Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

On the Hamiltonian and geometric structure of Langmuir circulation

  • The Craik-Leibovich equation (CL) serves as the theoretical model for Langmuir circulation. We show that the CL equation can be reduced to the dual space of a certain Lie algebra central extension. On this space, the CL equation can be rewritten as a Hamiltonian equation corresponding to the kinetic energy. Additionally, we provide an explanation of the appearance of this central extension structure through an averaging theory for Langmuir circulation. Lastly, we prove a stability theorem for two-dimensional steady flows of the CL equation. The paper also contains two examples of stable steady CL flows.

    Citation: Cheng Yang. On the Hamiltonian and geometric structure of Langmuir circulation[J]. Communications in Analysis and Mechanics, 2023, 15(2): 58-69. doi: 10.3934/cam.2023004

    Related Papers:

    [1] Sergey A. Rashkovskiy . Quantization of Hamiltonian and non-Hamiltonian systems. Communications in Analysis and Mechanics, 2023, 15(2): 267-288. doi: 10.3934/cam.2023014
    [2] Long Ju, Jian Zhou, Yufeng Zhang . Conservation laws analysis of nonlinear partial differential equations and their linear soliton solutions and Hamiltonian structures. Communications in Analysis and Mechanics, 2023, 15(2): 24-49. doi: 10.3934/cam.2023002
    [3] Tianfang Wang, Wen Zhang . Existence and concentration of homoclinic orbits for first order Hamiltonian systems. Communications in Analysis and Mechanics, 2024, 16(1): 121-146. doi: 10.3934/cam.2024006
    [4] Andrea Brugnoli, Ghislain Haine, Denis Matignon . Stokes-Dirac structures for distributed parameter port-Hamiltonian systems: An analytical viewpoint. Communications in Analysis and Mechanics, 2023, 15(3): 362-387. doi: 10.3934/cam.2023018
    [5] Panyu Deng, Jun Zheng, Guchuan Zhu . Well-posedness and stability for a nonlinear Euler-Bernoulli beam equation. Communications in Analysis and Mechanics, 2024, 16(1): 193-216. doi: 10.3934/cam.2024009
    [6] Richard Cushman . Normalization and reduction of the Stark Hamiltonian. Communications in Analysis and Mechanics, 2023, 15(3): 457-469. doi: 10.3934/cam.2023022
    [7] Wang Xiao, Xuehua Yang, Ziyi Zhou . Pointwise-in-time $ \alpha $-robust error estimate of the ADI difference scheme for three-dimensional fractional subdiffusion equations with variable coefficients. Communications in Analysis and Mechanics, 2024, 16(1): 53-70. doi: 10.3934/cam.2024003
    [8] Henryk Żoła̧dek . An example in Hamiltonian dynamics. Communications in Analysis and Mechanics, 2024, 16(2): 431-447. doi: 10.3934/cam.2024020
    [9] Anthony Bloch, Marta Farré Puiggalí, David Martín de Diego . Metriplectic Euler-Poincaré equations: smooth and discrete dynamics. Communications in Analysis and Mechanics, 2024, 16(4): 910-927. doi: 10.3934/cam.2024040
    [10] Shuyue Ma, Jiawei Sun, Huimin Yu . Global existence and stability of temporal periodic solution to non-isentropic compressible Euler equations with a source term. Communications in Analysis and Mechanics, 2023, 15(2): 245-266. doi: 10.3934/cam.2023013
  • The Craik-Leibovich equation (CL) serves as the theoretical model for Langmuir circulation. We show that the CL equation can be reduced to the dual space of a certain Lie algebra central extension. On this space, the CL equation can be rewritten as a Hamiltonian equation corresponding to the kinetic energy. Additionally, we provide an explanation of the appearance of this central extension structure through an averaging theory for Langmuir circulation. Lastly, we prove a stability theorem for two-dimensional steady flows of the CL equation. The paper also contains two examples of stable steady CL flows.



    In 1938, Langmuir [1] reported his observation of windrows of seaweeds in the Sargasso Sea. When a wind blows over a water surface steadily, small objects, such as seaweeds or bubbles, floating on the water will align with the wind direction. This is called Langmuir circulation. Since its discovery, this fascinating phenomenon has sparked a lot of research, both experimental and theoretical. In 1976, Craik and Leibovich [2] derived the CL equation as a theoretical model for Langmuir circulation.

    According to the Craik-Leibovich theory, Langmuir circulation arises due to the interaction between the flow and the fast oscillating fluid surface. The corresponding averaged system in a 3-dimensional domain with fixed boundaries is the CL equation:

    {vt+(v,)v+curlv×Vs=p,(v+Vs)n=0, (1.1)

    where Vs is obtained by averaging the oscillating surface and is referred to as the Stokes drift, while n represents the outer normal vector of the boundary.

    The Hamiltonian formulation of the classical CL equation was studied in the works of Holm [3] and Vladimirov [4]. In the present paper, we will discuss the Hamiltonian structure of the CL equation from a geometric point of view based on [5]. It turns out that, on the dual space of a certain Lie algebra central extension, the CL equation can be rewritten as a Hamiltonian equation corresponding to the kinetic energy. Then, one can generalize the CL equation to any Riemannian manifold with boundaries.

    The equation (1.1) was first derived by Craik and Leibovich using the averaging method. Later, Vladimirov and his coauthors [4] developed a multiscale averaging method to study Langmuir circulation. In [5], the author carried out a general averaging theory on a principal bundle related to incompressible free boundary hydrodynamics problems to obtain the CL equation. This perturbation theory also clarifies the origin of the Stokes drift.

    Central extension structures also appear in other mathematical equations. We would like to mention the superconductivity equation and the βplane equation. For details of these examples, please refer to Section 2.

    The rest of the paper is organized as follows. Some preliminaries about the Euler equation and central extensions of Lie algebras are given in Section 2. In Section 3, we first generalize the CL equation to any Riemannian manifold with boundaries. Then, we prove the central extension structure of the CL equation along with its Hamiltonian structure, and we obtain a broad class of invariant functionals. The averaging theory for Langmuir circulation which explains the appearance of the central extension structure is presented in Section 4. Lastly, Section 5 discusses the stability of two-dimensional steady flows of the Craik-Leibovich equation.

    In his seminal paper [6], Arnold studied the geodesics on Lie groups with one-sided invariant Riemannian metric. He showed that the geodesic equation can be reduced to the Euler equation on the dual space of Lie algebra.

    More specifically, consider a Lie group G which can be finite or infinite dimensional. Let g be its Lie algebra with g being the dual of g. The geodesic equation can be formulated as a Hamiltonian equation on TG with a kinetic Hamiltonian. Use the inertia operator K:gg, and one can write the kinetic Hamiltonian as a function on g: H(μ)=12K1μ,μ for μg. The corresponding Hamiltonian equation on g is given by

    dμdt=adK1μμ. (2.1)

    Definition 2.1. Equation (2.1) is called the Euler equation.

    One can also write the kinetic energy E(v)=12v,Kv on the Lie algebra g. Then, by the right (or left) translation, one obtains a right (or left)-invariant metric on the Lie group G. Arnold proved in [6] that for the volume-preserving diffeomorphism group equipped with the right-invariant L2-metric, equation (2.1) is the incompressible Euler equation in hydrodynamics.

    Let h be a Lie algebra and H be a vector space. One can define a bilinear, antisymmetric map: ˆω:h×hH. If it satisfies the identity

    ˆω([u,v],w)+ˆω([v,w],u)+ˆω([w,u],v)=0,u,v,wh,

    then ˆω is a Lie algebra 2-cocycle. One can use the 2-cocycle to define a new bracket: for u,vh and a,bH,

    [(u,a),(v,b)]=([u,v],ˆω(u,v)).

    Definition 2.2. A new Lie algebra ˆh=hH with the Lie bracket [,] is called a central extension of h by H.

    Example 2.1. Consider a 3-dimensional compact manifold M, dvol is a volume form on it, and B is a divergence-free vector field. It is known that the Lie algebra of the volume-preserving diffeomorphism group Diffdvol(M) is formed by all the divergence-free vector fields [7]. We denote the Lie algebra by Xdvol(M). If M has boundaries, then the elements in Xdvol(M) should be tangent to the boundaries. One can define a 2-cocycle ˆωB by

    ˆωB(v,w)=M(iviwiBdvol)dvolforv,wXdvol(M).

    Let ˆXdvol(M) be the central extension of Xdvol(M) through the 2-cocycle ˆωB(v,w), and let ˆXdvol(M) be its dual space. Then, the Euler equation on ˆXdvol(M) coincides with the 3-dimensional superconductivity equation [8,9,10]

    ut+uu+u×B=p,u=0. (2.2)

    Example 2.2. Another interesting equation appears in the study of rotating 2D fluids. Let u be the velocity field, and in the 2D case, its vorticity ω is a function. The stream function ϕ satisfies u=(yϕ,xϕ). Then, the motion of fluids is governed by the β-plane equation:

    tω={ϕ,ω}βxϕ, (2.3)

    where βR, and the last term in the equation is the effect of the Coriolis force.

    All the symplectic vector fields on a 2-dimensional manifold D form a Lie algebra Xsymp(D). Zeitlin [11] considered a certain central extension ˆXsymp(D), whose dual is ˆXsymp(D), and demonstrated that equation (2.3) is the Euler equation on ˆXsymp(D).

    In this section, we will define a central extension related to the geometric structure of Langmuir circulation.

    Definition 3.1. On a Lie algebra g, we first define the shifted 2-cocycle ˆωVs:g×gR for a fixed vector Vsg by

    ˆωVs(u,v)=aduK(Vs),v, (3.1)

    where u,vg, and K is a map from g to g.

    Remark 3.1. Because aduK(Vs),v=K(Vs),[u,v], the shifted 2-cocycle ˆωVs is a 2-coboundary.

    By means of the shifted 2-cocycle ˆωVs, one can get a new Lie algebra ˆgVs, which is the central extension of g by R. We derive the Euler equation on its dual ˆgVs.

    Proposition 3.1. The Euler equation on ˆgVs is

    ddtμ=adK1μ{μaK(Vs)}. (3.2)

    Proof. Since

    ad(X,b)(μ,a),(Y,c)=(μ,a),([X,Y],ˆωVs(X,Y))=μ,[X,Y]+aˆωVs(X,Y)=adXμ,YaadXK(Vs),Y=adXμaadXK(Vs),Y,

    take X=K1μ to get the Euler equation on ˆgVs

    ddtμ=adK1μ{μaK(Vs)}.

    Remark 3.2. As mentioned in Section 2, the Euler equation is Hamiltonian, and the corresponding Hamiltonian function is H(μ)=12K1μ,μ.

    Let M be an n-dimensional Riemannian manifold with boundary M. The group Diffdvol(M) is the group of all volume-preserving diffeomorphisms on M. Its Lie algebra Xdvol(M) consists of all the divergence-free vector fields on M that are tangent to the boundary M. The dual space Xdvol(M) of the Lie algebra is the space of 1-forms on M modulo the exact 1-forms. We now introduce the generalized CL equation.

    Theorem 3.1. ([5]) The generalized CL equation on the space Xdvol(M) is

    ddt[u]=Lv+Vs[u], (3.3)

    where v+VsXdvol(M), and [u]=[vb]Xdvol(M).

    Proof. Let u=vb, and the equation (3.3) becomes

    ddtu=Lv+Vsu+df.

    From the identities

    Lv(vb)=(vv)b+12dv,v

    and

    (curlvVs)=iVsicurlvμ=iVsdvb=LVsu,

    we obtain an equation which can be seen as the generalization of the CL equation on M,

    vt+vv+p=Vs×curlv. (3.4)

    Also, the condition v+VsXdvol(M) gives us the boundary condition. In dimension 3, the equation (3.4) is the classical CL equation.

    Remark 3.3. Note that velocity fields v and Vs do not have to be elements of Xdvol(M), but their sum v+VS is an element of Xdvol(M), which is the boundary condition.

    We show that the CL equation (3.3) is also Hamiltonian on the dual space Xdvol(M).

    Corollary 3.1. Consider the function H(u)=12([u+Vbs],K1[u+Vbs]) defined on Xdvol(M). The Hamiltonian equation for this function is the CL equation (3.3).

    Proof. The Hamiltonian equation on Xdvol(M) is

    dds[u]=LδHδ[u][u].

    For the Hamiltonian function H=12([u+Vbs],K1[u+Vbs]), the functional derivative is δHδ[u]=K1[u+Vbs])=v+Vs. Therefore, we obtain equation (3.3).

    Now, set [u]=[u+Vbs], and equation (3.3) becomes

    ddt[u]=LK1[u]{[u][Vbs]}. (3.5)

    Consider a 2-cocycle ˆωVs on Xdvol(M) defined by

    ˆωVs(u,v)=LuVbs,v.

    Then, using it, one can define a Lie algebra ˆXdvol(M), which is a central extension of Xdvol(M) by R.

    Theorem 3.2. ([5]) Equation (3.5) is the Euler equation on ˆXdvol(M).

    Proof. Let a=1 in equation (3.2), and we obtain equation (3.5).

    Remark 3.4. One can also express the 2-cocycle ˆωVs as the integral of a 2-form dVbs on M. Indeed,

    MdVbs(u,v)dvol=MiudVbs,vdvol=MLuVbs,vdvol+MdiuVbs,vdvol=LuVbs,v=ˆωVs(u,v),

    where the equation MdiuVbs,vdvol=0 holds since vXdvol(M).

    Next, we present the first integrals of the CL equation. The following corollary follows immediately from the fact that the action of the group Diffdvol(M) can be viewed as a change of variables that preserves dvol on M.

    Corollary 3.2. (1) For a (2k+1)-dimensional manifold M, the first integral of Equation (3.3) is I([u])=Mu(du)k.

    (2) For a 2k-dimensional manifold M and an arbitrary function h, the first integral of Equation (3.3) is

    Ih([u])=Mh((du)kdvol)dvol.

    Remark 3.5. Note that in dimension 3 (i.e., k=1), the integral I([u])=Mudu is the Eulerian mean helicity discussed in [3].

    Remark 3.6. The moment [u]=[vb] is transferred by the flow corresponding to the velocity field v+Vs (Kelvin's theorem for the CL equation). For the CL equation, we provide two equivalent definitions of isovorticed fields, corresponding to equation (3.3) and (3.5), respectively.

    Definition 3.2. For equation (3.3), two vector fields u1 and u2 are isovorticed if curlu1 can be transferred to curlu2 by a volume-preserving diffeomorphism, and they satisfy the same boundary condition: (u1+Vs)n=(u2+Vs)n=0.

    Definition 3.2'. For equation (3.5), two vector fields u1,u2Xdvol(M) are isovorticed if curl(u1Vs) can be transferred to curl(u2Vs) by a volume-preserving diffeomorphism.

    The central extension structure of the CL equation arises from the process of averaging. Specifically, let MH be a circle bundle with the base H being a Lie group. Consider a Hamiltonian function H(x,y) for (x,y)TxM. After averaging with respect to the circle action S1, we obtain the averaged system on the reduced manifold, which is TH equipped with a reduced symplectic form ωσ. The averaged system is Hamiltonian. We denote the averaged Hamiltonian function by ˉH(ˉx,ˉy) for (ˉx,ˉy)TˉxH. The reduced symplectic form ωσ is

    ωσ=ω0σπHdˉa,

    where ω0=dˉxdˉy, and πH is the pullback of the projection πH:THH. Here, σR is a value of the momentum map of the circle action. The 1-form ˉa is the averaged connection 1-form on H (see Theorem 6.7 in [12] for more details).

    Furthermore, if the original Hamiltonian H(x,y) is H-invariant, then we can perform another reduction on the cotangent bundle (TH,ωσ) ("the reduction by stages"). Let h be the dual of the Lie algebra h of H, and then this reduction gives us a certain Poisson structure on h, given by (see Theorem 7.2.1 in [13])

    {f,g}σ(m)=m,[δfδm,δgδm]cσ(e)(δfδm,δgδm) (4.1)

    for mh, where f and g are smooth functions on h. The 2-form cσ on H satisfies ωσ=πHcσ, and cσ(e) represents taking the value at the identity element eH.

    For a nonzero value σ of the momentum map of the circle action, we can define a 2-cocycle c:h×hR by c:=1σcσ(e). By means of c, we obtain the central extension ˆh of the Lie algebra h. The Poisson bracket (4.1) can be regarded as the natural Poisson bracket on the dual of ˆh.

    This explains the origin of the central extension structure that appears in the Hamiltonian formulation of the CL equation.

    The vorticity equation of the incompressible CL flow corresponding to equation (3.5) is

    ωt+{v,ωcurlVs}=0, (5.1)

    where ω=curlv is the vorticity. Note that the velocity field v is divergence-free, and (vVs) satisfies the CL equation (1.1). Therefore, the steady solution of equation (5.1) is given by

    {v,curl(vVs)}=0. (5.2)

    Let us consider the following variational problem:

    Problem 1. Suppose that the central extension group ^Diffdvol(M), which corresponds to the central extension of the Lie algebra Xdvol(M) described in Section 3.2, exists. Given a vector field u0Xdvol(M), we aim to find the critical points of the kinetic function K(u)=12u,u on the set S={uXdvol(M)(u,1)=Adg(u0,1),g^Diffdvol(M)}, where Ad is the group adjoint action.

    It turns out that the steady CL flows are the critical points of this variational problem.

    Theorem 5.1. The steady CL flows that satisfy equation (5.2) coincide with the critical points of variational problem 1.

    Proof. Let (u,b)Xdvol(M)R. The variation δ(v,a) of a field (v,a) under the adjoint action of (u,b) is given by

    δ(v,a)=[(u,b),(v,a)]=([u,v],Vs,[u,v]).

    Suppose vXdvol(M) is a critical point of Problem 1. Then, the first variation of E taken at v should be 0, so we have

    0=δE=(v,1),δ(v,1)=(v,1),({v,u},Vs,[u,v])=(v,1),{v,u}Vs,curl(u×v)=u,v×curlucurlVs,u×v=u,v×curlvu,v×curlVs=u,v×curl(vVs).

    So, we have {v,curl(vVs)}=0.

    Let ˆg=gR be a one-dimensional central extension of an arbitrary Lie algebra with 2-cocycle ˆω. We introduce the bilinear operation B:g×gg defined by

    [v1,v2],v3=B(v3,v1),v2, (5.3)

    where vig,i=1,2,3. Using this operation B, one can rewrite the Euler equation in its Lie algebra form [7]:

    dvdt=B(v,v), (5.4)

    where vg.

    Then, we define an operator w:gg induced from the 2-cocycle ˆω by ˆω(u,v)=w(u),v for any u,vg.

    Proposition 5.1. On the central extension ˆg, the Euler equation is given by

    dvdt=B(v,v)+aw(v). (5.5)

    Thus, the equilibrium point (ve,ae)ˆg satisfies

    B(ve,ve)+aew(ve)=0. (5.6)

    Proof. We can compute the bilinear operation ˆB:ˆg׈gˆg of the central extension ˆg as follows:

    ˆB((v3,a3),(v1,a1)),(v2,a2)=[(v1,a1),(v2,a2)],(v3,a3)=([v1,v2],ˆω(v1,v2)),(v3,a3)=[v1,v2],v3+a3ˆω(v1,v2)=B(v3,v1)+a3w(v1),v2,

    where (vi,ai)ˆg,i=1,2,3. So, the Euler equation on ˆg takes the form

    d(v,a)dt=ˆB((v,a),(v,a))

    and becomes equation (5.5). (For brevity, here we omit the second equation dadt=0.)

    The Lie algebra ˆg is foliated by the coadjoint orbits. Next, we prove a stability theorem for the equilibrium points on ˆg.

    Theorem 5.2. Assume that the equilibrium point (ve,ae)ˆg is a regular point of the coadjoint foliation. Consider a test quadratic form T(ve,ae):

    T(ve,ae)(ξ)=B(ve,ζ)+aew(ζ),B(ve,ζ)+aew(ζ)+[ζ,ve],B(ve,ζ)+aew(ζ), (5.7)

    where ξ=B(ve,ζ)+aew(ζ)g. If for all nonzero ξg we have T(ve,ae)(ξ)>0 or T(ve,ae)(ξ)<0, then the equilibrium solution (ve,ae)ˆg of equation (5.5) is Lyapunov stable.

    Proof. To prove this, we use the second variation of the kinetic function K(u)=12u,u on the leaf of this coadjoint foliation of ˆg. As shown by Arnold (see, e.g., [7]), the second variation is given by

    2δ2K(ve,ae)(ξ)=ˆB(ve,ζ),ˆB(ve,ζ)+[ζ,ve],ˆB(ve,ζ), (5.8)

    where ξ=ˆB(ve,ζ)g. Note that the quadratic form δ2K does not depend on the choice of ζ but only on ξ=ˆB(ve,ζ).

    Thanks to the computation in Proposition 5.1, we have ˆB(ve,ζ)=B(ve,ζ)+aew(ζ). Substituting this into (5.8), we obtain the test quadratic form (5.7). The Lyapunov stability of the equilibrium point (ve,ae) then follows from a revised Lagrange's theorem in chapter §Ⅱ.3 of [7].

    Let D be a 2-dimensional domain with boundary, and dA is an area form. The velocity field ve=ψe is a stationary solution of equation (5.1), and ψe stands for the stream function of the shifted velocity field veVs.

    We can now prove the following theorem, which provides an a priori estimate for 2-dimensional steady flows of the CL equation:

    Theorem 5.3. Consider a 2-dimensional domain D with an area form dA. Assume that (i) ψe=F(Δψe) for some function F, and (ii) there exist two constants c1 and c2 such that

    0<c1ψeΔψec2<. (5.9)

    Let ψ(x,y,t)=ψe+h(x,y,t) be the stream function corresponding to a different solution of the CL equation such that Dψdl=Dψedl. Then, for the perturbation h=h(x,y,t), we have the following inequality:

    h22+c1Δh22h022+c2Δh022, (5.10)

    where h0=h(x,y,0), and 22 denotes the square of the L2norm, which is given by u22=D(u,u)dA for a vector field u and f22=Df2dA for a function f.

    Proof. By the assumption of Theorem 5.3, we have ψe=F(Δψe). Let the function P be the primitive of F, i.e., P=F. Then, P(Δψe)=ψeΔψe. Again by the assumption, we have c1P(ω)c2, which gives

    c1η22P(ω+η)P(ω)P(ω)ηc2η22.

    This implies

    h22+2D(P(Δψe+Δh)P(Δψe)P(Δψe)Δh)dAh22+c1Δh22, (5.11)
    h022+2D(P(Δψe+Δh0)P(Δψe)P(Δψe)Δh0)dAh022+c2Δh022. (5.12)

    Introduce a functional

    C(h)=h222+D(P(Δψe+Δh)P(Δψe)P(Δψe)Δh)dA.

    Then, the left-hand sides of (5.11) and (5.12) are 2C(h(t)) and 2C(h(0)), respectively. Therefore, if we could prove

    C(h(t))=C(h(0)), (5.13)

    then the theorem will follow immediately from (5.11), (5.12) and (5.13).

    To prove equation (5.13), we construct the following invariant functional according to the conservation of kinetic energy and vorticity:

    Γ(ψ)=ψ222+DP(Δψ)dA,

    where ψ+Vs=ψ. The first variation of Γ at the equilibrium solution ψe is

    δΓψe(h)=D((h,ψe)+P(Δψe)Δh)dA=D(ψeΔh+P(Δψe)Δh)dA+Dψehndl.

    Since P(Δψe)=F(Δψe)=ψe, and Dψehndl=0, we obtain δΓψe(h)=0.

    For another functional ˜Γ(h):=Γ(ψe+h)Γ(ψe), we have

    ˜Γ(h(t))=˜Γ(h(0))

    and

    ˜Γ(h)=δΓψe(h)+C(h),

    and these two equalities imply (5.13). This completes the proof of the theorem.

    Remark 5.1. Consider the leaf of the coadjoint foliation of ˆXdvol(D) which contains the equilibrium point (ve,1). Then, (v,1)ˆXdvol(D) is on this leaf if and only if v is isovorticed to the equilibrium field ve in the sense of Definition 3.2. The second variation of K(v)=12D(v,v)dA on the leaf is

    δ2Kve(ξ)=12D((ξ,ξ)+ψeΔψe(curlξ)2)dA, (5.14)

    where ξ stands for a variation field at ve.

    Next, we prove equation (5.14). According to equation (5.8), we have

    2δ2Kve(ξ)=D((ξ,ξ)+(ξ,[ζ,ve]))dA, (5.15)

    where ξ=B(ve,ζ)+w(ζ)=B(veVs,ζ). The last term of this equation is

    D(ξ,[ζ,ve])dA=D(ξ,curl(ζ×ve))dA=D(curlξ,(ζ×ve))dA. (5.16)

    Since ve=ψe, and veVs=ψe, one gets

    curlξ=LζΔψe=(ζ,Δψe),
    ζ×ve=ζ×(ψe)=(ζ,ψe).

    Thus,

    ζ×ve=ψeΔψecurlξ. (5.17)

    By equation (5.15), (5.16) and (5.17), we prove the required equation (5.14).

    It is evident from the expression of (5.14) that the assumption (ⅱ) in Theorem 5.3 guarantees that the second variation δ2K is positive definite.

    In the following, we give two examples of stable steady CL flows.

    Example 5.1. Let D={(x,y)|0y1} be the domain, and let us=(us(y),0) be the Stokes drift velocity. A shear flow in this domain has velocity ue=(u(y),0). It is easy to verify that this is a stationary flow. Assume that ue=ψe,ueus=ψe, and we have

    ψeΔψe=u(uus).

    Hence, this shear flow is stable if there exist two constants c1 and c2 such that 0<c1ψeΔψec2<. This implies u(y)us(y) has no inflection point.

    Example 5.2. Let A={1r2} be the domain, where r=x2+y2. Consider a velocity field ue and a drift velocity field us. If there are two functions f and g such that the stream functions ψe and ψe of ue and ueus satisfy ψe=f(r) and ψe=g(r), then the flow is steady, and we have

    ψeΔψe=f(g+g/r).

    By Theorem 5.3, this flow is stable if f(g+g/r) is positive defined.

    The author would like to express his gratitude to Boris Khesin for many fruitful discussions. He also is thankful to Xiaoping Yuan for valuable suggestions and encouragement.

    The authors declare there is no conflict of interest.



    [1] I. Langmuir, Surface Motion of Water Induced by Wind, Science, 87 (1938), 119–123. https://doi.org/10.1126/science.87.2250.119 doi: 10.1126/science.87.2250.119
    [2] D. D. Craik, S. Leibovich, A rational model for Langmuir circulations, J. Fluid. Mech., 73 (1976), 401–426. https://doi.org/10.1017/s0022112076001420 doi: 10.1017/s0022112076001420
    [3] D. D. Holm, The Ideal Craik-Leibovich Equations, Physica D, 98 (1996), 415–441. https://doi.org/10.1016/0167-2789(96)00105-4 doi: 10.1016/0167-2789(96)00105-4
    [4] V. A. Vladimirov, M. R. E. Proctor, D. W. Hughes, Vortex dynamics of oscillating flows, Arnold Math J., 1 (2015), 113–126. https://doi.org/10.1007/s40598-015-0010-x doi: 10.1007/s40598-015-0010-x
    [5] C. Yang, Multiscale method, Central extensions and a generalized Craik-Leibovich equation, J. Geom. Phys., 116 (2017), 228–243. https://doi.org/10.1016/j.geomphys.2017.02.004 doi: 10.1016/j.geomphys.2017.02.004
    [6] V. I. Arnold, Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l'hydrodynamique des uides parfaits, Ann. Inst. Fourier., 16 (1966), 319–361. https://doi.org/10.5802/aif.233 doi: 10.5802/aif.233
    [7] V. I. Arnold, B. A. Khesin, Topological methods in hydrodynamics, Springer-Verlag, New York, 1998. https://doi.org/10.1007/b97593
    [8] B. A. Khesin, Yu. V. Chekanov, Invariants of the Euler equations for ideal or barotropic hydrodynamics and superconductivity in D dimensions, Physica D, 40 (1989), 119–131. https://doi.org/10.1016/0167-2789(89)90030-4 doi: 10.1016/0167-2789(89)90030-4
    [9] C. Roger, Extensions centrales d'algèbres et de groupes de Lie de dimension infinie, algèbre de Virasoro et généralisations, Rep. Math. Phys., 35 (1995), 225–266. https://doi.org/10.1016/0034-4877(96)89288-3 doi: 10.1016/0034-4877(96)89288-3
    [10] C. Vizman, Geodesics on extensions of Lie groups and stability: the superconductivity equation, Phys. Lett. A, 284 (2001), 23–30. https://doi.org/10.1016/s0375-9601(01)00279-1 doi: 10.1016/s0375-9601(01)00279-1
    [11] V. Zeitlin, Vorticity and waves: geometry of phase-space and the problem of normal variables, Physics. Letters. A., 164 (1992), 177–183. https://doi.org/10.1016/0375-9601(92)90699-m doi: 10.1016/0375-9601(92)90699-m
    [12] C. Yang, B. Khesin, Averaging, symplectic reduction, and central extensions, Nonlinearity, 33 (2020), 1342–1365. https://doi.org/10.1088/1361-6544/ab5cdf doi: 10.1088/1361-6544/ab5cdf
    [13] J. E. Marsden, G. Misiolek, J. Ortega, M. Perlmutter, T. S. Ratiu, Hamiltonian Reduction by Stages, Springer-Verlag, New York, 2007. https://doi.org/10.1007/978-3-540-72470-4
  • This article has been cited by:

    1. Xue Wang, Guoxian Chen, A positivity-preserving well-balanced wet-dry front reconstruction for shallow water equations on rectangular grids, 2024, 198, 01689274, 295, 10.1016/j.apnum.2024.01.012
  • Reader Comments
  • © 2023 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(2086) PDF downloads(123) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog