on june 23, 2018 dispersionofdiscontinuous...

21
rspa.royalsocietypublishing.org Research Cite this article: Chen G, Olver PJ. 201 Dispersion of discontinuous periodic waves. Proc R Soc A 469: 20120407. http://dx.doi.org/10.1098/rspa.2012.0407 Received: 9 July 2012 Accepted: 17 September 2012 Subject Areas: applied mathematics, differential equations, mathematical modelling Keywords: dispersion, Talbot effect, wave, fractal, splitting method Author for correspondence: Peter J. Olver e-mail: [email protected] Dispersion of discontinuous periodic waves Gong Chen and Peter J. Olver School of Mathematics, University of Minnesota, Minneapolis, MN 55455, USA The dynamic evolution of linearly dispersive waves on periodic domains with discontinuous initial profiles is shown to depend remarkedly upon the asymptotics of the dispersion relation at large wavenumbers. Asymptotically linear or sublinear dispersion relations produce slowly changing waves, while those with polynomial growth exhibit dispersive quantization, a.k.a. the Talbot effect, being (approxi- mately) quantized at rational times, but a non- differentiable fractal at irrational times. Numerical experiments suggest that such effects persist into the nonlinear regime, for both integrable and non- integrable systems. Implications for the successful modelling of wave phenomena on bounded domains and numerical challenges are discussed. 1. Introduction The most common device used to model physical wave phenomena is to approximate the fully nonlinear fluid mechanics problem in a targeted physical regime by first suitably rescaling the original variables and then expanding the rescaled system in powers of an appropriate small parameter. Truncation at low (typically first) order produces a mathematically simpler system, whose solutions, one hopes, closely track the originals, at least over an initial time span. At the linear level (at least in the homogeneous, isotropic regime), conservative wave dynamics is entirely determined by the dispersion relation, which relates the temporal frequency of an oscillatory complex exponential solution to its wavenumber (spatial frequency). The standard modelling paradigm can thus, at the linear level, be viewed as replacing the full dispersion relation by a suitable approximation. However, in most perturbative models, while the approximate dispersion relation is designed to fit its physical counterpart well at small wavenumbers, e.g. they have the same Taylor polynomial at the origin, they often have little in common at the high-frequency c 2012 The Author(s) Published by the Royal Society. All rights reserved. 3 on August 20, 2018 http://rspa.royalsocietypublishing.org/ Downloaded from

Upload: ledang

Post on 20-Aug-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

rspa.royalsocietypublishing.org

ResearchCite this article: Chen G, Olver PJ. 201Dispersion of discontinuous periodic waves.Proc R Soc A 469: 20120407.http://dx.doi.org/10.1098/rspa.2012.0407

Received: 9 July 2012Accepted: 17 September 2012

Subject Areas:applied mathematics, differential equations,mathematical modelling

Keywords:dispersion, Talbot effect, wave, fractal,splitting method

Author for correspondence:Peter J. Olvere-mail: [email protected]

Dispersion of discontinuousperiodic wavesGong Chen and Peter J. Olver

School of Mathematics, University of Minnesota, Minneapolis,MN 55455, USA

The dynamic evolution of linearly dispersive waveson periodic domains with discontinuous initialprofiles is shown to depend remarkedly uponthe asymptotics of the dispersion relation at largewavenumbers. Asymptotically linear or sublineardispersion relations produce slowly changing waves,while those with polynomial growth exhibit dispersivequantization, a.k.a. the Talbot effect, being (approxi-mately) quantized at rational times, but a non-differentiable fractal at irrational times. Numericalexperiments suggest that such effects persist intothe nonlinear regime, for both integrable and non-integrable systems. Implications for the successfulmodelling of wave phenomena on bounded domainsand numerical challenges are discussed.

1. IntroductionThe most common device used to model physicalwave phenomena is to approximate the fully nonlinearfluid mechanics problem in a targeted physical regimeby first suitably rescaling the original variables andthen expanding the rescaled system in powers of anappropriate small parameter. Truncation at low (typicallyfirst) order produces a mathematically simpler system,whose solutions, one hopes, closely track the originals,at least over an initial time span.

At the linear level (at least in the homogeneous,isotropic regime), conservative wave dynamics is entirelydetermined by the dispersion relation, which relatesthe temporal frequency of an oscillatory complexexponential solution to its wavenumber (spatialfrequency). The standard modelling paradigm canthus, at the linear level, be viewed as replacing thefull dispersion relation by a suitable approximation.However, in most perturbative models, while theapproximate dispersion relation is designed to fit itsphysical counterpart well at small wavenumbers, e.g.they have the same Taylor polynomial at the origin,they often have little in common at the high-frequency

c© 2012 The Author(s) Published by the Royal Society. All rights reserved.

3

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

2

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

range. For example, the free boundary problem governing water waves has a dispersion relationthat is asymptotically proportional to a square root of the wavenumber. On the other hand, inthe shallow water regime, the unidirectional Korteweg–de Vries model’s dispersion relation iscubic, and of the wrong sign, and the competing regularized long wave (RLW) or Benjamin–Bona–Mahony (BBM) model’s is asymptotically zero, while that of some standard bidirectionalBoussinesq systems is asymptotically constant. Provided one restricts attention to smoothsolutions, or to unbounded domains, the discrepancies in the dispersion of high-frequency modesdo not seem to play a very noticeable role. However, on bounded domains, rougher solutions,which have more substantial high-frequency components, are affected in a significant mannerowing to subtle and as yet poorly understood resonances among the varyingly propagatingFourier modes.

In Olver [1], the second author observed that, for integral polynomial dispersion relations,i.e. those whose coefficients are integer multiples of a common real number λ, the solution inquestion has the remarkable behaviour that, at irrational times (relative to the ratio β = �/λ

between the length � of the interval and the dispersion constant λ), it is a continuous, butfractal, non-differentiable function, whereas at rational times, the solution is piecewise constant!While previously unnoticed1 by experts in dispersive wave theory, the phenomenon of dispersivequantization had, in fact, been first discovered in the early 1990s, in the contexts of optics andquantum mechanics, by Berry and various co-workers [2–6], who named it the Talbot effect inhonour of a striking 1836 optical experiment by the photographic pioneer Talbot [7]. Experimentalconfirmations of the Talbot effect in optics and in quantum revival are described in Berry et al. [5].Rigorous analytical results and estimates justifying the Talbot effect can be found in the work ofKapitanski & Rodnianski [8], Rodnianski [9], Oskolkov [10,11] and Taylor [12].

The initial aim of the present paper was to extend these studies to basic linear differentialand integro-differential equations, depending on a single spatial variable, which have non-polynomial dispersion relations. Our main conclusion is that the large wavenumber asymptoticsof the dispersion relation plays the dominant role governing the qualitative features exhibited bytheir rough solutions on periodic domains. In analogy with the basic Riemann problem of gasdynamics [13], we focus on the specific initial data provided by a step function. The resultingFourier series solution is easily constructed, and of independent interest. For example, when thedispersion relation is polynomial, the solution at rational times has the form of a Gauss or Weylexponential sum, of fundamental importance in analytic number theory since the work of Hardy& Littlewood [14] and Vinogradov [15]. Indeed, as explained in Ford [16], because the behaviourof such exponential sums has implications for the size of the zero-free region of the Riemannzeta function in the critical strip, further refinement of known estimates could provide a route toproving the Riemann hypothesis!

We will also initiate the investigation of such effects in the nonlinear regime. The revolutionarydiscovery of the soliton was sparked by the original movie of Zabusky & Kruskal [17], whichdisplayed a numerical simulation of the solution to a particular periodic initial-boundary valueproblem for the Korteweg–de Vries equation with small nonlinearity. (By contrast, the celebratedstudies of Lax & Levermore [18], Lax et al. [19], Venakides [20,21] are concerned with the smalldispersion regime and convergence to shock wave solutions of the limiting nonlinear transportequation.) Because Zabusky and Kruskal’s selected initial data were a smoothly varying cosine,the Talbot fractalization effect was (perhaps fortunately!) not observed. (And, technically, theelastically interacting waves that emerge from the initial cosine profile are not true solitons, in thatthese exist only for the full line problem, but, rather, cnoidal waves embedded in a hyperellipticfinite gap solution [22,23].) Our numerical experiments with step function initial data stronglyindicate that the dispersive quantization/fractalization effect is present in the periodic boundaryvalue problems for both the integrable nonlinear Korteweg–de Vries and Schrödinger equations,as well as non-integrable models of a similar nature, and extends well beyond the weaklynonlinear regime.

1The work of Oskolkov [11], which analyses both the linear Schrödinger equation and the linearized Korteweg–de Vriesequation on periodic domains, is a notable exception.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

3

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

One consequence of these studies is that, contrary to the conventional wisdom, when dealingwith nonlinear wave models on bounded intervals, the principal source of analytic difficulty maybe, counterintuitively, not the nonlinear terms, but rather the poorly understood behaviour oflinearly dispersive partial differential equations. Our investigations imply that the qualitativebehaviour of the solution to the periodic problem depends crucially on the asymptotic behaviourof the dispersion relation at large wavenumbers, reinforcing Benjamin et al.’s critique [24] that theKorteweg–de Vries equation, say, is an unsatisfactory model for surface waves, because its cubicdispersion relation effectively transmits the high-frequency modes in the wrong direction, withunboundedly negative phase velocity and group velocity, inciting unphysical interactions withother solution components. And indeed, in the periodic problem, this shortcoming is observedas the apparently number-theoretic resonant interaction of high-frequency modes spawnedby the initial data serve to produce a-physical fractalization and quantization effects in theengendered solution.

2. Dispersion relations and wave modelsConsider a linear, scalar, constant coefficient partial differential equation for a function u(t, x) oftime t and a single spatial variable x. Recall [13] that the dispersion relation ω(k) relates temporalfrequency ω to wavenumber (spatial frequency) k of an oscillatory complex exponential solutionof the form

u(t, x) = ei(kx−ωt). (2.1)

The differential equation will be called purely dispersive if the resulting function ω(k) is real.(Complex dispersion relations indicate the presence of dissipative or viscous effects that dampout solutions, or, alternatively, ill-posedness, depending on the sign of their imaginary part.)Given a real dispersion relation, the phase velocity cp = ω(k)/k prescribes the speed of an individualoscillatory wave, while, as a consequence of the method of stationary phase, wave packets andenergy move with the group velocity cg = dω/dk [13]. These differ when the dispersion relationis nonlinear, causing initially localized disturbances to spread out in a dispersive fashion—aphysical effect that can be seen, for example, by throwing a rock into a pond and observingthat the individual waves move at a different speed, namely the phase velocity, than the overalldisturbance, which moves at the group velocity. The dispersion relation of nonlinear equations isthat of their linearization.

For example, the basic transport equation

ut + cux = 0 (2.2)

has linear dispersion relation ω(k) = ck, whereas the one-dimensional wave equation

utt = c2uxx (2.3)

has bidirectional dispersion relation ω(k) = ±ck. Linearity of these dispersion relations has theconsequence that all Fourier modes travel with a common speed, thereby producing travellingwaves of unchanging form. The simplest quadratic dispersion relation, ω(k) = k2, appears in thefree space Schrödinger equation

i ut = uxx, (2.4)

which is fundamental to quantum mechanics [25]. The linear beam equation

utt + uxxxx = 0, (2.5)

which models small vibrations of a thin elastic beam, and has bidirectional quadratic dispersion:ω(k) = ±k2. The third-order linear wave model

ut + uxxx = 0, (2.6)

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

4

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

Dh

y = h + h(t, x)a

h

c = ÷gh

Figure 1. Water waves.

which is sometimes known as the Airy partial differential equation, because it admits solutionsin terms of Airy functions [26] or, alternatively, the linear Korteweg–de Vries equation, has cubicdispersion relation ω(k) = k3.

A fertile source of interesting wave models is the free boundary problem for water waves,meaning the dynamics of an incompressible, irrotational fluid under gravity. Let us review,following [13,27,28], the derivation of models in the shallow water regime. For simplicity, werestrict our attention to two-dimensional flow, with x indicating the horizontal coordinate andy the vertical. (See Yuen & Lake [29] for the deep water regime, and [27,30] for extensions tothree-dimensional models.)

Suppose that the water sits on a flat, impermeable horizontal bottom, which we take as y =0, and has undisturbed depth h. We assume further that the surface elevation at time t is thegraph of a single-valued function y = h + η(t, x), i.e. there is no wave overturning. The periodicproblem of interest here takes η to be a periodic function of x with mean 0, while, in contrast,solitary waves require η(t, x) → 0 as |x| → ∞. The motion of the fluid is prescribed by the velocitypotential φ(t, x, y), which is defined within the domain occupied by the fluid Dη = {0 < y < h +η(t, x)} (figure 1). In the absence of surface tension, the free boundary problem governing thedynamical evolution of the water is then given by

φt + 12 φ2

x + 12 φ2

y + gη = 0,ηt = φy − ηxφx,

}y = h + η(t, x),

φxx + φyy = 0, 0 < y < h + η(t, x)

and φy = 0, y = 0,

⎫⎪⎪⎪⎪⎪⎬⎪⎪⎪⎪⎪⎭

(2.7)

in which g is the gravitational constant. As shown in Whitham [13, §13.4], the dispersion relationfor the linearized water wave system, obtained by omitting the nonlinear terms from the freesurface conditions, is

ω(k)2 = gk tanh(hk). (2.8)

In the shallow water regime, one introduces the rescaling

x �−→ �x, y �−→ hy, t �−→ �

ct, η �−→ aη, φ �−→ ga�

cφ, (2.9)

in which � represents a typical wavelength, h the undisturbed depth, c = √gh the leading order

wave speed and a the wave amplitude. Substituting (2.9) into the free boundary value problem(2.7), one formally solves the resulting rescaled Laplace equation for the potential φ(t, x, y), asa power series in the vertical coordinate y, in terms of the horizontal velocity u(t, x) = φx(t, x, θ)

at a fraction 0 ≤ θ ≤ 1 of the undisturbed depth h. Substituting the result into the free boundary

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

5

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

conditions results in a system of equations of Boussinesq type, which is then truncated to, say,first order in the small parameters

α = ah

, β = h2

�2 , (2.10)

representing the ratio of undisturbed depth to height of the wave, and the square of the ratio ofheight to wavelength. We are interested in the regime of long waves in shallow water, in whichboth α and β are small, but of comparable magnitudes. As in [27,28,31], the resulting Boussinesqsystems are all of the general form

ηt + ux + α(ηu)x + β(auxxx − bηxxt) = 0

and ut + ηx + αuux + β(cηxxx − duxxt) = 0,

}(2.11)

where the parameters a, b, c, d, depend on the depth, and are subject to the physical constraints

a + b = 16 (3θ2 − 1), c + d = 1

2 (1 − θ2), a + b + c + d = 1. (2.12)

The particular system

ut + ηx + αuux = 0, ηt + ux + α(ηu)x + 13 βuxxx = 0, (2.13)

valid for θ = 1, i.e. on the free surface, was derived by Whitham [32] and Broer [33], and shownto be completely integrable (in fact, tri-Hamiltonian) by Kaup [34] and Kupershmidt [35]. Bonaet al. [36] subsequently found that this system is, in fact, linearly ill-posed—although the questionof nonlinear ill-posedness is, apparently, still open. The corresponding second-order models canfound in earlier studies [27,28].

The second stage is to restrict the bidirectional Boussinesq system (2.12) to a submanifold ofapproximately unidirectional solutions, building on the usual factorization of the second-orderwave equation (2.3) into right- and left-moving waves. Substituting the right-moving ansatz

η = u + 14 αu2 + ( 1

2 θ2 − 13 )βuxx + · · · (2.14)

and truncating to first order reduces the Boussinesq system to the Korteweg–de Vries equation

ut + ux + 32 αuux + 1

6 βuxxx = 0, (2.15)

for unidirectional propagation of long waves in shallow water. In the long wave regime, β � 1,the nonlinear effect is negligible, and the equation reduces to the linear equation

ut + ux + 16 βuxxx = 0, (2.16)

which, in turn, can be mapped to the third-order Airy equation (2.6) by passing to a movingcoordinate frame and then rescaling. Alternatively, because to leading order ut ≈ −ux, one canreplace the third derivative term, uxxx ≈ −uxxt, leading to the RLW/BBM model [24,37],

ut + ux + 32 αuux − 1

6 βuxxt = 0. (2.17)

Observe that both first-order models (2.15) and (2.17) are independent of the depth parameter θ ,which appears only in the second- and higher-order terms in α, β. Inverting the unidirectionalconstraint (2.14) leads to the same Korteweg–de Vries equation and RLW/BBM equations forthe surface elevation η. Again, this is only true up to first order, and the higher-order terms areslightly different [28]. Because the preceding unidirectional models do not admit the exact waterwave dispersion relation (2.8), Whitham [32] proposed the alternative model

ut + L[u] + 32 αuux = 0, (2.18)

in which the Fourier integral operator

L[u] = 12π

∫∞

−∞

∫∞

−∞ω(k) eik(x−ξ)u(t, ξ) dξ dk

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

6

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

exactly reproduces the rescaled water wave dispersion relation given by (2.8) when g = h = 1.However, Benjamin et al. [24] argue that the linear dispersion in Whitham’s model is insufficientlystrong to counteract the tendency to shock formation.

Let us also list a few important integrable nonlinear evolution equations, from a variety ofphysical sources. The nonlinear Schrödinger equation

iut = uxx + |u|2u (2.19)

arises in nonlinear optics as well as in the modulation theory for water waves [29,38].Its linearization has quadratic dispersion relation ω(k) = k2. Integrability was established byZakharov & Shabat [39], and the fact that it admits localized soliton solutions that preserve theirshape under collision plays an important role in transmission of signals through optical fibres[40]. The integrable Benjamin–Ono equation

ut + uux + H[uxx] = 0, (2.20)

in which H is the Hilbert transform, given by the Cauchy principal value integral

H[u] =∫∞

−∞− u(ξ)dξ

ξ − x, (2.21)

arises as a model for internal waves [41]. Its linearization has quadratic dispersion relation ω(k) =k2 sign k, which is the odd counterpart of the quadratic Schrödinger dispersion.

Another integrable model is the Boussinesq equation

utt = uxx + 32 α(u2)xx ± 1

3 βuxxxx, (2.22)

also derived by Boussinesq from the water wave problem. While the Boussinesq equation admitswaves propagating in both directions, it is not, in fact, equivalent to any of the bidirectionalBoussinesq systems (2.11), and, indeed, its derivation from the water wave problem relies onthe unidirectional hypothesis. The plus sign is the ‘bad Boussinesq’, which is unstable, whereasthe minus sign is the stable ‘good Boussinesq’. However, both versions admit solutions thatblow up in finite time; see McKean [42] for a detailed analysis of solutions to the periodicproblem, and also [43,44] for further results on stability and blow-up of solutions. An alternative,regularized version

utt = uxx + 32 α(u2)xx ± 1

3 βuxxtt (2.23)

was subsequently investigated by Whitham [13]. The integrable Boussinesq equation (2.23) andits regularization (2.2) were proposed as model for DNA dynamics by Muto [45] and Scott [46].The associated periodic boundary value problems can thus be viewed as a model for the dynamicsof DNA loops, whereas individual DNA strands require suitably adapted boundary conditionsat each end.

In table 1 below, we summarize dispersion relations arising from water waves (in units so thatg = h = 1) and some of the more important shallow water models, rescaled so that α = β = 1. Inthe first three cases, taking the positive square root corresponds to right-moving waves. The finalcolumn lists their leading-order asymptotics (omitting constant multiples) at large wavenumber.

3. Linear dispersion in periodic domainsIn general, let L be a scalar, constant coefficient (integro-)differential operator with purelyimaginary Fourier transform (FT) L(k) = iϕ(k). The associated scalar evolution equation

∂u∂t

= L[u] (3.1)

has real dispersion relation ω(k) = iL(k) = −ϕ(k).We subject (3.1) to periodic boundary conditions on the interval −π ≤ x ≤ π , (or, equivalently,

work on the unit circle: x ∈ S1) with initial conditions u(0, x) = f (x). To construct the solution, we

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

7

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

Table 1. Shallow water dispersion relations.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

water waves√k tanh k |k|1/2 sign k

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Boussinesq system, regularized Boussinesq equationk√

1 + 13 k

2sign k

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Boussinesq equation k√1 + 1

3 k2 k2 sign k

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Korteweg–de Vries k − 16 k

3 k3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

RLW/BBMk

1 + (1/6)k2k−1

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

expand the initial condition in a Fourier series:

f (x) ∼∞∑

k=−∞ck eikx, where ck = 1

∫π

−π

f (x) e−ikx dx. (3.2)

The solution to the periodic initial-boundary value problem then has time-dependentFourier series

u(t, x) ∼∞∑

k=−∞ck ei[kx−ω(k)t]. (3.3)

In particular, the fundamental solution u = F(t, x) takes a delta function as its initial data, u(0, x) =δ(x), and hence has the following Fourier expansion:

F(t, x) ∼ 12π

∞∑k=−∞

ei[kx−ω(k)t]. (3.4)

The solution (3.3) can then be rewritten as a convolution integral with the fundamental solution:

u(t, x) =∫π

−π

F(t, x − ξ)f (ξ) dξ . (3.5)

The following result underlies the dispersive quantization effect for equations with ‘integralpolynomial’ dispersion relations.

Definition 3.1. A polynomial P(k) = cmkm + · · · + c1k + c0 will be called integral if all itscoefficients are integers: cj ∈ Z, j = 0, . . . , m.

Theorem 3.2. Suppose that the dispersion relation of the evolution equation (3.1) is a multiple of anintegral polynomial: ω(k) = λP(k), for some 0 �= λ ∈ R. Let β = 2π/λ. Then, at every rational time t =βp/q, with p and 0 �= q ∈ Z, the fundamental solution (3.4) is a linear combination of q periodically extendeddelta functions concentrated at the rational nodes xj = 2π j/q for j ∈ Z.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

8

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

As in Olver [1], this result is an immediate consequence of the following well-known lemma:

Lemma 3.3. Let 0 < q ∈ Z. The coefficients of a complex Fourier series (3.2) are q-periodic in theirindices; so ck+q = ck for all k, if and only if the series represents a linear combination of the q periodicallyextended delta functions concentrated at the rational nodes xj = 2π j/q for j ∈ Z :

f (x) =∑

− 12 q<j≤ 1

2 q

ajδ

(x − 2π j

q

),

for certain a0, . . . , aq−1 ∈ C.

Applying theorem 3.2 to the superposition formula (3.5) produces the followingintriguing corollary.

Corollary 3.4. Under the assumptions of theorem 3.2, at a rational time t = βp/q, the solution profileto the periodic initial-boundary value problem is a linear combination of ≤ q translates of its initial datau(0, x) = f (x), i.e.

u(

βpq

, x)

=∑

− 12 q<j≤ 1

2 q

aj

(βpq

)f(

x − 2π jq

).

In the case of the ‘Riemann problem’, the initial data are the unit step function:

u(0, x) = f (x) = σ(x) ={

0, −π < x < 0,

1, 0 < x < π .(3.6)

Under the assumption that the dispersion relation is odd, ω(−k) = −ω(k), the Fourier expansionof the resulting solution is

u�(t, x) ∼ 12

+ 2π

∞∑j=0

sin[(2j + 1)x − ω(2j + 1)t]2j + 1

, (3.7)

whose spatial derivative is the difference of two translates of the fundamental solution, namely∂u�/∂x = F(t, x) − F(t, x − π). Non-odd dispersion relations produce complex solutions, whosereal part is given by the more complicated expression

12

+ 1π

∞∑j=0

sin[(2j + 1)x − ω(2j + 1)t] + sin[(2j + 1)x + ω(−2j − 1)t]2j + 1

. (3.8)

In the following figures, we display graphs of the solution (3.7) at some representativetimes, for various odd dispersion relations, most of which are associated with water wavesmodels. The figures are ordered by their asymptotic order α at large wavenumber: ω(k) =O(|k|α) as |k| → ∞. We did not use numerical integration to obtain these plots; rather, theywere produced in MATHEMATICA by summing the first 150 terms in the Fourier series (3.7),which was adequate to capture the essential detail. (In all the cases we looked at, summingthe first 1000 terms does not appreciably change the solution graphs.) In some plots, a tinyresidual Gibbs effect can be seen at the jump discontinuities. By comparing the various profiles,we conclude that the qualitative behaviour of the solution (3.7) depends crucially on theasymptotic behaviour of the dispersion relation ω(k) for large wavenumber. However, a fullunderstanding of the various qualitative behaviours cannot be gleaned from these few staticsolution profiles, and so the reader is strongly encouraged to watch the corresponding moviesposted at http://www.math.umn.edu/∼olver/mvdql.html. Each movie shows the dynamicalbehaviour of the solutions (3.7) for various dispersion relations over a range of times; theassociated figures at the end of the paper are selected frames extracted from the correspondingmovie. In some cases, the solution evolves at a glacially slow pace; so a representativecollection of sub-movies are posted. Again, we emphasize that all dispersion relations usedin the figures and movies are odd functions of k. More general dispersion relations, which

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

9

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

t = 1 t = 5 t = 10

t = 50 t = 100 t = 1000

Figure 2. RLW/BBM dispersion:ω = k/(1 + (1/6)k2). (Online version in colour.)

produce complex oscillatory solutions whose real part is given by the more complicatedformula (3.8), exhibit additional qualitative features that we will investigate more thoroughlyelsewhere.

Here is an attempt to convey in words the behaviours that can observed in the movies, but aresometimes less evident in the plots. In all cases considered, the large wavenumber asymptotics isfixed by a certain power of the wavenumber

ω(k) ∼ |k|α as |k| → ∞. (3.9)

In our experiments, the overall qualitative behaviour seems to be entirely determined bythe asymptotic exponent α, although the intricate details of the solution profile do depend uponthe particularities of the dispersion relation. As α varies, there is a continual change from one typeof qualitative behaviour to the next. Over the range −∞ < α < ∞, there appear to be five mainregions, which are roughly demarcated as follows.

— α ≤ 0. Large-scale oscillations: the two initial discontinuities, at 0 and π , remain stationaryand begin to produce oscillating waves moving to the left. After this initial phase,the solution settles into what is essentially a stationary up- and down-motion, withvery gradually accumulating waviness superimposed. Two examples are the dispersionrelations associated with the RLW/BBM equation (2.17), with α = −1, shown in figure 2,and the regularized Boussinesq equation (2.23) as well as some versions of the Boussinesqsystem (2.11), for which α = 0, shown in figure 3. Interestingly, the former becomes waviersooner, whereas in the latter, even after time t = 1000, the smaller scale oscillations are stillrather coarse, and numerical limitations prevented us from following it far enough to seevery high-frequency oscillations appear.

— 0 < α < 1. Dispersive oscillations: each of the two initial discontinuities changes into aslowly moving ‘seed’ that continually spawns a right-moving oscillatory wave train.After a while, the wave train emanating from one seed interacts with the other, andthe seeds follow along these increasingly rapid oscillations. Much later, the solutionhas assumed a slightly fractal wave form superimposed over a slowly oscillating ocean,similar to ripples on a swelling sea that moves up and down while gradually changingform. Examples include the full water wave dispersion relation, plotted in figure 4, andthe square root dispersion graphed in figure 5. Observe that both have the same largewavenumber asymptotics, and the corresponding solutions have very similar qualitativebehaviour, while differing in their fine details.

— α = 1. Slowly varying travelling waves: as α approaches 1, the oscillatory waves acquire anincreasingly noticeable motion to the right. Indeed, if the dispersion relation is exactlylinear, ω(k) = k, the solution is a pure travelling wave u(t, x) = f (x − t), which moves

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

10

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

t = 1 t = 5 t = 10

t = 50 t = 100 t = 1000

Figure 3. Regularized Boussinesq dispersion:ω = k/√1 + 1

3 k2. (Online version in colour.)

t = 1 t = 2 t = 5

t = 10 t = 20 t = 35

t = 50 t = 75 t = 100

Figure 4. Water wave dispersion:ω = √k tanh k sign k. (Online version in colour.)

t = 2 t = 5 t = 10

t = 20 t = 50 t = 100

Figure 5. Square root dispersion:ω = √|k| sign k. (Online version in colour.)

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

11

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

t = 0.5 t = 1 t = 5

t = 50 t = 500 t = 5000

Figure 6. Asymptotically linear dispersion:ω = √|k(k + 1)| sign k. (Online version in colour.)

t = 0.05 t = 0.1 t = 0.2

t = 0.5 t = 1 t = 2

Figure 7. Three-halves dispersion:ω = |k|3/2 sign k. (Online version in colour.)

unchanged with unit speed. Having asymptotically, but not exactly linear, dispersionrelation superimposes slowly varying oscillations on top of the travelling wave. Anexample appears in figure 6, based on the asymptotically linear dispersion relationω(k) = √|k(k + 1)|, sign k associated with, for example, the equation2

utt = uxx + iux. (3.10)

— 1 < α < 2. Oscillatory becoming fractal: in this range, each of the initial discontinuitiesproduces an oscillatory wave train propagating to the right. After the wave trainencounters the other discontinuity, the oscillations become increasingly rapid over theentire interval, and, after a while, the entire solution acquires a fractal profile, whichdisplays an overall motion to the right while its small scale features vary rapidly andseemingly chaotically. An example is provided in figure 7, based on the dispersionrelation ω(k) = |k|3/2 sign k corresponding to the equation

utt = −iuxxx. (3.11)

— α ≥ 2. Fractal/quantized: once α reaches 2, the solution exhibits the fractalization/quantization phenomenon, which provably happens when the dispersion relation is anintegral polynomial, e.g. ω(k) = km, where 2 ≤ m ∈ Z. In such cases, theorem 3.2 impliesthat the solution quantizes at rational times, meaning that, when t = βp/q, it is piecewiseconstant on intervals of length 2π/q, and, as proved by Oskolkov [11], continuous butnowhere differentiable at irrational times. For non-polynomial dispersion with integralasymptotic exponent 2 ≤ α ∈ Z, most frames in the movie show a fractal profile, but

2The figure can be viewed as graphing the real part of the right-moving component of the solution to the periodic Riemanninitial value problem.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

12

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

t = p130

t = p115

t = p110

t = 0.1 t = 0.2 t = 0.3

Figure 8. Benjamin–Ono dispersion:ω = |k|2 sign k. (Online version in colour.)

t = p130

t = p115

t = p110

t = 0.1 t = 0.2 t = 0.3

Figure 9. Integrable Boussinesq dispersion:ω = k√1 + 1

3 k2. (Online version in colour.)

occasionally the profile will abruptly quantize, with jump discontinuities punctuatingnoticeably less fractal subgraphs. We presume that, as in the polynomial case, thetimes at which the solution (approximately) quantizes are densely embedded in thetimes at which it has a continuous, fractal profile. However, a rigorous proof of thisobservation appears to be quite difficult, requiring delicate Fourier estimates. Examplesinclude the linearized Korteweg–de Vries equation (2.6), and linearized Benjamin–Onoequation (2.20), both of which precisely quantize, as seen in figures 8 and 11, andthe linearization of the integrable Boussinesq equation (2.2), which appears to do soapproximately, as in figure 9. On the other hand, if 2 ≤ α �∈ Z is not an integer (or at leastnot very close to an integer), only fractal solution profiles are observed; see figure 10 forthe case ω(k) = |k|5/2 sign k corresponding to the equation

utt = iuxxxxx. (3.12)

Observe that, as an immediate consequence of corollary 3.4, for integral polynomialdispersion relations, and, presumably, more general relations in this region, thequantization effect persists under the addition of noise to the initial data, although thesmall jumps at rational numbers with large denominators would be overwhelmed bythe noise, whereas at irrational steps one ends up with a ‘noisy fractal’.

Let us outline a justification of our observation that the quantization and fractalizationphenomena hold for dispersion relations that are asymptotically polynomial at largewavenumber. Assume that

ω(k) ∼ λP(k) + O(k−1) as |k| −→ ∞, (3.13)

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

13

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

t = p130

t = p115

t = p110

t = 0.1 t = 0.2 t = 0.3

Figure 10. Five-halves dispersion:ω = |k|5/2 sign k. (Online version in colour.)

t = p130

t = p115

t = p110

t = 0.1 t = 0.2 t = 0.3

Figure 11. Korteweg–de Vries dispersion:ω = k − 16 k

3. (Online version in colour.)

where 0 �= λ ∈ R, and P(k) is an integral polynomial. Let us rewrite the corresponding complexFourier series solution in the form

u(t, x) = b0 +∞∑

k=−∞k �=0

bk

kei[kx−ω(k)t] = b0 +

∞∑k=−∞

k �=0

bk

kei[kx−p(k)t] +

∞∑k=−∞

k �=0

rk eikx,

We assume that the coefficients bk, which are prescribed by the initial data, are uniformlybounded, |bk| ≤ M, as is the case with the step function. With this assumption, rk = O(k−2), andhence the final series represents a uniformly and absolutely convergent Fourier series, whosesum is a continuous function. Thus, the discontinuities of u(t, x) will be determined by the initialseries, which, by theorem 3.2, exhibits jump discontinuities and quantization at rational times.We conclude that solutions to linear equations with such asymptotically integral polynomialdispersion relations will exhibit quantization at the rational times t = βp/q, where β = 2π/λ.A rigorous proof of fractalization at irrational times will require a more detailed analysis.

A similar argument explains why the discontinuities in the solution remain (almost) stationary,when the asymptotic exponent α < 0. Formally, suppose ω(k) = |k|αμ(k), where α < 0 and μ(k) =O(1). Then, the Fourier series solution has the form

u(t, x) = b0 +∞∑

k=−∞k �=0

bk

kei[kx−ω(k)t] = b0 +

∞∑k=−∞

k �=0

bk

keikx[1 − it|k|αμ(k)].

Again, assuming that the coefficients bk are uniformly bounded, the remainder terms are of orderk−1+α with α < 0, and hence represent an uniformly convergent series whose sum is continuous.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

14

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

RLW/BBM

w(k) = k1.1 sign k w(k) = k5/4 sign k

w(k) = k3/2 sign k

w(k) = k5/2 sign k

water waves

Benjamin–Ono

integrable Boussinesq

w(k) = k4 sign kKorteweg–deVries

Figure 12. Plots of u(t, 0) for representative dispersion relations. (Online version in colour.)

We conclude that the discontinuities of u(t, x) are completely determined by those of the leadingterm, which are stationary.

Finally, in figure 12, we display graphs of the temporal evolution of the solution at the origin,u(t, 0), for a representative subset of the dispersion relations we’ve considered. In the first twofigures, for the RLW/BBM and water wave dispersion, we graph on the long time interval0 ≤ t ≤ 100, while in the other plots, the time interval is 0 ≤ t ≤ 2π . When the asymptotic power(3.9) governing the high wavenumber dispersion satisfies α ≤ 1, the time plot appears smooth.Following the non-rigorous analysis presented in Berry [2, §2], for α ≥ 1, the fractal dimension ofthe time plot should be 2 − 1/(2α). This seems correct as stated when α ≥ 2, as these plots becomeincreasingly fractal with increasing asymptotic dispersion power α. However, for α near 1, theplot appears to be initially smooth for a short time, which is followed by the appearance ofincreasingly rapid oscillations. After some further time has elapsed, the graph eventually assumesa fractal form.

4. Fractalization and quantization in nonlinear systemsIn this final section, we turn our attention to the nonlinear regime. Basic numerical techniques willbe used to approximate the solutions to the Riemann initial value problem on a periodic domain.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

15

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

While we do not have any rigorous theoretical results to justify the resulting observations, ournumerical studies strongly indicate that the variety of qualitative solution features observed inthe underlying linear problem persist, at the very least, into the weakly nonlinear regime. Inparticular, the fractalization and quantization of solutions to the Airy evolution equation areobserved in the numerical solutions to both the integrable Korteweg–de Vries equation andits non-integrable quartic counterpart. Similar phenomena have been found in nonlinearSchrödinger equations, but these will be reported on elsewhere.

Our numerical algorithms are based on a standard operator splitting method [47]. Considerthe initial value problem for an evolution equation

ut = K[u], u(t0, x) = u0(x), (4.1)

where K is a differential operator (linear or nonlinear) in the spatial variable. Here, the system isalways supplemented by periodic boundary conditions, but the numerical scheme is of completegenerality. Assuming the basic existence and uniqueness of solutions to the initial value problem,at least over a certain time interval 0 < t < t�, we denote the resulting solution to (4.1), i.e. theinduced flow, by u(t, · ) = ΦK(t)u0.

Operator splitting relies on writing the spatial operator as a sum

K[u] = L[u] + N[u] (4.2)

of two simpler operators, which can each be more readily solved numerically. In all cases treatedhere, L[u] represents the linear part of the evolution equation (4.1), whereas the nonlinear termsappear in N[u]. For example, writing the Korteweg–de Vries equation in the reduced form3

ut = K[u] = −uxxx + uux, (4.3)

we split the right-hand side into the sum of the linear part L[u] = −uxxx, with cubic dispersion,and the nonlinear part given by the inviscid Burgers’ operator N[u] = uux.

For numerical purposes, we are interested in approximating the solution un(x) = u(tn, x) atthe mesh points 0 = t0 < t1 < t2 < · · · which, for simplicity, we assume to be equally spaced, with�t = tn − tn−1 fixed. (Of course, we will also discretize the spatial variable.) We use u�(tn, x) =u�

n (x) to denote the numerical approximation to the solution to the original evolution equation attimes tn = n�t, which is obtained by successively applying the flows ΦL and ΦN corresponding,respectively, to the linear and nonlinear parts in the splitting (4.2). The simplest splitting algorithmis the Godunov scheme:

u�n+1 = ΦL(�t)ΦN(�t)u�

n , n = 0, 1, 2, . . . , u�0 = u0. (4.4)

In favourable situations, for a suitable choice of norm and appropriate conditions on the initialdata, it can be proved that the Godunov scheme is first-order convergent:

‖u�n (x) − u(tn, x)‖ = O(�t), as �t → 0.

The convergence can be improved to second order by use of the Strang splitting scheme:

u�n+1 = ΦN( 1

2 �t)ΦL(�t)ΦN( 12 �t)u�

n , n = 0, 1, 2, . . . , u�0 = u0. (4.5)

In this case,

‖u�n (x) − u(tn, x)‖ = O(�t2), as �t → 0,

again under appropriate assumptions. In our numerical experiments, the results of the Godunovand Strang splitting are very similar, and so we display only the Godunov versions in the figures.

3The model version (2.15) can be changed into the reduced form by applying a combination of scaling and Galilean boost.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

16

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

0 1 2 3 4 5 6 7−0.5

t = 0.015

0

0.5

1.0

1.5

0 1 2 3 4 5 6 7t = 0.03

−1.0

−0.5

0

0.5

1.0

1.5

0 1 2 3 4 5 6 7t = 0.15

−1.0−0.5

00.51.01.52.0

0 1 2 3 4 5 6 7

t = 0.3

−1.0−0.5

00.51.01.52.0

0 1 2 3 4 5 6 7

t = 1.5

−0.5

0

0.5

1.0

1.5

0 1 2 3 4 5 6 7

t = 3.0

−1.0−0.5

00.51.01.52.0

Figure 13. Korteweg–de Vries equation: irrational times. (Online version in colour.)

0 1 2 3 4 5 6 7−1.0

t = 0.01p

t = 0.5p t = 1.5pt = p

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7t = 0.1p t = 0.25p

−0.5

0

0.5

1.0

1.5

−0.4−0.2

00.20.40.60.81.01.2

−0.20

0.20.40.60.81.01.2

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7−0.2

00.20.40.60.81.01.2

−0.20

0.20.40.60.81.01.2

−0.20

0.20.40.60.81.01.2

Figure 14. Korteweg–de Vries equation: rational times. (Online version in colour.)

In particular, to solve the initial value problem for the Korteweg–de Vries equation (4.3), weapply the Godunov splitting scheme (4.4) to its linear and nonlinear parts

ut = L[u] = −uxxx, ut = N[u] = ∂x(12 u2), (4.6)

each of which is subject to periodic boundary conditions. Periodicity and discretization ofthe spatial variable enables us to apply the fast Fourier transform (FFT) to exactly solve thediscretized linear equation. On the other hand, the nonlinear inviscid Burgers’ equation is first

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

17

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

written in conservative form as above. As in the linear case, we use FFT to write the numericalapproximation in the form of a Fourier series, apply convolution to square the numerical solutionand then differentiate by multiplication by the frequency variable. The resulting system ofordinary differential equations is mildly stiff, and is integrated by applying the backward Eulermethod, based on the implicit midpoint rule, using fixed point iteration to approximate thesolution to the resulting nonlinear algebraic system to within a prescribed tolerance. Because thetime step is small, difficulties with the formation of shock waves and other discontinuities do notcomplicate the procedure.

Our results, at rational and irrational times, are plotted in figures 13 and 14. These clearlysuggest the presence of both quantization and fractalization phenomena in the nonlinear system.It is striking that these appear, not just, as one might expect, in the weakly nonlinear regime, butin a fully nonlinear framework in which the coefficients of both the linear and nonlinear termsin (4.3) have comparable magnitude. Indeed, recent results of Erdogan et al. [48], imply that, forperiodic boundary conditions, the temporal evolution of high-frequency data for the Korteweg–de Vries equation and other nonlinear water wave models is nearly linear over long time intervals.So, while it may happen that these effects may eventually be overcome by the nonlinearity, anyserious investigation will require very accurate long time numerical integration schemes. Thecorresponding movies have been posted at http://www.math.umn.edu/∼olver/mvdqn.html.Our numerical results are also in agreement with a theorem of Erdogan & Tzirakis [49,corollary 1.5], that, when the initial data u(0, · ) lies in the Sobolev space Hs with s > − 1

2 , thenthe difference between solutions of the Korteweg–de Vries equation and its linearization can becontrolled, meaning that the difference lies in Hr for any r < min{3s + 1, s + 1}, and has at mostpolynomially growing Hr norm.

To verify that these phenomena are not restricted to integrable models such as theKorteweg–de Vries equation, in figures 15 and 16, we plot the solution to the non-integrablequartic Korteweg–de Vries equation

ut = −uxxx + u3ux = −uxxx + ( 14 u4)x. (4.7)

The numerical solution algorithm is again based on Godunov splitting, using the sameFourier-based algorithms to integrate the individual linear and nonlinear parts.

Keep in mind that the coefficients in both the Korteweg–de Vries and quartic Korteweg–deVries equations can be rescaled to any convenient non-zero values, and thus quantization andfractalization of the solution can be expected to appear in any version, perhaps with the overallmagnitude of the effect governed by the relative size of the initial discontinuities. Thus, the Talboteffect appears to be rather robust, and hence of importance to a wide range of linear and nonlineardispersive equations on periodic domains.

5. Further directionsWe conclude by listing a few of the possibly fruitful directions for further research.

— Our numerical integration schemes are fairly crude, and it would be worth implementingmore sophisticated algorithms. See Zuazua [50] for a recent survey on the numerics ofdispersive partial differential equations. In the linear case, the complicated behavioursexhibited by the (partial) Fourier series sums would serve as a good testing ground forthe application of numerical methods for dispersive waves to rough data.

— While the explicit formulas for the solutions are rather simple Fourier series, theamazing variety of observed behaviours, solution profiles, and fascinating spatial andtemporal patterns indicates that formulation of theorems and rigorous proofs will bevery challenging. Progress may well rely on delicate estimates of the type used in theanalysis of number-theoretic exponential sums of Gauss and Weyl type [15]. One is alsoreminded of Bourgain’s celebrated analysis of the periodic problems for the nonlinear

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

18

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7

−0.5

t = 0.01 t = 0.02 t = 0.1

0

0.5

1.0

1.5

−0.5

0

0.5

1.0

1.5

−1.0

−0.5

0

0.5

1.0

1.5

2.0

−0.4−0.2

00.20.40.60.81.01.21.4

−0.4−0.2

00.20.40.60.81.01.21.41.6

t = 0.2 t = 1 t = 2

−1.0

−0.5

0

0.5

1.0

1.5

2.0

Figure 15. Quartic Korteweg–de Vries equation: irrational times. (Online version in colour.)

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7−0.5

0

0.5

1.0

1.5

t = 01p t = 0.1p t = 0.25p

t = 0.5p t = p t = 1.5p

−0.4−0.2

00.20.40.60.81.01.2

−0.20

0.20.40.60.81.01.2

0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7−0.2

00.20.40.60.81.01.2

−0.4−0.2

00.20.40.60.81.01.21.4

−0.4−0.2

00.20.40.60.81.01.21.4

Figure 16. Quartic Korteweg–de Vries equation: rational times. (Online version in colour.)

Schrödinger and Korteweg–de Vries equations [51,52], which involves similarly subtlenumber-theoretic analysis.

— Zabusky and Kruskal’s numerical experimentations that led to the discovery of thesoliton were motivated by the fact that the Korteweg–de Vries and Boussinesq equationsare continuum limits of the nonlinear mass-spring chains, whose surprising lack ofthermalization came to light in the seminal work of Fermi et al. [53]. It would beinteresting to see how the foregoing dispersive effects might be manifested in the originalFermi–Pasta–Ulam chains when the initial displacement includes a sharp transition.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

19

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

— An interesting and apparently difficult question is to determine the fractal dimensionof the graphs of solutions at fixed times that are given by such slowly convergingFourier series. The wide variety of individual space plots in the accompanying figuresindicates that this is considerably more subtle than the Fourier series that were rigorouslyanalysed by Chamizo & Córdoba [54], or those of Weierstrass–Mandelbrot form thatwere investigated by Berry & Lewis [55]. Indeed, for fixed t, all of our series have thesame slow O(1/k) rate of decay of their Fourier coefficients, and hence the fractal natureof their graphs depends essentially on the detailed dispersion relation asymptotics.Further, when the large wavenumber asymptotic power is in the range 1 < α < 2, thegraphs appear to be (mostly) smooth over some initial time interval, then increasinglyoscillatory, only later apparently achieving a fractal nature. Using Besov space methods,Rodnianski [9] rigorously proves that the graphs of real and imaginary parts of roughsolutions to the linear Schrödinger equation have fractal dimension 3

2 at a densesubset of irrational times, reconfirming some of Berry’s observations [2]. We do notknow how difficult it would be to extend Rodnianski’s techniques to more generaldispersion relations.

— We have concentrated on the periodic boundary value problem for linearly dispersivewave equations. The behaviour under other boundary conditions, e.g. u(t, 0) = ux(t, 0) =u(t, 2π) = 0, is not so clear because, unlike the Schrödinger equation, these boundaryvalue problems are not naturally embedded in the periodic version. Fokas & Bona [56–58]have developed a new solution technique for initial-boundary value problems for linearand integrable nonlinear dispersive partial differential equations based on novel complexintegral representations. It would be instructive to investigate how the effects of thedispersion relation in the periodic and other problems are manifested in this approach.

— Another important direction would be to extend our analysis to linearly dispersiveequations in higher space dimensions. Berry [2] gives a non-rigorous argument that, atleast for the linearized Schrödinger equation, a fractal Talbot effect persists for higherdimensional boundary value problems. Other important examples worth investigatinginclude the integrable Kadomtsev–Petviashvili and Davey–Stewartson equations [41],and non-integrable three-dimensional surface wave models found in [27,30].

We thank Michael Berry, Jerry Bona, Adrian Diaconu, Dennis Hejhal, Svitlana Mayboroda and KonstantinOskolkov for supplying references and helpful advice. We also thank the anonymous referees for helpfuland thought-provoking comments. The work of second author is supported in part by NSF grant no.DMS–1108894.

References1. Olver PJ. 2010 Dispersive quantization. Am. Math. Mon. 117, 599–610. (doi:10.4169/

000298910X496723)2. Berry MV. 1996 Quantum fractals in boxes. J. Phys. A 29, 6617–6629. (doi:10.1088/0305-

4470/29/20/016)3. Berry MV, Bodenschatz E. 1999 Caustics, multiply-reconstructed by Talbot interference. J. Mod.

Optics 46, 349–365.4. Berry MV, Klein S. 1996 Integer, fractional and fractal Talbot effects. J. Mod. Optics 43,

2139–2164. (doi:10.1080/09500349608232876)5. Berry MV, Marzoli I, Schleich W. 2001 Quantum carpets, carpets of light. Phys.World 14, 39–44.6. Hannay JH, Berry MV. 1980 Quantization of linear maps on a torus–fresnel diffraction by a

periodic grating. Physica D 1, 267–290. (doi:10.1016/0167-2789(80)90026-3)7. Talbot HF. 1836 Facts related to optical science. No. IV. Phil. Mag. 9, 401–407.8. Kapitanski L, Rodnianski I. 1999 Does a quantum particle know the time? In Emerging

applications of number theory, vol. 109 (eds D Hejhal, J Friedman, MC Gutzwiller & AMOdlyzko). IMA Volumes in Mathematics and its Applications, pp. 355–371. New York, NY:Springer.

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

20

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

9. Rodnianski I. 2000 Fractal solutions of the Schrödinger equation. Contemp. Math. 255, 181–187.(doi:10.1090/conm/255/03981)

10. Oskolkov KI. 1998 Schrödinger equation and oscillatory Hilbert transforms of second degree.J. Fourier Anal. Appl. 4, 341–356. (doi:10.1007/BF02476032)

11. Oskolkov KI. 1992 A class of I.M. Vinogradov’s series and its applications in harmonicanalysis. In Progress in approximation theory, vol. 19. Springer Series Comput. Math.,pp. 353–402. New York, NY: Springer.

12. Taylor M. 2003 The Schrödinger equation on spheres. Pacific J. Math. 209, 145–155.(doi:10.2140/pjm.2003.209.145)

13. Whitham GB. 1974 Linear and nonlinear waves. New York, NY: John Wiley & Sons.14. Hardy GH, Littlewood JE. 1914 Some problems of Diophantine approximation. II. The

trigonometric series associated with the elliptic ϑ-functions. Acta Math. 37, 193–238.(doi:10.1007/BF02401834)

15. Vinogradov IM. 2004 The method of trigonometrical sums in the theory of numbers. Mineola, NY:Dover Publication.

16. Ford K. 2002 Recent progress on the estimation of Weyl sums. In Modern problems of numbertheory and its applications: current problems, Part II (eds VN Chubarikov & GI Arkhipov),pp. 48–66. Mosk. Gos. Univ. im. Lomonosova, Mekh.-Mat. Fak., Moscow. Seehttp://www.ams.org/mathscinet/search/publdoc.html?pg1=IID&s1=325647&vfpref=html&r=34&mx-pid=1985940.

17. Zabusky NJ, Kruskal MD. 1965 Interaction of ‘solitons’ in a collisionless plasma and therecurrence of initial states. Phys. Rev. Lett. 15, 240–243. (doi:10.1103/PhysRevLett.15.240)

18. Lax PD, Levermore CD. 1983 The small dispersion limit of the Korteweg–de Vries equationI–III. Commun. Pure Appl. Math. 36, 253–290. (doi:10.1002/cpa.3160360302)

19. Lax PD, Levermore CD, Venakides S. 1993 The generation and propagation of oscillationsin dispersive initial value problems and their limiting behavior. In Important developments insoliton theory (eds AS Fokas & VE Zakharov), pp. 205–241. Berlin, Germany: Springer.

20. Venakides S. 1985 The zero dispersion limit of the Korteweg–de Vries equation with nontrivialreflection coefficient. Commun. Pure Appl. Math. 38, 125–155. (doi:10.1002/cpa.3160380202)

21. Venakides S. 1987 The zero dispersion limit of the Korteweg–de Vries equation with periodicinitial data. Trans. Am. Math. Soc. 301, 189–225. (doi:10.1090/S0002-9947-1987-0879569-7)

22. Lax PD. 1975 Periodic solutions to the KdV equation. Commun. Pure Appl. Math. 28, 141–188.(doi:10.1002/cpa.3160280105)

23. McKean HP, van Moerbeke P. 1975 The spectrum of Hill’s equation. Invent. Math. 30, 217–274.(doi:10.1007/BF01425567)

24. Benjamin TB, Bona JL, Mahony JJ. 1972 Model equations for long waves in nonlineardispersive systems. Phil. Trans. R. Soc. Lond. A 272, 47–78. (doi:10.1098/rsta.1972.0032)

25. Thaller B. 2000 Visual quantum mechanics. New York, NY: Springer.26. Olver PJ. In press. Introduction to partial differential equations. Upper Saddle River, NJ: Pearson.27. Bona JL, Chen M, Saut J-C. 2002 Boussinesq equations and other systems for small-amplitude

long waves in nonlinear dispersive media. I. Derivation and linear theory. J. Nonlinear Sci. 12,283–318. (doi:10.1007/s00332-002-0466-4)

28. Olver PJ. 1984 Hamiltonian perturbation theory and water waves. Contemp. Math. 28, 231–249.(doi:10.1090/conm/028/751987)

29. Yuen HC, Lake BM. 1980 Instabilities of waves on deep water. Ann. Rev. Fluid Mech. 12,303–334. (doi:10.1146/annurev.fl.12.010180.001511)

30. Nwogu O. 1993 Alternative form of Boussinesq equations for nearshore wave propagation.J. Waterway Port Coastal Ocean Eng. 119, 618–638. (doi:10.1061/(ASCE)0733-950X(1993)119:6(618))

31. Bona JL, Smith R. 1976 A model for the two-way propagation of water waves in a channel.Math. Proc. Camb. Phil. Soc. 79, 167–182. (doi:10.1017/S030500410005218X)

32. Whitham GB. 1967 Variational methods and applications to water waves. Proc. R. Soc. Lond. A299, 6–25. (doi:10.1098/rspa.1967.0119)

33. Broer LJF. 1975 Approximate equations for long water waves. Appl. Sci. Res. 31, 377–395.(doi:10.1007/BF00418048)

34. Kaup DJ. 1975 A higher-order water-wave equation and the method for solving it. Prog. Theor.Phys. 54, 396–408. (doi:10.1143/PTP.54.396)

35. Kupershmidt BA. 1985 Mathematics of dispersive waves. Commun. Math. Phys. 99, 51–73.(doi:10.1007/BF01466593)

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from

21

rspa.royalsocietypublishing.orgProcRSocA469:20120407

..................................................

36. Bona JL, Chen M, Saut J-C. 2004 Boussinesq equations and other systems for small-amplitudelong waves in nonlinear dispersive media. II. The nonlinear theory. Nonlinearity 17, 925–952.(doi:10.1088/0951-7715/17/3/010)

37. Peregrine DH. 1966 Calculations of the development of an undular bore. J. Fluid Mech. 25,321–330. (doi:10.1017/S0022112066001678)

38. Zakharov VE. 1968 Stability of periodic waves of finite amplitude on the surface of a deepfluid. J. Appl. Mech. Tech. Phys. 2, 190–194.

39. Zakharov VE, Shabat AB. 1974 A scheme for integrating the nonlinear equations ofmathematical physics by the method of the inverse scattering problem. I. Funct. Anal. Appl.8, 226–235. (doi:10.1007/BF01075696)

40. Hasegawa A, Matsumoto M. 2003 Optical solitons in fibers, 3rd edn. New York, NY: Springer.41. Drazin PG, Johnson RS. 1989 Solitons: an introduction. Cambridge, UK: Cambridge University

Press.42. McKean HP. 1981 Boussinesq’s equation on the circle. Commun. Pure Appl. Math. 34, 599–691.

(doi:10.1002/cpa.3160340502)43. Bona JL, Sachs RL. 1988 Global existence of smooth solutions and stability of

solitary waves for a generalized Boussinesq equation. Commun. Math. Phys. 118, 15–29.(doi:10.1007/BF01218475)

44. Sachs RL. 1990 On the blow-up of certain solutions of the ‘good’ Boussinesq equation. Appl.Anal. 36, 145–152. (doi:10.1080/00036819008839928)

45. Muto V. 2011 Soliton oscillations for DNA dynamics. Acta Appl. Math. 115, 5–15. (doi:10.1007/s10440-010-9564-8)

46. Scott AC. 1985 Soliton oscillations in DNA. Phys. Rev. A 31, 3518–3519. (doi:10.1103/PhysRevA.31.3518)

47. Holden H, Karlsen KH, Risebro NH, Tao T. 2011 Operator splitting for the KdV equation.Math. Comp. 80, 821–846. (doi:10.1090/S0025-5718-2010-02402-0)

48. Erdogan, Tzirakis N, Zharnitsky V. 2011 Nearly linear dynamics of nonlinear dispersivewaves. Physica D 240, 1325–1333. (doi:10.1016/j.physd.2011.05.009)

49. Erdogan MB, Tzirakis N. In press. Global smoothing for the periodic KdV evolution. Inter.Math. Res. Notes. (doi:10.1093/imrn/rns189)

50. Zuazua E. 2005 Propagation, observation, and control of waves approximated by finitedifference methods. SIAM Rev. 47, 197–243. (doi:10.1137/S0036144503432862)

51. Bourgain J. 1993 Exponential sums and nonlinear Schrödinger equations. Geom. Funct. Anal.3, 157–178. (doi:10.1007/BF01896021)

52. Bourgain J. 1997 Periodic Korteweg de Vries equation with measures as initial data. SelectaMath. 3, 115–159. (doi:10.1007/s000290050008)

53. Fermi E, Pasta J, Ulam S. 1974 Studies of nonlinear problems. I. Los Alamos Report LA 1940,1955. In Nonlinear wave motion, vol. 15 (ed. AC Newell), Lectures in Applied Mathematics,pp. 143–156. Providence, RI: American Mathematical Society.

54. Chamizo F, Córdoba A. 1999 Differentiability and dimension of some fractal Fourier series.Adv. Math. 142, 335–354. (doi:10.1006/aima.1998.1792)

55. Berry MV, Lewis ZV. 1980 On the Weierstrass–Mandelbrot fractal function. Proc. R. Soc. Lond.A 370, 459–484. (doi:10.1098/rspa.1980.0044)

56. Bona JL, Fokas AS. 2008 Initial-boundary-value problems for linear and integrablenonlinear dispersive partial differential equations. Nonlinearity 21, T195–T203. (doi:10.1088/0951-7715/21/10/T03)

57. Fokas AS. 2008 A unified approach to boundary value problems. CBMS–NSF Conference Series inApplied Math., vol. 78. Philadelphia, PA: SIAM.

58. Fokas AS, Spence EA. 2012 Synthesis, as opposed to separation, of variables. SIAM Rev. 54,291–324. (doi:10.1137/100809647)

on August 20, 2018http://rspa.royalsocietypublishing.org/Downloaded from