Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2014 Dec 8;470(2172):20140322. doi: 10.1098/rspa.2014.0322

Variational necessary and sufficient stability conditions for inviscid shear flow

M Hirota 1,, P J Morrison 2, Y Hattori 1
PMCID: PMC4241005  PMID: 25484600

Abstract

A necessary and sufficient condition for linear stability of inviscid parallel shear flow is formulated by developing a novel variational principle, where the velocity profile is assumed to be monotonic and analytic. It is shown that unstable eigenvalues of Rayleigh's equation (which is a non-self-adjoint eigenvalue problem) can be associated with positive eigenvalues of a certain self-adjoint operator. The stability is therefore determined by maximizing a quadratic form, which is theoretically and numerically more tractable than directly solving Rayleigh's equation. This variational stability criterion is based on the understanding of Kreĭn signature for continuous spectra and is applicable to other stability problems of infinite-dimensional Hamiltonian systems.

Keywords: flow stability, variational method, Kreĭn signature, continuum Hamiltonian Hopf bifurcation

1. Introduction

In this paper, ideas from the theory of Hamiltonian systems are used to obtain both necessary and sufficient stability conditions by a variational procedure. The proposed procedure is of general utility, but the treatment here will be confined to plane parallel inviscid shear flow (e.g. [1]). In this section, we give an overview of the underlying basis for the procedure in terms of a finite-dimensional Hamiltonian framework, and then place the present contribution in the context of the many previous results for shear flow.

For some Hamiltonian systems, the sign of the curvature of the potential energy function provides a necessary and sufficient condition for stability. This is referred to as Lagrange's theorem, which is the crux of many fluid and plasma stability results including the ‘δW’ energy principle of ideal magnetohydrodynamics (MHD) [2]. For a more general class of Hamiltonian systems, definiteness of the Hamiltonian Hessian matrix evaluated at the equilibrium point of interest provides only a sufficient condition for stability. This is sometimes referred to as Dirichlet's theorem, which is the crux of many sufficient conditions for stability in the fluid and plasma literature. The essential reason that Dirichlet's theorem does not provide a necessary condition for stability is the possible existence of negative energy modes. Negative energy modes are modes of undamped oscillation, for which the Hamiltonian decreases when the mode is excited, i.e. the second variation of the Hamiltonian evaluated on the mode is negative. For stable (non-degenerate) Hamiltonian systems of n degrees-of-freedom the linear dynamics can be brought by a canonical transformation into the following normal form:

H=α=1nσαωα2(pα2+qα2), 1.1

where (q1,q2,…,qn;p1,p2,…,pn) are the canonically conjugate coordinates, ωα are positive real numbers representing the mode frequencies, and σα∈{±1} are the signatures of the mode, often called Kreĭn signature, with +1 and −1 corresponding to positive and negative energy modes, respectively. Evidently, systems with both positive and negative energy modes are linearly stable but do not have a definite Hessian matrix. (See [3] for review.)

An advantage afforded by Lagrange's criterion over Dirichlet's is the powerful Rayleigh–Ritz variational method [4], which underlies the MHD and other energy principles. With this method, one needs to only produce a trial function that makes the Rayleigh quotient negative in order to show instability and, also in this way, threshold parameter values for the transition to instability can be determined. When this method is applicable, it is of great utility because linear stability conditions for interesting fluid and plasma systems are generally difficult to derive theoretically. However, it only applies to a restricted class of Hamiltonian systems for which the eigenvalue problem is self-adjoint, i.e. systems with steady-state bifurcations to instability through zero frequency that have pure exponential growth upon transition, and it is known that systems with shear flow are not self-adjoint and have instead Kreĭn bifurcations [5] to overstability, i.e. unstable eigenvalues with both non-zero real and imaginary parts. Such bifurcations are often called Hamiltonian–Hopf bifurcations [58] and can be viewed as a resonance between positive and negative energy modes leading to instability.

Thus, we are led to re-examine Dirichlet's principle and seek an alteration that affords the utility of the Rayleigh–Ritz variational method for investigation of Kreĭn bifurcations. For finite-dimensional Hamiltonian system written in the normal form of (1.1), it is evident that Iα=pα2+qα2 is a constant of motion for each α. From these constants of motion, one can construct a constant of motion with definite Hessian simply by flipping the signs of the signature in the normal form Hamiltonian. Unfortunately, a priori knowledge of the existence of such a definite constant of motion is generally not at hand, and one must actually solve the eigenvalue problem in order to be informed of its existence. This is the essential reason Dirichlet's theorem does not give both necessary and sufficient conditions for stability and a Rayleigh-like variational method is not at hand.

However, there are two related discoveries that we can exploit to circumvent this deficiency for problems with continuous spectra, such as those related to the Vlasov equation, MHD and the plane shear flow problem considered here. The first is the infinite sequence of constants of motion discovered in [9], which were elaborated on in [10] and used in the present context in [11]. The second is the discovery of a Kreĭn-like signature for the continuous spectra of Vlasov–Poisson equilibria in [12,13], which was applied to plane shear flow in [14] and extended to a large class of systems in [1517]. These constants of motion in conjunction with the definition of signature allow the construction of a quadratic form, which we will call Q, the variation of which can be used in a manner akin Rayleigh's principle for ascertaining necessary and sufficient conditions for stability. A version of the quadratic form Q was previously given in [11], but it was not used to obtain sufficient conditions for instability. We note in passing that the discovery of signature for the continuous spectrum has also led to rigorous Kreĭn-like theorems [1820], where discrete eigenvalues emerge from the continuous spectrum, termed Continuum Hamiltonian Hopf bifurcations. (See also [21,22] on infinite-dimensional Hamiltonian systems and many other contributions in the recent book [23].)

There have been many significant contributions to the classical plane parallel shear flow problem; thus, it is important to put our contribution into perspective, which we do here. The most famous condition is Rayleigh's criterion [24] that stipulates the existence of an inflection point in the velocity profile is necessary for instability, a criterion that was improved by Fjørtoft [25]. These criteria were obtained by direct manipulation of Rayleigh's equation (see equation (2.4)), which governs linear disturbance about a base shear flow. The first allusion to Hamiltonian structure for this system appears in the works of Arnold [2628], who obtained more general sufficient conditions for stability by making use of additional invariants. This idea was anticipated in the plasma physics literature [29], where the additional invariants were referred to as generalized entropies; today, the generalized entropies are referred to as Casimir invariants and the general procedure is called the energy-Casimir method (e.g. [3,3032]). In [14], it was shown explicitly that all of the above criteria amount to a version of Dirichlet's theorem for this infinite-dimensional Hamiltonian system, where it was also shown how to explicitly map the system into the infinite-dimensional version of the normal form of (1.1). In this way, a signature for the continuous spectrum was first defined for this system, by paralleling the analogous procedure for the Vlasov–Poisson system [12,13]. Arnold also introduced an important kind of constrained variation he termed isovortical perturbations (e.g. [33]), which are a special case of the dynamically accessible variations of [34,35] that are generated by Poisson brackets [36,3]. Isovortical perturbations together with a more general Dirichlet-like sufficient condition for shear flow due to Barston [11] play important roles for obtaining the results of the present paper. The general idea of these earlier works is to obtain improved sufficient conditions for stability by constructing a suitable quadratic form (or the Lyapunov function) with the aid of additional invariants. This idea is powerful especially for integrable systems (e.g. [37,38]), in which abundant constants of motion are available not only for linearized but also for nonlinear dynamics.

We also note that prior to our variational approach, necessary and sufficient stability conditions were obtained for certain classes of shear flows using two other non-Hamiltonian approaches. One is the perturbation expansion around a neutrally stable eigenmode, which was pioneered by Tollmien [39] and developed by many authors [4043], and the other is analysis based on the Nyquist method [44,45] (applied to MHD in [46]). These two approaches, however, require detailed probing of Rayleigh's equation to obtain information under specific conditions. Our variational approach is not only consistent with these earlier results, but also advantageous in that we do not have to solve Rayleigh's equation in a rigorous manner. Namely, we can prove the instability by simply finding a test function (in the appropriate function space) that makes our quadratic form Q positive. This is useful because explicit solutions for Rayleigh's equation are generally not available for a given velocity profile. Moreover, in numerical calculation, one can obtain stability boundaries more efficiently from this variational problem (i.e. the maximization of Q), compared to solving Rayleigh's equation. We emphasize that our approach is not limited to shear flow with rather simple velocity profiles, but the same idea is applicable to various fluid and plasma stability problems to which it can be of practical use.

Our paper is organized as follows. In §2, Rayleigh's equation is first introduced, and in §3, the notion of isovortical variation is reviewed. Here, we describe the quadratic form Q, which provides the necessary and sufficient conditions if the velocity profile satisfies the assumptions of analyticity and monotonicity. In particular, we present the main theorem of this work (theorem 3.1), in which the quadratic form Q is given explicitly. Then, in §4, the proof of the main theorem is given. Here, we first focus on restricting perturbations to the appropriate function space, and then perform the spectral decomposition in a rigorous manner, which largely reproduces the well-known spectral properties of Rayleigh's equation (e.g. [47]). Next, we calculate the signature of Q by applying techniques [14,16,17] for the action–angle representation of continuous spectrum, where the positive signature of Q indeed predicts the existence of unstable eigenvalues. Finally, in §4, we show that the function space (the search space on which Q should be maximized) can be extended to a larger one, which is actually beneficial for solving the variational problem more effectively. Section 5 contains a demonstration that our variational criterion reproduces the earlier results of the Nyquist method [44,45] and the perturbation analysis of the neutral modes [3943], while §6, contains several numerical examples that demonstrate of our theorem. We summarize in §7.

2. Rayleigh equation

We consider the linear stability of inviscid parallel shear flow U=(0,U(x)) on a domain (x,y)[L,L]×[,] bounded by two walls at xL, where the flow is assumed to be incompressible and two-dimensional. By introducing the z-component of the vorticity disturbance as w(x,t) eiky+c.c. for a wavenumber k>0, the linearized vorticity equation is written as follows:

itw=kUw+kUGw=:kLw, 2.1

where the prime (′) indicates the x derivative, and the convolution operator G is defined by

(Gw)(x,t)=LLg(x,s)w(s,t)ds, 2.2

with a kernel,

g(x,s)={sinhk(sL)sinhk(x+L)ksinh2kLx<ssinhk(s+L)sinhk(xL)ksinh2kLs<x. 2.3

The stream function ϕ(x,t) is therefore given by ϕ=Gw or w=G1ϕ=ϕ+k2ϕ. By assuming an exponential behaviour ϕ(x,t)=ϕ^(x)eiωt with a complex frequency ωC, the eigenvalue problem for (2.1) is known as Rayleigh's equation [24]

(cU)(ϕ^k2ϕ^)+Uϕ^=0 2.4

and

ϕ^(L)=ϕ^(L)=0, 2.5

where c=ω/k is a complex phase speed. If this equation has a non-trivial solution ϕ^ for c with a positive imaginary part, Imc>0, the linearized system (2.1) is spectrally unstable due to an exponentially growing eigenmode.

3. Variational stability criterion

Hamiltonian structure of the linearized vorticity equation (2.1) is highly related to its adjoint equation for ξ(x,t) [14,16,33]

itξ=kUξ+kG(Uξ)=:kLξ, 3.1

where L is the adjoint operator of L with respect to the inner product,

ξ¯,η=LLξ(x)¯η(x)dxforξ,ηL2+iL2. 3.2

Here, ξ¯ denotes the complex conjugate of ξ, and we consider the function space for disturbances w and ξ to be the complex Hilbert space L2+iL2 on [−L,L] that is defined by the norm ξL2+iL22=ξ¯,ξ.

Since the relation UL=LU holds, w=−U′′ξ is found to be a solution of (2.1) if ξ is a solution of (3.1). This solution must vanish at positions where U′′=0, and hence it does not span the whole space L2+iL2 when U′′=0 somewhere. This class of perturbations belonging to the range of U′′ is said to be isovortical because the vorticity disturbance (w) is induced by a displacement (ξ) of the fluid while preserving the conservation law of circulation [33] (see also appendix A). In this manner, Arnold [33] derived a constant of motion

δ2H=LLξ¯U[Uξ+G(Uξ)]dx. 3.3

Arnold showed that this is the second variation of the energy with respect to the isovortical variation, while in [14] it was shown that this quantity is in fact the Hamiltonian for the linear Hamiltonian dynamics and there the diagonalizing transformation to action–angle variables was first obtained. In a frame moving at a velocity U*, Arnold replaced U by UU* in δ2H to obtain

δ2H=LLξ¯U[(UU)ξ+G(Uξ)]dx=LLw¯(UUU+G)wdx, 3.4

which is also a constant of motion, while the last expression may not be well-defined when U′′ becomes zero and w is not isovortical. In [14], it was shown explicitly that δ2H* is in fact the second variation of the full Hamiltonian in the inertial frame boosted by velocity U* by adding the appropriate total momentum.

Thus, we have a version of Dirichlet's theorem, where the shear flow U(x) is stable in the sense of Lyapunov, if there exists UR such that the quadratic form δ2H* is either positive or negative definite, i.e. ∃ϵ>0 such that δ2Hϵw¯,w or δ2Hϵw¯,w. For example, when U′′≠0 everywhere, one can make δ2H* positive definite by choosing U* such that (UU*)/U′′>0 holds everywhere, which reproduces the Rayleigh criterion [24]. When U(x) has only one inflection point xI (i.e. U′′(xI)=0), the choice U*=UI:=U(xI) is made by Arnold. Then, δ2H* is positive definite if (UUI)/U′′>0 holds everywhere, in agreement with the Fjørtoft criterion [25]. These facts imply that this variational criterion of Arnold [2628] applies to a larger class of flow profiles than Rayleigh–Fjørtoft's stability theorem. However, all these criteria, including a generalization by Barston [11], are still sufficient conditions for stability and, hence, are indeterminate when δ2H* is indefinite, as discussed in §1 there could be negative energy modes.

In this work, we obtain an improved variational criterion, but this requires introducing the following assumptions on U(x).

Assumption —

  • (A 1) U(x) is an analytic (i.e. regular) and bounded function on [−L,L].

  • (A 2) U(x) is strictly monotonic [i.e. U′(x)≠0 for all x] and, if U′′(xI)=0 at x=xI, then U′′′(xI)≠0.

The last statement implies that the inflection point xI must be a simple zero of U′′(x) and the sign of U′′(x) must change at x=xI. We expect that it is not difficult to relax these restrictions on U(x) except for the monotonicity. To simplify our mathematical arguments, we will not pursue generalization in the present work, but we do remark upon this point in our summary of §7.

Our main result is that a necessary and sufficient condition for spectral stability is attained by the following variational criterion.

Theorem 3.1 —

Let U(x) satisfy (A 1) and (A 2). Denote the inflection points of U by xIn, n=1,2,…,N, and define a quadratic form Q=ξ,Hξ as

Q=νLLξn=1N[UUIn+UG]Uξdx, 3.5

where UIn=U(xIn) and either ν=1 or ν=−1 is chosen such that

νUn=1N(UUIn)0 3.6

holds for all x∈[−L,L]. Equation (2.1) is spectrally stable if and only if

Q=ξ,Hξ0for allξL2. 3.7

In this theorem, we have introduced Q on the real Hilbert space L2 defined by the norm ξL22=ξ,ξ, which indicates that in practice the search space of this variational criterion is the half of L2+iL2 since H is a real self-adjoint operator. Actually, the stability condition (3.7) can be replaced by Q=ξ¯,Hξ0 for all ξL2+iL2, and we will prove the latter by regarding ξL2+iL2 as a solution of (3.1).

This Q=ξ¯,Hξ is equivalent to the constant of motion derived by Barston [11] (except for the coefficient ν), and Q=−νδ2H* with U*=UI for the case of single inflection point. Hence, our theorem claims that Arnold–Barston's stability criteria (Dirichlet sufficient stability conditions) are in fact necessary and sufficient when U(x) satisfies (A 1) and (A 2).

We remark that Q=ξ¯,Hξ no longer represents the second variation of the energy for the case of multiple infection points. Actually, it belongs to the class of infinite number of constants of motion introduced in [911].

Proposition 3.2 —

Let f(c) be any real polynomial of cR. Then,

Qf=LLξ¯Uf(L)ξdx=LLξ¯f(L)(Uξ)dxR 3.8

is a constant of motion for equation (3.1).

Proof. —

Using UL=LU and Lf(L)=f(L)L, we can directly show that Qf is real and dQf/dt=0. ▪

Therefore, we have specifically chosen f(c)=νn=1N(cUIn) to generate Q of theorem 3.1.

The proof of theorem 3.1 is given in the next section and our strategy is as follows. First, we reduce the function space L2+iL2 to a smaller one that will be denoted by X+iX, to which unstable eigenfunctions must belong. Then, we decompose the spectrum σC of the operator kL into the neutrally stable part σcR (which is mostly a continuous spectrum) and the remaining part σσcCR (which is a set of pairs of growing and damping eigenvalues). By proving that Q≤0 for all the neutrally stable disturbance ξ belonging to σc, we will claim that Q>0 for some ξX indicates the existence of at least one unstable eigenvalue, ωσσc that has a growth rate Imω>0.

4. Proof of theorem 3.1

(a). Reduction to isovortical disturbance

For the purpose of seeking unstable eigenmodes, we restrict the function space of disturbance to X+iX, where

X=H01H2. 4.1

As usual, we denote by Hn the real Sobolev space on [−L,L];

Hn={ξL2|jnxjξL2<} 4.2

and H01 denotes the subspace of H1 in which the boundary conditions, ξ(−L)=ξ(L)=0, are imposed on ξH01. The restriction from L2+iL2 to X+iX is feasible when U(x) is a sufficiently smooth function. In this work, we simply assume (A 1) is sufficient for the following:

Proposition 4.1 —

Let U(x) satisfy (A 1). Given the initial condition ξ(x,0)=ξ0(x)∈X+iX, the solution ξ(x,t) of (3.1) belongs to X+iX for all t. Moreover, w=−U′′ξX+iX is a solution of (2.1).

Proof. —

By noting that G:L2X is one to one and onto, we find that L is a bounded operator on X+iX and, hence, the solution ξ=eikLtξ0 belongs to X+iX for all t. Using the property UL=LU, it is obvious that w=−U′′ξ is automatically a solution of (2.1) and also belongs to X+iX. ▪

When U′′ vanishes somewhere on [−L,L], the function space of w=−U′′ξ is further restricted to the range of U′′ (i.e. the isovortical disturbance). We can find that all unstable eigenfunctions must belong to this space as follows.

Proposition 4.2 —

Let U(x) satisfy (A 1). Equation (2.1) is spectrally stable if and only if the adjoint equation (3.1) for ξX+iX is spectrally stable.

Proof. —

If cC and w^=ϕ^+k2ϕ^L2+iL2 satisfy Rayleigh's equation with a growth rate Im c>0, then (cU)≠0 holds everywhere and

ξ^=1cUGw^X+iX 4.3

is found to be an eigenfunction of the adjoint equation (3.1) with the same eigenvalue c. Hence, the adjoint equation on X+iX is also spectrally unstable.

Conversely, if c and ξ^X+iX satisfy the adjoint eigenvalue problem with Im c>0, then Uξ^ is not identically zero and w^=Uξ^ satisfies the Rayleigh equation with the same c. ▪

(b). Spectral decomposition

Next, we investigate the spectrum σC of the operator kL. For a given initial condition ξ(x,0)=ξ0(x)∈X+iX, let Ξ(x,Ω)∈X+iX be the solution of

(ΩkL)Ξ(x,Ω)=ξ0(x), 4.4

for ΩCσ. Then, the solution of (3.1) is formally represented by the Dunford integral (or the inverse Laplace transform)

ξ(x,t)=12πiΓ(σ)Ξ(x,Ω)eiΩtdΩ, 4.5

where Γ(σ) is a path of integration that encloses all the spectrum σC of kL counterclockwise (as shown in figure 1). In terms of Φ(x,Ω)=−(Ω/kU)Ξ(x,Ω), equation (4.4) is transformed into

E(Ω)Φ(x,Ω)=1k(ξ0k2ξ0) 4.6

and

Φ(L,Ω)=Φ(L,Ω)=0, 4.7

where

E(Ω)=2x2+k2kUΩkU. 4.8

Suppose that we have solved

E(Ω)Φ<(x,Ω)=0,Φ<(L,Ω)=0,Φ<(L,Ω)=1 4.9

and

E(Ω)Φ>(x,Ω)=0,Φ>(L,Ω)=0,Φ>(L,Ω)=1, 4.10

to obtain two linearly independent solutions Φ<(x,Ω) and Φ>(x,Ω). Then, by using the method of Green's function, the solution of (4.6) and (4.7) can be expressed as

Φ(x,Ω)=1W(Ω)LLΦG(x,s,Ω)1k[ξ0(s)k2ξ0(s)]ds, 4.11

where

ΦG(x,s,Ω)=Φ>(s,Ω)Φ<(x,Ω)Y(sx)+Φ<(s,Ω)Φ>(x,Ω)Y(xs), 4.12

with Y (x) being the Heaviside function, and

W(Ω)=Φ<(x,Ω)Φ>(x,Ω)+Φ<(x,Ω)Φ>(x,Ω)=Φ>(L,Ω)=Φ<(L,Ω) 4.13

is the Wronskian. When Ω avoids the range of kU, denoted by

σc={kU(x)R|x[L,L]}, 4.14

the operator E(Ω) is non-singular and both Φ< and Φ> are regular functions of Ω. Therefore, the spectrum σ of the operator kL is composed of a continuous spectrum σc [47,48] and some eigenvalues ωjC, j=1,2,… (i.e. point spectra) that satisfy W(ωj)=0. Owing to the property W(Ω)=W(Ω¯)¯, the eigenvalues always exist as pairs of growing (ωj) and damping (ω¯j) ones when Im ωj>0, a spectral property of Hamiltonian systems (cf. [1820]). To be more precise, σc is mostly the continuous spectrum because some real eigenvalues (ωjR) may exist in it. In the remainder of this subsection, we will ascertain these spectral properties of kL, which will be summarized explicitly in proposition 4.6.

Figure 1.

Figure 1.

Schematic view of contour integral Γ(σ)=Γ(σc)∪Γ(σσc) in the case of σσc={ω1,ω2,ω¯1,ω¯2}.

For the purpose of showing the existence or non-existence of such eigenvalues, we will frequently use the following lemma.

Lemma 4.3 —

Let U(x) satisfy (A 1) and (A 2). For any ωR, the general solution Φ(x,ω) of E(ω)Φ(x,ω)=0 has at most one zero on each of the following intervals:

  • (i) [−L,L] if there is no critical layer, i.e. ωR that satisfies ω=kU(x)∀x∈[−L,L],

  • (ii) [−L,xc] and [xc,L] if there is a critical layer xc∈[−L,L] that satisfies ω=kU(xc).

Proof. —

(i) A consequence of E(ω)Φ(x,ω)=0 is the following identity (see the Appendix of [45]):

[Φ(ΦUΦUc)]x1x2=x1x2[(ΦUΦUc)2+k2Φ2]dx, 4.15

which is valid for any solution Φ and subinterval [x1,x2]⊆[−L,L]. This identity follows directly from Rayleigh's equation by multiplying by Φ, manipulating and integrating. If there are two zeros (i.e. xz1 and xz2) that satisfy Φ(xz1,ω)=Φ(xz2,ω)=0, choosing x1=xz1 and x2=xz2 implies

xz1xz2[(ΦUΦUc)2+k2Φ2]dx=0, 4.16

which requires Φ to be the trivial solution Φ≡0. Thus, any (non-trivial) solution has at most one zero on [−L,L].

(ii) Consider the interval [xc,L] (the same argument goes for [−L,xc]). The identity (4.15) again holds for xc<x1<x2L and, hence, there is at most one zero on (xc,L]. In the neighbourhood of xc, the solution Φ is expressed by a linear combination of the Frobenius series solutions, in which Φ(xc,ω) is always bounded (see (4.21)). If Φ(xc,ω)≠0, the lemma is automatically proved (although both sides of (4.15) go to infinity as x1xc). If Φ(xc,ω)=0, the identity (4.15) is still valid for x1=xc and we can prove that there is no other zero (xz2) except for xz1=xc on [xc,L], in the same manner as (i). ▪

It is well known from Tollmien's argument [3941] that, as k varies, an unstable eigenvalue Im c>0 of Rayleigh's equation emerges through a neutrally stable eigenvalue c=U(xI)R where U′′(xI)=0. In other words, the neutrally stable eigenvalue cR can exist only if U′′(xc)=0 at the corresponding critical layer xc=U−1(c). Unfortunately, this argument is not always true especially for non-monotonic profiles of U(x) (see [42,43,45] for mathematical justification in the shear flow context and [18,19] for a discussion of k≠0 bifurcations in the Vlasov context). In this work, we simply assume the monotonicity (A 2) and verify Tollmien's argument as follows.

Proposition 4.4 —

Let U(x) satisfy (A 1) and (A 2). Denote the inflection points of U by xIn, n=1,2,…,N and define UIn:=U(xIn). Then, the function W(ω±i0) of ωR can vanish only at ω=kUIn, n=1,2,…,N, and moreover

limΩω±i0n=1N(ΩkUIn)W(Ω)<. 4.17

Proof. —

For ωRσc, there is no critical layer and, from lemma 4.3, the solution Φ<(x,ω) does not have zero on [−L,L] except for x=−L. Hence, W(ω)=Φ<(L,ω)≠0.

For ωσc, there is only one critical layer xc=U−1(ω/k). Since U(x) is analytic, Φ<(x,Ω) can be expressed by a linear combination of the Frobenius series solutions (so-called Tollmien's inviscid solutions) around xc. By taking account of the branch cut of the logarithmic function, it is written in the limit Ωω±i0 as

Φ<(x,ω±i0)=Cr(ω)Φ1(x,ω)+Cs(ω){Φ2(x,ω)+U(xc)U(xc)Φ1(x,ω)[log|xxc|πiY(xxc)]}, 4.18

where Φ1(x,ω) and Φ2(x,ω) are real and regular functions with Φ1(xc,ω)=Φ2′(xc,ω)=0 and Φ1′(xc,ω)=Φ2(xc,ω)=1 [49]. From definition (4.9), the coefficients Cr(ω) and Cs(ω) are found to be real and so is Φ<(x,ω±i0) on [−L,xc] since Y (xxc)≡0 for x<xc. Lemma 4.3 shows that Φ<(x,ω±i0) has no zero on [−L,xc] other than x=−L and hence Cs(ω)>0 (since Φ1(xc,ω)=0).

If ωkUIn (i.e. xcxIn), then U′′(xc)≠0 holds and Φ<(x,ω±i0) possesses the imaginary part on [xc,L]. Lemma 4.3 again shows that this imaginary part has no zero on [xc,L] other than x=xc. Therefore, we conclude that Im W(ω±i0)=Im Φ<(L,ω±i0)≠0 for all ωR{kUIn|n=1,2,,N}.

If ω=kUIn, then Φ<(x,kUIn) is a real and regular function on the whole domain [−L,L] due to U′′(xc)=0 and has at most one zero on [xc,L]. Therefore, W(kUIn)=0 may occur only when this zero corresponds to x=L, and as such the zero must be simple, namely, (4.17) holds. ▪

Besides the singularity stemming from the zeros of W(Ω), Φ(x,Ω) has also the following essential singularity along the continuous spectrum σc.

Proposition 4.5 —

Let U(x) satisfy (A 1) and (A 2). For all ξ0X+iX and ωR,

Ψ(x,ω±i0)H01+iH01, 4.19

where

Ψ(x,Ω):=Φ(x,Ω)n=1N(ΩkUIn). 4.20

Proof. —

For any fixed s∈[−L,L], the function ∂ΦG/∂x(x,s,ω±i0) is regular almost everywhere except that it has a logarithmic singularity, log|xxc|, and discontinuities, Y (xxc) and Y (xs). Hence,

LL|ΦGx(x,s,ω±i0)|dx<. 4.21

Since ξ0′′−k2ξ0L2+iL2, the following convolution integral is also square-integrable:

limΩ=ω±i0[W(Ω)Φ(x,Ω)]=LLΦGx(x,s,ω±i0)1k[ξ0(s)k2ξ0(s)]dsL2+iL2, 4.22

that is,

limΩ=ω±i0[W(Ω)Φ(x,Ω)]H1+iH1. 4.23

In combination with (4.17) and the boundary condition ΦL,Ω)=0, the proposition is proven. ▪

In summary, the spectrum is decomposed as follows.

Proposition 4.6 —

Let U(x) satisfy (A 1) and (A 2). The spectrum σC of kL on X+iX is composed of

  • (a) point spectra σσc={ωj,ω¯jC|Imωj0,W(ωj)=0,j=1,2,,Np}, where Np is finite.

  • (b) point spectra σI={kUInR|W(kUIn±i0)=0,n=1,2,,N}.

  • (c) continuous spectrum σcσI.

where (c) always exists while (a) and (b) can be empty. The residual spectrum is always empty.

Proof. —

The finiteness of Np is proved in [47]. The existence of (b) follows from proposition 4.4. The proof of (c) and the absence of the residual spectrum are relegated to appendix B, since the subsequent discussion will not refer to these facts. ▪

(c). Signature of Q

Now, let us substitute expression (4.5) into the quadratic form Q=ξ¯,Hξ, which is a constant of motion for equation (3.1). By using the property of the resolvent (ΩkL)1 [16,17], we obtain

Q=12πiΓ(σ)h(Ω)dΩ 4.24

where h:CC is given by

h(Ω)=νLLξ¯0[kUΩkUΨ(x,Ω)]dx. 4.25

Upon decomposing the spectrum σC into σcR and others σσcCR and, accordingly, deforming the contour Γ(σ) into Γ(σc) and Γ(σσc) (figure 1), we obtain Q=Q|σc+Q|σσc with

Q|σc=12πiΓ(σc)h(Ω)dΩ=σch^(ω)dω, 4.26

where

h^(ω)=12πi[h(ω+i0)+h(ωi0)]. 4.27

The existence of this limit for all ωσc is guaranteed by proposition 4.5 and ξ0X+iX.

Then, we can prove the following inequality.

Proposition 4.7 —

Q|σc=σch^(ω)dω0 4.28

for all solutions ξX+iX of (3.1) with initial data ξ0X+iX.

Proof. —

To observe the signature of the function h^(ω) more explicitly, we rewrite h(Ω) as

h(Ω)=νkp(Ω)LLΨ(x,Ω¯)¯E(Ω)Ψ(x,Ω)dxνp(Ω)kLL(|ξ0|2+k2|ξ0|2)dx, 4.29

where we have put p(Ω)=n=1N(Ω/kUIn). Since p(Ω) is an regular function of Ω, we may neglect the second term on the right-hand side when calculating h^(ω). As shown in Proposition 10 of [16], the first term is further transformed into

kLLΨ(x,Ω¯)¯E(Ω)Ψ(x,Ω)dx=kF(x,Ω)¯,E(Ω)F(x,Ω)kG(x,Ω)¯,E(Ω)G(x,Ω)+p(Ω)G(x,Ω)¯,ξ0k2ξ0+p(Ω)ξ0k2ξ0¯,G(x,Ω), 4.30

where

F(x,Ω)=12[Ψ(x,Ω)Ψ(x,Ω¯)] 4.31

and

G(x,Ω)=12[Ψ(x,Ω)+Ψ(x,Ω¯)]. 4.32

In the limit of Ωω±i0, the relations F(ω+i0)=−F(ω−i0) and G(ω+i0)=G(ω−i0) hold. Using the formula,

E(ω+i0)+E(ωi0)=kU(x)ω+i0kU(x)kU(x)ωi0kU(x)=2πiU(xc)|U(xc)|δ(xxc), 4.33

we finally obtain

h^(ω)=νkU(xc)p(ω)|U(xc)|[|F(xc,ω+i0)|2+|G(xc,ω+i0)|2], 4.34

where xc=U−1(ω/k) should be read as a function of ω. According to the definition (3.6) of ν, this expression indicates that h^(ω) is negative for all ωσc and we conclude that Q is negative semi-definite on the continuous spectrum. ▪

If Q=ξ¯,Hξ>0 for some ξX+iX, then σσc must not be null and there exists at least one pair of complex eigenvalues, say ωj and ω¯j with Im ωj>0, which correspond to growing and damping modes, respectively. Since H is actually a real self-adjoint operator, the condition Q=ξ,Hξ>0 for some ξX comes to the same conclusion. Thus, we have proved the instability if Q>0 for some ξX.

Conversely, if the flow is unstable due to the presence of several pairs of complex eigenvalues σσc={ωj,ω¯jC|Imωj>0,j=1,2,}, the constant of motion Q must be indefinite in the corresponding eigenspaces, as shown in [6,7,50] for a Hamiltonian function. Indeed, the solution ξ is subject to the following modal decomposition:

ξ=j(ajξ^jeiωjt+bjξ^¯jeiω¯jt)+, 4.35

where ξ^j is the eigenfunction for ωj and aj,bjC are the mode amplitudes which depend on ξ0. As usual, the eigenfunctions {ξ^j,ξ^¯j|j=1,2,} constitute a non-orthogonal basis and its dual basis is provided by the eigenfunctions {w^j,w^¯j|j=1,2,} of L, where w^j=Uξ^j holds from proposition 4.2. This leads to the following orthogonality relations (see Sec. III of [16] for details):

ξ^l,Uξ^j=ξ^¯l,Uξ^j=ξ^¯j,Uξ^j=0(lj),ξ^j,Uξ^j0. 4.36

By substituting this modal decomposition into Q and using the above orthogonality, we obtain

Q|σσc=νj[ajb¯jp(ωj)ξ^j,Uξ^j+c.c.], 4.37

whose sign is clearly indefinite. For example, by setting either (aj,bj)=(1,1) or (aj,bj)=(1,−1), we can make Q|σσc>0. Thus, we have proved that the equation (2.1) is spectrally stable if and only if Q=ξ¯,Hξ0 for all ξX+iX, namely, if and only if Q=ξ,Hξ0 for all ξX.

(d). Extension of search space

Our remaining task is to extend the search space from X to L2. Maximization of Q=ξ,Hξ on L2 is, in practice, more tractable than that on X, since the variational problem λmax=maxQ/ξL22 simply searches the maximum eigenvalue λmax of the self-adjoint operator H. Let us consider the eigenvalue problem (λH)ξ^=0 and rewrite H in the form of

H=νUn=1N(UUIn)+R, 4.38

where R represents the sum of all operators that involve at least one multiplication of G,

R=νl=1N(j=1Nl(UUIj))UG(n=Nl+2N[(UUIn)+UG])U, 4.39

and hence R:L2X. It follows from the condition (3.6) that H has a continuous spectrum for the negative side, min[νUn=1N(UUIn)]λ0. On the other hand, for λ>0, the eigenvalue problem is non-singular and can be rewritten as follows:

ξ^=1λνUn=1N(UUIn)Rξ^. 4.40

Since Rξ^X, this eigenfunction ξ^ inevitably belongs to X. If H has such a positive discrete eigenvalue, the corresponding eigenfunction ξ^X directly proves the instability Q=ξ^,Hξ^>0. Conversely, if Q≤0 for all ξL2, then obviously Q≤0 for all ξXL2. Therefore, we may replace the search space X by L2; thus, the proof of theorem 3.1 is completed.

We can further extend this idea as follows:

Corollary 4.8 —

The stability condition (3.7) in theorem 3.1 can be replaced by

Q=w,Hvw0for allwL2, 4.41

where w=−U′′ξ and, hence,

Hv=νUn=1N[UUIn+UG]. 4.42

Proof. —

Since (ν/U)n=1N(UUIn)<0 follows from the assumptions, the operator Hv is found to be bounded; ∃C>0 such that w,Hvw<CwL22 for all wL2. Suppose that we find a function w^L2 that makes Q positive

0<w^,Hvw^w^L22<C. 4.43

Then, consider a sequence ξmL2, m=1,2,,, that satisfies w^+UξmL20 as m. Since ξm,Hξmw^,Hvw^ as m, ξm,Hξm also becomes positive when m is sufficiently large.

On the other hand, if w,Hvw0 for all wL2, then obviously ξ,Hξ=Uξ,HvUξ0 for all ξL2. ▪

We will actually adopt corollary 4.8 in the subsequent sections, because this variational problem for wL2 is more beneficial than that for ξL2, both analytically and numerically. This fact is evident from the corresponding eigenvalue problem

(λHv)w^=0. 4.44

The operator Hv, which is again written as

Hv=νUn=1N(UUIn)+1UR1U 4.45

has a continuous spectrum, but it is remarkable that the upper edge of this continuous spectrum, λu=max[(ν/U)Πn=1N(UUIn)], is separated from the origin (λu<0). Owing to this property, the variational problem for wL2 is useful for investigating the stability boundary at λ=0 without suffering from any singularity.

5. Comparison with existing results

In this section, we explore several alternative representations of our variational stability criterion by assuming that we have somehow solved Rayleigh's equation under specific conditions. As a consequence of this exploration, we reproduce existing stability theorems and gain a clear-cut understanding of the onset of instability.

(a). Single inflection point

First consider the case of a single inflection point with the condition (UUI)/U′′<0 for all x∈[−L,L], since the opposite case (UUI)/U′′>0 is always stable. According to corollary 4.8, we maximize Q with respect to wL2, where the corresponding eigenvalue problem (4.44) is simply

λw^=UUIUw^+Gw^. 5.1

We are interested in whether a positive eigenvalue λ>0 exists or not, for its existence is the necessary and sufficient condition for instability. By focusing on λ>λu, where λu=min[(UIU)/U]<0, the eigenvalue problem is transformed into

ϕ^k2ϕ^+1λ+(UIU)/Uϕ^=0 5.2

and

ϕ^(L)=ϕ^(L)=0, 5.3

using ϕ^=Gw^. If (5.2) is viewed as ϕ^+f(x,λ,k)ϕ^=0, we can apply Sturm's oscillation theorem [51] to this equation with respect to both λ and k. As the two parameters λ>λu and k>0 increase, f(x,λ,k) decreases everywhere on [−L,L] and hence the general solution ϕ^ becomes less oscillatory (i.e. the interval between any two zeros of ϕ^ expands). When k2≥1/(λ−λu), it becomes non-oscillatory, f(x,λ,k)≤0, and unable to satisfy (5.3). It follows that the eigenvalue λ is bounded by

λ<1k2+λu. 5.4

If k2>λu1, no positive eigenvalue λ>0 exists and, hence, the flow U is stable for such large k.

Since marginal stability occurs at λ=0 in (5.2), we analyse the equation

EIϕ^:=ϕ^+k2ϕ^UUIUϕ^=0. 5.5

If this solution is somehow available, we obtain the following stability criterion.

Corollary 5.1 —

If U(x) satisfies (A 1) and (A 2) and has a single inflection point xI, and ϕc(x) denotes the solution of

EIϕc=0,ϕc(L)=0,ϕc(L)=1, 5.6

then (2.1) is spectrally stable if and only if ϕc(L)≥0.

Proof. —

According to lemma 4.3, ϕc does not have zero on [−L,xI] other than x=−L and has at most one zero on [xI,L]. Note that, by increasing λ from 0, the general solution ϕ^ of (5.2) becomes less oscillatory than ϕc. If ϕc(L)<0, ϕc(x) has one zero on [xI,L] and hence there must be one eigenvalue λ∈[0,1/k2u] for which ϕ^ satisfies (5.2) and (5.3).

Conversely, if ϕc(L)≥0, then ϕc(x) does not have zero on −L<x<L and the solution ϕ^ of (5.2) cannot satisfy the boundary condition (5.3) when λ>0, i.e. there is no positive eigenvalue λ>0. ▪

In particular, we can obtain an analytical solution ϕc for the case of k0 as

ϕc(x)=[U(L)UI][U(x)UI]Lx1[U(s)UI]2ds. 5.7

Then, the necessary and sufficient stability condition ϕc(L)≥0 becomes

1U(s)[U(s)UI]|LL+LLU(s)U2(s)[U(s)UI]ds0, 5.8

which agrees with the result of Rosenbluth & Simon [44]. (Note, the typographical error in the final eqn (4) of this paper, in which w3 should be replaced by w2.)

Another equivalent approach is to solve the equation EIϕc=0 with boundary conditions ϕc(−L)=ϕc(L)=0 and with a derivative jump at x=xI

ϕc(xI+0)=ϕc(xI0)andα:=ϕc(xI+0)ϕc(xI0). 5.9

In other words, we solve EIϕc=αδ(xxI) or

ϕc+G(UUIUϕc)=αg(x,xI), 5.10

where g is defined in (2.3). By introducing a normalization LL(ϕc+k2ϕc)dx=1 for ϕc, we can determine α as

α=1+LLUUIUϕcdx, 5.11

and arrive at the integral equation (5.12). This approach reproduces the stability criterion obtained by Balmforth & Morrison [45].

Corollary 5.2 —

If U(x) satisfies (A 1) and (A 2) and has a single inflection point xI, and ϕc(x) denotes the solution of

ϕc(x)+g(x,xI)+LL[g(x,s)g(x,xI)]U(s)UIU(s)ϕc(s)ds=0, 5.12

then (2.1) is spectrally stable if and only if

1+LLUUIUϕcdx<0, 5.13

Proof. —

According to lemma 4.3, ϕc(x) does not have zero on −L<x<L and its sign should be always positive ϕc(x)>0 due to the normalization. If α>0, we can eliminate this derivative jump by increasing λ from 0, since the general solution ϕ^ of (5.2) becomes less oscillatory than ϕc. Therefore, there must be an eigenvalue λ∈[0,1/k2u] for which ϕ^ satisfies (5.2) and (5.3) without the derivative jump.

Conversely, if α≤0, this derivative jump gets large as λ increases from 0 and, hence, there is no positive eigenvalue λ>0. ▪

(b). Multiple inflection points

Here, we address the problem of multiple inflection points. Recall from proposition 4.4 that neutrally stable eigenmodes may exist only at the frequencies ω=kUIn, n=1,2,…,N. In the same manner as for the case of a single inflection point, we consider the equations for the neutrally stable eigenmodes

EInϕ^c:=ϕ^c+k2ϕ^cUUInUϕ^c=0 5.14

and

ϕ^c(L)=ϕ^c(L)=0, 5.15

for every inflection point xIn, n=1,2,…,N. Since these equations do not have non-trivial solutions for general k, we seek them for some characteristic values of k, in the same spirit as Tollmien's approach [3943].

Proposition 5.3 —

Let U(x) satisfy (A 1) and (A 2). For each inflection point xIn, there is at most one critical wavenumber kn>0 at which the equation

EIn|knϕ^c=0,ϕ^c(L)=ϕ^c(L)=0, 5.16

has a non-trivial solution ϕ^c, where EIn|kn denotes the operator EIn at k=kn.

Proof. —

According to lemma 4.3, the solution ϕc(x) of EInϕc=0 satisfying ϕc(−L)=0 and ϕc′(−L)=1 has at most one zero on −L<xL. This ϕc(x) becomes less oscillatory as k increases from 0 to and eventually has no zero for k2>max[U/(UInU)]. Therefore, there exists at most one value kn of k for which ϕc(L)=0 holds. ▪

Without loss of generality, let us focus on an inflection point xI1 and assume that there is a critical wavenumber k1>0 for it. Namely, we have a solution w^c=ϕ^c+k12ϕ^cL2 that satisfies

w^c+UUI1UG|k1w^c=0,or(L|k1UI1)w^c=0. 5.17

Now, we again invoke corollary 4.8 and consider the self-adjoint eigenvalue problem (4.44). The above neutrally stable eigenfunction w^c clearly corresponds to the marginally stable eigenfunction (λ=0) of (4.44) at k=k1, namely, Hv|k1w^c=0.

Let us continuously change the parameter k in the neighbourhood of k1 and investigate how an eigenvalue λ and an eigenfunction w^ deviate from λ=0 and w^=w^c, respectively. By differentiating the identity,

0=LLw^(λHv)w^dx, 5.18

with respect to k and setting k=k1, we obtain

0=LLw^c(λk|k1Hvk|k1)w^cdx=LLw^c[λk|k1νULk|k1(UI1UI2)(UI1UI3)(UI1UIN)]w^cdx, 5.19

where (5.17) has been used. Since

Lk=UGk=2kUGG, 5.20

we obtain

λk|k1w^cL22=2k1ν(UI1UI2)(UI1UI3)(UI1UIN)ϕ^cL22. 5.21

Similar relations are available for the other critical wavenumbers k2,k3,…,kN if they exist. In the view of condition (3.6), one can distinguish the sign of ∂λ/∂k|kn from (5.21) as follows;

sgnλk|kn=sgn[U(xIn)U(xIn)]=sgn[(U2)(xIn)], 5.22

which agrees with Tollmien and Lin's result [3941]. In other words, if the absolute value of the background vorticity |U′(x)| has a local maximum (or minimum) at x=xIn, then a positive eigenvalue λ>0 emerges at k=kn as k decreases (or increases).

We note that there is no positive eigenvalue λ>0 of (4.44) in the limit of k. As k continuously changes from to 0, the number of positive eigenvalues increases (or decreases) by one when k passes through kn that is associated with the inflection point xIn satisfying (U2)′′(xIn)<0 (or >0). We can summarize these facts into the following stability criterion.

Corollary 5.4 —

Let U(x) satisfy (A 1) and (A 2). Suppose that, for every inflection points xIn, n=1,2,…,N, the critical wavenumbers kn> 0, n=1,2,…,N, are either solved or proven to be non-existent according to proposition 5.3. Then, equation (2.1) is spectrally unstable if and only if N+N>0, where

N+: number of the critical wavenumbers kn that satisfy k<kn and (U2)′′(xIn)<0,

N: number of the critical wavenumbers kn that satisfy k<kn and (U2)′′(xIn)>0.

When N+N is positive, it corresponds to the number of positive eigenvalues λ of (4.44). This number cannot be greater than the number of the inflection points xIn satisfying (U2)′′(xIn)<0, i.e. the number of local maxima of |U′(x)|.

A similar result to corollary 5.4 is shown by Lin [42,43] as a rigorous justification of Tollmien's method. While he treats a larger class of flows than ours, his criterion is sufficient but not necessary for instability in the presence of multiple inflection points [43]. Balmforth & Morrison [45] have also discussed the case of multiple inflection points in the same manner as corollary 5.2, where the derivative jump αn is evaluated for each inflection point xIn and then αn<0 (or αn>0) corresponds to k<kn (or k>kn). However, in this work the importance of sgn(U2)′′(xIn) was not observed.

6. Numerical tests

Finally, we exhibit numerical results to illustrate the practicability of our method. For three velocity profiles U(x), we compare the results of two different numerical codes: one code solves the Rayleigh equation (2.4) directly for complex eigenvalues c=ω/kC, while the other code solves for the eigenvalues λ12,…, of the self-adjoint operator Hv in descending order. In fact, the maximum eigenvalue λ1 can be more easily determined by the variational problem, λ1=maxQ/wL22, and the flow U(x) is spectrally unstable if and only if λ1>0 (see corollary 4.8).

The first example is

U(x)=tanh(x),x[,], 6.1

which is well known to be unstable for 0<k<1. The result is shown in figure 2, where we also plot λ~1=maxQ/ξL22 for comparison (the damping eigenvalue Im c<0 is not plotted since its presence is trivial). As expected from the results of §4d, λ=0 is the upper edge of the continuous spectrum of H. Since the eigenfunction ξ^1 becomes singular, i.e. ξ^1L2, as λ~1+0, the curve of λ~1 is tangent to the marginal line λ=0 and the critical wavenumber k=1 is not so evident. On the other hand, the upper edge of the continuous spectrum of Hv is less than zero, λu=max[tanh(x)/tanh(x)]=0.5<0, and hence the maximum eigenvalue λ1 of Hv smoothly intersect with λ=0 at k=1 in figure 2. Thus, for the purpose of drawing the stability boundary, the variational principle with respect to the norm wL2 is seen to be numerically efficient and accurate.

Figure 2.

Figure 2.

Growth rate Im c (where Re c≡0), λ1=maxQ/wL22 and λ~1=maxQ/ξL22 versus wavenumber k for the shear flow U(x)=tanh(x).

The second example is

U(x)=x+5x3+1.62tanh[4(x0.5)],x[1,1], 6.2

which was previously addressed by Balmforth & Morrison [45]. This flow has three inflection points

xI1=0.069,UI1=1.65,xI2=0.622,UI2=2.55andxI3=0.665,UI3=3.07,} 6.3

at which (U2)′′ is positive, negative and positive, respectively. Only for xI2 and xI3, do the critical wavenumbers k2≃1.2 and k3≃0.4 exist. As predicted in corollary 5.4, the instability occurs only for finite wavenumbers k3<k<k2. In figure 3, the positive signature of the maximum eigenvalue λ1 certainly agrees with this unstable regime. In practice, our variational approach can directly prove the instability at a fixed k without knowing the existence of nor the values k1,k2 and k3.

Figure 3.

Figure 3.

(a,b) Growth rate (Im c) and phase speed (Re c) versus wavenumber k for the shear flow U(x)=x+5x3+1.62tanh[4(x0.5)]. The dashed line is λ1=maxQ/wL22.

The third example is

U(x)=x0.02+sin[8(x0.02)]16,x[1,1], 6.4

which has five inflection points

xI1=0.765,UI1=0.785,xI2=0.373,UI2=0.393,xI3=0.020,UI3=0.0,xI4=0.413,UI4=0.393andxI5=0.805,UI5=0.785.} 6.5

For this example, there exist three critical wavenumbers k1, k3 and k5 for the inflection points xI1,xI3 and xI5, all of which have (U2)′′ negative. Therefore, three unstable eigenvalues emerge at k1,k3 and k5 with different phase speeds UI1,UI3 and UI5, respectively. Thus, three eigenvalues λ12 and λ3 of our variational problem completely predict the onsets of instabilities, as shown in figure 4.

Figure 4.

Figure 4.

(a) Growth rate (Im c) and (b) phase speed (Re c) versus wavenumber k for the shear flow U(x)=x0.02+sin[8(x0.02)]/16. The dashed lines are eigenvalues λ123 of Hv.

7. Summary

We have investigated the linear stability of inviscid plane parallel shear flow (Rayleigh's equation) as a typical example of an infinite-dimensional and non-self-adjoint eigenvalue problem that originates upon linearizing a Hamiltonian system. By assuming monotonicity and analyticity of the shear profile, a necessary and sufficient condition for spectral stability was obtained in the form of a variational criterion (theorem 3.1). Our theory is based on (i) the existence of the infinite number of constants of motion Qf (proposition 3.2), whose definition includes an arbitrary real polynomial f(c) and (ii) the rigorous derivation of the Kreĭn signature (i.e. the signature of δ2H) for the continuous spectrum. Since the energy, δ2H=Qf with f(c)=c, is generally indefinite due to the presence of both positive and negative energy modes, we have chosen a special f(c) such that Q=Qf becomes negative semi-definite Q|σc≤0 (proposition 4.7) for the neutrally stable spectrum σcR, which is mostly the continuous spectrum in the present case. Then, a positive signature of the quadratic form Q=ξ,Hξ implies existence of an unstable eigenmode. Since H is self-adjoint, we were able to prove instability must occur if some test function ξ (virtual displacement) exists that makes Q positive, which is analytically and numerically easier to do than solving Rayleigh's equation. Moreover, the singularity at the stability boundary (due to the continuous spectrum) was shown to be removed technically by maximizing Q with respect to the vorticity disturbance wL2, instead of the displacement ξL2. However, we remark that, unlike the Rayleigh–Ritz method, neither maxQ/ξL22 nor maxQ/wL22 are quantitatively related to the maximum growth rate of instability.

Our variational criterion is an improvement of previous sufficient stability criteria [33,11]. Given that Rayleigh's equation has been solved under a specific condition, we have also reproduced the earlier results of the Nyquist method [44,45] and Tollmien's analysis of the neutral modes [3941].

In this paper, we have imposed the assumptions (A 1) and (A 2) on velocity profile U(x) to simplify the discussion. The relaxation of these assumptions is possible to some extent, but it would be difficult to overcome the following difficulties: (i) if analyticity is not assumed and U′′ is only continuous, special care is needed for piecewise-linear regions of U. In such a region, say [x1,x2], all points are regarded as inflection points and we expect that the variational problem would become a minimax problem like minxI[x1,x2]maxξL2Q>0, which is not so analytically tractable. (ii) If monotonicity is not assumed, a serious difficulty arises when the sign of U′′ is not identical at the locations of multiple critical layers for a phase speed c=ω/kR. Since at this frequency ω=kc belongs to degenerate multiple continuous spectra whose signature is indefinite, our technique for constructing Q breaks down.

In conclusion, we note that our variational approach will be applicable to rather simple equilibrium profiles which are free from the above difficulties. However, there is a large class of fluid and plasma systems with existing sufficient stability criteria (e.g. MHD [52,53] with flow) that have Kreĭn-like signature (or action–angle variables) for a continuous spectrum. This is the key ingredient needed for constructing the quadratic form. Thus, our techniques are available for a large class of applications governed by other dynamical systems. We will report our additional results in future publications.

Acknowledgements

The authors would like to thank Z. Yoshida, Y. Fukumoto and G. Hagstrom for fruitful discussions.

Appendix A. Brief review of the isovortical variation

Consider the vorticity equation ∂tw=curl(u×w) where w=curl u. For a given displacement vector field ξ, let us generate a variation w0wϵ around a steady-state w0=curl u0 by solving

ϵwϵ=curl(ξ×wϵ),wϵ|ϵ=0=w0, A 1

in terms of a small parameter ϵR. This so-called isovortical variation automatically preserves the Kelvin's circulation law (or the topology of the vorticity w0), where we usually expand wϵ as

wϵ=w0+ϵcurl(ξ×w0)+ϵ22curl[ξ×curl(ξ×w0)]+O(ϵ3) A 2
=:w0+ϵδw0+ϵ22δ2w0+O(ϵ3). A 3

For the steady-state w0=(0,0,U′(x)) and two-dimensional motion ξ=(ξx,ξy,0) discussed in this paper, the first variation δw0=curl(ξ×w0) is reduced to δw0=(0,0,−U′′ξx). Then, the second variation of the Hamiltonian H=|u|2/2d3x around u0 results in (3.3).

If the dynamics of ξ is taken into account, it must be related to δu0=P(ξ×w0) (where P is the projection operator to the solenoidal vector field) by

tξ+(u0)ξ(ξ)u0=δu0+O(ϵ) A 4

(see Sec. 6 of [54] or [52,53]). When O(ϵ) is neglected in the linear analysis, this is indeed the adjoint equation of the linearized vorticity equation. In this paper, its x component ∂tξx+Uyξx=δu0x corresponds to (3.1).

Appendix B. Proof of σcσI being the continuous spectrum

Let ω=kcσcσI, namely, ω=kU(xc) and W(ω±i0)≠0. Using lemma 4.3, one finds Φ<(xc,ω)≠0 and Φ>(xc,ω)≠0, which implies that there exists ξ0X+iX such that Φ(xc,ω±i0)≠0. Then, Ξ(x,Ω)=(ΩkL)1ξ0=kΦ(x,Ω)/(ΩkU) becomes singular Ξ(x,ω±i0)∉X+iX as Ωω±i0, namely, the resolvent operator (ωkL)1 is unbounded.

Moreover, the range of ωkL is dense in X+iX as shown below. Therefore, ω is not in the residual spectrum but the continuous spectrum.

Lemma B.1 —

Let ω=kcσcσI. For any given η0X+iX and ϵ>0,

ηX+iXs.t.(ωkL)ηη0X+iX<ϵ. B 1

Proof. —

Define a neighbourhood of xc as Bϵ1:=[xcϵ1,xc+ϵ1] with 0<ϵ1R. Let us consider ξ0=GθX+iX where θL2+iL2 is given by

θ:={η0+k2η0on[L,L]Bϵ1,0onBϵ1. B 2

As in (4.11), we generate Φ(x,Ω) with this ξ0 and define ϕ*(x,ω) as

ϕ(x,ω)=Φ(xc,ωi0)Φ(x,ω+i0)Φ(xc,ω+i0)Φ(x,ωi0)Φ(xc,ωi0)Φ(xc,ω+i0)H01+iH01, B 3

which indeed exists due to proposition 4.5. Then, ϕ* satisfies ϕ*(xc,ω)=0 and, from (4.6),

ϕ+k2ϕUcUϕ=1kθ. B 4

In the neighbourhood Bϵ1, this ϕ* must be the regular Frobenius series solution ϕ*(x,ω)=constΦ1(x,ω) since the right-hand side of (B.4) is zero and ϕ*(xc,ω)=0 (recall (4.18), where Φ1(xc,ω)=0 and Φ2(xc,ω)=1). If we set

η=ϕcU, B 5

then this η is still regular on Bϵ1 and hence ηH01+iH01. Moreover, ηX+iX because ϕ*′′∈L2+iL2 follows from (B.4). The relation (B.4) is transformed into (ωkL)η=Gθ, which implies (ωkL)ηX+iX. Therefore, we obtain (ωkL)ηη0=G(θ+η0k2η0), where

θ+η0k2η0L2+iL22ϵ1supx[L,L]|η0k2η0|. B 6

By adopting the definition X+iX:=G1L2+iL2 for simplicity, the required result is obtained by making ϵ1 small such that 2ϵ1supx[L,L]|η0k2η0|<ϵ. ▪

Funding statement

This work was supported by a grant-in-aid for scientific research from the Japan Society for the Promotion of Science (no. 25800308). P.J.M. was supported by U.S. Dept. of Energy Contract no. DE-FG05-80ET-53088.

References

  • 1.Drazin PG, Howard LN. 1966. Hydrodynamic stability of parallel flow of inviscid fluid. Adv. Appl. Mech. 9, 1–89 (doi:10.1016/S0065-2156(08)70006-1) [Google Scholar]
  • 2.Bernstein IB, Frieman EA, Kruskal MD, Kulsrud RM. 1958. An energy principle for hydromagnetic stability problems. Proc. R. Soc. Lond. A 244, 17–40 (doi:10.1098/rspa.1958.0023) [Google Scholar]
  • 3.Morrison PJ. 1998. Hamiltonian description of the ideal fluid. Rev. Mod. Phys. 70, 467–521 (doi:10.1103/RevModPhys.70.467) [Google Scholar]
  • 4.Rayleigh JWS. 1945. (reissue of 1877 version). The theory of sound 2nd edn. New York, NY: Dover. [Google Scholar]
  • 5.Kreĭn MG. 1950. A generalization of some investigations on linear differential equations with periodic coefficients. Dokl. Akad. Nauk SSSR A 73, 445–448. [Google Scholar]
  • 6.Moser JK. 1958. New aspects in the theory of stability of Hamiltonian systems. Comm. Pure Appl. Math. 11, 81–114 (doi:10.1002/cpa.3160110105) [Google Scholar]
  • 7.MacKay R. 1986. Stability of equilibria of Hamiltonian systems. In Nonlinear Phenomena and Chaos (ed. Sarkar S.), pp. 254–270 Bristol, UK: Adam Hilger. [Google Scholar]
  • 8.Morrison PJ, Kotschenreuther M. 1990. The free energy principle, negative energy modes and stability. In Nonlinear World: IV International Workshop on Nonlinear and Turbulent Processes in Physics (eds Bar'yakhtar VG, Chernousenko VM, Erokhin NS, Sitenko AB, Zakharov VE.), pp. 910–932 Singapore: World Scientific. [Google Scholar]
  • 9.Oberman C, Kruskal M. 1965. Some constants of the linearized motion of Vlasov plasmas. J. Math. Phys. 6, 327–335 (doi:10.1063/1.1704284) [Google Scholar]
  • 10.Case KM. 1965. Constants of the linearized motion of Vlasov plasmas. Phys. Fluids 8, 96–101 (doi:10.1063/1.1761106) [Google Scholar]
  • 11.Barston EM. 1991. On the linear stability of inviscid incompressible plane parallel flow. J. Fluid Mech. 233, 157–163 (doi:10.1017/S0022112091000435) [Google Scholar]
  • 12.Morrison PJ, Pfirsch D. 1992. Dielectric energy versus plasma energy, and Hamiltonian action–angle variables for the Vlasov equation. Phys. Fluids B 4, 3038–3057 (doi:10.1063/1.860415) [Google Scholar]
  • 13.Morrison PJ. 2000. Hamiltonian description of Vlasov dynamics: action–angle variables for the continuous spectrum. Trans. Theory Stat. Phys. 29, 397–414 (doi:10.1080/00411450008205881) [Google Scholar]
  • 14.Balmforth NJ, Morrison PJ. 2002. Hamiltonian description of shear flow. In Large-scale atmosphere–ocean dynamics II (eds Norbury J, Roulstone I.), pp. 117–142 Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 15.Morrison PJ. 2003. Hamiltonian description of fluid and plasma systems with continuous spectra. In Nonlinear processes in geophysical fluid dynamics (eds Velasco Fuentes OU, Sheinbaum J, Ochoa J.), pp. 53–69 Dordrecht, The Netherlands: Kluwer. [Google Scholar]
  • 16.Hirota M, Fukumoto Y. 2008. Energy of hydrodynamic and magnetohydrodynamic waves with point and continuous spectra. J. Math. Phys. 49, 083101 (doi:10.1063/1.2969275) [Google Scholar]
  • 17.Hirota M. 2010. Action–angle representation of waves and instabilities in flowing plasmas. J. Plasmas Fusion Res. SERIES 9, 463–470. [Google Scholar]
  • 18.Hagstrom GI, Morrison PJ. 2010. On Krein-like theorems for noncanonical Hamiltonian systems with continuous spectra: application to Vlasov–Poisson. Trans. Theory Stat. Phys. 39, 466–501 (doi:10.1080/00411450.2011.566484) [Google Scholar]
  • 19.Morrison PJ, Hagstrom GI. 2014. Continuum Hamiltonian Hopf bifurcation I. In Nonlinear physical systems—spectral analysis, stability and bifurcations (eds Kirillov O, Pelinovsky D.), pp. 247–282 London, UK: ISTE; Hoboken, NJ: John Wiley. [Google Scholar]
  • 20.Hagstrom GI, Morrison PJ. 2014. Continuum Hamiltonian Hopf bifurcation II. In Nonlinear physical systems—spectral analysis, stability and bifurcations (eds Kirillov O, Pelinovsky D.), pp. 283–310 London, UK: ISTE; Hoboken NJ: John Wiley. [Google Scholar]
  • 21.Grillakis M. 1990. Analysis of the linearization around a critical point of an infinite dimensional Hamiltonian system. Comm. Pure Appl. Math. 43, 299–333 (doi:10.1002/cpa.3160430302) [Google Scholar]
  • 22.Kapitula T, Kevrekidis PG, Sandstede B. 2004. Counting eigenvalues via the Krein signature in infinite-dimensional Hamiltonian systems. Physica D 195, 263–282 (doi:10.1016/j.physd.2004.03.018) [Google Scholar]
  • 23.Kirillov O, Pelinovsky D. 2014. Nonlinear physical systems—spectral analysis, stability and bifurcations. London, UK: ISTE; Hoboken, NJ: John Wiley. [Google Scholar]
  • 24.Rayleigh JWS. 1880. On the stability, or instability, of certain fluid motions. Proc. Lond. Math. Soc. 11, 57–72. [Google Scholar]
  • 25.Fjø rtoft R. 1950. Application of integral theorems in deriving criteria of stability for laminar flows and for the baroclinic circular vortex. Geofys. Publ. 17, 1–52. [Google Scholar]
  • 26.Arnold VI. 1966. An a priori estimate in the theory of hydrodynamic stability. Izv. Vyssh. Uchebn. Zaved. Mat. [Sov. Math. J.] 5, 3–5. [Google Scholar]
  • 27.Arnold VI. 1966. Sur la geometrie differentielle des groupes de Lie de dimension infinie et ses applications a l'hydrodynamique des fluides parfaits Ann. Inst. Fourier 16, 319–361 (doi:10.5802/aif.233) [Google Scholar]
  • 28.Arnold VI. 1969. The Hamiltonian nature of the Euler equations in the dynamics of a rigid body and of an ideal fluid. Usp. Mat. Nauk. [Sov. Math. Usp.] 24, 225–226 (doi:10.5802/air.233) [Google Scholar]
  • 29.Kruskal MD, Oberman C. 1958. On the stability of plasma in static equilibrium. Phys. Fluids 1, 275–280 (doi:10.1063/1.1705885) [Google Scholar]
  • 30.Hazeltine RD, Holm DD, Marsden JE, Morrison PJ. 1984. Generalized Poisson brackets and nonlinear Liapunov stability-application to reduced MHD. In International Conference on Plasma Physics Proceedings 1 (eds Tran MQ, Sawley ML.), p. 203 Lausanne: Ecole Polytechnique Federale de Lausanne. [Google Scholar]
  • 31.Holm DD, Marsden JE, Ratiu T, Weinstein A. 1985. Nonlinear stability of fluid and plasma equilibria. Phys. Rep. 123, 1–116 (doi:10.1016/0370-1573(85)90028-6) [Google Scholar]
  • 32.Morrison PJ, Eliezer S. 1986. Spontaneous symmetry breaking and neutral stability in the noncanonical Hamiltonian formalism. Phys. Rev. A 33, 4205–4214 (doi:10.1103/PhysRevA.33.4205) [DOI] [PubMed] [Google Scholar]
  • 33.Arnold VI. 1978. Mathematical methods of classical mechanics. Berlin, Germany: Springer. [Google Scholar]
  • 34.Morrison PJ, Pfirsch D. 1989. Free-energy expressions for Vlasov equilibria. Phys. Rev. A 40, 3898–3910 (doi:10.1103/PhysRevA.40.3898) [DOI] [PubMed] [Google Scholar]
  • 35.Morrison PJ, Pfirsch D. 1990. The free energy of Maxwell–Vlasov equilibria. Phys. Fluids B 2, 1105–1113 (doi:10.1063/1.859246) [Google Scholar]
  • 36.Morrison PJ. 1982. Poisson brackets for fluids and plasmas. In Mathematical methods in hydrodynamics and integrability in related dynamical systems (eds Tabor M, Treve Y.), pp. 13–46 AIP Conference Proceedings no. 88 New York, NY: AIP. [Google Scholar]
  • 37.Maddocks J, Sachs R. 1993. On the stability of KdV multi-solitons. Comm. Pure Appl. Math. 46, 867–901 (doi:10.1002/cpa.3160460604) [Google Scholar]
  • 38.Bottman N, Nivala M, Deconinck B. 2011. Elliptic solutions of the defocusing NLS equation are stable. J. Phys. A 44, 285201 (doi:10.1088/1751-8113/44/28/285201) [Google Scholar]
  • 39.Tollmien W. 1935. Ein allgemeines Kriterium der Instabilitat laminarer Gescgwindigkeits-verteilungen. Nachr. Wiss Fachgruppe, Göttingen, Math. Phys. 1, 79–114. [Google Scholar]
  • 40.Lin CC. 1945. On the stability of two-dimensional parallel flows Part II—stability in an inviscid fluid. Quart. Appl. Math. 3, 218–234. [Google Scholar]
  • 41.Lin CC. 1955. The theory of hydrodynamic stability. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 42.Lin Z. 2003. Instability of some ideal plane flows. SIAM J. Math. Anal. 35, 318–356 (doi:10.1137/S0036141002406266) [Google Scholar]
  • 43.Lin Z. 2005. Some recent results on instability of ideal plane flows. Contemp. Math. 371, 217–229 (doi:10.1090/conm/371/06857) [Google Scholar]
  • 44.Rosenbluth MN, Simon A. 1964. Necessary and sufficient conditions for the stability of plane parameter inviscid flow. Phys. Fluids 7, 557–558 (doi:10.1063/1.1711237) [Google Scholar]
  • 45.Balmforth NJ, Morrison PJ. 1999. A necessary and sufficient instability condition for inviscid shear flow. Stud. Appl. Math. 102, 309–344 (doi:10.1111/1467-9590.00113) [Google Scholar]
  • 46.Chen XL, Morrison PJ. 1991. A sufficient condition for the ideal instability of shear flow with parallel magnetic field. Phys. Fluids B 3, 863–865 (doi:10.1063/1.859841) [Google Scholar]
  • 47.Rosencrans SI, Sattinger DH. 1966. On the spectrum of an operator occurring in the theory of hydrodynamic stability. J. Math. Phys. 45, 289–300. [Google Scholar]
  • 48.Case KM. 1960. Stability of inviscid plane Couette flow. Phys. Fluids 3, 143–148 (doi:10.1063/1.1706010) [Google Scholar]
  • 49.Drazin PG, Reid WH. 1981. Hydrodynamic stability. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 50.Whittaker ET. 1947. A treatise on the analytical dynamics of particles and rigid bodies, 4th edn. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 51.Ince EL. 1926. Ordinary differential equations. New York, NY: Dover. [Google Scholar]
  • 52.Frieman E, Rotenberg M. 1960. On hydromagnetic stability of stationary equilibria. Rev. Mod. Phys. 32, 898–902 (doi:10.1103/RevModPhys.32.898) [Google Scholar]
  • 53.Andreussi T, Morrison PJ, Pegoraro F. 2013. Hamiltonian magnetohydrodynamics: Lagrangian, Eulerian, and dynamically accessible stability-theory. Phys. Plasmas 20, 092104 (doi:10.1063/1.4819779) [Google Scholar]
  • 54.Fukumoto Y, Hirota M. 2008. Elliptical instability of a vortex tube and drift current induced by it. Phys. Scr. T132, 014041 (doi:10.1088/0031-8949/2008/T132/014041) [Google Scholar]

Articles from Proceedings. Mathematical, Physical, and Engineering Sciences / The Royal Society are provided here courtesy of The Royal Society

RESOURCES