Processing math: 51%
Research article

The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces

  • Received: 11 November 2020 Accepted: 17 February 2021 Published: 24 March 2021
  • MSC : primary 35Qxx; secondary 76D05, 76D07, 35M32, 76N10

  • This study typically emphasizes analyzing the geometrical singularities of weak solutions of the mixed boundary value problem for the stationary Stokes and Navier-Stokes system in two-dimensional non-smooth domains with corner points and points at which the type of boundary conditions change. The existence of these points on the boundary generally generates local singularities in the solution. We will see the impact of the geometrical singularities of the boundary or the mixed boundary conditions on the qualitative properties of the solution including its regularity. The solvability of the underlying boundary value problem is analyzed in weighted Sobolev spaces and the regularity theorems are formulated in the context of these spaces. To compute the singular terms for various boundary conditions, the generalized form of the boundary eigenvalue problem for the stationary Stokes system is derived. The emerging eigenvalues and eigenfunctions produce singular terms, which permits us to evaluate the optimal regularity of the corresponding weak solution of the Stokes system. Additionally, the obtained results for the Stokes system are further extended for the non-linear Navier-Stokes system.

    Citation: Yasir Nadeem Anjam. The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces[J]. AIMS Mathematics, 2021, 6(6): 5647-5674. doi: 10.3934/math.2021334

    Related Papers:

    [1] Junling Sun, Xuefeng Han . Existence of Sobolev regular solutions for the incompressible flow of liquid crystals in three dimensions. AIMS Mathematics, 2022, 7(9): 15759-15794. doi: 10.3934/math.2022863
    [2] Yasir Nadeem Anjam . Singularities and regularity of stationary Stokes and Navier-Stokes equations on polygonal domains and their treatments. AIMS Mathematics, 2020, 5(1): 440-466. doi: 10.3934/math.2020030
    [3] Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363
    [4] Benedetta Ferrario, Christian Olivera . Lp-solutions of the Navier-Stokes equation with fractional Brownian noise. AIMS Mathematics, 2018, 3(4): 539-553. doi: 10.3934/Math.2018.4.539
    [5] Shaoliang Yuan, Lin Cheng, Liangyong Lin . Existence and uniqueness of solutions for the two-dimensional Euler and Navier-Stokes equations with initial data in H1. AIMS Mathematics, 2025, 10(4): 9310-9321. doi: 10.3934/math.2025428
    [6] Qingkun Xiao, Jianzhu Sun, Tong Tang . Uniform regularity of the isentropic Navier-Stokes-Maxwell system. AIMS Mathematics, 2022, 7(4): 6694-6701. doi: 10.3934/math.2022373
    [7] Cuiman Jia, Feng Tian . Regularity of weak solution of the compressible Navier-Stokes equations with self-consistent Poisson equation by Moser iteration. AIMS Mathematics, 2023, 8(10): 22944-22962. doi: 10.3934/math.20231167
    [8] Jae-Myoung Kim . Blow-up criteria for the full compressible Navier-Stokes equations involving temperature in Vishik Spaces. AIMS Mathematics, 2022, 7(8): 15693-15703. doi: 10.3934/math.2022859
    [9] Linlin Tan, Meiying Cui, Bianru Cheng . An approach to the global well-posedness of a coupled 3-dimensional Navier-Stokes-Darcy model with Beavers-Joseph-Saffman-Jones interface boundary condition. AIMS Mathematics, 2024, 9(3): 6993-7016. doi: 10.3934/math.2024341
    [10] Kaile Chen, Yunyun Liang, Nengqiu Zhang . Global existence of strong solutions to compressible Navier-Stokes-Korteweg equations with external potential force. AIMS Mathematics, 2023, 8(11): 27712-27724. doi: 10.3934/math.20231418
  • This study typically emphasizes analyzing the geometrical singularities of weak solutions of the mixed boundary value problem for the stationary Stokes and Navier-Stokes system in two-dimensional non-smooth domains with corner points and points at which the type of boundary conditions change. The existence of these points on the boundary generally generates local singularities in the solution. We will see the impact of the geometrical singularities of the boundary or the mixed boundary conditions on the qualitative properties of the solution including its regularity. The solvability of the underlying boundary value problem is analyzed in weighted Sobolev spaces and the regularity theorems are formulated in the context of these spaces. To compute the singular terms for various boundary conditions, the generalized form of the boundary eigenvalue problem for the stationary Stokes system is derived. The emerging eigenvalues and eigenfunctions produce singular terms, which permits us to evaluate the optimal regularity of the corresponding weak solution of the Stokes system. Additionally, the obtained results for the Stokes system are further extended for the non-linear Navier-Stokes system.



    Let DR2 be a 2-dimensional bounded domain, whose boundary D comprises the corner points and points at which the type of boundary conditions change. Note that a point PD is said to be a corner point if there exists a neighborhood η(P) of P such that Dη(P) is diffeomorphic to an angle κ intersected with the unit circle. For simplicity, we are considering a bounded plane polygonal domain (see Figure 1) with corner points (ωπ) and points (ω=π) where the boundary conditions change. The boundary points where the boundary conditions change are also referred to as corner points or vertices. The obtained results for a polygonal domain can be extended to a 2-dimensional bounded domain, i.e. (Lipschitz continuous) C0,1 with corner points. We considered one point as a special case of interest of corner points with an angle ω=π on one side of the domain D, where the Neumann boundary condition, the Dirichlet boundary condition, respectively, is prescribed.

    Figure 1.  Schematic illustration of the polygonal domain with vertices P1,...,PN.

    For the polygonal domain D with the vertices P1,...,PN, we introduce the following notations. Let PN+1=P1, J={1,...,N}, Γi(iJ) is the open edge connecting the vertices Pi+1 and Pi, Γ0=ΓN, and ωi(iJ) is the interior angle made by Γi1,Γi. Let JD={iJ: on Γi the Dirichlet boundary conditions are prescribed} and JN={iJ: on Γi the Neumann boundary conditions are prescribed}. Let N denote the set of the boundary points of D, i.e., N={Pi}, iJ.

    We assume that JD, JN are non-empty disjoint sets and J=JDJN. Moreover, let Γ0, Γ1 be given by Γ0=iJD¯Γi, Γ1=iJN¯Γi. We have Γ0Γ1= and D=¯Γ0¯Γ1.

    We consider the Navier-Stokes equations for an incompressible, viscous fluid, i.e.

    {ut+jujxju+qνΔu=finD,divu=0inD, (1.1)

    where u=(u1,u2) is the velocity vector field with the cartesian components u1,u2, ν is the viscosity parameter of the fluid flow, i.e., (ν>0), q is the hydrostatic pressure and f is a given volume force density.

    If the viscosity coefficient ν is sufficiently large, then the flow can be described by the following stationary Stokes system on a domain D:

    {νΔu+q=finD,divu=0inD. (1.2)

    The following mixed boundary conditions are considered on the boundary D:

    u=h1 on Γ0, (1.3)
    S[u,q]n=h2 on Γ1, (1.4)

    where n=(n1,n2) is the unit outward normal vector to the boundary and S[u,q] is the hydrostatic stress tensor with the cartesian components

    S[u,q]=qδij+ν(uixj+ujxi). (1.5)

    Here, δij is the Kronecker symbol. Further note that if the second equation of (1.2) becoming divu=g for a given function g satisfying the property Dgdx=0, then a particular regularity of g is required for proving the regularity of the pressure function or for handling the non-zero boundary data. Generally, for incompressible flows the function g is set equal to zero to satisfy the incompressibility condition. For simplicity, we are considering g equal to zero. Therefore, for a smooth boundary, smooth given data and boundary conditions, the system (1.2) has a smooth solution [41]. The system (1.2) with the boundary conditions (1.3)-(1.4) is known as the stationary Stokes system with mixed boundary conditions [30,36].

    The Navier-Stokes equations or even the Stokes equations are solved for Dirichlet boundary conditions [7,10,11,15,17] but this is not common in some situations like finite channel flow models [17,28]. Usually, these boundary conditions are used in the upstream of the channel and on the fixed walls but not downstream of the channel, because the downstream velocity depends on the flow in a channel which is unknown. The situation becomes more intricate when the boundary of the domain has corners or edges and the Neumann boundary conditions are applied on parts of the boundary [27,35]. Therefore, the second equation of (1.2) helps to characterize the different types of Neumann boundary conditions with the Green theorem. In numerical methods, the condition (1.4) is used on the downstream boundary [13].

    Presently, the corner singularity theory has been constructed for compressible viscous Stokes and Navier-Stokes systems for polygonal domains in 2-dimensions and polyhedral domains in 3-dimensions. The mathematical techniques to analyze the singular structure of the solutions near corners, edges, and cusps have been discussed in [12,16,21,22,26,27,37]. The key point of the corner singularity theory is the decomposition of the solution of the given problem into a regular part and a locally acting singular part which is a linear combination of explicit model singular solutions sm with unknown coefficients cm. The special singular functions sm rely on the geometry of the model problem, the differential operator, and the characteristic boundary conditions.

    In the singularity expansion method for the Stokes problem, the spectral problems related to the corner singularities of solutions to elliptic equations were discussed in [7,8]. Kweon [31] has considered zero Dirichlet boundary conditions to investigate the regularity results of the incompressible Navier-Stokes equations in a non-convex polygonal domain. [32,34] have extended these results for a non-convex polyhedral cylinder in R3 with inflow boundary conditions for compressible Navier-Stokes equations. The Helmholtz decomposition to obtain regularity results of the compressible Stokes system in a non-convex polygonal domain with no-slip boundary conditions is used in [33]. Recently, Anjam [4] has comprehensively discussed the singularities and regularity results of the stationary Stokes and Navier-Stokes equations on polygonal domains with convex or non-convex corners.

    This study typically emphasizes analyzing the boundary singularities and regularity results of the stationary Stokes and Navier-Stokes system in two-dimensional non-smooth domains with corner points and points at which the type of boundary conditions change. We use the theory developed by Kondratˊiev [24,25] and further extended by [39] for scalar problems in the context of weighted Sobolev spaces. The solvability of the considered boundary value problem is analyzed in the context of these weighted Sobolev spaces and the regularity theorems are formulated. To compute the singular terms for various boundary conditions, the Fourier transform is used to obtain the generalized form of the boundary eigenvalue problem for the stationary Stokes system. The emerging eigenvalues and eigenfunctions produce singular terms, which permits us to evaluate the optimal regularity of the corresponding weak solution of the stationary Stokes system.

    The main result for the stationary Stokes system is presented in Theorem 3, and the regularity results that are direct consequences of Theorem 3 are given in Section 4. Moreover, it is proved that the weak solution (u,q) of the underlying boundary value problem belongs to W2γ,2(D)2×W1γ,2(D), where γ is an arbitrarily small positive real number that depends on the apex angle ω0. Additionally, the obtained results for the Stokes system are further extended for the non-linear Navier-Stokes system. It is proved by using the local diffeomorphism theorem that the solution of the Navier-Stokes equations has similar regularity results as the solution of the generalized Stokes problem near the corner points if the norm of the body force is sufficiently small. So far the problem has been ignored with such type of boundary conditions and domain.

    The rest of this paper is organized as follows: Section 2 is devoted to present the weak formulation of the Stokes problem and introduce some function spaces. In Section 3, we determine a parametric boundary eigenvalue problem with a complex parameter ξ, the stationary Stokes system is being considered for various combinations of Dirichlet, Neumann, and mixed boundary conditions. The main regularity and expansion theorem for the stationary Stokes system is given in Theorem 3. The transcendental equations for various conditions whose zeros are the eigenvalues of the operator ˆU(ξ) are derived. Further, the distribution of eigenvalues and eigenfunctions are discussed. In Section 4, some regularity results for the stationary Stokes system are investigated. The results of Section 4 are further extended for non-linear Navier-Stokes system in Section 5. Section 6 is devoted to conclusions.

    In this section, we consider the weak formulation of the stationary Stokes problem (1.2)-(1.4). The variational formulation, solvability, and the uniqueness of the solution are offered. For a weak solution, we have restricted ourself to the homogenous boundary conditions. Denote

    E(D)={uC(¯D)2;divu=0,¯suppuΓ0=},

    where suppu={x|u(x)0}, and suppuD. Let Vm,p be a closure of E(D) in the norm of Wm,p(D)2,1p< and m0 (m need not be an integer), it is a Banach space with the norm of Wm,p(D)2. For simplicity, we represent V0,2 and V1,2, respectively, as H and V. Usually, these spaces are used to find the solution of the Navier-Stokes equations with homogeneous Dirichlet boundary conditions, and they are closed subspaces of the spaces L2(D)2 and W1,2(D)2. They are Hilbert spaces with the scalar products

    (u,v)H=Duvdx,((u,v))V=Duvdx=Duixjvixjdx. (2.1)

    Let fH. A pair (u,q)V×L2(D) is called weak solution for the problem (1.2)-(1.4) if it satisfies

    a(u,v)+b(q,v)=(f,v) vV,b(u,p)=0 pL2(D), (2.2)

    where

    a(u,v)=νDuvdxandb(q,v)=Dq(divv)dx.

    Equation (2.2) is the weak formulation of the boundary value problem (1.2)-(1.4), which is obtained by multiplying the first equation of (1.2) by a test function vV and the second equation by pL2(D). The pressure q is a scalar function, such that, the pair (u,q) satisfy the Eqs (1.2)-(1.4) in domain D in the sense of distributions. The bilinear form a(.,.) is elliptic and continuous, whereas the bilinear form b(.,.) is continuous and verifies the inf-sup condition (see [15,20,41]). It is proved in [6,38] that for every fH, there exists a unique weak solution (u,q) of the boundary value problem (1.2)-(1.4) and for it the following estimate holds:

    uV+qL2(D)cfH, (2.3)

    where c=c(D). Hence, we have to analyze the smoothness of the weak solution (u,q) and see how it depends on the sizes of the angles ωi,i=1,...,N.

    Remarks 1. If the given data on the right-hand sides of (1.2)-(1.4) are smoother, for example, fL2(D)2, h1[H32(Γ0)]2 and h2[H12(Γ1)]2, if the domain is sufficiently smooth and the boundary conditions do not change their types, then it is proved (see [41]) that the weak solution (u,q)[H2(D)]2×[H1(D)]. On the other hand, the same regularity result does not hold, if the domain has corner points or points at which the type of boundary conditions changes (see [16,23]). As a matter of fact, in these cases, the regularity can be described by a decomposition of the two-dimensional solution field

    u(x1,x2)=(u1(x1,x2)u2(x1,x2)q(x1,x2)), (2.4)

    into singular and regular parts of the form

    u=using+ureg,=j,krξj,kkΦj,k(ξj,k,rk,θk)+ureg. (2.5)

    Here, the regular part ureg belongs to [H2(D)]2×[H1(D)], the corner points are shown by k with the equivalent polar coordinates (rk,θk), the exponents ξj,k are the eigenvalues of a Sturm-Liouville problem, and Φj,k are the corresponding generalized eigenvector fields.

    To investigate the regularity results of the weak solution (u,q) of the corresponding boundary value problem, firstly, we introduce some function spaces in line with [1,14,24,39].

    Let N be the set of the corner points and of points where the type of boundary conditions change (shortly called the singular boundary points), i.e., ND. Denote

    CN={vC(¯D),suppv¯N=},

    where the suppv is bounded. Let α=(α1,...,αN) be an Ntuple of real numbers which satisfying 0<αi<1 for 1iN. Therefore, the weight function is characterized by

    Φα+m(x)=Ni=1(ri(x))αi+m,

    where m is an any integer and ri(x)=dist(x,Pi). We denote by Dβv be the multi-index notation for higher-order derivatives and in cartesian coordinates is defined as

    Dβv=|β|vxβ11xβ22, β=(β1,β2), |β|=β1+β2.

    Let Wm,pα(D) be the weighted Sobolev spaces and is the closure of CN(D) equipped with the norm

    vWm,pα(D)=(|β|mD|x|p(αm+|β|)|Dβv|pdx)1p. (2.6)

    Analogously to the factor spaces, the trace spaces are also defined as

    Wm1p,pα(D)=Wm,pα(D)/Wm,p0,α(D), (2.7)

    where Wm,p0,α(D) is the closure of C0(D) with respect to the norm of Wm,pα(D). This approach is classical for domains with conical points. Principally, for 1p<, we denote Lpα(D)=W0,pα(D).

    For negative integers m, i.e., mZ,m<0, we describe the spaces Wm,pα(D) as the closure of the set CN(D) equipped with the subsequent norm

    uWm,pα(D)=supvWm,qα(D),v0|Duvdx|/vWm,qα(D), (2.8)

    where q=pp1 is known as the inverse of p. The dual space of Wm,pα(D) is given as Wm,qα(D). Similarly for the trace spaces, defining the space Wm+1q,pα(D) for m<0,mZ, as the dual space to Wm1q,qα(D). Therefore, the consequent continuous imbeddings are considered directly from the definition of the above spaces

    Wm,pα(D)Wm1,pα1(D), (2.9)
    Wm1p,pα(D)Wm11p,pα1(D). (2.10)

    For a bounded plane domain D, we have the subsequent continuous imbeddings

    Wm,pα(D)Wm,pα1(D)if α1>α, (2.11)

    and

    Wm,p2α2(D)Wm,p1α1(D), (2.12)

    provided that p2p1 and α2+2p2<α1+2p1.

    Let Q={(τ,θ):<τ<,0<θ<ω0} denote the infinite strip with positive width ω0. For any real h>0 and for an integer m0, the spaces are defined as

    Wmh(Q)={uL2(Q):|β|mQe2hτ|Dβu|2dτdθ<},

    where

    uWmh(Q)=(|β|mQe2hτ|Dβu|2dτdθ)12.

    In this section, we will see the occurrence of the singular terms of the solution of the mixed boundary value problem for the stationary Stokes problem near the corners and the structure which they have. Further, the distribution of the eigenvalues and eigenfunctions are given.

    Assume that D is a polygonal domain. To show that the weak solution (u,q) of the underlying boundary value problem is regular, we have to investigate its behaviour near the corner points Pi(iJ). Let us consider the point PN as origin and denote ωN=ω0(0,2π). An appropriate infinite differentiable cut-off function χ(|x|)=χ(r) depending on the distance r from the point PN is defined as

    χ(r)={1 for0rϵ,0 forr2ϵ.

    The number ϵ is so small that PN is the only corner point of the domain D that lies inside the circle {x:|x|2ϵ}. We multiply the both sides of (1.2) and (1.3)-(1.4) by the smooth cut-off function χ, then substitute (v,p)=(χu,χq) in (1.2) and likewise in (1.3)-(1.4). The derivatives are considered in the distribution sense. Thus, the boundary value problem is set into an infinite cone

    S={(r,θ):0<r<,0<θ<ω0},

    and coincides with the original problem near the PN. The Stokes system (1.2) becomes

    {νΔv+p=FinS,divv=GinS, (3.1)

    where F=χf2νχuνuΔχ+qχ and G=uχ. The behavior of (v,p) near the corner point PN illustrates the regularity of the solution (u,q) in the neighborhood of the point PN. If we suppose that the right-hand side in (1.2) is fL2(D)2, then FL2(S)2 and GH1(S). Besides, the following boundary conditions are prescribed on the subsequent edges ΓS,0(θ=0) and ΓS,ω0(θ=ω0) of the cone (see Figure 2). Just one condition is considered per edge to differentiate between the mixed boundary conditions. Therefore, the obtained boundary conditions are:

    Dirichlet boundary conditions:

    v=H1 on ΓS,0,ΓS,ω0ifΓS,0,ΓS,ω0Γ0, (3.2)

    where χh1=H1.

    Figure 2.  Infinite cone S with opening angle ω0.

    Neumann boundary conditions:

    S[v,p]n=H2 on ΓS,0,ΓS,ω0,ifΓS,0,ΓS,ω0Γ1, (3.3)

    where χh2+νn(χu+uχ)=H2, and the notation () denotes the vector direct product between two vectors.

    Mixed boundary conditions:

    {v=H1on ΓS,0 if ΓS,0Γ0,S[v,p]n=H2onΓS,ω0 if ΓS,ω0Γ1. (3.4)

    It is observed that the right-hand sides of the obtained boundary conditions have similar smoothness as the original problem in the domain D. To analyze the regularity results of the boundary value problem (3.1)-(3.4), we rewrite the operators in polar coordinates. Hence, the transformed form is

    ν(2vrr2+1rvrr+1r22vrθ2vrr22r2vθθ)+pr=Fr,ν(2vθr2+1rvθr+1r22vθθ2vθr2+2r2vrθ)+1rpθ=Fθ,1rr(rvr)+1rθvθ=G, (3.5)

    where (vr,vθ) are the polar components of the velocity vector ¯v, (Fr,Fθ) are the polar components of ¯F and are given by

    ¯v=(vrvθ)=A(v1v2),¯F=(FrFθ)=A(F1F2),A=(cosθsinθsinθcosθ).

    Correspondingly, the boundary conditions (3.2)-(3.4) emerge as

    ¯v|θ=0,ω0=(vr,vθ)T|θ=0,ω0=¯H1, (3.6)
    {1rvrθ+vθr1rvθ|θ=0,ω0=¯H2r,p+2ν(1rvθθ+1rvr)|θ=0,ω0=¯H2θ, (3.7)
    {vr|θ=0=¯H1r,vθ|θ=0=¯H1θ,1rvrθ+vθr1rvθ|θ=ω0=¯H2r,p+2ν(1rvθθ+1rvr)|θ=ω0=¯H2θ, (3.8)

    and ¯Hm=(¯Hmr,¯Hmθ)T, where m=1 for Dirichlet and m=2 for Neumann boundary conditions. They hold in the infinite cone where ¯v(r,θ)=v(x1,x2),¯p(r,θ)=p(x1,x2),¯F(r,θ)=F(x1,x2) and ¯G(r,θ)=G(x1,x2).

    Now, the variable τ is introduced by the relation r=eτ. Accordingly, the system (3.5) is set on the infinite strip with width ω0 as

    ν(2˜vττ2+2˜vτθ2˜vτ2˜vθθ)+˜pτ˜p=~Fτ in ˉS,ν(2˜vθτ2+2˜vθθ2˜vθ+2˜vτθ)+˜pθ=~Fθ in ˉS,˜vττ+˜vτ+˜vθθ=˜G in ˉS. (3.9)

    Here, ˉS={(τ,θ):<τ<,0<θ<ω0} and ˜v=¯v(eτ,θ),˜p=eτ¯p(eτ,θ),˜F=e2τ¯F(eτ,θ) and ˜G=eτ¯G(eτ,θ). The Dirichlet, Neumann and mixed boundary conditions also yield the transformed form with the boundary data ˜Hl+1=elτ¯Hl+1(eτ,θ),l=0,1 as

    ˜v|θ=0,ω0=(˜vτ,˜vθ)T|θ=0,ω0=˜H1, (3.10)
    {±ν(˜vτθ+˜vθτ˜vθ)|θ=0,ω0=˜H2τ,±(˜p+2ν(˜vθθ+˜vτ))|θ=0,ω0=˜H2θ, (3.11)
    {˜vτ|θ=0=˜H1τ,˜vθ|θ=0=˜H1θ,ν(˜vτθ+˜vθτ˜vθ)|θ=ω0=˜H2τ,˜p+2ν(˜vθθ+˜vτ)|θ=ω0=˜H2θ. (3.12)

    To obtain the boundary eigenvalue value problem, the complex Fourier transform with respect to τ is introduced as

    F[v](ξ)=ˆv(ξ)=(2π)12eiξτ˜v(τ)dτ, ξC, (3.13)

    and the inverse Fourier transform is

    F1[v](ξ)=˜v(τ)=(2π)12+ih+iheiξτˆv(ξ)dξ. (3.14)

    It defines an isomorphic mapping, i.e.,

    F[v](ξ)={˜v(τ):e2hτ|˜v(τ)|2dτ<}L2(R+ih), (3.15)

    for ξ=s+ih, where h=constant, R+ih={ξC:Imξ=h}. Therefore, the subsequent Parseval identity holds

    e2hτ|˜v(τ)|2dτ=+ih+ih|ˆv(ξ)|2dξ. (3.16)

    We have

    F(dmdτm˜v(τ))(ξ)=(iξ)mF(˜v(τ))(ξ). (3.17)

    Moreover, it is noted that if h1<h2 and the following properties are satisfied

    +e2h1τ|˜v(τ)|2dτ<,+e2h2τ|˜v(τ)|2dτ<, (3.18)

    then ˆv(ξ) is holomorphic in the strip h1<Imξ<h2. Therefore, the relationship between the Fourier transform and the Mellin transform for any αC is given by

    Reα=Imξ,Imα=Reξ.

    Now, by applying (3.13) to (3.9)-(3.8) with respect to τ, the two-point boundary value problem for the unknown functions (ˆvτ,ˆvθ,ˆp) is obtained. It depends on the complex parameter ξ and holds on the interval I=(0,ω0). The transformed form of the problem (3.9) is given by

    ν(2ˆvτθ2(1+ξ2)ˆvτ2ˆvθθ)+(1+iξ)ˆp=ˆFτ,ν(2ˆvθθ2(1+ξ2)ˆvθ+2ˆvτθ)+ˆpθ=ˆFθ,(1+iξ)ˆvτ+ˆvθθ=ˆG. (3.19)

    For complex parameter ξ, we have ˆvW2,2(I)2, ˆpW1,2(I), ˆFL2(I)2 and ˆGW1,2(I). Let ˆL(ξ) denote the matrix differential operator analogous to the system (3.19) and maps W2,2(I)2×W1,2(I)L2(I)2×W1,2(I). Therefore, one has

    ˆL(ξ)(ˆv,ˆp)=(ˆF,ˆG) on I=(0,ω0), (3.20)

    where

    ˆL(ξ)=(ν[2θ2(1+ξ2)]2νθ(1iξ)2νθν[2θ2(1+ξ2)]θ(1+iξ)θ0). (3.21)

    The operator ˆL(ξ) is considered for all parameter ξC with various combinations of the boundary conditions to analyze the qualitative properties of the solution of the underlying problem near the corner points. Additionally, the Fourier transformed form of the boundary conditions is also expressed as follows:

    {ˆvτ(ξ,θ)|θ=0,ω0=ˆH1τ,ˆvθ(ξ,θ)|θ=0,ω0=ˆH1θ. (3.22)
    {±ν(ˆvτθ(1iξ)ˆvθ)|θ=0,ω0=ˆH2τ,±(ˆp+2ν(ˆvθθ+ˆvτ))|θ=0,ω0=ˆH2θ. (3.23)
    {ˆvτ(ξ,θ)|θ=0=ˆH1τ,ˆvθ(ξ,θ)|θ=0=ˆH1θ,ν(ˆvτθ(1iξ)ˆvθ)|θ=ω0=ˆH2τ,ˆp+2ν(ˆvθθ+ˆvτ)|θ=ω0=ˆH2θ. (3.24)

    Additionally, the matrix boundary operators for different kinds of boundary conditions can be written as:

    For Dirichlet boundary conditions

    ˆBDD1(ξ)|θ=0=(100010), ˆBDD2(ξ)|θ=ω0=(100010). (3.25)

    For Neumann boundary conditions

    ˆBNN1(ξ)|θ=0=(νθν(1iξ)02ν2νθ1),ˆBNN2(ξ)|θ=ω0=(νθν(1iξ)02ν2νθ1). (3.26)

    For mixed boundary conditions

    ˆBDN1(ξ)|θ=0=(100010),ˆBDN2(ξ)|θ=ω0=(νθν(1iξ)02ν2νθ1). (3.27)

    Analogously, the operator ˆB[..](ξ) is used below to define the general transformed form of the matrix boundary operators for different kinds of boundary conditions

    {ˆB[..](ξ)(ˆv,ˆp)}=(ˆH1,ˆH2) on I=(0,ω0). (3.28)

    Accordingly, the generalized form of the operator pencil ˆU(ξ) for the two-point boundary value problem can be written as

    ˆU(ξ)=[ˆL(ξ),{ˆB[..](ξ)}]. (3.29)

    Thus, the operator ˆU(ξ) maps W2,2(I)2×W1,2(I) into L2(I)2×W1,2(I)×C2×C2. Note that ˆU(ξ) can be defined for every boundary point in the sense of [2,3]. So, ˆU(ξ)(θ,ξ)=0 is used to describe a generalized eigenvalue problem and the solvability of these type of problems is discussed in [26]. The operator ˆU(ξ) realizes an isomorphism for all ξC apart from some isolated points (known as the eigenvalues of ˆU(ξ)). So, the resolvent R(ξ)=[ˆU(ξ)]1 is an operator-valued, meromorphic function of ξ has poles of finite multiplicity. The eigenvalues of ˆU(ξ) are obtaining with the determinant method, which means that the nontrivial solution of the generalized eigenvalue problem leads a transcendental equation whose zeros are the eigenvalues of ˆU(ξ).

    Besides, the properties of the operator ˆU(ξ) in the neighborhood of the corner points can be obtained by the properties of the special operator ˆU0(ξ) which is explicitly given by the principal parts of the matrix differential operator ˆL(ξ) and the matrix boundary operators ˆB[..](ξ). Hence, we have

    ˆL0(ξ)=(ν(2θ2ξ2)2νθiξ2νθν(2θ2ξ2)θiξθ0), (3.30)

    and

    ˆB0DD1(ξ)|θ=0=(100010), ˆB0DD2(ξ)|θ=ω0=(100010). (3.31)
    ˆB0NN1(ξ)|θ=0=(νθνiξ02ν2νθ1),ˆB0NN2(ξ)|θ=ω0=(νθνiξ02ν2νθ1). (3.32)

    Likewise, we can write for mixed boundary conditions. To compute the eigenvalues ξμ (generally referred for multiple eigenvalues) and the corresponding eigenfunctions, we proceed as.

    Definition 1. A complex number ξ=ξ0 is known as the eigenvalue of ˆU(ξ) if there exists a nontrivial solution i.e., ˆu(.,ξ0)0, which is holomorphic at ξ0, and ˆU(ξ0)ˆu(θ,ξ0)=0. ˆu(θ,ξ0) is called an eigenfunction of ˆU(ξ0) corresponding to the eigenvalue ξ0. The set of fields {ˆu0(θ,ξ0),ˆu0,1(θ,ξ0),...,ˆu0,s(θ,ξ0)} with ˆu0,0=ˆu0 is said to be a Jordan chain corresponding to the eigenvalue ξ0, if the equation

    σq=01q!(ξ)qˆU(ξ)ˆu0,mq(θ,ξ)|ξ=ξ0=0form=1,2,...,s,

    is satisfied. The number s+1 is called the length of the Jordan chain.

    Remarks 2. It is noted [24,25,26] that if the complex number ξ is not an eigenvalue of the operator ˆU(ξ), then ˆU(ξ) is an isomorphism between the spaces W2,2(I)2×W1,2(I) and L2(I)2×W1,2(I)×C2×C2.

    Lemma 1. Let lh={ξC:Imξ=h}. If no eigenvalues of ˆU(ξ) lies on the line lh, then the system (3.19) and (3.22)-(3.24) admits a unique solution (ˆv,ˆp)W2,2(I)2×W1,2(I) provided (ˆF,ˆG,ˆH1,ˆH2)L2(I)2×W1,2(I)×C2×C2, and it holds for all ξlh:

    ˆv2W2,2(I)2+ˆp2W1,2(I)c{ˆF2L2(I)2+ˆG2W1,2(I)+l=0,1|ξ|32l|ˆHl+1|2}, (3.33)

    with the constant c is independent of ξ.

    Proof. A similar theorem is proved in [[19], Theorem 4.9]. So, we omit its proof.

    Therefore, the Lemma 1 provides us an opportunity to prove the following theorem of the solvability of the problem (3.1)-(3.4).

    Theorem 1. Let FW0,2α(S)2, GW1,2α(S) and Hl+1W2l12α(Γl)2,l=0,1. If the line Imξ=h=α1 contains no eigenvalue of the operator ˆU(ξ), then the problem (3.1)-(3.4) admits a uniquely determined solution (v,p)W2,2α(S)2×W1,2α(S) and satisfies the following estimate

    vW2,2α(S)2+pW1,2α(S)c{FW0,2α(S)2+GW1,2α(S)+l=0,1Hl+1W2l12α(Γl)2}, (3.34)

    where c>0 is independent of v and F.

    Proof. We prove this theorem by following the idea of Kondratˊiev [24]. Suppose that the line Imξ=h=α1 contains no eigenvalue of the operator ˆU(ξ). First of all, we prove that the right-hand sides functions of the system (3.9) are Fourier transform in the sense of (3.15). We know from (3.1) that FL2(S)2, GW1,2(S). Further note that for all α0, FW0,2α(S)2 and GW1,2α(S). Since, FW0,2α(S)2, we have

    S|F(x)|2|x|2αdx=ˉSe2(τα+τ)|˜F(τ,θ)|2dτdθ<, (3.35)

    where h=α1 for all α0 and it is meaningful in the sense of (3.15). Therefore, the Fourier transform of ˜F(τ,θ)=(˜Fτ,˜Fθ) is meaningful in the half plane h=Imξ1 for almost all θ(0,ω0).

    Analogously, for GW1,2α(S), we have

    S|G(x)|2|x|2(α1)dx=ˉSe2τ(α1)|˜G(τ,θ)|2dτdθ<, (3.36)

    where h=α1 for all α0. Therefore, the function ˜G(τ,θ) is also Fourier transformable in the half plane h=Imξ1 for almost all θ(0,ω0) in the sense of (3.15).

    Now, the construction of the singular vector functions can be explained by the observations from [25,26,39]. The main question is the inverse Fourier transform of the right-hand sides of (3.19) and (3.22)-(3.24) or simply (3.29) which can be read as follows using the formula (3.14):

    ˜vj,h(τ,θ)=(2π)12+ih+iheiξτˆU1(ξ)[(ˆF,ˆG),(ˆH1,ˆH2)]dξ. (3.37)

    Using the Cauchy theorem yields

    ˜vj,h(τ,θ)=(2π)12limn{n+iδn+iheiξτˆU1(ξ)[(ˆF,ˆG),(ˆH1,ˆH2)]dξ+n+iδn+iδeiξτˆU1(ξ)[(ˆF,ˆG),(ˆH1,ˆH2)]dξ+n+ihn+iδeiξτˆU1(ξ)[(ˆF,ˆG),(ˆH1,ˆH2)]dξ}+12π2πiNν=1Res(eiξτˆU1(ξ)[(ˆF,ˆG),(ˆH1,ˆH2)])|ξ=ξν. (3.38)

    The first integral and the third integral tend to zero as n by [24]. The second integral yields vj,h(x)W2,2α(S)2. The calculation of the residue gives the singular terms. If the operator ˆU(ξ) contains no eigenvalue on the line Imξ=α1,α0, then the residue vanishes and the inverse Fourier transform

    ˜vj,h(τ,θ)=(2π)12+ih+iheiξτˆvj,h(ξ,θ)dξ=vj,h(x)W2,2α(S)2,

    for j=1,2 exists. Thus, ˜vj,h(x) is the uniquely determined solution from W2,2α(S)2 of the underlying boundary value problem. An analogous result, we obtain for the pressure ˜pW1,2α(S), where h=α1.

    Now, using Lemma 1 and (3.16)-(3.17), we can get (3.34).

    Lemma 2. [29]. Let ˆU1(ξ) be the inverse operator of ˆU(ξ). ˆU1(ξ) is a meromorphic operator-valued function with poles which are the eigenvalues of ˆU(ξ). The order m of a pole ξ0 is the largest of the lengths of the Jordan chains corresponding to ξ0. Moreover, the operator ˆU1(ξ) has the following expansion in the neighborhood of ξ0:

    ˆU1(ξ)=qm(ξξ0)m+...+q1(ξξ0)+Γ(ξ), (3.39)

    where qi:i=1,...,m are the finite-dimensional operators which do not depend on ξ and Γ(ξ) is holomorphic. The operator qm behaves into the space of eigenfunctions of ˆU(ξ) corresponding to ξ0, while the operators qm1,...,q1 behave into the subspaces of the corresponding associate functions.

    The next theorem describes the expansion and the regularity of the problem (3.1)-(3.4).

    Theorem 2. Let (v,p)W2,2α(S)2×W1,2α(S) be a solution of the problem (3.1)-(3.4) for every FW0,2α(S)2W0,2α1(S)2, GW1,2α(S)W1,2α1(S) and Hl+1W2l12α(Γl)2W2l12α1(Γl)2,l=0,1, α1<α. Assume that no eigenvalue of ˆU(ξ) lies on the lines Imξ=h1=α11 and Imξ=h=α1. If ξ1,ξ2,...,ξM are the eigenvalues of ˆU(ξ) in the strip α11<Imξ<α1, then the solution (v,p) admits the following expansion

    (v,p)=[Mμ=1Iμρ=1κμρ1κ=0cμ,ρ,κΨμ,ρ,κ(r,θ)]+[vreg(r,θ),preg(r,θ)], (3.40)

    where Ψμ,ρ,κ(r,θ) are the corresponding singular functions given by

    Ψμ,ρ,κ(r,θ)=(vμ,ρ,κ(r,θ),pμ,ρ,κ(r,θ)),

    with

    vμ,ρ,κ(r,θ)=riξμκj=0(ilogr)jj!ϕρ,κjμ(θ),
    \begin{equation*} {p}_{\mu,\,\rho,\,\kappa}(r,\, \theta) = r^{i\xi_{\mu}-1}\,\sum\limits_{j = 0}^{\kappa}\frac{(i\,\log r)^{j}}{j!}\,{\psi}_{\mu}^{\rho,\, \kappa-j}(\theta). \end{equation*}

    The regular part \Big({\bf v}_{reg}(r, \, \theta), {p}_{reg}(r, \, \theta)\Big)\in {\mathcal {W}_{\alpha_1}^{2, \, 2}(S)^2}\times {\mathcal {W}_{\alpha_1}^{1, \, 2}(S)} and satisfies the following estimate

    \begin{equation} \Vert {{\bf v}_{reg}}\Vert_{\mathcal {W}_{\alpha_1}^{2,\, 2}(S)^2}+\Vert p_{reg}\Vert_{\mathcal {W}_{\alpha_1}^{1,\, 2}(S)}\leq c\Big\{\Vert{\bf F}\Vert_{\mathcal {W}_{\alpha_1}^{0,\, 2}(S)^2}+\Vert{{G}}\Vert_{\mathcal {W}_{\alpha_1}^{1,\, 2}(S)}+\sum\limits_{l = 0,\,1} \Vert {\bf H_{l+1}}\Vert_{\mathcal {W}_{\alpha_1}^{2-l-\frac{1}{2}}(\Gamma^l)^2}\Big\}. \end{equation} (3.41)

    Proof. It follows from Theorem 1 that the solution ({\bf v}, p)\in {\mathcal {W}_{\alpha}^{2, \, 2}(S)^2}\times {\mathcal {W}_{\alpha}^{1, \, 2}(S)} of the problem (3.1)-(3.4) is uniquely determined and specified by the formula (3.37). The use of Cauchy theorem yields (3.38). It is already stated in Theorem 1 that the first and third integrals in (3.38) are tending to zero as for {n\rightarrow \infty} . The second integral produces that {\bf v}_{reg}(\mbox{x})\in \mathcal {W}_{\alpha_1}^{2, \, 2}(S)^2 , {p}_{reg}(\mbox{x})\in \mathcal {W}_{\alpha_1}^{1, \, 2}(S) is the uniquely determined solution of (3.1)-(3.4) and the estimate (3.41) is valid. This statement follows from Theorem 1.

    Now, we should calculate the residue in (3.38). Lemma 2 provide us that \hat{\mathcal{U}}^{-1}(\xi) is a meromorphic operator-valued function with poles which are the eigenvalues of \hat{\mathcal{U}}(\xi) , and \hat{\mathcal{U}}^{-1}(\xi) has the expansion in the form of (3.39). Moreover (\hat{{\bf F}}, \, \hat{G}) is holomorphic respecting \xi in the strip \alpha_{1}-1 < \mathrm{Im}\, {\xi} < \alpha-1 . Therefore, we can write

    \begin{equation} [\hat{{\bf F}}, \hat{G},\hat{{\bf H}}^{1}, \hat{{\bf H}}^{2}\big] = \sum\limits_{m = 0}^{\infty}\,b_m(\theta)\,(\xi-\xi_\mu)^m, \end{equation} (3.42)

    in the neighborhood of \xi_\mu , where the coefficients b_m(\theta) are elements of L^{2}(I)^2\times W^{1, 2}(I)\times \mathbb{C}^{2}\times \mathbb{C}^{2} . Further, we have

    \begin{equation} e^{i\xi\tau} = e^{i\xi_\mu\tau}\big[1+i(\xi-\xi_\mu)\tau+...+\frac{[i(\xi-\xi_\mu)\tau]^n}{n!}+...\big]. \end{equation} (3.43)

    From (3.39), (3.42) and (3.43), it follows that

    \begin{equation} \begin{gathered} e^{i\xi\tau}\,{\hat{\mathcal{U}}}^{-1}(\xi)\big[(\hat{{\bf F}}, \hat{G}),(\hat{{\bf H}}^{1}, \hat{{\bf H}}^{2})\big] = e^{i\xi_\mu\tau}\big[1+i(\xi-\xi_\mu)\tau+...+\frac{[i(\xi-\xi_\mu)\tau]^n}{n!}+...\big]\\ \cdot\big[\frac{q_{k_{\mu_1}}}{(\xi-\xi_\mu)^{k_{\mu_1}}}+...+\frac{q_1}{(\xi-\xi_\mu)}+\Gamma(\xi)\big] \big[\sum\limits_{m = 0}^{\infty}\,b_m(\theta)\,(\xi-\xi_\mu)^m\big]. \end{gathered} \end{equation} (3.44)

    Therefore, we conclude that

    \begin{equation} \begin{gathered} Res\,\Big[e^{i\xi\tau}\,{\hat{\mathcal{U}}}^{-1}(\xi)\big[(\hat{{\bf F}}, \hat{G}),(\hat{{\bf H}}^{1}, \hat{{\bf H}}^{2})\big]\Big]\Big\vert_{\xi = \xi_{\mu}} = e^{i\xi_\mu\tau}\big[{q_1}{b_1(\theta)}+...+q_{k_{\mu_1}}b_{k_{\mu_1}}(\theta)\big]\\ +e^{i\xi_\mu\tau}\frac{(i\tau)^{1}}{1!}\big[q_2 b_{1}(\theta)+...+q_{k_{\mu_1}}b_{k_{\mu_1}-1}(\theta)\big]+...+ e^{i\xi_\mu\tau}\frac{(i\tau)^{{k_{\mu_1}-1}}}{({k_{\mu_1}-1})!}\big[q_{k_{\mu_1}}b_{1}(\theta)\big]. \end{gathered} \end{equation} (3.45)

    Now, substituting r = e^{\tau} , applying Lemma 2 and ([27], Theorem 1.1.5), we obtain (3.40).

    Now, we derive our fundamental regularity and expansion theorem for the mixed boundary value problem for the stationary Stokes system in a two-dimensional bounded domain with corner points. By considering the substitution \mbox{Re}\, \alpha = -\mbox{Im}\, \xi-2 , it improves the theorems ([26], Theorem 8.2.1 and Theorem 8.2.2) which are based on the Mellin transform and used for the solvability of the elliptic systems.

    Theorem 3. (Regularity and expansion theorem): Let \alpha_{1} and \alpha_{2} be real numbers and satisfying \alpha_{1}-1 < \alpha_{2} < \alpha_{1} . Let a pair ({\bf u}, q)\in \mathcal {W}_{\alpha_{1}}^{m, \, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha_{1}}^{m-1, \, 2}(\mathcal{D}) be a solution of the stationary Stokes system (1.2) with the homogenous Dirichlet, Neumann, and mixed boundary conditions (1.3)-(1.4) and {\bf f}\in \mathcal {W}_{\alpha_{2}}^{m_{1}-2, \, p}(\mathcal{D})^{2}\cap \mathcal {W}_{\alpha_{1}}^{m-2, \, 2}(\mathcal{D})^{2} , where 1\leq p < \infty , m_{1}\geq m \geq 2 and \alpha_{1}\geq\alpha_{2}\geq0 . Then, the following implications hold:

    1. If the strip \alpha_{2}+\frac{2}{p}-m_{1}\leq \mathrm{Im}\, {\xi}\leq \alpha_{1}+1-m is free of eigenvalues of the operator \hat{\mathcal{U}}(\xi) , then the solution ({\bf u}, \, q)\in \big[\mathcal {W}_{\alpha_{2}}^{m_{1}, \, p}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha_{2}}^{m_{1}-1, \, p}(\mathcal{D})\big] and holds the following estimate

    \begin{equation*} \Vert {\bf u}\Vert_{\mathcal {W}_{\alpha_{2}}^{m_{1},\, p}(\mathcal{D})^{2}}+\Vert q\Vert_{\mathcal {W}_{\alpha_{2}}^{m_{1}-1,\, p}(\mathcal{D})}\leq c(\nu,\mathcal{D})\,\Vert{\bf f}\Vert_{\mathcal {W}_{\alpha_{2}}^{m_{1}-2,\, p}(\mathcal{D})^{2}}. \end{equation*}

    2. Let \xi_{1}, \xi_{2}, ..., \xi_{M} are the eigenvalues of the operator \hat{\mathcal{U}}(\xi) , and suppose that no eigenvalue lie on the lines \mathrm{Im}\, {\xi} = \alpha_{2}+\frac{2}{p}-m_{1} and \mathrm{Im}\, {\xi} = \alpha_{1}+1-m . If the eigenvalues \xi_{1}, \xi_{2}, ..., \xi_{M} are situated in the strip \alpha_{2}+\frac{2}{p}-m_{1} < \, \mathrm{Im}\, {\xi} < \alpha_{1}+1-m , then the solution ({\bf u}, q) admits the following expansion in the neighborhood P_{\delta} of the corner point P :

    \begin{equation} \big({\bf u}, q\big) = \chi(r)\,\Bigg[\sum\limits_{\mu = 1}^{M}\sum\limits_{\rho = 1}^{I_{\mu}}\sum\limits_{\kappa = 0}^{\kappa_{\mu\rho}-1} c_{\mu,\,\rho,\,\kappa}\,{\Psi}_{\mu,\,\rho,\,\kappa}(r,\, \theta)\Bigg]+\Big[{\bf u}_{reg}(r,\, \theta),\,{q}_{reg}(r,\, \theta)\Big], \end{equation} (3.46)

    with \big({\bf u}_{reg}(r, \, \theta), {q}_{reg}(r, \, \theta)\big)\in \mathcal {W}_{\alpha_{2}}^{m_{1}, \, p}{(P_{\delta})}^{2}\times \mathcal {W}_{\alpha_{2}}^{m_{1}-1, \, p}{(P_{\delta})} . Here, M is the number of eigenvalues of the operator \hat{\mathcal{U}}(\xi) in the strip, the constants c_{\mu, \, \rho, \, \kappa} depend on the data and the singular functions, I_{\mu} = \mathrm{dim}\, \mathrm{Ker}\, \hat{\mathcal{U}}(\xi_{\mu}) , \kappa_{\mu\rho} is the length of the Jordan chains of \hat{\mathcal{U}}(\xi_{\mu}) , and the corresponding singular functions are given by

    \begin{equation*} {\Psi}_{\mu,\,\rho,\,\kappa}(r,\, \theta) = \Big({\bf u}_{\mu,\,\rho,\,\kappa}(r,\, \theta),\, {q}_{\mu,\,\rho,\,\kappa}(r,\, \theta)\Big), \end{equation*}

    with

    \begin{equation} \begin{gathered} {\bf u}_{\mu,\,\rho,\,\kappa}(r,\, \theta) = r^{i\xi_{\mu}}\,\sum\limits_{j = 0}^{\kappa}\frac{(i\,\log r)^{j}}{j!}\,{\psi}_{\mu}^{\rho,\, \kappa-j}(\theta),\\[.5ex] {q}_{\mu,\,\rho,\,\kappa}(r,\, \theta) = r^{i\xi_{\mu}-1}\,\sum\limits_{j = 0}^{\kappa}\frac{(i\,\log r)^{j}}{j!}\,{\phi}_{\mu}^{\rho,\, \kappa-j}(\theta). \end{gathered} \end{equation} (3.47)

    It is noted from (3.46) and (3.47) that the eigenvalues \xi_{\mu} = 0 do not yield singularities in the development of the solution in the neighborhood P_{\delta} .

    It is recognized for elliptic boundary value problems that the eigenvalues of the operator \hat{\mathcal{U}}(\xi) which lies on the strip have a significant role in the regularity results. The assertions 1 and 2 of Theorem 3 represents the regularity and expansion of the solutions of the system (1.2)-(1.4) near the corner points.

    Remarks 3. The technique of Mellin transform, the method of special ansatzes, and spherical coordinates are used in [26,27] to obtain the generalized form of the boundary eigenvalue problem for the stationary Stokes system with Dirichlet and mixed boundary conditions. The existence of the generalized eigenvalues is discussed in a strip \mathrm{Re}\, \xi\in(0, \, 1) . Here, we use the Fourier transform technique to obtain the generalized form of the boundary eigenvalue problem for the stationary Stokes system with mixed boundary conditions. Moreover, the existence of the generalized eigenvalues in a strip \mathrm{Im}\, \xi\in(-1, \, 0) with various combinations of the boundary conditions that depend on the apex angle \omega_{0} are studied.

    Let (\hat{v}_{\tau}, \hat{v}_{\theta}, \hat{p}) be denoting the general solution of the system (3.19) by considering the right-hand side functions equal to zero, i.e., \hat{L}(\xi)(\hat{{\bf v}}, \hat{p}) = 0, where (\hat{v}_{\tau}, \hat{v}_{\theta}) stands the components of the velocity vector \hat{{\bf v}} and \hat{p} for the pressure function. On the other hand, If the system (3.19) has the non-zero right-hand side functions, then the general solution can be written as

    \begin{equation*} \Big[\hat{v}_{\tau},\, \hat{v}_{\theta},\, \hat{p}\Big]^{T} = I_{hom}+I_{p}(\xi,\, \theta). \end{equation*}

    Here, I_{p} denote the particular solution corresponding to the non-zero right-hand side functions and I_{hom} is the homogenous or general solution for zero right-hand sides. Here, our interest is to find the I_{hom} solution.

    For simplicity, we substitute \xi = -iz from [18] into (3.19), then a system of linear ordinary differential equations is obtained that depends on the complex parameter z .

    A system of linear ordinary differential equations in \theta with complex parameter z is considered

    \begin{equation} \begin{split} -\nu\Big(\frac{\partial^{2} \hat{v}_{\tau}}{\partial \theta^{2}}-(1-z^{2})\hat{v}_{\tau}-{2}\frac{\partial \hat{v}_{\theta}}{\partial \theta}\Big)-(1-z)\hat{p} & = \hat{F}_{\tau}, \\[.5ex] -\nu\Big(\frac{\partial^{2}\hat{v}_{\theta}}{\partial \theta^{2}}-(1-z^{2})\hat{v}_{\theta}+{2}\frac{\partial \hat{v}_{\tau}}{\partial \theta}\Big)+\frac{\partial \hat{p}}{\partial \theta} & = \hat{F}_{\theta}, \\[.5ex] (1+z)\hat{v}_{\tau}+\frac{\partial \hat{v}_{\theta}}{\partial \theta} & = \hat{G}. \end{split} \end{equation} (3.48)

    Furthermore, the system (3.48) provides a linear homogenous fourth-order ordinary differential equation with constant complex coefficients, i.e.,

    \begin{equation} \dfrac{d^{4}\hat{v}_{\theta}}{d\theta^{4}}+2(1+z^{2})\dfrac{d^{2}\hat{v}_{\theta}}{d\theta^{2}}+(z^{2}-1)^{2}\hat{v}_{\theta} = 0. \end{equation} (3.49)

    It is noted from the general theory of ordinary differential equations that (3.49) gives four independent solutions, and the general form of the fundamental solution for (z\neq 0, \, \pm 1) or (\xi\neq 0, \, \pm i) can be written as

    \begin{equation} \begin{split} \begin{pmatrix} \hat{v}_{\tau}\\\hat{v}_{\theta}\\\hat{p} \end{pmatrix} & = B_{1}\begin{pmatrix} \sin(z+1)\theta\\ \cos(z+1)\theta \\0 \end{pmatrix}+B_{2}\begin{pmatrix} -\cos(z+1)\theta\\ \sin(z+1)\theta \\0 \end{pmatrix}\\ &+B_{3}\begin{pmatrix} (z-1)\,\cos(z-1)\theta\\ -(z+1)\,\sin(z-1)\theta \\4\nu\, z\,\cos(z-1)\theta \end{pmatrix}+B_{4}\begin{pmatrix} (z-1)\,\sin(z-1)\theta\\ (z+1)\,\cos(z-1)\theta \\4\nu \,z\,\sin(z-1)\theta \end{pmatrix}. \end{split} \end{equation} (3.50)

    Therefore, it is necessary to consider the other cases for various values of z or \xi , and the general forms of their fundamental solutions are described as follows:

    Case 1. (For z = 0 ; or \xi = 0 ):

    \begin{equation} \begin{split} \begin{pmatrix} \hat{v}_{\tau}\\\hat{v}_{\theta}\\\hat{p} \end{pmatrix} & = B_{1}\begin{pmatrix} \sin\theta\\ \cos\theta \\0 \end{pmatrix}+B_{2}\begin{pmatrix} -\cos\theta\\ \sin\theta \\0 \end{pmatrix}\\&+B_{3}\begin{pmatrix} -\cos\theta+\theta\,\sin\theta\\ \theta\,\cos\theta \\-2\nu\,\cos\theta \end{pmatrix}+B_{4}\begin{pmatrix} -\sin\theta-\theta\,\cos\theta\\ \theta\,\sin\theta \\-2\nu\,\sin\theta \end{pmatrix}. \end{split} \end{equation} (3.51)

    Case 2. (For z = -1 ; or \xi = i ):

    \begin{equation} \begin{split} \begin{pmatrix} \hat{v}_{\tau}\\\hat{v}_{\theta}\\\hat{p} \end{pmatrix} = B_{1}\begin{pmatrix} 1\\ 0 \\0 \end{pmatrix}+B_{2}\begin{pmatrix} \cos 2\theta\\ 0 \\2\nu\,\cos\theta \end{pmatrix}\,+B_{3}\begin{pmatrix} \sin 2\theta\\ 0 \\2\nu\,\sin\theta \end{pmatrix}+B_{4}\begin{pmatrix} 0\\ 1 \\0 \end{pmatrix}. \end{split} \end{equation} (3.52)

    Case 3. (For z = 1 ; or \xi = -i ):

    \begin{equation} \begin{split} \begin{pmatrix} \hat{v}_{\tau}\\\hat{v}_{\theta}\\\hat{p} \end{pmatrix} = B_{1}\begin{pmatrix} \sin 2\theta\\ \cos 2\theta \\0 \end{pmatrix}+B_{2}\begin{pmatrix} \cos 2\theta\\ -\sin 2\theta \\0 \end{pmatrix}\,+B_{3}\begin{pmatrix} 0\\ 0 \\1 \end{pmatrix}+B_{4}\begin{pmatrix} 0\\ 1 \\0 \end{pmatrix}. \end{split} \end{equation} (3.53)

    The coefficients {\bf B} = (B_1, B_2, B_3, B_4)^T would be determined according to the types of boundary conditions. In line with the above cases, we can substitute \xi = -iz into Dirichlet, Neumann, and mixed boundary conditions. As well, their transformed forms also depend on the complex parameter z .

    To evaluate the eigenvalues and corresponding eigenfunctions of the stationary Stokes system, the solution (3.50) with various combinations of the boundary conditions is considered to obtain a system of four linear homogeneous equations with our unknowns B_{1}, \, B_{2}, \, B_{3} and B_{4} . The resulting matrix of coefficients of these equations depends on the complex parameter z , and a nontrivial solution exists if the determinant of the resulting matrix of coefficients vanishes (see [5]). Further, it produces the transcendental equations whose roots are the eigenvalues, namely, \xi_{\mu} wherein ( {\mu} is used for multiple eigenvalues, i.e., {\mu} = 1, ..., M ). To compute these results, we proceed as follows.

    Dirichlet boundary conditions (DD)

    It means that the Dirichlet boundary conditions are given on both sides of the corner point. The determinant method is used from [5] to obtain a system of linear homogeneous equations. A non-trivial solution exists if the determinant \det{D_{DD}(z)} of the corresponding system of the matrix of coefficients vanishes. So, the computation leads to the transcendental equation

    \begin{equation} 2z^{2}\,\sin^{2}(\omega_{0})+\cos(2z\omega_{0})-1 = 0. \end{equation} (3.54)

    The roots of (3.54) are the eigenvalues of the operator \hat{\mathcal{U}}_{DD}(\xi) = \big[\hat{L}(\xi), \{\hat{B}_{[DD]}(\xi)\}\big] .

    Neumann boundary conditions (NN)

    It means that the Neumann boundary conditions are given on both sides of the corner point. Therefore, the computation leads to the transcendental equation

    \begin{equation} \sin^{2}(z\omega_{0})-z^{2}\,\sin^{2}(\omega_{0}) = 0, \quad\ z\neq 0. \end{equation} (3.55)

    The eigenvalues of \hat{\mathcal{U}}_{NN}(\xi) = \big[\hat{L}(\xi), \, \big\{\hat{B}_{[NN]}(\xi)\big\}\big] are the roots of (3.55).

    Mixed boundary conditions (DN)

    It means that the Dirichlet or Neumann boundary condition is given on one side of the corner point, and the other condition is given on the other side. Similar to the latter cases, the obtained equation for this case is

    \begin{equation} \sin^{2}(z\omega_{0})+z^{2}\,\sin^{2}(\omega_{0})-1 = 0. \end{equation} (3.56)

    The roots of (3.56) are the eigenvalues of \hat{\mathcal{U}}_{DN}(\xi) = \big[\hat{L}(\xi), \, \{\hat{B}_{[DN]}(\xi)\}\big] .

    Remarks 4. Due to symmetry, the same results can be obtained if the versed boundary conditions are used which means that the Neumann condition is at \theta = 0 and the Dirichlet condition is at \theta = \omega_{0} . Therefore, this case of boundary conditions is not considered here.

    Remarks 5. Consequently, the poles of \mathcal{R}(\xi) are the numbers of -iz_{n} , where z_{n} are the roots of the Eqs (3.54)-(3.56).

    The following theorem describes the distribution of the eigenvalues of the corresponding boundary value problem for Dirichlet, Neumann, and mixed boundary conditions for various cases of the values of z .

    Theorem 4. Let \xi = -iz be an eigenvalue of the operator \hat{\mathcal{U}}(\xi) , and satisfies the following transcendental equations for Dirichlet, Neumann, and mixed boundary conditions. Then

    (ⅰ) for the Dirichlet problem (3.48) and (3.22), z satisfies the equation

    \begin{equation} 2z^{2}\sin^{2}(\omega_{0})+\cos(2z\omega_{0})-1 = 0, \end{equation} (3.57)

    (ⅱ) for the Neumann problem (3.48) and (3.23), z satisfies the equation

    \begin{equation} \sin^{2}(z\omega_{0})-z^{2}\sin^{2}(\omega_{0}) = 0, \end{equation} (3.58)

    (ⅲ) for the mixed problem (3.48) and (3.24), z satisfies the equation

    \begin{equation} \sin^{2}(z\omega_{0})+z^{2}\sin^{2}(\omega_{0})-1 = 0. \end{equation} (3.59)

    It is easily examined from (3.29) and (3.57)-(3.59) that the zeros of these equations are symmetric with respect to the origin and the real axis lies in the complex plane. Therefore, the eigenvalues of the operator \hat{\mathcal{U}}(\xi) are positioned in the complex plane symmetrically with respect to the origin and the imaginary axis.

    Proof. For Dirichlet boundary conditions The Eq (3.57) is studied in (3.54) with Dirichlet boundary conditions and is satisfied for z\neq0, \, \pm 1 . Furthermore, the various cases of z are considered with Dirichlet boundary conditions.

    For z = 0 , the general solution is taken from (3.51) with Dirichlet boundary conditions and a system of linear homogeneous equations \Sigma\, {\bf B} = 0 is obtained, wherein the symbol \Sigma denotes the matrix of coefficients. It follows from the existence of the non-trivial solution which means that the determinant of the corresponding matrix of coefficients for the linear homogeneous equations is zero. We have

    \begin{equation} \det(\Sigma) = -\sin^{2}(\omega_{0})+\omega_{0}^{2} > 0, \end{equation} (3.60)

    which implies that {\bf B} = 0 and zero is not an eigenvalue.

    For z = \pm 1 , the general solutions of the system (3.48) is given in (3.52) and (3.53). By using Dirichlet boundary conditions, a system of linear homogeneous equations is obtained. Moreover, the determinant of the matrix of coefficients is zero and the non-trivial solution exists. However, we are not interested in the null space analogous to the eigenvalues \xi = \pm i .

    For Neumann boundary conditions The Eq (3.58) is calculated for Neumann boundary conditions in (3.55) and is satisfied for z\neq0, \, \pm 1 .

    For z = 0 , we consider the general solution given in (3.51) with the Neumann boundary conditions, and a system of linear homogeneous equations is obtained. By the use of the method of the determinant, we get zero determinant of the matrix of coefficients. Further, the rank of the matrix of coefficients is two and (0, \, 0, \, 0, \, 1)^{T} and (0, \, 0, \, 1, \, 0)^{T} are two linearly independent solutions. Thus, the corresponding eigenfunctions are {\bf e}_{1} = (\cos\theta, -\sin\theta)^{T} and {\bf e}_{2} = (\sin\theta, \cos\theta)^{T} which represent the translation in x and y directions.

    For z = 1 , the general solution is given in (3.53). The determinant of the matrix of coefficients for this case is zero and the non-trivial solution exists.

    For z = -1 , the general solution is given in (3.52). The use of the Neumann boundary conditions gives the following determinant \det(\Sigma) = 2\sin^{2}(\omega_{0}) of the corresponding matrix of coefficients. Since \omega_{0}\in(0, \, 2\pi] , for \omega_{0} = \pi, \, 2\pi , a non-trivial solution exists.

    For mixed boundary conditions The Eq (3.59) is studied in (3.56) with the mixed boundary conditions and is satisfied for z\neq0, \, \pm 1 .

    For z = 0 , the general solution is given in (3.51) and the mixed boundary conditions are used to get a system of linear homogenous equations. The determinant of the corresponding system is \det(\Sigma) = 1 > 0 , and is not an eigenvalue of \hat{\mathcal{U}}(\xi) .

    For z = 1 , the general solution is given in (3.53) and the determinant of the matrix of coefficients for this case is \det(\Sigma) = -\cos 2(\omega_{0}) .

    For z = -1 , the general solution is given in (3.52) and the determinant of the matrix of coefficients is \det(\Sigma) = -\cos \omega_{0}-\cos 2\omega_{0}+1 .

    Consequently, z = \pm1 , are the eigenvalues of the corresponding problem if the corresponding determinants are equal to zero.

    Let ({\bf u}, q)\in W^{1, \, 2}(\mathcal{D})^{2}\times L^{2}(\mathcal{D}) be the unique weak solution of the stationary Stokes problem. It follows from the theory of Kondrat \acute{i} ev [24] that the pair ({\bf u}, q)\in \mathcal {W}_{\gamma+1}^{2, \, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\gamma+1}^{1, \, 2}(\mathcal{D}) , where \gamma is a small positive real number. To obtain further qualitative regularity results, the theory proposed by [39] is employed and we will analyze that the weak solution ({\bf u}, q)\in \mathcal {W}_{\gamma+1}^{2, \, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\gamma+1}^{1, \, 2}(\mathcal{D}) .

    Firstly, the case of L^{2} -data is considered for the direct consequences of Theorem 3 and the observations of Section 3. Let \omega_{0DN} denote the maximal angle from the set of all angles of such corner points wherein types of boundary conditions change. Analogously, \omega_{0DD} and \omega_{0NN} are the maximal angle of those corner points which have similar types of boundary conditions, i.e., Dirichlet-Dirichlet and Neumann-Neumann, on both adjacent sides of the corner point. If no such type of points exists or the angles \omega_{0DD}, \, \omega_{0NN} < \pi , then strongly we can set \omega_{0DD}, \, \omega_{0NN} = \pi as the minimum value.

    The following propositions hold to formulate the regularity results of the weak solution ({\bf u}, q) of the stationary Stokes system with various combinations of the boundary conditions.

    Lemma 3. Suppose that if the strip \alpha-1\leq \mathrm{Im}\, {\xi} < \epsilon, \, \alpha > 0, is free of the zeros of the equations

    \begin{equation} \hat{\mathcal{U}}_{DD}(\xi) = (i\xi)^{2}\,\sin^{2}(\omega_{0})-\sin^{2}(i\xi\omega_{0}) = 0, \,\, \xi\neq 0, \end{equation} (4.1)

    and

    \begin{equation} \hat{\mathcal{U}}_{NN}(\xi) = \sin^{2}(i\xi\omega_{0})-(i\xi)^{2}\,\sin^{2}(\omega_{0}) = 0,\,\, \xi\neq 0, \end{equation} (4.2)

    for any arbitrary small \epsilon > 0 , then the solution ({\bf u}, q)\in \mathcal {W}_{\alpha}^{2, \, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha}^{1, \, 2}(\mathcal{D}) and satisfies the following estimate

    \begin{equation} \Vert {\bf u}\Vert_{\mathcal {W}_{\alpha}^{2,\, 2}(\mathcal{D})^{2}}+\Vert q\Vert_{\mathcal {W}_{\alpha}^{1,\, 2}(\mathcal{D})}\leq c(\mathcal{D})\Vert{\bf f}\Vert_{L_{2}(\mathcal{D})^{2}}. \end{equation} (4.3)

    Proof. To show that no eigenvalue of (4.1) and (4.2) lies in the strip \alpha-1\leq \mbox{Im}\, {\xi} < \epsilon for an angle \omega_{0} , where \alpha is a small positive real number. Firstly, we are considering the case of Dirichlet boundary conditions \hat{\mathcal{U}}_{DD}(\xi) for an apex angle \omega_{0} = \frac{\pi}{2} and \omega_{0} = {\pi} , respectively, or for an any arbitrary angle \omega_{0}\in (0, \pi] . We note that no eigenvalue of (4.1) is found that lie on the line h = \mbox{Im}\, {\xi} = t for t > -1 . Consequently, (\tilde{{\bf v}}, \tilde{p})\in[\mathcal {W}_{\alpha}^{2, \, 2}{(S)}]^{2}\times [\mathcal {W}_{\alpha}^{1, \, 2}{(S)}] for a small positive real number \alpha . In addition, the singularities appear for this case at corners with an apex angle \omega_{0} is greater than \pi . (See Figure 3). In all the following graphs, the red lines indicate the pure imaginary values, while the black lines indicate the complex parts of the complex ones.

    Figure 3.  Distribution of eigenvalues for D-D and N-N boundary conditions.

    Analogously, the case of Neumann boundary conditions \hat{\mathcal{U}}_{NN}(\xi) is considered. There is no eigenvalue of (4.2) that lie on the line h = \mbox{Im}\, {\xi} = t for t > -1 , for an angle \omega_{0} = \frac{\pi}{2} and \omega_{0} = {\pi} , respectively, or for any arbitrary angle \omega_{0}\in (0, \pi] . It produces singularities when the apex angle \omega_{0} is greater than \pi . Hence, it follows that the solution ({\bf u}, q)\in \mathcal {W}_{\alpha}^{2, \, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha}^{1, \, 2}(\mathcal{D}) .

    Additionally, we have a bounded domain \mathcal{D} and the corresponding continuous imbeddings (2.9) and (2.12). Let we have

    \begin{equation*} \mathcal {W}_{\alpha}^{2,\, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha}^{1,\, 2}(\mathcal{D})\hookrightarrow \mathcal {W}_{0}^{2,\, \frac{2}{1+\alpha}}(\mathcal{D})^{2}\times \mathcal {W}_{0}^{1,\, \frac{2}{1+\alpha}}(\mathcal{D}), \end{equation*}

    and

    \begin{equation*} W^{2,\, \frac{2}{1+\alpha}}(\mathcal{D})^{2}\times W^{{1,\, \frac{2}{1+\alpha}}}(\mathcal{D})\hookrightarrow W^{2-\alpha,\, 2}(\mathcal{D})^{2}\times W^{1-\alpha,\, 2}(\mathcal{D}). \end{equation*}

    Clearly,

    \begin{equation*} \mathcal {W}_{0}^{2,\, \frac{2}{1+\alpha}}(\mathcal{D})^{2}\times \mathcal {W}_{0}^{1,\, \frac{2}{1+\alpha}}(\mathcal{D}) = W^{2,\, \frac{2}{1+\alpha}}(\mathcal{D})^{2}\times W^{1,\, \frac{2}{1+\alpha}}(\mathcal{D}). \end{equation*}

    So,

    \begin{equation*} \mathcal {W}_{\alpha}^{2,\, 2}(\mathcal{D})^{2}\times \mathcal {W}_{\alpha}^{1,\, 2}(\mathcal{D})\hookrightarrow W^{2-\alpha,\, 2}(\mathcal{D})^{2}\times W^{1-\alpha,\, 2}(\mathcal{D}). \end{equation*}

    Finally, we obtain that the weak solution of the Stokes system is

    \begin{equation*} ({\bf u}, q)\in \big[W^{2-\alpha,\, 2}(\mathcal{D})^{2}\times W^{1-\alpha,\, 2}(\mathcal{D})\big], \end{equation*}

    where \alpha is a small positive real number. Accordingly, the estimate (4.3) can be followed directly from Theorem 3.

    Analogously, for the case of mixed boundary conditions, we obtain:

    Lemma 4. Suppose that if no eigenvalues of the mixed boundary condition

    \begin{equation} \hat{\mathcal{U}}_{DN}(\xi) = (i\xi)^{2}\,\sin^{2}(\omega_{0})-\cos^{2}(i\xi\omega_{0}) = 0, \end{equation} (4.4)

    lie on the line \mathrm{Im}\, {\xi} = h = \gamma-1 , then the solution ({\bf u}, q)\in W^{2-\gamma, \, 2}(\mathcal{D})^{2}\times W^{1-\gamma, \, 2}(\mathcal{D}) , where \gamma depends on the angle \omega_{0} .

    Proof. For any arbitrary angle \omega_{0}\in(0, 2\pi) , no eigenvalues of the mixed boundary condition (4.4) lie on the line \mbox{Im}\, {\xi} = h = \gamma-1 for h\geq -\frac{1}{4} . Thus, no eigenvalues of \hat{\mathcal{U}}_{DN}(\xi) that lie in the strip -\frac{1}{4}\leq \mbox{Im}\, {\xi} < \epsilon for \gamma\geq\frac{3}{4} , where any arbitrary small \epsilon > 0 . So, we obtain that the regularity result ({\bf u}, q)\in W^{\frac{5}{4}, \, 2}(\mathcal{D})^{2}\times W^{\frac{1}{4}, \, 2}(\mathcal{D}) . Besides, the singularities appear at a corner with an apex angle greater than \frac{\pi}{4} (See Figure 4).

    Figure 4.  Distribution of eigenvalues for Dirichlet-Neumann boundary conditions.

    Respectively, for an angle \omega_{0} = \frac{\pi}{2} , we have h = \mbox{Im}\, {\xi} > -\frac{1}{2} . So, we obtain that the regularity result ({\bf u}, q)\in W^{2-\gamma, \, 2}(\mathcal{D})^{2}\times W^{1-\gamma, \, 2}(\mathcal{D}) for \gamma > \frac{1}{2} . Accordingly, a similar regularity result is obtained for an angle \omega_{0} = \pi .

    Finally, we obtain that the regularity result ({\bf u}, q)\in W^{2-\gamma, \, 2}(\mathcal{D})^{2}\times W^{1-\gamma, \, 2}(\mathcal{D}) for an arbitrary small positive number \gamma that depends on the apex angle \omega_{0} .

    It is well-known that for every {\bf f}\in V^{\ast} , where V^{\ast} is the dual space of V , a unique weak solution ({\bf u}, q)\in V\times L^{2}(\mathcal{D}) of some generalized steady Stokes problem exists.

    Therefore, the following lemma describes the regularity results for L^{p} -data.

    Lemma 5. Suppose that

    (ⅰ) for \hat{\mathcal{U}}_{DD}(\xi) and \hat{\mathcal{U}}_{NN}(\xi) boundary conditions, no-eigenvalues lie on the strip -\mu\leq \mathrm{Im}\, {\xi} < \epsilon , for 0\leq \mu < 1 , and \epsilon > 0 is an arbitrary small positive number,

    (ⅱ) for mixed boundary conditions, \xi_{0} is the only eigenvalue of \hat{\mathcal{U}}_{DN}(\xi) that lie within the strip -\mu\leq \mathrm{Im}\, {\xi} < \epsilon , for 0\leq \mu < 1 , and additionally, we suppose that this is a simple eigenvalue.

    Then for every {\bf f}\in L^\frac{2}{2-\mu}(\mathcal{D})^{2} , the weak solution ({\bf u}, q) of the stationary Stokes system is contained in \big[W^{1+\mu, \, 2}(\mathcal{D})\big]^{2}\times \big[W^{\mu, \, 2}(\mathcal{D})\big] .

    Proof. The statement can be followed directly from Theorem 3 by considering p = \frac{2}{2-\mu} , \alpha_{2} = 0 and \alpha_{1} = 1+\epsilon .

    Since {\bf f}\in L^\frac{2}{2-\mu}(\mathcal{D})^{2}\subset V^{\ast} , the similar result is obtained by applying the same process used in Lemmas 3-4.

    We consider the steady Navier-Stokes equations

    \begin{equation} \left\{\begin{array}{cll} -\nu\Delta {\bf u}+ ({\bf u}\cdot \nabla)\,{\bf u}+\nabla\, q = {\bf f} & \mbox{in} \quad\mathcal{D},\\[.5ex] {\mbox{div}}\, {\bf u} = 0 & \mbox{in} \quad \mathcal{D}, \end{array}\right. \end{equation} (5.1)

    with the homogenous mixed boundary conditions (1.3)-(1.4). If the given right-hand sides have a sufficiently small norm, then we prove by using the local diffeomorphism theorem that the Navier-Stokes equations have similar regularity results as the solution of the generalized Stokes problem near the corner points. Denote

    \begin{equation*} \mathcal{E}(\mathcal{D}) = \Big\{{\bf u}\in C^{\infty}(\overline{\mathcal{D}})^{2};\, \mbox{div}\,{\bf u} = 0, \,\overline{\mbox{supp}{\bf u}}\cap\Gamma^{0} = \emptyset\Big\}. \end{equation*}

    Additionally, we denote H and V are the closures of \mathcal{E}(\mathcal{D}) equipped with the norms of L^{2}(\mathcal{D})^{2} and W^{1, \, 2}(\mathcal{D}) . Recall that V and H are the Hilbert spaces and their scalar products are given in (2.1).

    Definition 2. Let {\bf f}\in V^{\ast} . {\bf u} is called the weak solution of the problem (5.1) with the homogenous mixed boundary conditions (1.3)-(1.4), if {\bf u}\in V and satisfies

    \begin{equation} \begin{split} \nu\big(({\bf u}, {\bf v})\big)+b\big({\bf u}, {\bf u}, {\bf v}\big) = \big({\bf f}, {\bf v}\big) \quad \forall {\bf v}\in V. \end{split} \end{equation} (5.2)

    Further, b\big({\bf u}, {\bf v}, {\bf w}\big) describes the trilinear continuous form for every {\bf u}, {\bf v}, {\bf w}\in V by

    \begin{equation} b\big({\bf u}, {\bf v}, {\bf w} \big) = \int_{\mathcal{D}}{u}_{j}\cdot\frac{\partial{{v}_{i}}}{\partial x_{j}}\cdot {w}_{i}\,d{\mbox{x}}. \end{equation} (5.3)

    Definition 3. Let H represent the closure of \mathcal{E}(\mathcal{D}) in the norm L^{2}(\mathcal{D})^{2} and describe the Banach space

    \begin{equation*} \mathfrak{M} = \Big\{{\bf u};\, \mathrm{there\, exists} \,{\bf f}\in H\, \mathrm{such\, that} \,\nu\big(({\bf u}, {\bf v})\big) = \big({\bf f}, {\bf v}\big)\, \mathrm{for\, all}\, {\bf v}\in V\Big\}. \end{equation*}

    Note that, \mathfrak{M}\hookrightarrow \hookrightarrow L^{\infty}(\mathcal{D}) , \mathfrak{M}\hookrightarrow \hookrightarrow V and V\hookrightarrow \hookrightarrow H .

    For the solvability of the problem (5.1), the subsequent theorem is considered which is known as the local diffeomorphism theorem (see [9]).

    Theorem 5. Let \mathcal{M} be a mapping from \mathcal{X} into \mathcal{Y} which belongs to C^{1} in some neighbourhood \mathrm{W} of point v_{0}\in \mathcal{X} , where \mathcal{X} and \mathcal{Y} are Banach spaces. If the Fr \acute{e} chet derivative \acute{\mathcal{M}}(v_{0}): X \rightarrow Y is continuous, one-to-one and onto \mathcal{Y} , then there exists a neighbourhood \mathrm{U} of point v_{0} such that \mathrm{U}\subset \mathrm{W} and a neighbourhood \mathrm{V} of point \mathcal{M}(v_{0}) such that \mathrm{V}\subset \mathcal{Y} . So, the mapping \mathcal{M} is one-to-one from \mathrm{W} onto \mathrm{V} .

    Let {\bf h}\in H . Then the Lax-Milgram theorem and the Lemmas 3-4 yields that there exists a uniquely determined {\bf w}\in \mathfrak{M} , such as

    \begin{equation} \nu\big(({\bf w}, {\bf v})\big) = \big({\bf h}, {\bf v}\big) \quad \forall {\bf v}\in V. \end{equation} (5.4)

    Now, the operator \mathcal{Q}:\mathfrak{M}\rightarrow H is described by

    \begin{equation} (\mathcal{Q}({\bf w}), {\bf v}) = \nu\big(({\bf w}, {\bf v})\big) \quad \forall {\bf v}\in V. \end{equation} (5.5)

    Note that the mapping \mathcal{Q} is one-to-one. Moreover, we define an operator \mathcal{R}:\mathfrak{M}\rightarrow H which is given as

    \begin{equation} (\mathcal{R}({\bf u}), {\bf v}) = (\mathcal{Q}({\bf w}), {\bf v})+b\big({\bf u}, {\bf u}, {\bf v}\big)\quad \forall {\bf v}\in V. \end{equation} (5.6)

    Further, the pertinent problem (5.1) can be considered as the single operator equation \mathcal{R}({\bf u}) = {\bf f} .

    Let {\bf u} be a fixed point in \mathfrak{M} , and \mathcal{Z}_{\bf u}:\mathfrak{M}\rightarrow H be a linear operator which is described as follows

    \begin{equation} \begin{split} (\mathcal{Z}_{\bf u}({\bf w}), {\bf v})& = b\big({\bf u}, {\bf w}, {\bf v}\big)+b\big({\bf w}, {\bf u}, {\bf v}\big),\\[.5ex] & = \big(({\bf u}\nabla){\bf w}, {\bf v}\big)+\big(({\bf w}\nabla){\bf u}, {\bf v}\big)\quad \forall {\bf v}\in V. \end{split} \end{equation} (5.7)

    Lemma 6. Let {\bf u} be some arbitrary fixed element in \mathfrak{M} . The operator \mathcal{B}_{\bf u} is given by

    \begin{equation} (\mathcal{B}_{\bf u}({\bf w}), {\bf v}) = (\mathcal{Q}({\bf w}), {\bf v})+(\mathcal{Z}_{\bf u}({\bf w}), {\bf v})\quad \forall {\bf v}\in V, \end{equation} (5.8)

    is the Fr \acute{e} chet derivative of \mathcal{R} at the point {\bf u} and \mathcal{B}_{\bf u}\in C(\mathfrak{M}) .

    Proof. Since

    \begin{equation} \Vert\mathcal{R}({\bf u}+{\bf w})-\mathcal{R}({\bf u})- \mathcal{B}_{\bf u}({\bf w})\Vert_{H} = \Vert b\big({\bf w}, {\bf w}, \cdot\big)\Vert_{H}, \end{equation} (5.9)

    and

    \begin{equation} \Vert b\big({\bf w}, {\bf w}, \cdot\big)\Vert_{H}\leq C\Vert {\bf w}\Vert_{\mathfrak{M}}^2\quad \mbox{holds}\quad\mbox{for all}\quad {\bf w}\in\mathfrak{M}. \end{equation} (5.10)

    We get

    \begin{equation} \lim\limits_{\Vert {\bf w}\Vert_{\mathfrak{M}}\rightarrow0}\frac{\Vert\mathcal{R}({\bf u}+{\bf w})-\mathcal{R}({\bf u})- \mathcal{B}_{\bf u}({\bf w})\Vert_{H}}{\Vert {\bf w}\Vert_{\mathfrak{M}}}\leq \lim\limits_{\Vert {\bf w}\Vert_{\mathfrak{M}}\rightarrow0}\, C{\Vert {\bf w}\Vert_{\mathfrak{M}}} = 0, \end{equation} (5.11)

    the smoothness of \mathcal{B}_{\bf u}\in C(\mathfrak{M}) is obvious.

    Lemma 7. Let {\bf u} = 0 . Then \mathcal{B}_{\bf u} = \mathcal{B}_{0} = \mathcal{Q} is one-to-one.

    The more information to prove that \mathcal{B}_{\bf u} is one-to-one and onto H , we refer ([40], Theorem 5.5F) and [6].

    Theorem 6. Let \mathcal{P}_1 be a continuous one-to-one operator from \mathcal{X} onto \mathcal{Y} , and \mathcal{P}_2 be a compact linear operator from \mathcal{X} into \mathcal{Y} , where \mathcal{X} and \mathcal{Y} are Banach spaces. Therefore, the subsequent statements are equivalent:

    1. \mathcal{P}_1+\mathcal{P}_2 is one-to-one;

    2. \mathcal{P}_1+\mathcal{P}_2 is onto \mathcal{Y} .

    Theorem 7. Let {\bf f}\in H , the norm of {\bf f} is sufficiently small. Then a uniquely determined {\bf u}\in \mathfrak{M} exists, such that

    \begin{equation} \begin{split} \nu\big(({\bf u}, {\bf v})\big)+b\big({\bf u}, {\bf u}, {\bf v}\big) = \big({\bf f}, {\bf v}\big) \quad \forall {\bf v}\in V. \end{split} \end{equation} (5.12)

    Proof. By (5.8), \mathcal{B}_{\bf u} is the sum of operators \mathcal{Q} and \mathcal{Z}_{\bf u} . It is noted from the above results that \mathcal{Q}:\mathfrak{M}\rightarrow H is the one-to-one operator and onto H , and \mathcal{Z}_{\bf u} is a compact operator. Furthermore, from Lemma 7, \mathcal{B}_{\bf u} is one-to-one. These facts and Theorem 6 produces that \mathcal{B}_{\bf u} is onto H . Hence, the Theorem 5 yields a unique solution.

    In this article, we have studied the boundary singularities and regularity of the weak solution of the mixed boundary value problem for the stationary Stokes and Navier-Stokes system in a two-dimensional non-smooth domain with corner points and points at which the type of boundary conditions change. The solvability of the considered boundary value problem has been analyzed in the context of the weighted Sobolev spaces with Kondrat \acute{i} ev type weights and the regularity theorems are formulated. To compute the singular terms for various boundary conditions, the complex Fourier transform has been used to obtain the generalized form of the boundary eigenvalue problem for the stationary Stokes system. The emerging eigenvalues and eigenfunctions produce singular terms, which permits us to evaluate the optimal regularity of the corresponding weak solution of the stationary Stokes system.

    The main regularity and expansion theorem for the stationary Stokes system is presented in Theorem 3. We have discussed the regularity results of the corresponding boundary value problem for the case of L^2 and L^p -data which are the direct consequences of Theorem 3. It is seen for the case of Dirichlet and Neumann boundary conditions that if the domain \mathcal{D} has reentrant corners (\omega_i > \pi: i = 1, 2, ...N) , then the weak solution ({\bf u}, q) of the considered problem produces singularities. On the other hand, for the case of mixed conditions, the singularities appear at corners with (\omega_i > \frac{\pi}{4}: i = 1, 2, ...N) . Moreover, it is observed that if singularities exist, then splitting the solution into a singular part which defines a linear combination of explicit model singularity functions s_m for the Stokes operator with corresponding unknown coefficients C_m and a regular part that belongs to H^2\times H^1 . Finally, it is proved that the weak solution ({\bf u}, q) of the underlying boundary value problem belongs to W^{2-\gamma, \, 2}\mathcal{(D)}^{2}\times W^{1-\gamma, \, 2}\mathcal{(D)} , where \gamma is an arbitrarily small positive real number that depends on the apex angle \omega_{0} .

    Additionally, we have extended the obtained results for the Stokes system for the non-linear Navier-Stokes system. We have proved this by using the local diffeomorphism theorem that the solution of the Navier-Stokes equations has similar regularity results as the solution of the generalized Stokes problem near the corner points if the given body force has a sufficiently small norm. To prove this, an operator \mathcal{R} relating to the Navier-Stokes equations is defined and has shown that it is Fréchet differentiable at the point {\bf u} = 0 . Furthermore, the Fréchet derivative of \mathcal{R} at the point {\bf u} is agreed with the Stokes problem.

    Presently, the Stokes and Navier-Stokes equations with the Navier-slip boundary conditions and the free-boundary problems in domains with corners have very interesting phenomena. The issues regarding their existence and regularity are considered for smooth domains but theoretical results for the corner singularity decomposition are still not obtained. Therefore, these issues are numerically interesting. In future works, it is important to show the unique existence of the approximations for the regular parts and coefficients, and to derive their error estimates. On the other hand, it is also observed that the non-stationary compressible Stokes and Navier-Stokes equations on polygonal domains could be considered.

    The author declare that no conflict of interest exist.



    [1] R. A. Adams, J. J. Fournier. Sobolev spaces, Vol. 140, Academic Press, 2003.
    [2] S. Agmon, A. Douglis, L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions I, Commun. Pure Appl. Math., 12 (1959), 623–727. doi: 10.1002/cpa.3160120405
    [3] S. Agmon, A. Douglis, L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions II, Commun. Pure Appl. Math., 17 (1964), 35–92. doi: 10.1002/cpa.3160170104
    [4] Y. N. Anjam, Singularities and regularity of stationary Stokes and Navier-Stokes equations on polygonal domains and their treatments, AIMS Mathematics, 5 (2020), 440–466. doi: 10.3934/math.2020030
    [5] I. Babu\check{s}ka, B. Q. Guo, J. E. Osborn, Regularity and numerical solution of eigenvalue problems with piecewise analytic data, SIAM J. Num. Comp., 26 (1989), 1534–1560. doi: 10.1137/0726090
    [6] M. Bene\check{{s}}, P. Ku\check{{c}}era, Solutions of the Navier-Stokes equations with various types of boundary conditions, Arch. Math., 98 (2012), 487–497. doi: 10.1007/s00013-012-0387-x
    [7] H. J. Choi, J. R. Kweon, The stationary Navier-Stokes system with no-slip boundary condition on polygons: corner singularity and regularity, Commun. Part. Diff. Eq., 38 (2013), 1235–1255. doi: 10.1080/03605302.2012.752386
    [8] N. Chorfi, Geometric singularities of the Stokes problem, Abstr. Appl. Anal., 2014 (2014), 1–8.
    [9] P. G. Ciarlet, Mathematical elasticity (studies in mathematics and its applications), Elsevier, 1988.
    [10] M. Dauge, Stationary Stokes and Navier-Stokes systems on two or three-dimensional domains with corners. part I. linearized equations, SIAM J. Math. Anal., 20 (1989), 74–97. doi: 10.1137/0520006
    [11] M. Dauge, Elliptic boundary value problems on corner domains: smoothness and asymptotics of solutions, Vol. 1341, 2006.
    [12] M. Durand, Singularities in elliptic problems. In: Singularities and constructive methods for their treatment, Springer, Berlin, Heidelberg, 1985.
    [13] J. Fabricius, Stokes flow with kinematic and dynamic boundary conditions, Quart. Appl. Math., 77 (2019), 525–544. doi: 10.1090/qam/1534
    [14] S. Fucik, O. John, A. Kufner, Function spaces, Springer, Netherlands, 1977.
    [15] V. Girault, P. A. Raviart, Finite element methods for Navier-Stokes equations: theory and algorithms, Vol. 5, Springer Science and Business Media, 2012.
    [16] P. Grisvard, Behavior of the solutions of an elliptic boundary value problem in a polygonal or polyhedral domain, Numerical solution of partial differential equations, III, (1976), 207–274.
    [17] P. Grisvard, Elliptic problems in nonsmooth domains, Vol. 2, 2–2, Pitman Advanced Pub. Program, Boston, 1985.
    [18] B. Q. Guo, I. Babu\check{s}ka, On the regularity of elasticity problems with piecewise analytic data, Adv. Appl. Math., 14 (1993), 307–347. doi: 10.1006/aama.1993.1016
    [19] B. Q. Guo, C. Schwab, Analytic regularity of Stokes flow on polygonal domains in countably weighted Sobolev spaces, J. Comput. Appl. Math., 190 (2006), 487–519. doi: 10.1016/j.cam.2005.02.018
    [20] Y. Hou, S. Pei, On the weak solutions to steady Navier-Stokes equations with mixed boundary conditions, Math. Zeit., 291 (2019), 47–54. doi: 10.1007/s00209-018-2072-7
    [21] S. Itoh, N. Tanaka, A. Tani, On some boundary value problem for the stokes equations in an infinite sector, Anal. Appl., 4 (2006), 357–375. doi: 10.1142/S0219530506000826
    [22] V. V. Katrakhov, S. V. Kiselevskaya, A singular elliptic boundary value problem in domains with corner points. I. Function spaces, Diff. Eq., 42 (2006), 395–403.
    [23] R. B. Kellogg, J. E. Osborn, A regularity result for the Stokes problem in a convex polygon, J. Func. Anal., 21 (1976), 397–431. doi: 10.1016/0022-1236(76)90035-5
    [24] V. A. Kondrat\acute{i}ev, Boundary value problems for elliptic equations in domains with conical or angular points, Trudy Mos. Matem. Obsh., 16 (1967), 209–292.
    [25] V. A. Kondrat\acute{i}ev, O. A. Oleinik, Boundary-value problems for partial differential equations in non-smooth domains, Russ. Math. Surveys, 38 (1983), 1–86.
    [26] V. A. Kozlov, V. G. Maz'ya, J. Rossmann, Elliptic boundary value problems in domains with point singularities, American Mathematical Society, 1997.
    [27] V. A. Kozlov, V. G. Maz'ya, J. Rossmann, Spectral problems associated with corner singularities of solutions to elliptic equations, Vol. 85, American Mathematical Society, 2001.
    [28] S. Kr\breve{a}cmar, J. Neustupa, A weak solvability of a steady variational inequality of the Navier-Stokes type with mixed boundary conditions, Nonlinear Anal-Theor, 47 (2001), 4169–4180. doi: 10.1016/S0362-546X(01)00534-X
    [29] S. G. Krejn, V. P. Trofimov, Holomorphic operator-valued functions of several complex variables, Funct. Anal. i Prilo\breve{z}en., 3 (1969), 85–86.
    [30] P. Ku\check{c}era, Basic properties of solution of the non-steady Navier-Stokes equations with mixed boundary conditions in a bounded domain, Annali dell' Univ. di Ferrara. Sezione VII. Scie. Matem., 55 (2009), 289–308. doi: 10.1007/s11565-009-0082-4
    [31] J. R. Kweon, Regularity of solutions for the Navier-Stokes system of incompressible flows on a polygon, J. Diff. Eq., 235 (2007), 166–198. doi: 10.1016/j.jde.2006.12.008
    [32] J. R. Kweon, Edge singular behavior for the heat equation on polyhedral cylinders in R^{3}, Potential Anal., 38 (2013), 589–610. doi: 10.1007/s11118-012-9288-7
    [33] J. R. Kweon, The compressible Stokes flows with no-slip boundary condition on non-convex polygons, J. Math. Fluid Mech., 19 (2017), 47–57. doi: 10.1007/s00021-016-0264-7
    [34] O. S. Kwon, J. R. Kweon, Compressible Navier-Stokes equations in a polyhedral cylinder with inflow boundary condition, J. Math. Fluid Mech., 20 (2018), 581–601. doi: 10.1007/s00021-017-0336-3
    [35] I. Lasiecka, K. Szulc, A. \dot{Z}ochowski, Boundary control of small solutions to fluid-structure interactions arising in coupling of elasticity with Navier-Stokes equation under mixed boundary conditions, Nonlinear Anal-Real, 44 (2018), 54–85. doi: 10.1016/j.nonrwa.2018.04.004
    [36] V. Maza, J. Rossmann, Mixed boundary value problems for the stationary Navier-Stokes system in polyhedral domains, Arch. Rati. Mech. Anal., 194 (2009), 669–712. doi: 10.1007/s00205-008-0171-z
    [37] S. A. Nazarov, A. Novotny, K. Pileckas, On steady compressible Navier-Stokes equations in plane domains with corners, Math. Ann., 304 (1996), 121–150. doi: 10.1007/BF01446288
    [38] M. Orlt, A. M. S\ddot{a}ndig, Regularity of viscous Navier-Stokes flows in nonsmooth domains. Boundary value problems and integral equations in nonsmooth domains, Lecture Notes in Pure and Applied Mathematics, 167 (1995), 185–201.
    [39] A. M. S\ddot{a}ndig, Some applications of weighted Sobolev spaces, Vieweg+Teubner Verlag, 1987.
    [40] A. E. Taylor, Introduction to functional analysis, John Wiley and Sons, London, 1958.
    [41] R. Temam, Navier-Stokes equations: Theory and numerical analysis, Elsevier North-Holland, 1979.
  • This article has been cited by:

    1. Viktor A. Rukavishnikov, Alexey V. Rukavishnikov, On the Properties of Operators of the Stokes Problem with Corner Singularity in Nonsymmetric Variational Formulation, 2022, 10, 2227-7390, 889, 10.3390/math10060889
  • Reader Comments
  • © 2021 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(2689) PDF downloads(111) Cited by(1)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog