Galloping instability of viscous shock waves

https://doi.org/10.1016/j.physd.2008.03.008Get rights and content

Abstract

Motivated by physical and numerical observations of time oscillatory “galloping”, “spinning”, and “cellular” instabilities of detonation waves, we study Poincaré–Hopf bifurcation of traveling-wave solutions of viscous conservation laws. The main difficulty is the absence of a spectral gap between oscillatory modes and essential spectrum, preventing standard reduction to a finite-dimensional center manifold. We overcome this by direct Lyapunov–Schmidt reduction, using detailed pointwise bounds on the linearized solution operator to carry out a nonstandard implicit function construction in the absence of a spectral gap. The key computation is a space-time stability estimate on the transverse linearized solution operator reminiscent of Duhamel estimates carried out on the full solution operator in the study of nonlinear stability of spectrally stable traveling waves.

Introduction

Motivated by physical and numerical observations of time-oscillatory “galloping” or “pulsating” instabilities of detonation waves [76], [9], [22], [76], [3], [1], [19], [20], we study the Poincaré–Hopf bifurcation of viscous shock waves in one spatial dimension. Our main result is to obtain a rigorous (nonlinear) characterization in terms of spectral information. The complementary problem of verifying this spectral information has been studied already in [65], [67],where it was shown that the transition from stability to instability of viscous detonation waves generically involves a Poincaré–Hopf type bifurcation in the spectral configuration of the linearized operator about the wave. By essentially the same analysis we obtain also a corresponding multi-dimensional result applying to planar viscous shock fronts traveling in a cylinder of finite cross-section, with artificial periodic boundary conditions. This gives a simplified mathematical model for the time-oscillatory instabilities observed in detonation waves moving within a duct, which, besides the longitudinal galloping instabilities described above, also include transverse “cellular” or “spinning” instabilities. The method of analysis appears to be of general application, in particular, with suitable elaboration, to extend to the original motivating case of viscous detonation waves of the reactive Navier–Stokes equations with physical viscosity.

From a mathematical standpoint, the main issue in our analysis is that the linearized equations for a standing shock wave have no spectral gap between convective modes corresponding to the essential spectrum of the linearized operator L of the wave and the oscillatory modes corresponding to pure imaginary point spectrum. This prohibits the usual PDE analysis by center manifold reduction to a finite-dimensional subspace; likewise, at the linearized level, decay to the center subspace is at a time-algebraic rather than a time-exponential rate, so that oscillatory and other modes are strongly coupled. We overcome this difficulty by the introduction of a nonstandard Implicit Function Theorem framework suitable for an infinite-dimensional Lyapunov–Schmidt reduction in the absence of the spectral gap (Section 2.1), augmented by a finite-dimensional “weak” Implicit Function Theorem suitable for finite-dimensional bifurcation analysis in situations of limited regularity (Section 2.3). The latter, based on the Brouwer Fixed-point Theorem rather than the standard Contraction-mapping construction, seems of interest in its own right.

The study of bifurcation from stability is a natural followup to our previous work on the stability of viscous shock and detonation waves; see, e.g., [97], [73], [74], [40], [65], [67], [66]. Interestingly, the key estimate needed to apply our bifurcation framework turns out to be a space-time stability estimate on the transverse linearized solution operator (meaning the part complementary to oscillatory modes) quite similar to estimates on the full solution operator arising in the stability analysis of viscous traveling waves. To carry out this estimate requires rather detailed pointwise information on the Green function of the linearized equations for the wave; see Sections 1.6.4 Refined analysis, 5 Proof of. Indeed, we use the full power of the pointwise semigroup techniques developed in [97], [71], [73].

The equations of compressible gas dynamics in one spatial dimension, like many equations in continuum mechanics, take the form of hyperbolic conservation laws ut+f(u)x=0. (Euler equations) or hyperbolic–parabolic conservation laws ut+f(u)x=(B(u)ux)x. (Navier–Stokes equations), depending whether second-order transport effects–in this case, viscosity and heat-conduction–are neglected or included. Here, x,tR1 are spatial location and time, uRn is a vector of densities of conserved quantities — mass, momentum, and energy in the case of gas dynamics, fRn is a vector of corresponding fluxes, and BRn×n is a matrix of transport coefficients.

Such equations are well known to support traveling-wave solutions of the form u(x,t)=ū(xst){uxst0,u+xst>0 (discontinuous) and u(x,t)=ū(xst),limz±ū(z)=u± (smooth), respectively, known as ideal and viscous shock waves. These waves may be observed physically; indeed, the corresponding physical objects appear to be quite stable [5].

Similarly, the equations of reacting, compressible gas dynamics take the form ut+f̃(u,z)xqKϕ(u)z=0zt+(v(u,z)z)x+Kϕ(u)z=0 (reactive Euler, or Zeldovich–von Neumann–Doering equations (ZND)) or ut+f̃(u,z)xqKϕ(u)z=(B̃(u,z)ux)x,zt+(v(u,z)z)x+Kϕ(u)z=(D̃(u,z)zx)x (reactive Navier–Stokes equations (rNS)), where x,tR1, uRn is as before, zRr is a vector of mass fractions of different reactant species, vR1 is fluid velocity, and f̃, B̃ and D̃ are flux vectors and transport matrices depending (through z) on the chemical makeup of the gas, with f̃(u,0)=f(u), B̃(u,0)=B(u). The matrix K models reaction dynamics, q is a constant heat release coefficient (q>0 corresponding to an exothermic reaction), and ϕR1 is an ignition function serving to “turn on” the reaction: zero for temperatures below a certain critical temperature and positive for temperatures above.

Eqs. (1.5), (1.6) support traveling wave solutions u(x,t)=ū(xst) analogous to (1.3), (1.4), known as ideal and viscous detonation waves, which may likewise be observed physically. However, in contrast to the shock wave case, detonations appear to be rather unstable. What is typically observed is not the planar, steadily progressing solution (1.7), but rather a nearby solution varying time-periodically about (1.7): that is, an apparent bifurcation with exchange of stability. In the typical experimental setting of a detonation moving along a duct, these may be longitudinal “galloping”, or “pulsating”, instabilities for which the planar structure is maintained, but the form of the profile changes time-periodically, or they may be transverse instabilities in which the front progresses steadily in the longitudinal (i.e., axial) direction, but develops time-oscillatory structure in the transverse (cross-sectional) directions of a form depending on the geometry of the cross-section: “cellular” instabilities for a polygonal (e.g. rectangular) cross-section; “spinning” instabilities for a circular cross-section, with one or more “hot spots”, or “combustion heads”, moving spirally along the duct.

Stability of shocks and detonations may be studied within a unified mathematical framework; see, e.g., [17], [18], [56], [57], [16], [8], [21], [61], [91], [58], [68], [69], [70], [23], [77], [14] in the inviscid case (1.1), (1.5), and [86], [30], [31], [53], [54], [75], [62], [64], [32], [87], [27], [97], [11], [12], [13], [10], [59], [93], [94], [95], [96], [72], [73], [74], [33], [34], [35], [38], [4], [65], [67], [50], [41], [42], [6], [81], [24], [66] in the viscous case (1.2), (1.6). In particular, stability in each case has been shown to reduce to spectral considerations accessible by standard normal-modes analysis (see Section 1.2). However, they are studied with different motivations, and yield different results. Historically, it appears that instability motivated the physical study, which has focused on the detonation case and spectral stability criteria [5]. By contrast, the mathematical study has focused on stability and the somewhat simpler shock wave case, with the main (difficult!) issue being to establish full nonlinear stability assuming that spectral stability holds in the form of a suitable “Lopatinski” or “Evans” condition [70], [77], [93], [34]. Regarding bifurcation, there is strong evidence that the detonation instabilities described above correspond to a Poincaré–Hopf bifurcation (indeed, our description above benefits much by hindsight). In particular, Erpenbeck [18] and Bourlieux, Majda, and Roytburd [9] have carried out formal asymptotics in support of this viewpoint for the one-dimensional (galloping, or pulsating) case, in the context of the ZND equations. More recently, Kasimov and Stewart [52] have carried out a definitive study in the multidimensional case by numerical linearized normal modes analysis, also for the ZND equations, in which they demonstrate that the onset of such instabilities indeed corresponds to crossing of the imaginary axis by a conjugate pair of eigenvalues toward the unstable (positive real part) side, with, moreover, excellent correspondence between observed nonlinear oscillations and the associated normal modes. Depending whether the maximally unstable mode is in the longitudinal or transverse direction, the oscillatory behavior is seen to be of galloping or spinning (cellular) type. However, up to now, no rigorous analysis of these phenomena has been carried out. To fill this gap is the object of the present study.

The main difficulty in the analysis of shock or detonation waves is the lack of a spectral gap between stationary or oscillatory modes and the essential spectrum of the linearized operator for the wave. In the inviscid, hyperbolic case (1.1), (1.5), the linearized problem has essential spectrum filling the imaginary axis. In the viscous case (1.2), (1.6), the essential spectrum is, in contrast, tangent at the origin to the imaginary axis.

For example, without loss of generality taking a standing-wave solution u=ū(x), s=0 (i.e., working in coordinates moving with the wave), and linearizing (1.2) about ū, we obtain, taking BI for simplicity, ut=Luuxx(Au)x,Adf(ū(x)), where the asymptotically constant-coefficient operator L is what we have called the linearized operator for the wave. A standard result of Henry ([37], Theorem A.2, chapter 5) on the spectrum of operators with asymptotically constant coefficients asserts that the rightmost (i.e. largest real part) envelope of the essential spectrum with respect to any Lp is the envelope of the union of the rightmost envelopes of the spectra of the limiting, constant-coefficient operators at ±: in this case, L±x2A±x,A±df(u±). The L2 spectra of L± may be computed by Fourier transform to be the curves traced out by dispersion relations λj(ξ)iaj±ξξ2,ξR1, where aj± are the eigenvalues of A±. These eigenvalues are real and nonzero under the physical assumptions that (1.1) be hyperbolic in a neighborhood of u± and the shock be noncharacteristic. Likewise [86], L has always an L2 zero-eigenvalue with eigenfunction ū(x), associated with translation of the wave.

In the absence of a spectral gap, standard stability and bifurcation theorems do not apply, and so we must carry out a refined analysis.

The spectral configuration just described translates in the xt plane, in the stable case that there exist no other eigenvalues λ of L in the nonnegative complex half-plane λ0, to the following description developed in [62], [64], [97], [96], [73] of the Green function G(x,t;y) associated with tL. A point source, or delta-function initial datum, originating at y>0 will propagate initially as an approximate superposition of Gaussians with constant mass (total integral), centered along hyperbolic characteristics dx/dt=aj+ determined by the asymptotic system at +. Those propagating in the positive direction will continue out to +; those propagating in the negative direction will continue until they strike the shock layer at approximately x=0, whereupon they will be transmitted and reflected along outgoing characteristic directions dx/dt=ak, ak<0 and dx/dt=ak+, ak+>0. In addition, there will be deposited at the shock layer a certain amount of mass in the stationary eigenmode ū(x) (integral ū(x)dx=u+u0), corresponding to translation of the background wave.

We refer the reader to [97], [73] for further discussion and details. For the present purpose, it suffices to note that the Green function may be modeled qualitatively as the sum of terms K(x,t;y)t1/2e(xyat)2/4t and J(x,t;y)ū(x)errfn((yat)/2t1/2) propagating with noncharacteristic speed a<0, where errfn(z)12πzeξ2dξ, the first modeling moving heat kernels (Gaussian signals), and the second the excitation of the zero-eigenfunction by incoming signals from y0.

From this description, we see explicitly that decaying modes decay time-algebraically and not exponentially, in agreement with the lack of spectral gap. Moreover, the low-frequency/large-time–space behavior is approximately hyperbolic, propagating along characteristics associated with the endstates u±.

Further information may be obtained at the spectral level by the comparison of Evans and Lopatinski determinants D and Δ for the viscous and inviscid problems (1.2), (1.1). The Evans function D(λ), defined as a Wronskian of functions spanning the decaying manifolds of solutions of the eigenvalue equation (Lλ)u=0 associated with L at x+ and x is an analytic function with domain containing {λ0}, whose zeros away from the essential spectrum correspond in location and multiplicity with eigenvalues of L with respect to any Lp. Its behavior is also closely linked with that of the resolvent kernel of L,i.e., the Laplace transform with respect to time of the Green function G; see [2], [27], [97], [98], [93], [94] for history and further details. The corresponding object for the inviscid linearized problem is the Lopatinski determinant Δ(λ) defined in [58], [68], [69], [77], a homogeneous function of degree one: in the one-dimensional case, just linear.

An important relation between these two objects, established in [27], [98], [78], is the low-frequency expansion D(λ)=γΔ(λ)+o(|λ|) quantifying the above observation that low-frequency behavior is essentially hyperbolic, where γ is a constant measuring transversality of the profile ū as a connecting orbit of equilibria u± in the associated traveling-wave ODE, nonvanishing for transversal connections.

A corresponding expansion DrNS(λ)=γΔCJ(λ)+o(|λ|) holds for the Evans function DrNS associated with detonation wave solutions of (1.6), where ΔCJ denotes the Lopatinski condition, not for (1.5), but for the still simpler inviscid–instantaneous-reaction-rate Chapman–Jouguet model in which detonations are modeled by piecewise constant solutions (1.3) across which combustion proceeds instantaneously; see [93], [65], [67], [50], [14] for further details. That is, the low-frequency behavior is not only hyperbolic but “instantaneous”, with small-scale details of the profile structure lost. Both (1.13), (1.14) extend to the corresponding multidimensional problem of a planar shock moving in Rd (different from the finite-cross-sectional case considered here).

As calculated in [68], [65], respectively, neither Δ nor ΔCJ vanishes for λ0 for an ideal-gas equation of state.1 Likewise, γ does not vanish for any choice of parameters for ideal gas dynamics by a well-known result of Gilbarg [29], or, by results of [28] for reactive gas dynamics with an ideal gas equation of state in the ZND-limit |B|0. From these observations, combined with (1.13), (1.14), we may deduce that D and DrNS have for all choices of physical parameters precisely one zero at λ=0, corresponding to translation-invariance of the background equations.

In particular, as physical parameters are varied, starting from a stable viscous traveling-wave, transition to instability, signalled by passage from the stable complex half-plane {λ<0} to the unstable complex half-plane {λ>0} of one or more eigenvalues of the linearized operator about the wave, cannot occur through passage of a real eigenvalue through the origin, but rather must occur through the passage of one or more nonzero complex conjugate pairs through the imaginary axis: that is, a Poincaré–Hopf-type configuration. This observation, made in [65], [67], gives a rigorous corroboration at the spectral level of the numerical observations of Kasimov and Stewart [52]. What remains is to convert this spectral information into a rigorous nonlinear existence result.

From physical/numerical observations, one expects for an ideal gas equation of state that such a transition seldom or never occurs for shocks but frequently occurs for detonations. However, from a mathematical point of view, the situation for shocks and detonations appears to be entirely parallel. We may thus phrase a common mathematical problem:

(P)Let ūε denote a family of traveling-wave (either shock or detonation) solutions indexed by bifurcation parameter εR1, with associated linearized operators L(ε) transitioning at ε=0 from stability to instability as follows.

Assuming the “generic” spectral situation that

  • (i)

    the L2 essential spectrum of each L(ε) is contained in {λ<0}{0},

  • (ii)

    the translational zero-eigenvalue of each L(ε) is simple in the sense that the associated Evans function Dε vanishes at λ=0 with multiplicity one, and

  • (iii)

    a single complex conjugate pair of L2 eigenvalues of L(ε) (zeros of Dε) λ±(ε)=γ(ε)±iτ(ε),γ(0)=0,τ(0)0,(dγ/dε)(0)>0crosses the imaginary axis with positive speed, and the rest of the L2 point spectrum of L(ε) is located in {λ<0}{0},

show that there occurs a Poincaré–Hopf bifurcation from the family of traveling wave solutions ūε to nearby time-oscillatory solutions.

We examine this problem in a series of increasingly realistic contexts.

Let us first recall the situation considered in [88], of a one-parameter family of standing-wave solutions ūε(x) of a smoothly-varying family of equations ut=F(ε,u)uxxF(ε,u,ux) (possibly shifts F(ε,u,ux)f(u,ux)s(ε)ux of a single equation written in coordinates xxs(ε)t moving with traveling-wave solutions of varying speeds s(ε)), with linearized operators L(ε)F/u|u=ūε for which a spectral gap may be recovered in an appropriate exponentially-weighted norm fHη22j=02(d/dx)jf(x)Lη22,fLη2eη(1+|x|2)1/2f(x)L2,η>0. We call this the weighted norm condition.

This approach, introduced by Sattinger [86], applies to the case (see Section 1.2.1) that signals in the far field are convected under the linearized evolution equation ut=L(ε)u inward toward the background profile, hence time-exponentially decaying in the weighted norm Hη2 penalizing distance from the origin. The method encompasses both (1.2), (1.6) in the scalar case u, zR1, with artificial viscosity B=D=I. It has also interesting applications to certain reaction–diffusion systems and the related Poincaré–Hopf phenomenon of “breathers” [79], [44], [43]. However, it does not apply to either shock or detonation waves in the system case uRn, n2.

In this case, we have the following rather complete result, including stability along with bifurcation description.

Proposition 1.1 [88]

Let ūε ,(1.15)be a family of traveling-waves and systems satisfying the weighted norm condition and assumptions (P) , with FC4 . Then, for a0 sufficiently small and C>0 sufficiently large, there are C1 functions ε(a) , ε(0)=0 , and T(a) , T(0)=2π/τ(0) , and a C1 family of solutionsua(x,t)=ua(xσat,t)of(1.15)with ε=ε(a) , where ua(,t) is time-periodic with period T(a) and σa is a constant drift, such thatC1aua(,t)ūε(a)Hη2Cafor all t0 . Up to fixed translations in x , t , for ε sufficiently small, these are the only nearby solutions of this form, as measured in Hη2 . Moreover, solutions ua are time-exponentially phase-asymptotically orbitally stable with respect to Hη2 if dε/da>0 , in the sense that perturbed solutions converge time-exponentially to a specific shift in x and t of the original solution, and unstable if dε/da<0 .2

Proposition 1.1 was established in [88] by center-manifold reduction, with the main issue being to accommodate the underlying group invariance of translation. The basic idea is to coordinatize u as (v,α), where α parameterizes the group invariance, then work on the quotient space v, reducing the problem from relative to a standard Poincaré–Hopf bifurcation, afterward recovering the location α (a function of a and t) driven by the solution on the quotient space, by quadrature. The integral of the periodic driving term yields a periodic part that may be subsumed in the profile and a drift σat. See [88] for further discussion and details.

Remark 1.2

A consequence of (1.17) (by Sobolev embedding) is |ua(x,t)ūε|Ceη|x|. That is, the existence result in weighted norm space includes quite strong information on the structure of the wave. On the other hand, the stability result is somewhat weakened by the appearance of a spatial weight, being restricted to exponentially decaying perturbations.

The purpose of the present paper is to extend the analysis initiated in [88] for scalar models to the more physically realistic system case. For simplicity, we restrict to the somewhat simpler case of viscous shock solutions of systems with artificial viscosity BI; however, the method of analysis in principle applies also in the general case; see discussion, Section 1.7.

Specifically, consider a one-parameter family of standing viscous shock solutions u(x,t)=ūε(x),limz±ūε(z)=u±ε(constant for fixed ε) , of a smoothly-varying family of conservation laws ut=F(ε,u)uxxF(ε,u)x,uRn, with associated linearized operators L(ε)Fu|u=ūε=xAε(x)+x2, denoting Aε(x)Fu(ε,ūε(x)),A±εlimz±Aε(z)=Fu(ε,u±ε).

We take ūε to be of standard Lax type, meaning that the hyperbolic convection matrices A+ε and Aε at plus and minus spatial infinity have, respectively, np positive and p1 negative real eigenvalues for 1pn, where p is the characteristic family associated with the shock: in other words, there are precisely n1 outgoing hyperbolic characteristics in the far field.

This is the only type occurring for gas dynamics with a standard (e.g., ideal gas) equation of state; for reacting gas dynamics, the corresponding object is a strong detonation, which is the only (nondegenerate) type occurring in the ZND limit B,D0 [28]. The special features of Lax-type shocks (resp. strong detonations), as compared to more general undercompressive shocks (resp. weak detonations) that can occur in other settings, turn out to be important for the analysis (see Remark 3.10), in sharp contrast with the generality of [88]. Note, for the system case n2, that there is at least one outgoing characteristic, so that the weighted norm methods of the previous section do not apply. In particular, we see no way to construct a center manifold for this problem, and suspect that one may not exist; the slow (time-algebraic) decay rate of outgoing modes evident in our description of the Green function in Section 1.2.1 suggests an essential obstacle to such construction. On the other hand, the conservative form of system (1.20) implies that relative mass u(x)dx is conserved for all time for perturbations u=ũūε, ũ satisfying (1.20). Thus, at least at a formal level, there is a convenient invariant subspace of (1.20) consisting of perturbations with zero excess mass, on which the zero eigenvalue associated with translational invariance is removed. For, recall that ū(x)dx=u+u0, so that the associated zero-eigenfunction is not in the subspace. On the other hand, nonzero eigenfunctions always have zero mass [97], since λϕ=L(ε)ϕ implies λϕ(x)dx=0 by divergence form of L(ε), hence the crossing nonzero imaginary eigenvalues λ±(ε) in (P) persist. Thus, the spectral scenario (P) translates in the zero-mass subspace to a standard Poincaré–Hopf scenario, with no additional zero-eigenvalue, for which the translational group-invariance need not be taken into account. Recall, that this was the main issue in the analysis of the scalar case.

In short, the two problems (Model analysis I vs. Model analysis II) have essentially complementary mathematical difficulties, hence little technical contact. Accordingly, the analysis has a quite different flavor in the system case, depending on both the full, pointwise Green function bounds of [73] and the special, conservative structure of the equations.

Our result in this case, and the main result of the paper, is as follows.

Theorem 1.3

Let ūε ,(1.20)be a family of traveling-waves and systems satisfying assumptions (P) , with FC2 . Assume further that ūε is of Lax type, and that the eigenvalues of A±ε , as defined in(1.22), are real, nonzero and simple. Then, for a0 sufficiently small and C>0 sufficiently large, there are C1 functions ε(a) , ε(0)=0 , and T(a) , T(0)=2π/τ(0) , and a C1 family of solutions ua(x,t) of(1.20)with ε=ε(a) , time-periodic of period T(a) , such thatC1asupxR(1+|x|)|ua(x,t)ūε(a)(x)|Ca,for all t0.Up to fixed translations in x , t , for ε sufficiently small, these are the only nearby solutions as measured in norm fX1(1+|x|)f(x)L(x)that are time-periodic with period T[T0,T1] , for any fixed 0<T0<T1<+ . Indeed, they are the only nearby solutions of the more general form(1.16).

Note that the statement of Theorem 1.3 is considerably weaker than that of Proposition 1.1, asserting no stability information, and only the relatively weak structural information of algebraic decay (1.23) at ±, as compared to the exponential bound (1.18). As suggested by our formal “zero-mass” discussion, the periodic solutions constructed have zero mean drift σ, in contrast to (1.16). However, somewhat surprisingly, they do not appear to have zero excess mass. See Remark 1.8, Remark 2.10 for further discussion and explanation.

One-dimensional traveling-wave solutions (1.4) of (1.2) with BI may alternatively be viewed as the restriction to one dimension of a planar viscous shock solution u(x,t)=ū(x1st) of a multidimensional system of viscous conservation laws ut+fj(u)xj=Δxu,uRn,xRd,tR+ on the whole space. Likewise, traveling-wave solutions (1.24) may be viewed as planar traveling-wave solutions of (1.25) on an infinite cylinder C{x:(x1,x̃)R1×Ω},x̃=(x2,,xd) with bounded, cross-section ΩRd1, under artificial Neumann boundary conditionsu/x̃νΩ=0 forx̃Ω, or, in the case that Ω is reflection symmetric, periodic boundary conditionsu(x1,x̃)=u(x1,x̃) forx̃Ω.

We take this as a simplified mathematical model for flow in a duct, in which we have neglected boundary-layer phenomena along the wall Ω in order to isolate the oscillatory phenomena of our main interest.

Consider a one-parameter family of standing planar viscous shock solutions ūε(x1) of a smoothly-varying family of conservation laws ut=F(ε,u)Δxuj=1dFj(ε,u)xj,uRn in a fixed cylinder C, with periodic boundary conditions (typically, shifts Fj(ε,u)xjfj(u)xjs(ε)ux1 of a single Eq. (1.25) written in coordinates x1x1s(ε)t moving with traveling-wave solutions of varying speeds s(ε)), with linearized operators L(ε)F/u|u=ūε. Let A±1(ε)limz±Fu1(ε,ūε). We take ūε, considered as a shock wave in one dimension, to be of standard Lax type (an assumption on A±1(ε), as in the previous subsection). For simplicity, we take Ω=Td1, i.e., Ω=[0,1]d1 with periodic boundary conditions (1.26). Then, we have the following result generalizing Theorem 1.3.

Theorem 1.4

Let ūε ,(1.27)be a family of traveling-waves and systems satisfying assumptions (P) , with FC2 , Ω=Td1 . Assume further that ūε is of Lax type, and that the eigenvalues of A±1(ε) , as defined in(1.28), are real, nonzero and simple. Then, for a0 sufficiently small and C>0 sufficiently large, there are C1 functions ε(a) , ε(0)=0 , and T(a) , T(0)=2π/τ(0) , and a C1 family of solutions ua(x1,t) of(1.27)with ε=ε(a) , time-periodic of period T(a) , such thatC1asupx1R(1+|x1|)|ua(x1,t)ūε(a)(x1)|Ca,for allt0.Up to fixed translations in x , t , for ε sufficiently small, these are the only nearby solutions as measured in norm fX1(1+|x1|)f(x)L(x)that are time-periodic with period T[T0,T1] , for any fixed 0<T0<T1<+ . Indeed, they are the only nearby solutions of the more general form(1.16).

Remark 1.5

Extension to a circular cross-section Ω{|x̃|R} follows by Bessel function expansion as in [52]. Extensions to general cross-sectional geometries and or Neumann boundary conditions are interesting directions for further investigation.

We now briefly discuss the ideas behind the proofs of Theorem 1.3, Theorem 1.4. Under the spectral assumptions (P) (stated in Section 1.2.3), there exist smooth (in ε) L(ε)-invariant projections onto the two-dimensional eigenspace Σε of L(ε) associated with the pair of crossing eigenvalues λ±(ε)=γ(ε)±iτ(ε) and its complement Σ̃ε. Projecting onto these subspaces, and rewriting in polar coordinates the flow on Σε, we may thus express (1.20) in standard fashion as ṙ=γ(ε)r+Nr(ε,r,θ,v),θ̇=τ(ε)+Nθ(ε,r,θ,v),v̇=L̃(ε)v+Nv(ε,r,θ,v), where the “transverse linearized operator” L̃(ε) is the restriction of L(ε) to Σ̃ε, and Nj are higher-order terms coming from the nonlinear part of (1.20): Nr and Nv quadratic order in r,v and their derivatives, and Nθ linear order; in particular Nj(ε,0,θ,0)0 for j=r,θ,v. We are precisely interested in the case that L̃ has no spectral gap, i.e., σ(L)θ<0 for any θ>0. One may think for example of the finite-dimensional case that L̃(0) has additional pure-imaginary spectra besides λ±(0)=±iτ(0). Thus, the center manifold may be of higher dimension, and one cannot follow the usual course of reduction to a center manifold involving only r, θ, ε and applying the standard, two-dimensional Poincaré–Hopf Theorem. Instead, we proceed by a direct analysis, combining the two-dimensional Poincaré return map construction with Lyapunov–Schmidt reduction.

Specifically, truncating |v|C0r, C01, in the arguments of Nr, Nθ, we obtain |Nθ|Cr, and therefore θ̇τ(0)/2>0 for ε, r sufficiently small. Thus, in seeking periodic solutions (r,θ,v)(T)=(r,θ,v)(0), we may eliminate θ, solving for T(ε,a,b) as a function of initial data (a,b)(r,v)(0) and the bifurcation parameter ε, with T(0,0,0)=2π/τ(0), and seek solutions (1.30) as fixed points (a,b)=(r,v)(T(ε,a,b)) of the Poincaré return map (r,v)(T(ε,,)).

Using Duhamel’s formula/variation of constants, we may express (1.31) in a standard way (see, e.g., [36]) as 0=f(ε,a,b)=(eγ(ε)T(ε,a,b)1)a+N1,0=g(ε,a,b)=(eL̃(ε)T(ε,a,b)Id)b+N2, where N10T(ε,a,b)eγ(ε)(T(ε,a,b)s)Nr(ε,r,θ,v)(s)ds,N20T(ε,a,b)eL̃(ε)(T(ε,a,b)s)Nv(ε,r,θ,v)(s)ds are, formally, quadratic order terms in a, b. Note that (r,θ,v) are functions of (ε,a,b) as well as s, through the flow of (1.29), with (r,v)C(a,b) for any “reasonable” norm in the sense that (1.29) are locally well-posed (in practice, no restriction). By quadratic dependence of Nj, we have, evidently, that (ε,a,b)=(ε,0,0) is a solution of (1.32) for all ε.

Continuing in this standard fashion, we should next like to perform a Lyapunov–Schmidt reduction, using the Implicit Function Theorem to solve the g equation for b in terms of (ε,a), i.e., to find a function b=B(ε,a),B(0,0)=0, satisfying g(ε,a,B(ε,a))0: equivalently, to reduce to the nullcline of g.

If this were possible, from |N2|C|(a,b)|2 in (1.32)(ii), we would find, further, that |B(ε,a)|C|a|2, from which straightforward Duhamel/Gronwall estimates on (1.29)(i),(ii) would yield |v(s)|C1r2 for s[0,T], justifying a posteriori the truncation |v|C0r, of Nθ performed in the first step, provided that (ε,a,b) are taken sufficiently small: specifically, small enough that r remains less than or equal to C0/C1 for s[0,T].

Substituting into the f equation, we would then obtain a reduced, scalar bifurcation problem 0=f(ε,a)f(ε,a,B(ε,a)),f(ε,0)=0 for all ε, that would be solvable in the usual way (i.e., by dividing out a and applying the Implicit Function Theorem a second time, using ε(a1f)(0,0)=ε(γ/τ)(0)=εγ(0)/τ(0)0 to solve for ε as a function of a; see, e.g., [36], or Section 2.3.1). The Poincaré return map system (1.32) would then be solved, and a solution of (1.32) would generate a periodic solution of (1.29).

Let us ask ourselves now under what circumstances this basic reduction procedure may actually be carried out. Formal differentiation of (1.32) yields bg(0,0,0)=e2π(L̃/τ)(0)Id. In the finite-dimensional ODE case, the usual condition of application of the (standard) Implicit Function Theorem is thus that (e2π(L̃/τ)(0)Id) be invertible, i.e., that L̃(0) have no “resonant” oscillatory modes, i.e., pure imaginary eigenvalues that are integer multiples of the crossing modes λ±(0)=±iτ(0). Note that this includes interesting cases for which L̃ has no spectral gap, and thus that do not yield to the standard center-manifold reduction. The analogous criterion in the infinite-dimensional (e.g., PDE) setting is that all possible resonant modes niτ(0), nZ, lie in the resolvent set of L̃. This is encouraging, and shows that the requirements of Lyapunov–Schmidt reduction are much less than those for center-manifold reduction: in particular, only spectral separation (from both ±iτ(0) and their integer multiples niτ(0)) rather than a spectral gap is needed. Unfortunately, in the setting (1.20), (1.27) of our interest we do not even have spectral separation, since the essential spectra of L̃(ε) accumulates at λ=0 for every ε. In this case, or others for which separation fails, we must follow a different approach.

Alternatively, we may rewrite (1.32)(ii) formally as a fixed-point equation b=(IdeL̃(ε)T(ε,a,b))1N2(ε,a,b), then apply the Contraction-mapping Principle to carry out an Implicit Function construction “by hand” (reminiscent of, but not exactly the standard proof of the Implicit Function Theorem). In the standard case that |eL̃t| is exponentially decaying (spectral gap), we may rewrite (1.34) using Neumann expansion as b=j=0ejL̃(ε)T(ε,a,b)N2(ε,a,b). The basis for our analysis is the simple observation that, more generally, provided the series on the righthand side converges (conditionally) for (a,b) bounded, then (i) (by scaling argument) the righthand side is contractive in b for (ε,a,b) sufficiently small, and (ii) (by standard telescoping sum argument, applying (IdeL̃(ε)T(ε,a,b)) to (1.34)) the resulting solution b=B(ε,a) guaranteed by the Contraction-mapping Principle is in fact a solution of the original (1.32)(ii). See Section 2 for further details.

In the context of (1.20), by divergence form of the equation, N2=(n2)x, with n2=O(|(r,v)|2). For simplicity of discussion, model n2 as a bilinear form in (r,v), so that (by quadratic scaling) n2L1 for vL2, from which we obtain N2=nx, nL1 for bL2. Thus, replacing the righthand side of (1.35) by its continuous approximant N2+T1TeL̃(ε)tN2dt=N2+T1T(eL̃(ε)tx)ndt,nL1, we see that convergence (as a map from L2 to L2) reduces, roughly, to a space-time stability estimate +(TG̃y(x,t;y)dt)n(y)dyL2(x)C,nL1, where G̃ is the Green kernel associated with transverse solution operator eL̃t. Estimate (1.36) is quite similar to those arising for the full solution operator (again, through Duhamel’s principle; see, for example, [96], [74]) in the study of nonlinear stability of spectrally stable waves. That (1.36) holds for all nL1, is equivalent to supyTG̃y(,t;y)dtL2C. (One implication is a consequence of the triangle inequality; for the other take a sequence converging to a Dirac function in L1.) This would hold, for example, if supyTG̃y(,t;y)dtL2C.

Note that the neglected first term N2, corresponding to j=0 in (1.35), is C2 from L2 to L2 by the smoothing action of (1.33)(ii); see Section 4.

In the arguments above, we have used strongly the specific structure of the nonlinearity, both quadratic dependence and divergence form, since the L2-operator norm |eL̃t|L2L2 does not decay in t. However, this is still not enough. For, recall the one-dimensional discussion of Section 1.2.1, modeling G̃ qualitatively as the sum of a convected heat kernel K as in (1.10) and an error-function term J as in (1.11). From the standard bounds Ky(,t;y)L2=Ct3/4, Jy(,t;y)L2Ct1/2, hence (the support of K and J being essentially disjoint) TG̃y(x,t;y)dt is not absolutely convergent in L2(x), and the bound (1.38) does not hold.

Instead, we must show conditional convergence, using detailed knowledge of the propagator G̃ to identify cancellation in TG̃y(x,t;y)dt. For example, using Ky=a1(KtKyy), we find that TtKy(x,s;y)ds=a1(TtKt(x,s;y)dsTtKyy(x,s;y)ds)=a1(K(x,s;y)|TtTtKyy(x,t;y)ds) is the sum of a1K(x,T;y)L2(x) with terms a1K(x,t;y)t1/4 and a1TtKyy(x,t;y)dsTts5/4ds that are respectively decaying and absolutely convergent in L2(x), hence converges uniformly with respect to y in L2(x).

A similar cancellation argument yields convergence of the error-function term, along with uniform boundedness with respect to y. Convergence of the error-function term is not uniform with respect to y, a detail which necessitates further technicalities: in particular, the weighted estimate |v(x)|C(1+|x|)1 leading eventually to (1.23). Likewise, there are further issues associated with ε-regularity of the solution, needed for bifurcation analysis of the reduced equation. However, the main idea is contained in calculation (1.39).

The multi-dimensional case goes similarly, with the computation of the critical neutral, zero transverse wave number reducing to the one-dimensional case, and all others to the case of a spectral gap.

Remark 1.6

It is readily checked that the above arguments go through in the general case that n2 is quadratic order for |v|LC and not bilinear in (v,r), substituting vL2L (n2L1) for vL2, working in the L2L norm in place of L2, and substituting for (1.37) the estimate supyTG̃y(x,t;y)dtL2L(x)C. Similar cancellation estimates are important in the study of asymptotic behavior of stable viscous shock waves [63], [87], [64], [82], [40].

Remark 1.7

As in the result of Proposition 1.1 obtained by center manifold reduction, the family of periodic solutions obtained, up to translation, lies tangent to the (r,θ)-plane corresponding to linearized oscillatory behavior.

Remark 1.8

As noted earlier, operator S=eL̃T by the divergence form of the equations preserves the property of zero mass, udx=0, hence each finite approximant j=0NSj to (IS)1 also preserves zero mass. However, udx=0 is not closed with respect to L2, as may be seen by the example K(x,s;y)|TtK(x,T;y) in L2 as t; that is, mass may “escape at infinity” in the limiting process. Comparing with (1.39), we see that the principal term in limNj=0NSjx is exactly of this form, hence the L2-limit (IS)1 does not preserve zero mass, even though each finite approximant does. Accordingly, there is no reason that the solution b=b(a,ε)=(IS)1N2(b,a,ε) obtained by reduction should have zero mass: in the model case n=u2, G=K, it can be seen that it does not. Thus, the idea of carrying out a simplified bifurcation analysis in the invariant subspace of zero-mass perturbations, though appealing at a formal level, in general cannot succeed. See also Remark 2.10.

Theorem 1.3, Theorem 1.4 together with the spectral observations of [65], [67] give rigorous validation in a simplified context of the formal and numerical observations of [9], [52]. An interesting problem for future work is to extend these results to the originally-motivating case of detonation waves of the full, reacting compressible Navier–Stokes equations.

We expect that our one-dimensional analysis will extend in a straightforward fashion, combining tools developed in [73], [74], [93], [66] to treat, respectively, nonreacting gas dynamics with physical, partial viscosity and reacting gas dynamics with artificial viscosity. Likewise, we expect that we can readily treat flow in a cylinder for the physical equations with artificial Neumann or periodic boundary conditions. However, the treatment of physical, no-slip boundary conditions (presumably associated with characteristic viscous boundary layers) involves technical and philosophical difficulties beyond the scope of the present analysis, as yet unresolved even for nonreacting, incompressible flows: for example, even the construction of a background traveling profile becomes problematic in this setting. We point out that viscous boundary effects (since also viscosity) are also neglected in the ZND setting of [9], [52].

A second natural direction for future investigation is the question of the stability of the periodic waves whose existence we have established here. In the absence of a spectral gap, our method of analysis does not directly yield stability as in the case of center manifold reduction, but at best partial information on the location of the point spectrum associated with oscillatory modes, with stability presumably corresponding to the standard condition dε/da>0. The hope is that we could combine such information with an analysis like that carried out for stationary waves in [97], [73], adapted from the autonomous to the time-periodic setting: that is, a generalized Floquet analysis in the PDE setting and in the absence of a spectral gap. We consider this a quite exciting direction for further development of the theory.

Bifurcation in the absence of a spectral gap has been considered by a number of authors in different settings; in particular, it has been studied systematically by Ioss et al. [45], [46], [47], [48], [49], [83], [84] in various contexts using an alternative “spatial dynamics” approach. It would be very interesting to investigate what results could be obtained by this technique in the context of viscous shock waves: more generally, to relate it at a technical level to the one used here. Considered in this larger context, the interest of the present approach is that it gives a simple and explicit connection between stability and bifurcation, in the spirit of center-manifold reduction and formal asymptotics, but adapted to the boundary case of zero spectral gap and time-algebraic decay. On the other hand, the price of this “direct” approach is that the needed estimates may in practice (as here) be rather delicate to obtain.

We mention also a recent work of Kunze and Schneider [60] in which they analyze pitchfork bifurcation in the absence of a spectral gap using Sattinger’s weighted-norm method, as described in Section 1.3, but in a situation where convection is outward, away from the profile layer. This entails the use of “wrong-way” exponentially decaying weights at the linearized level, introducing a spatially-exponentially growing multiplier in quadratic-order source terms, in combination with separate, compensating estimates at the nonlinear level. For similar arguments in the context of nonlinear stability, see, e.g., [80], [15]. This interesting approach has been used successfully in the shock-wave context to treat the scalar undercompressive case [15]; however, it does not appear to generalize to the system case.

Remarks 1.9

  • 1.

    Birtea et al. in [7] successfully carry out a rather complicated bifurcation analysis in the ODE setting without explicit knowledge of the background solution. Similar techniques might perhaps be useful in treating flow in a duct with physical, nonslip boundary conditions, for which description of the background flow is itself problematic.

  • 2.

    In carrying out our nonstandard Implicit Function construction for shock waves, we faced the problem of non-uniform convergence of series (1.35) due to lack of spatial localization. We remedied this problem by additional weighted-norm estimates. However, the example of another famous nonstandard Implicit Function construction, namely, Nash–Moser iteration, suggests the alternative approach of introducing a “localizing” step dual to the smoothing step in the Nash–Moser scheme. It would be interesting to see if this approach could also be carried out, thus avoiding the need for additional analysis associated with pointwise bounds. We note that Nash–Moser iteration combining temporal with the usual frequency cutoffs has been carried out by Klainerman [55] in the context of a nonlinear wave equation.

Plan of the paper. In Section 2, we formalize the reduction procedure set out in Section 1.6.4 as an abstract bifurcation framework suitable for application to general discrete dynamical systems in the absence of a spectral gap. In Section 3, we recall the pointwise estimates furnished by the methods of [97], [73] on the Green function G̃ associated with the transverse linearized solution operator eL̃t, then, in Sections 4 Return map construction, 5 Proof of use these to verify that the shock bifurcation problem indeed fits the hypotheses of our abstract framework. This establishes a Lipschitz version of Main Theorem 1.3; we improve this to C1 by a bootstrap argument in Section 5.4. In Section 6, we describe the extension to multidimensions, verifying Theorem 1.4.

Notes. Since the completion of this work, there have been several further developments. In [85], Sandstede and Scheel recover and somewhat sharpen our results using spatial dynamics techniques, answering the question posed in Section 1.7 of what results may be obtained by these methods, obtaining the additional information of exponential localization of solutions and exchange of spectral stability. In [89], [90], we extend our results to shock and detonation waves of systems with physical viscosity, at the same time greatly sharpening and simplifying the basic cancellation estimate. The latter yields also exponential localization and appears to shed light on the more general question of the technical relation between spatial dynamics methods and the “temporal dynamics” method used here; see Remark 1.10, [89].

Section snippets

Abstract bifurcation framework

We begin by formalizing the approach described in Section 1.6.4, providing an abstract framework for the analysis of bifurcations in the absence of a spectral gap for discrete dynamical systems in the general form (1.32): that is, the portion of our analysis occurring after the reduction via Poincaré return-map construction to a fixed-point problem. The specifics of the return-map construction are discussed separately in Section 4. We carry out the analysis at the level of Lipschitz regularity;

Linearized bounds

In the remainder of the paper, we apply the framework of Section 2 to the problem of bifurcation of viscous shock waves, first in the one-dimensional and then in the multi-dimensional case. We begin by assembling the relevant linearized estimates on the one-dimensional flow near a viscous shock.

Return map construction

We show in this Section how the question of existence and uniqueness of time-periodic solutions to (1.20) in a neighborhood of the stationary solution (1.19) can be formulated in the abstract framework of Section 2. We use the notations set out in Section 3.1, in particular Assumptions (H0)–(H2) and (Dε), and functional spaces X1,X2 and B1,B2, and the short-time bounds of Lemma 3.6.

Given a family of stationary solutions ūε(3.1) of the system (1.20), and a family of dynamic solutions ũε of the

Proof of Theorem 1.3

We start from the perturbation equations (4.2), where the linear operator L(ε) is assumed to satisfy Assumptions (H0)–(H2) and (Dε) (which are stated in Section 3.1 and are shown in Lemma 3.2 to be equivalent to the assumptions of Theorem 1.3).

Consider the Poincaré return map system (4.28) associated with the truncated system (4.17), derived from (4.2) by a proper choice of coordinatization and truncation. Periodic solutions of (4.17) are in one-to-one correspondence with fixed points of (4.28)

The multi-d case: Proof of Theorem 1.4

We now briefly describe the extension to the multidimensional case, model problem III, Section 1.5. Consider a one-parameter family of standing planar viscous shock solutions ūε(x1) of a smoothly-varying family of conservation laws ut=F(ε,u)Δxuj=1dFj(ε,u)xj,uRn in an infinite cylinder C{x:(x1,x̃)R1×Ω},x̃=(x2,,xd)ΩRd1 bounded and reflection-symmetric, with periodic boundary conditions u(x1,x̃)=u(x1,x̃) forx̃Ω, with associated linearized operators L(ε)F/u|u=ūε=j=1dxjAj(x1,ε)+Δx

Acknowledgements

Research of B.T. was partially supported under NSF grant number DMS-0505780. Research of K.Z. was partially supported under NSF grant number DMS-0300487. Thanks to Claude Bardos for pointing out the references [45], [46], and to Björn Sandstede and Arnd Scheel for general discussions of the spatial dynamics method. Thanks to Walter Craig for pointing out the Ref. [55], to Thierry Gallay for pointing out [60], and to Tudor Ratiu for pointing out [7].

K.Z. thanks the Ecole Polytechnique Fédérale

References (98)

  • D. Sattinger

    On the stability of waves of nonlinear parabolic systems

    Adv. Math.

    (1976)
  • J. Alexander et al.

    A topological invariant arising in the analysis of traveling waves

    J. Reine Angew. Math.

    (1990)
  • R.L. Alpert et al.

    Periodicity in exothermic hypersonic flows about blunt projectiles

    Acta Astron.

    (1972)
  • W.-J. Beyn et al.

    Stability of viscous profiles: Proofs via dichotomies

    J. Dynam. Differential Equations

    (2006)
  • A.A. Barmin et al.

    Stability of shock waves

    Adv. Mech.

    (1992)
  • B. Barker, J. Humpherys, K. Rudd, K. Zumbrun, Stability of viscous shocks in isentropic gas dynamics, 2006...
  • A.M. Blokhin et al.

    Stability of fast parallel and transversal MHD shock waves in plasma with pressure anisotropy

    Acta Mech.

    (1999)
  • A. Bourlioux et al.

    Theoretical and numerical structure for unstable one-dimensional detonations.

    SIAM J. Appl. Math.

    (1991)
  • L. Brin, Numerical testing of the stability of viscous shock waves, Doctoral thesis, Indiana University,...
  • L.Q. Brin

    Numerical testing of the stability of viscous shock waves

    Math. Comp.

    (2001)
  • L. Brin et al.

    Analytically varying eigenvectors and the stability of viscous shock waves

    Seventh Workshop on Partial Differential Equations, Part I (Rio de Janeiro, 2001)

    Mat. Contemp.

    (2002)
  • N. Costanzino et al.

    Existence and stability of curved multidimensional detonation fronts

    Indiana Univ. Math. J.

    (2007)
  • J. Dodd, Convection stability of shock profile solutions of a modified KdV–Burgers equation, Thesis, University of...
  • P. D’yakov

    Zhur. Eksptl. i Teoret. Fiz.

    (1954)
  • J.J. Erpenbeck

    Stability of step shocks.

    Phys. Fluids

    (1962)
  • J.J. Erpenbeck

    Nonlinear theory of unstable one–dimensional detonations

    Phys. Fluids

    (1967)
  • W. Fickett

    Detonation in miniature

  • W. Fickett et al.

    Detonation

    (1979)
  • Fickett et al.

    Flow calculations for pulsating one-dimensional detonations

    Phys. Fluids

    (1966)
  • H. Freistühler et al.

    Spectral stability of small shock waves

    Arch. Ration. Mech. Anal.

    (2002)
  • R. Gardner et al.

    Traveling waves of a perturbed diffusion equation arising in a phase field model

    Indiana Univ. Math. J.

    (1989)
  • R. Gardner et al.

    The Gap Lemma and geometric criteria for instability of viscous shock profiles

    Commun. Pure Appl. Math.

    (1998)
  • I. Gasser et al.

    A geometric singular perturbation analysis of detonation and deflagration waves

    SIAM J. Math. Anal.

    (1993)
  • D. Gilbarg

    The existence and limit behavior of the one-dimensional shock layer

    Amer. J. Math.

    (1951)
  • J. Goodman

    Nonlinear asymptotic stability of viscous shock profiles for conservation laws

    Arch. Ration. Mech. Anal.

    (1986)
  • J. Goodman

    Remarks on the stability of viscous shock waves

  • J. Goodman et al.

    Viscous limits for piecewise smooth solutions to systems of conservation laws

    Arch. Ration. Mech. Anal.

    (1992)
  • O. Gues et al.

    Multidimensional viscous shocks I: degenerate symmetrizers and long time stability

    J. Amer. Math. Soc.

    (2005)
  • O. Gues et al.

    Existence and stability of multidimensional shock fronts in the vanishing viscosity limit

    Arch. Ration. Mech. Anal.

    (2004)
  • O. Gues et al.

    Navier–Stokes regularization of multidimensional Euler shocks

    Ann. Sci. cole Norm. Sup. (4)

    (2006)
  • J. Hale et al.
  • D. Henry

    Geometric theory of semilinear parabolic equations

  • P. Howard et al.

    Pointwise asymptotic behavior of perturbed viscous shock profiles

    Adv. Differential Equations

    (2006)
  • P. Howard et al.

    Sharp pointwise bounds for perturbed shock waves

    J. Hyperbolic Differ. Equ.

    (2006)
  • J. Humpherys et al.

    Spectral stability of small amplitude shock profiles for dissipative symmetric hyperbolic–parabolic systems

    Z. Angew. Math. Phys.

    (2002)
  • T. Ikeda et al.

    Hopf bifurcation of travelling pulses in some bistable reaction–diffusion systems

    Methods Appl. Anal.

    (2000)
  • T. Ikeda et al.

    Pattern selection for two breathers

    SIAM J. Appl. Math.

    (1994)
  • G. Iooss, Travelling water-waves, as a paradigm for bifurcations in reversible infinite-dimensional “dynamical”...
  • G. Iooss

    On the standing wave problem in deep water

    J. Math. Fluid Mech.

    (2002)
  • Cited by (30)

    • Transverse bifurcation of viscous slow MHD shocks

      2021, Physica D: Nonlinear Phenomena
      Citation Excerpt :

      Our main physical conclusions are two. First, we make a mathematical connection between the spectral instability observed for slow inviscid MHD shocks [3,9] and “corrugation instabilities” observed in the astrophysics community [11,12], via a viscous bifurcation analysis as in [4–6]. Namely, we demonstrate transitions from stability to instability satisfying the bifurcation hypotheses proposed in [4] (parallel magnetic field case) and [6] (nonparallel case), implying bifurcation in a mode transverse to the direction of shock propagation, i.e., lying in the direction parallel to the front, the first examples for which these scenarios have been shown to occur.

    • Transverse steady bifurcation of viscous shock solutions of a system of parabolic conservation laws in a strip

      2014, Journal of Differential Equations
      Citation Excerpt :

      As described above, the value of the imaginary parts may be prescribed arbitrarily by choice of an appropriate moving coordinate frame in the y-direction; hence in some sense the two problems are equivalent. On the other hand, our technique differs from the ones in [19] and [14]: in their case, a Poincaré first return map is used, while here we fully explore the fact that the underlying shock profile is steady to set the system of equations as a time independent problem of the form (1.8). The solution is then obtained by quite different and more direct techniques, which yield also the additional information that the time-periodic solutions of [19] are, more precisely, spatially-periodic traveling waves (see the proof of Corollary 6.3).

    • Stability of viscous detonations for Majda's model

      2013, Physica D: Nonlinear Phenomena
    • Saddle-node bifurcation of viscous profiles

      2012, Physica D: Nonlinear Phenomena
      Citation Excerpt :

      For large amplitude shock waves in specific systems from gas dynamics spectral stability has been established in work of Zumbrun and coworkers by complementing analytical arguments with numerical computations; see e.g. [18]. Concerning bifurcations of profiles in systems of viscous conservation laws, Hopf bifurcations were studied by Texier and Zumbrun [19,20] as well as Sandstede and Scheel [21]. They assume the existence of a parametrized family of viscous shock waves with an associated pair of complex conjugate eigenvalues of the linearization crossing the imaginary axis as the parameter passes through the bifurcation point.

    • Nash-Moser iteration and singular perturbations

      2011, Annales de l'Institut Henri Poincare (C) Analyse Non Lineaire
    View all citing articles on Scopus
    View full text