetic fields and rotation axes, it has limitations in modeling variations in the magnetic fields and rotation axes over time and space, and in adapting to more complex geometrical shapes. Future research will explore the effects of complex variations in the magnetic fields and rotation axes on the boundary layer. It may extend the model to accommodate various geometries, including spherical and nonplanar, to solve more practical problems.

    Yifei Jia: Writing-original and draft; Guanglei Zhang and Kexue Chen: Writing-review and editing. All authors have read and agreed to the published version of the manuscript.

    The authors declare they have not used artificial intelligence (AI) tools in the creation of this article.

    This work was partially supported by the National Key R&D Program of China (2020YFA072500) and the Innovation Project of Excellent Doctoral Students of Xinjiang University (XJU2024BS038).

    All authors declare no conflict of interest in this article.



    [1] L. Ali, P. Kumar, H. Poonia, S. Areekara, R. Apsari, The significant role of Darcy-Forchheimer and thermal radiation on Casson fluid flow subject to stretching surface: a case study of dusty fluid, Mod. Phys. Lett. B, 38 (2024), 2350215. http://doi.org/10.1142/S0217984923502159 doi: 10.1142/S0217984923502159
    [2] L. Ali, R. Apsari, A. Abbas, P. Tak, Entropy generation on the dynamics of volume fraction of nano-particles and coriolis force impacts on mixed convective nanofluid flow with significant magnetic effect, Numer. Heat Tr. A Appl., 2024 (2024), 1–16. https://doi.org/10.1080/10407782.2024.2360652 doi: 10.1080/10407782.2024.2360652
    [3] D. Bresch, B. Desjardins, D. Gérard-Varet, Rotating fluids in a cylinder, Discrete Cont. Dyn. Syst., 11 (2004), 47–82. http://doi.org/10.3934/dcds.2004.11.47 doi: 10.3934/dcds.2004.11.47
    [4] S. Chandrasekhar, Hydrodynamic and hydromagnetic stability, International Series of Monographs on Physics, 1961.
    [5] J. Y. Chemin, B. Desjardins, I. Gallagher, E. Grenier, Mathematical geophysics. An introduction to rotating fluids and the Navier-Stokes equations, Oxford University Press, 2006.
    [6] B. Desjardins, E. Dormy, E. Grenier, Stability of mixed Ekman-Hartmann boundary layers, Nonlinearity, 12 (1999), 181–199. http://doi.org/10.1088/0951-7715/12/2/001 doi: 10.1088/0951-7715/12/2/001
    [7] E. Dormy, Modélisation numérique de la dynamo terrestre, Paris: Institut de Physique du Globe de Paris, 1997.
    [8] E. Dormy, P. Cardin, D. Jault, MHD flow in a slightly differentially rotating spherical shell, with conducting inner core, in a dipolar magnetic field, Earth Planet. Sc. Lett., 160 (1998), 15–30. http://doi.org/10.1016/S0012-821X(98)00078-8 doi: 10.1016/S0012-821X(98)00078-8
    [9] G. Duvaut, J. L. Lions, Inéquations en thermoélasticité et magnétohydrodynamique, Arch. Rational Mech. Anal., 46 (1972), 241–279. https://doi.org/10.1007/BF00250512 doi: 10.1007/BF00250512
    [10] M. A. Fahmy, M. O. Alsulami, A. E. Abouelregal, Sensitivity analysis and design optimization of 3T rotating thermoelastic structures using IGBEM, AIMS Math., 7 (2022), 19902–19921. http://doi.org/10.3934/math.20221090 doi: 10.3934/math.20221090
    [11] M. A. Fahmy, A nonlinear fractional BEM model for magneto-thermo-visco-elastic ultrasound waves in temperature-dependent FGA rotating granular plates, Fractal Fract., 7 (2023), 214. http://doi.org/10.3390/fractalfract7030214 doi: 10.3390/fractalfract7030214
    [12] G. P. Galdi, An introduction to the mathematical theory of the Navier-Stokes equations, Springer Tracts in Natural Philosophy, 1994.
    [13] D. Gérard-Varet, A geometric optics type approach to fluid boundary layers, Commun. Part. Diff. Eq., 28 (2003), 1605–1626. https://doi.org/10.1081/PDE-120024524 doi: 10.1081/PDE-120024524
    [14] H. P. Greenspan, The theory of rotating fluids, Cambridge University Press, 1969.
    [15] J. Pedlosky, Geophysical fluid dynamics, New York: Springer-Verlag, 1979.
    [16] F. Rousset, Large mixed Ekman-Hartmann boundary layers in magnetohydrodynamics, Nonlinearity, 17 (2004), 503–518. http://doi.org/10.1088/0951-7715/17/2/008 doi: 10.1088/0951-7715/17/2/008
    [17] F. Rousset, Stability of large amplitude Ekman-Hartmann boundary layers in MHD: the case of ill-prepared data, Commun. Math. Phys., 259 (2005), 223–256. https://doi.org/10.1007/s00220-005-1371-0 doi: 10.1007/s00220-005-1371-0
    [18] F. Rousset, Asymptotic behavior of geophysical fluids in highly rotating balls, Z. Angew. Math. Phys., 58 (2007), 53–67. https://doi.org/10.1007/s00033-006-5021-y doi: 10.1007/s00033-006-5021-y
    [19] M. Sermange, R. Temam, Some mathematical questions related to the MHD equations, Commun. Pur. Appl. Math., 36 (1983), 635–664. https://doi.org/10.1002/cpa.3160360506 doi: 10.1002/cpa.3160360506
    [20] M. E. Schonbek, T. P. Schonbek, E. Süli, Large-time behaviour of solutions to the magnetohydrodynamics equations, Math. Ann., 304 (1996), 717–757. https://doi.org/10.1007/BF01446316 doi: 10.1007/BF01446316
    [21] S. Ukai, The incompressible limit and the initial layer of the compressible Euler equation, J. Math. Kyoto Univ., 26 (1986), 323–331. http://doi.org/10.1215/kjm/1250520925 doi: 10.1215/kjm/1250520925
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(348) PDF downloads(46) Cited by(0)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog