9
Nonlinear Schrödinger invariants and wave statistics F. Fedele, 1,a Z. Cherneva, 2,b M. A. Tayfun, 3,c and C. Guedes Soares 2,d 1 School of Civil and Environmental Engineering, Georgia Institute of Technology, Savannah, Georgia 31407, USA 2 Centre for Marine Technology and Engineering (CENTEC), Technical University of Lisbon, Instituto Superior Técnico, Lisbon 1049-001, Portugal 3 Department of Civil Engineering, Kuwait University, Safat 13060, Kuwait Received 16 September 2009; accepted 26 January 2010; published online 17 March 2010 Third-order quasiresonant interactions among free waves and associated modulational instabilities can significantly affect the statistics of various surface features in narrowband waves. In particular, modulational instabilities tend to induce intermittent amplifications on the surface displacements, causing their statistics to deviate from the linear Gaussian and second-order models. Herein, we investigate the nature of such instabilities on the statistical and spectral characteristics of deep-water waves generated in a large wave basin. We analyze the spectral changes that occur as waves propagate along the basin, develop bounds on the spectrum bandwidth, and interpret various statistics based on third-order Gram–Charlier distributions. © 2010 American Institute of Physics. doi:10.1063/1.3325585 I. INTRODUCTION Surface waves are nonlinear in nature, in particular, rela- tively long-crested waves because of their tendency to be affected by the interactions among freely propagating el- ementary waves with random amplitudes and phases. Conse- quently, their dynamics, kinematics, and spectral evolution in space and time are often interpreted stochastically, using per- turbation models. If nonlinearities are not particularly sig- nificant, the statistics of various surface features tend to fol- low a Gaussian structure, modeled as the linear superposition of a large number of elementary wavelets with Rayleigh- distributed amplitudes and random phases. At this level of approximation, surface displacements are “symmetric” with respect to the mean sea level. As a result, wave crest and trough amplitudes are Rayleigh-distributed, and large waves are likely to be generated by the linear focusing of elemen- tary phases Lindgren, 1 Boccotti, 2,3 Fedele, 4 and Fedele and Tayfun 5 . At the next level of approximation to O in wave steepness , the perturbations of the Stokes equations lead to a variety of random models in which the first-order Gaussian structure is modified by second-order nonresonant bound waves see, e.g., Longuet-Higgings 6 . The latter are non- Gaussian and phase-coupled to the linear waves, making wave crests sharper and narrower and troughs shallower and more rounded. As a result, surface elevations are positively skewed, and the distributions of wave crests and trough am- plitudes asymmetrically deviate from the Rayleigh law Fedele and Tayfun 5 . Models constructed to reflect such second-order nonlinearities tend to describe the statistics of wave heights, and crest and trough amplitudes fairly accu- rately Fedele and Tayfun, 5 Tayfun and Fedele, 7 and Tayfun 8 . To O 2 , third-order multiple scale perturbation solu- tions of the Stokes equations show that energy is transferred via resonant and quasiresonant interactions to longer and also shorter scales where it is dissipated by breaking or vis- cous dissipation. The resulting sea state is referred to as “wave” or “weak” turbulence WT in analogy with the Kol- mogorov energy cascade in fluid turbulence Zakharov 9,10 . WT states ensue from the space-time evolution of a sea of weakly nonlinear coupled dispersive waves in accordance with the Zakharov equation, 9 valid for an arbitrary spectral width. In WT, an initial Gaussian field is weakly modulated as nonlinearities develop in time, leading to intermittency in the turbulent signal due to the formation of sparse but coher- ent structures. In recent numerical studies see, e.g., Onorato et al. 11 , it is speculated that the large wave crests observed during these localized events may explain the occurrence of abnormal, rogue or freak waves. These are unusually large waves that appear from nowhere in the open ocean, as has been identified in records of full scale waves Guedes Soares et al. 12,13 . Their frequency of occurrence significantly ex- ceeds the theoretical predictions based on linear Gaussian or second-order statistical models Socquet-Juglard et al., 14 Petrova et al., 15 and Dysthe et al. 16 . Up to date, rogue waves have systematically been ob- served in unidirectional narrowband waves mechanically generated in tanks Onorato et al., 17,18 Petrova and Guedes Soares, 19 Shemer and Sergeeva, 20 and Cherneva et al. 21 . Near-resonant interactions and associated Benjamin–Feir- type modulational instabilities 22 appear as the essential fea- tures of the evolutionary dynamics of such waves. Typically, an initially narrowband wave train can undergo intense modulations, attended by an asymmetrical growth of the spectral sidebands, enhancing the occurrence of larger waves Janssen 23 . As a result, the distribution of wave and crest a Electronic mail: [email protected]. b Electronic mail: [email protected]. c Elelctronic mail: [email protected]. d Electronic mail: [email protected]. PHYSICS OF FLUIDS 22, 036601 2010 1070-6631/2010/223/036601/9/$30.00 © 2010 American Institute of Physics 22, 036601-1

Nonlinear Schrödinger invariants and wave statistics

Embed Size (px)

Citation preview

Nonlinear Schrödinger invariants and wave statisticsF. Fedele,1,a� Z. Cherneva,2,b� M. A. Tayfun,3,c� and C. Guedes Soares2,d�

1School of Civil and Environmental Engineering, Georgia Institute of Technology, Savannah,Georgia 31407, USA2Centre for Marine Technology and Engineering (CENTEC), Technical University of Lisbon,Instituto Superior Técnico, Lisbon 1049-001, Portugal3Department of Civil Engineering, Kuwait University, Safat 13060, Kuwait

�Received 16 September 2009; accepted 26 January 2010; published online 17 March 2010�

Third-order quasiresonant interactions among free waves and associated modulational instabilitiescan significantly affect the statistics of various surface features in narrowband waves. In particular,modulational instabilities tend to induce intermittent amplifications on the surface displacements,causing their statistics to deviate from the linear Gaussian and second-order models. Herein, weinvestigate the nature of such instabilities on the statistical and spectral characteristics of deep-waterwaves generated in a large wave basin. We analyze the spectral changes that occur as wavespropagate along the basin, develop bounds on the spectrum bandwidth, and interpret variousstatistics based on third-order Gram–Charlier distributions. © 2010 American Institute of Physics.�doi:10.1063/1.3325585�

I. INTRODUCTION

Surface waves are nonlinear in nature, in particular, rela-tively long-crested waves because of their tendency to beaffected by the interactions among freely propagating el-ementary waves with random amplitudes and phases. Conse-quently, their dynamics, kinematics, and spectral evolution inspace and time are often interpreted stochastically, using per-turbation models. If nonlinearities are not particularly sig-nificant, the statistics of various surface features tend to fol-low a Gaussian structure, modeled as the linear superpositionof a large number of elementary wavelets with Rayleigh-distributed amplitudes and random phases. At this level ofapproximation, surface displacements are “symmetric” withrespect to the mean sea level. As a result, wave crest andtrough amplitudes are Rayleigh-distributed, and large wavesare likely to be generated by the linear focusing of elemen-tary phases �Lindgren,1 Boccotti,2,3 Fedele,4 and Fedele andTayfun5�.

At the next level of approximation to O��� in wavesteepness �, the perturbations of the Stokes equations lead toa variety of random models in which the first-order Gaussianstructure is modified by second-order nonresonant boundwaves �see, e.g., Longuet-Higgings6�. The latter are non-Gaussian and phase-coupled to the linear waves, makingwave crests sharper and narrower and troughs shallower andmore rounded. As a result, surface elevations are positivelyskewed, and the distributions of wave crests and trough am-plitudes asymmetrically deviate from the Rayleigh law�Fedele and Tayfun5�. Models constructed to reflect suchsecond-order nonlinearities tend to describe the statistics ofwave heights, and crest and trough amplitudes fairly accu-

rately �Fedele and Tayfun,5 Tayfun and Fedele,7 andTayfun8�.

To O��2�, third-order multiple scale perturbation solu-tions of the Stokes equations show that energy is transferredvia resonant and quasiresonant interactions to longer andalso shorter scales where it is dissipated by breaking or vis-cous dissipation. The resulting sea state is referred to as“wave” or “weak” turbulence �WT� in analogy with the Kol-mogorov energy cascade in fluid turbulence �Zakharov9,10�.WT states ensue from the space-time evolution of a sea ofweakly nonlinear coupled dispersive waves in accordancewith the Zakharov equation,9 valid for an arbitrary spectralwidth. In WT, an initial Gaussian field is weakly modulatedas nonlinearities develop in time, leading to intermittency inthe turbulent signal due to the formation of sparse but coher-ent structures. In recent numerical studies �see, e.g., Onoratoet al.11�, it is speculated that the large wave crests observedduring these localized events may explain the occurrence ofabnormal, rogue or freak waves. These are unusually largewaves that appear from nowhere in the open ocean, as hasbeen identified in records of full scale waves �Guedes Soareset al.12,13�. Their frequency of occurrence significantly ex-ceeds the theoretical predictions based on linear Gaussian orsecond-order statistical models �Socquet-Juglard et al.,14

Petrova et al.,15 and Dysthe et al.16�.Up to date, rogue waves have systematically been ob-

served in unidirectional narrowband waves mechanicallygenerated in tanks �Onorato et al.,17,18 Petrova and GuedesSoares,19 Shemer and Sergeeva,20 and Cherneva et al.21�.Near-resonant interactions and associated Benjamin–Feir-type modulational instabilities22 appear as the essential fea-tures of the evolutionary dynamics of such waves. Typically,an initially narrowband wave train can undergo intensemodulations, attended by an asymmetrical growth of thespectral sidebands, enhancing the occurrence of larger waves�Janssen23�. As a result, the distribution of wave and crest

a�Electronic mail: [email protected]�Electronic mail: [email protected]�Elelctronic mail: [email protected]�Electronic mail: [email protected].

PHYSICS OF FLUIDS 22, 036601 �2010�

1070-6631/2010/22�3�/036601/9/$30.00 © 2010 American Institute of Physics22, 036601-1

heights can deviate from the linear and second-order models.This is confirmed by the numerical simulations of Dysthe’smodified nonlinear Schrödinger �MNLS� equation �Dysthe,24

Socquet-Juglard et al.,14 and Dysthe et al.16�.In a recent study, Shemer and Sergeeva20 described the

evolution of narrowband nonlinear waves in a large wavechannel. Unidirectional waves generated at the wave makerfrom an initially narrowband Gaussian-shaped spectrumwere subsequently measured at numerous locations as theypropagated along the channel. The statistics of the unusuallylarge wave heights, and crest and trough amplitudes ob-served in these experiments are explained reasonably well bytheoretical approximations based on Gram–Charlier �GC�expansions �Tayfun and Fedele7�. Such approximations rep-resent Hermite series expansions of distributions describingnon-Gaussian random functions �Longuet-Higgins6�. Theyare related to the stochastic structure of waves only throughcertain key statistics such as the skewness and kurtosis ofsurface displacements whose closed-form solutions in termsof surface spectra follow from Zakharov’s WT model �see,e.g., Fedele4�.

In this study, we will analyze nonlinear waves generatedin a large wave basin at Marintek, Trondheim, Norway in1999. The surface elevations measured at several gaugesplaced along the basin display relatively strong nonlinearitiesand contain an adequately large population of freak wavesfor reliable statistical analyses and comparisons with the the-oretical models. In particular, we focus attention only on theeffects of quasiresonant interactions on various statistics ob-served during the experiments. In order to interpret the sta-tistics, we draw on the spatial version of the nonlinearSchrödinger �NLS� equation. Shemer et al.25 show that Dys-the’s MNLS is more accurate than the NLS model for de-scribing the evolution of groups of strongly nonlinear wavesgenerated in wave tanks. Nonetheless, the NLS model and itsproperties have not been fully explored in interpreting thespectral and statistical characteristics of random waves ofmoderate steepness, in particular, directly from experimentaltime series. Here, we attempt to do so. Our work reveals thatalthough the NLS model cannot model various nonlinearitiessuch as the O��� front-rear asymmetry of wave groups anddispersion effects as higher order models do, it does none-theless describe the surface statistics fairly well. Further, weexploit the integrals of motion to analyze how the surfacespectra change spatially along the wave basin, establishbounds on the spectrum bandwidth, and derive an analyticalexpression describing the spatial variation of the excess kur-tosis of surface displacements. Finally, we construct the em-pirical distributions describing wave heights, crest andtrough amplitudes observed, and compare these with the GCdistributions proposed by Tayfun and Fedele.7

II. NLS MODEL

We consider the spatial NLS equation valid for narrow-band waves in deep water �Mei26�. To O��2�, the surfacedisplacement � from the mean water level, observed at afixed point x in time t, can be expressed as

� = Re�aei� +kma2

2e2i� +

3km2 a3

8e3i�� , �1�

where �=��x , t�, Re�z� denotes the real part of z, a=a0B,B�� ,�� is the complex envelope, �=kmx−�mt+�, � is thewave phase uniformly distributed in �0,2�� initially at x=0,�=a0km is the wave steepness, a0 is the characteristic ampli-tude, cg is the group velocity corresponding to the spectral“mean” frequency �m, and wave number such that km

=�m2 /g. The complex envelope satisfies the damped version

of the NLS equation26

��B + i��2B + iB2B = − B , �2�

where

� = ��mt −x

cg�, � = �2kmx .

For a rectangular wave flume of width b, the viscous damp-ing coefficient can be expressed as �Kit and Shemer27 andShemer et al.28�

=2km

b� e

�mexp�− i�/4� , �3�

where e represents an effective viscosity coefficient.The complex envelope can be expressed in the form

a0Bei�=�1+ i�̂1 where the linear component �1 of the sur-face displacement and its Hilbert transform �̂1 are given by

�1 = a0Bcos�� + ��, �̂1 = a0Bsin�� + �� , �4�

where

B = ��12 + �̂1

2�1/2/a0, � = tan−1��̂1/�1� − � .

As a result, the surface displacement �1� can be rewritten as

� = �1 + �2 + �3, �5�

with

�2 =km

2��1

2 − �̂12�, �3 =

3km2

8��1

3 − 3�1�̂12� . �6�

Hereafter, we set a0=max��1�, and also let S�1and S� denote

the frequency spectra of �1 and �, respectively. The ordinarymoments of S�1

are mj �j=0,1 ,2 , . . .� so that �2=m0

variance of �1, �m=m1 /m0, and = �m0m2 /m12−1�1/2

spectral bandwidth. Also, because the statistics of � haspreviously been investigated elsewhere �Onorato et al.,17,18

Shemer and Sergeeva,20 and Cherneva et al.21�, we will focuson �1 and its statistics as affected by quasiresonant interac-tions and associated modulational instabilities. To do sobased on an observational time series of �, the second- andthird-order bound harmonics will have to be removed. Here,we follow a procedure similar to that described by Tayfun8

and solve Eq. �1� for �1 via inversion and then require��1

3�=0 �see Appendix A�. To convey a physical picture ofwhat ensues from this procedure, we show in Fig. 1 a partialtime series of � measured at one of the gauges at Marintek,the second- and third-order bound harmonics removed fromit, and the resulting free-wave profile �1.

036601-2 Fedele et al. Phys. Fluids 22, 036601 �2010�

III. NLS properties of �1

Consider the wave action, momentum and Hamiltoniandefined, respectively, by �cf. Ablowitz and Segur29�

A = �B2� ,

M = i�B��B� − B���B�/2, �7�

H = ���B2 − B4/2� ,

where the angle brackets represent averages with respect to�. If viscous damping is excluded from Eq. �3� by setting=0, the preceding expressions will represent the integralsof motion which do not depend on the spatial coordinate �.However, in the most general case, they do vary with � alongthe experimental basin because of wave breaking and/or vis-cous dissipation of high-frequency spectral components notdescribed by the NLS equation. To describe the spectral vari-ability of waves, we define a convenient measure for thebandwidth of S�1

as

�2 = �2 ���B2�

�B2�. �8�

Using the exponential form of B and Eq. �2�, the precedingdefinition can be rewritten as

�2 = �2 �B2 �2 + ���B�2�

�B2�. �9�

Physically, �=��� represents a relative measure for thedispersion or spread of spectral frequencies around �m, andso does �. In the NLS theory, � is not an invariant since itcan vary with �. Defining � has the advantage that it iseasily estimated directly from a time series, whereas the con-ventional bandwidth measure requires the spectral mo-ments of �1. Also, the analysis of the Marintek data willshow later on that � and are practically the same. Per-haps, more significantly, the NLS theory allows us to derivecertain spatially uniform upper and lower bounds for �

which are invariants as those in Eq. �7� if viscous dissipationis neglected. To do so, we use Eq. �7� and write Eq. �8� as

�2 = �2H

A+

�B4�2�B2�� . �10�

It readily follows then that given �,

L � � � U, �11�

where

L = ��H

A, U =

2A�1 +�1 +

4

A2H

A+

A

2�� . �12�

In general, these bounds are not sharp, but spatially invari-ant. Indeed, L holds only when H�0 and follows directlyfrom Eq. �10� by neglecting the non-negative spatially vary-ing term between parentheses �the term H /A is the only in-variant�, whereas U was derived by Thyagaraja30 and it isvalid for any H.

The surface spectrum can vary with � as waves propa-gate along the wave basin, but it is not expected to violatethe bounds in Eq. �12� and as �→� it asymptotically relaxestoward a statistically stationary state. This is what is ob-served in both numerical simulations and experiments�Socquet-Juglard et al.,14 Shemer and Sergeeva20�.

IV. MARINTEK EXPERIMENTS

Marintek data were obtained during a sequence of fiveexperiments run in a wave basin 80 m long and 50 m wide.Surface displacements were measured by ten capacitancewave gauges placed along the centerline of the basin. Thefirst gauge is at 10 m from a double-flap wave maker, and thesubsequent ones are placed at a uniform spacing of 5 malong the section where the water depth is 2 m. The spectrumgenerated at the wave maker is of the JONSWAP type withthe Phillips parameter 0.0178, a peak-enhancement factor of3, peak frequency �p=6.343 rad s−1, and it is bandlimited tofrequencies in �0,3�p�. For waves generated at the wavemaker, =0.298, kp=�p

2 /g=4.105 m−1 and steepness �kp

=0.072 so that principal wave components such as those as-sociated with the peak and mean frequencies are essentiallyin deep water. A more detailed description of these experi-ments is given by Cherneva et al.21

From a time series of � measured at a gauge, we firstestimate the free-wave component �1 via inversion, as de-scribed in Appendix A. Table I summarizes the ensemble

-2 -1 0 1 2-0.1

-0.05

0

0.05

0.1

t [s]

elevations[m]

1

2

3

η

η

η

η

FIG. 1. A segment of the full surface elevation � observed at x=45 m�gauge 8�, the corresponding free wave �1, and the second and third-ordercorrections ��2 and �3� removed from � to obtain �1.

TABLE I. �1: principal spectral parameters.

x �m� � �m� �m �s−1� km �m−1�

10 0.019 7.108 5.150 0.267

15 0.018 7.023 5.028 0.261

20 0.018 6.954 4.930 0.257

25 0.017 6.940 4.910 0.253

30 0.017 6.910 4.867 0.256

35 0.016 6.868 4.808 0.257

40 0.016 6.832 4.759 0.247

45 0.016 6.805 4.721 0.249

50 0.016 6.775 4.679 0.246

55 0.016 6.761 4.660 0.241

036601-3 NLS invariants and wave statistics Phys. Fluids 22, 036601 �2010�

average values of �, , �m and km for �1 obtained from thefive experimental series of � at gauges 1–10, where x de-notes the distance from the maker. They are similar in natureto those of � actually observed �see Cherneva et al.,21 TableI�. In either case, because some waves simulated in theseexperiments are rather small and high-frequency, they areprone to viscous damping as they propagate along the rela-tively long wave basin. This explains at least partially whythe parameters listed in Table I here and also in the work ofCherneva et al.21 tend to decrease steadily with distance fromthe wave maker.

The envelope B observed at each wave gauge is con-structed from Eq. �3�, using the values a0=0.074 m and�m=7.108 rad s−1 observed at gauge 1. This approach en-ables us to obtain from Eq. �7�, the spatial variations of A, M,and H, shown in Fig. 2. Evidently, all three averages tend todecrease somewhat with distance from the wave maker aswaves propagate along the basin, plausibly due to the variousdissipative effects previously mentioned. If we consider thedamped NLS model �2�, then the spatial damping of waveaction A is described by dA /d�=−2rA, where r

=km�2e /�m /b and follows from the real part of Eq. �6�.

Thus, A���=A�0�exp�−2r��, where r and thus e are easilyestimated from the observed values of A by regression analy-sis. This process leads us in the present case to the estimatesr�3.4�10−3 m−1 and e�1.5�10−3 m2 s−1.

As viscous effects dissipate relatively miniscule high-frequency waves as they propagate along the basin, the spec-trum bandwidth also gradually decreases, as shown inTable I. Figure 3 shows how � and the associated lower andupper bounds vary at different gauges along the channel. Thevariation of is also included in the same figure which sug-gests that it is practically the same as �. Clearly, � doesdecrease along the basin as does, but it does not violate thetheoretical bounds in Eq. �11�. This is confirmed in Fig. 4more explicitly by the gradual damping of relatively high-frequency components in the ensemble-averaged spectra S�

and S�1at increasing distances x=10, 35, and 45 m from the

wave maker, respectively. A slight downshifting of the spec-tral peak is also noticeable in this figure, plausibly due to

quasiresonant modulations. One would also expect to see awidening of the spectra for the same reason, as in the experi-ments of Shemer and Sergeeva20 with larger waves charac-terized by more intense modulations than here. In contrast,Figs. 3 and 4 both clearly show that this does not reallyhappen in this particular set of experiments, again possiblybecause of the viscous damping of high-frequency compo-nents. Evidently, spectral slopes tend to behave as �−5 closerto the wave maker, and eventually steepen to approximate a�−6 viscous range. The gradual downshift of the mean fre-quency �m of �1 with distance along the basin, as seen inTable I here and similarly for � �see Cherneva et al.,21 TableI�, is consistent with this interpretation. Other experimentscarried out at Marintek more recently also report quite simi-lar results �see, e.g., Onorato et al.18�.

0 10 20 30 40 50 60-0.05

0

0.05

0.1

0.15

distance from the wavemaker [m]

A,H,M

H

A

M

FIG. 2. Spatial variations of the averaged Hamiltonian H, wave action Aand momentum M. Each point represents an overall average of five experi-mental series with the range of values observed in separate series indicatedby vertical lines �not clearly visible for M�. The horizontal straight linescorrespond to the expected theoretical values along the channel in accord tothe NLS model.

0 10 20 30 40 50 600.1

0.15

0.2

0.25

0.3

0.35

0.4

distance from wavemaker [m]

∆U

∆ω∆L

FIG. 3. Free wave �1: the spatial variations of �, its lower and upperbounds and �points�, all representing the average values of fiveexperiments.

10-7

10-5

10-3

[m2 .s]

(a)-5

-6

ω

ωS η

10-7

10-5

10-3

[m2 .s]

(b)-5

-6

ω

ωS η

3 5 10 3010-7

10-5

10-3

�[rad.s-1]

[m2 .s]

(c)-5

-6

ω

ω

ω

S η

10-7

10-5

10-3[m

2 .s]

(d)-5

-6

ω

ωS1η

10-7

10-5

10-3

[m2 .s]

(e)-5

-6

ω

ωS1η

3 5 10 3010-7

10-5

10-3

[rad.s -1]

[m2 .s]

(f)-5

-6

ω

ω

ω

S1η

FIG. 4. Average spectra S� of the actual series � observed at �a� x=10 m�gauge 1�, �b� x=30 m �gauge 5�, and �c� x=45 m �gauge 8� from the wavemaker. Similarly, the spectra S�1

of the corresponding free-wave series �1

observed at �d� x=10 m, �e� x=30 m, and �f� x=45 m from the wavemaker.

036601-4 Fedele et al. Phys. Fluids 22, 036601 �2010�

V. EXCESS KURTOSIS OF �1

Closer to the wave maker, free-wave component �1 isquasi-Gaussian. As waves propagate away from the wavemaker, nonlinear interactions gradually build up, leading tointermittency. Larger waves occur more frequently and thesurface statistics tend to deviate from the linear Gaussianmodels. The natural presence of second- and third-order non-resonant bound interactions in the fully nonlinear surfacedisplacement � amplifies these further. In particular, previ-ous investigations by Onorato et al.,17,18 Shemer andSergeeva,20 Cherneva et al.,21 Petrova et al.,31 and otherssuggest that the frequency of occurrence of unusually largewaves increases noticeably, accompanied by an increase inthe excess kurtosis �40 of �. These results reflect the com-bined effects of resonant and nonresonant interactions asthey are coupled through Eq. �1�.

Here, we focus on the excess kurtosis �40 of �1 andexplore the effects of quasiresonant interactions only. In thiscontext, Mori and Janssen32 have previously derived an ana-lytical expression describing �40 during the temporal evolu-tion of spatially homogenous waves. In the present case, weconsider the spatial evolution of stationary waves based onthe spatial NLS equation �4�, which is more appropriate forwaves simulated in tanks. The spatial variation of �40 of �1

follows with relative ease from the stochastic formulation ofnonlinear wave groups, elaborated by Fedele8 for theZakharov equation.9 Using the latter formulation coupledwith the general theory of quasideterminism of Boccotti,2,3

Fedele8 derived an expansion of the highest nonlinear crestin terms of Rayleigh-distributed variables from the Zakharovequation directly. A comparison of coefficients in that expan-sion to those in the GC-type crest models �Tayfun andFedele7� leads to an analytical expression for �40 in terms ofS�1

. This expression is identical to the result obtained byJanssen23 and Mori and Janssen.32 For the NLS model of Eq.�2� with =0, it assumes the form

�40 = 24�km��2� � � S1S2S31 − cos�Wkmx�

W�6 dw1dw2dw3,

�13�

where W=w2+w12−w2

2−w32, w=−w1+w2+w3, Sj =S�1

�wj�,and wj = �� j /�m−1� for j=1, 2, and 3, S�1

represents theinitial spectrum at the wave maker where x=0, and the limitsof integration are from −� to �. For a Gaussian-shaped ini-tial spectrum

S�1��� =

�2

�2�2exp−

��/�m − 1�2

22 � , �14�

Eq. �13� is simplified to

�40�x� = 12�km�/�2J��� , �15�

where �=22kmx and

J��� =1

�2��3/2

�� � � exp−z1

2 + z22 + z3

2

2�1 − cos Z

Zdz1dz2dz3.

�16�

In the preceding expressions, Z= �z1−z3��z2−z3�, x is the dis-tance of a wave gauge from the wave maker, and km, �, and are the initial values at the wave maker. Carrying out theintegration in Eq. �16� leads to �see Appendix B�

�40�x� = 4�3 km�

�2��

6− Im�i sin−11 + 2i�

2��� ,

�17�

where Im�z� denotes the imaginary part of z. As �→�,J���→� /6�3, which agrees with the result obtained byMori and Janssen.32 Thus, at steady state, �40 is the same forboth the spatial and temporal cases, as it should be whenwaves become statistically homogenous. Figure 5 comparesEq. �17� to the Marintek data. The observed values of �40

compare fairly well with the theoretical predictions suffi-ciently away from the wave maker where free-wave interac-tions become significant. This agrees with the recent simula-tions of the Zakharov model by Annenkov and Shrira,33

clearly showing that kurtosis in narrowband waves is mostlydue to nonlinear quasiresonant interactions whereas it is al-most entirely due to bound harmonics in broadband windwaves.

VI. THEORETICAL DISTRIBUTIONS

Tayfun and Fedele7 describe various GC distributions forapproximating the statistics of nonlinear random waves.These can be used for describing the statistics of �1 just aswell as the fully nonlinear �. Various results relevant to thestatistics of � can be found in previous studies �Tayfun andFedele,7 Tayfun,8 Shemer and Sergeeva,20 and Chernevaet al.21�. Here, we consider the statistics of �1 for which thesteepness parameter defined by �= ��1

3� /3, one of two keyparameters in the GC-type distributions, is zero since �1 isderived from � so that ��1

3�=0. On this basis, the GC exceed-

0 10 20 30 40 50 600

0.1

0.2

0.3

0.4

0.5

0.6

distance from wave maker [m]

40λ

FIG. 5. Excess of kurtosis �40 of �1 �points� compared with the NLS theoryfrom Eq. �17� �continuous curve� based on km�=0.093 and �=0.267 at thewave maker.

036601-5 NLS invariants and wave statistics Phys. Fluids 22, 036601 �2010�

ance distribution describing the crest and trough amplitudesof �1 easily follows, by setting �=0 in the expressions ofTayfun and Fedele,7 as

QGC� �z� Pr�� � z� = exp−

z2

2��1 +

64z2�z2 − 4�� ,

�18�

where � stands for the wave crest or trough amplitude scaledwith � and

� �40 + 2�22 + �04, �19�

for simplicity, with �40 ��14�−3, �22 ��1

2�̂12�−1, and �04

��̂14�−3. Evidently, �1 is non-Gaussian, but its crest and

trough amplitudes have the same distribution �18�. The sym-metric amplifications imposed on them by quasiresonant in-teractions are reflected in the parameter �. In linear waves,��0 and Eq. �18� reduces to the Rayleigh exceedance dis-tribution given by

QR�z� = exp�− z2/2� . �20�

The exceedance distribution describing the statistics of largecrest-to-trough wave heights scaled with �, say h, in Gauss-ian seas are described fairly accurately by the theoreticalexpressions devised by Boccotti2,3 and Tayfun.34 For narrow-band waves, h�2� so that both expressions tend to the Ray-leigh limit of the form

Qh�z� = Pr�h � z� � exp�− z2/8� . �21�

Although second-order nonlinearities do not appear to affectthe statistics of wave heights significantly, third-order qua-siresonant interactions and associated modulational instabili-ties do so. The appropriate GC model accounting for thelatter effects on �1 follows, again by setting �=0 in theexpressions given by Tayfun and Fedele,7 as

QGC�z� = Qh�z��1 +�

1024z2�z2 − 16�� . �22�

Table II summarizes the ensemble averaged nontrivial valuesof the cumulants �nm of �1 and the resulting � observed atgauges 1–10. All other third and fourth-order cumulants notshown in this table are zero to two decimals. This essentiallyconfirms the relative validity of the procedure we have usedin removing the bound harmonics from � to obtain �1. Fur-

ther, the fourth-order moments and thus � tend to monotoni-cally increase with distance from the wave maker. This ten-dency is a clear indication of the progressive development ofquasiresonant modulations as waves propagate along the ba-sin. Also, notice that for �1, all values of �nm and � in TableI satisfy the equalities �40=3�22=�04 and �app=8�40 /3=�

very nearly, as suggested by Mori and Janssen.32 As a result,the theoretical expressions which require � can be expressedin simpler forms dependent solely on the excess kurtosis �40

of �1. Unfortunately, however, the same equalities do notgenerally hold for � �see Cherneva et al.,21 Table II�.

VII. COMPARISONS

The comparisons here will focus only on the measure-ments at gauges 1, 5, and 8, located at x=10, 30, and 45 mfrom the wave maker, respectively. At these locations, thethird-order quasiresonant interactions appear to be at theirinitial, intermediate and peak stages of development. Thedistributions of �1 observed at these gauges will be com-pared to the predictions from the third-order GC distribu-tions, represented by Eq. �18� for crest and trough ampli-tudes, and Eq. �22� for wave heights. For contrast, we willalso include the Rayleigh limits appropriate to linear waves,namely, Eqs. �20� and �21�, in the same comparisons.

The results shown in Fig. 6 suggest that the wave-heightexceedances observed compare with the predictions from Eq.�22� favorably, for the most part. It is noticed that at gauge 1,where the third-order modulations have not yet fully devel-oped, waves are largely Gaussian. At all gauges, the ob-served distributions of wave crests are practically the sameas those of trough amplitudes. This suggests that �1 has asymmetrical statistical structure with respect to the mean wa-ter level, unlike �. The statistics of � observed at x=45 m�gauge 8� in the work of Cherneva et al.21 and reproduced inFig. 7 here show that in contrast with the results in Fig. 6�c�for �1, � displays a significant crest-trough asymmetry. Evi-dently, the GC distributions describing wave crests andtrough amplitudes also agree favorably with the observedempirical distributions, albeit not altogether impressively inall cases. Note in particular that as third-order quasiresonantinteractions become significant, the observed distributionstend to deviate from the Rayleigh approximation for linearwaves, also as predicted by the theoretical expressions.

TABLE II. �1: nontrivial cumulants �mn, � and �app /�.

x �m� �40 �22 �04 � �app /�

10 0.141 0.047 0.144 0.380 0.986

15 0.246 0.080 0.235 0.643 1.027

20 0.312 0.099 0.283 0.793 1.050

25 0.367 0.121 0.360 0.970 1.011

30 0.505 0.163 0.476 1.307 1.031

35 0.528 0.177 0.533 1.415 0.995

40 0.514 0.170 0.506 1.360 1.008

45 0.578 0.191 0.566 1.526 1.010

50 0.528 0.172 0.504 1.377 1.025

55 0.522 0.172 0.513 1.380 1.012

036601-6 Fedele et al. Phys. Fluids 22, 036601 �2010�

VIII. CONCLUDING REMARKS

The present analysis and results on the properties of thetheoretical NLS model and its invariants should provide ad-ditional insight into the nonlinear physics and statistics ofmechanically generated waves. We have attempted to dem-onstrate this here with a detailed analysis of the free-wavecomponent �1 governed by the NLS equation in variouscomparisons with the empirical data gathered during theMarintek experiments. All this basically required the re-moval of the bound harmonics from the fully nonlinear timeseries � actually observed by way of a simple inversion pro-cedure based on the assumption that the coupling betweenquasiresonant and nonresonant wave components can be ex-pressed in a narrowband form, as in Eq. �1�.

Notably, we observe that the spectra of both � and �1

appear to be sensitive to viscous dissipation over the high-frequency range. The net effect is manifested as a decrease inthe spectral bandwidth and a downshift of the mean fre-

quency �m as waves propagate along the experimental basin.Further, we also observe a slight downshift of the spectralpeak, plausibly due to modulations as reported in other simi-lar experimental and numerical results.

One significant effect of quasiresonance modulations onthe statistics of �1 is a monotonic increase in the excesskurtosis along the basin. That in turn implies that wave crestsand trough amplitudes are amplified symmetrically relativeto the mean water level. As a result, their distributions pro-gressively deviate from the Rayleigh form as waves propa-gate away from the wave maker. Such deviations appear tobe reasonably well predicted by the third-order GC distribu-tions considered here.

Our results suggest that the classical NLS equation isable to describe the statistical properties of unidirectionalnonlinear waves reasonably well. Asymmetries of O��� ob-served in the experiments and predicted by the higher orderDysthe or Zakharov models do not appear to noticeably af-fect the surface statistics considered here.

ACKNOWLEDGMENTS

The second author has been supported by the PortugueseFoundation for Science and Technology with the Grant No.SFRH/BRD/30589/2006. The data analyzed here was ob-tained by the fourth author at Marintek, Trondheim in 1999as a part of the project Interactions between Waves and Cur-rents, partially funded by the European Union under ContractNo. ERBFMGECT980135 for access to Large Scale Facili-ties.

APPENDIX A: DERIVATION OF �1

We rewrite Eq. �1� as

� = �1 +�

2��1

2 − �̂12� +

3�2

8��1

3 − 3�1�̂12� , �A1�

where ��1 and represents a parameter to be determined sothat ��1

3�=0. From an inversion of the preceding expression,we obtain

0 2 4 6 8 10 1210−6

10−4

10−2

100

Q

z

QhQR

QGCQ+GCQ−

GC

FIG. 7. Same as Fig. 6�c� except for � observed at x=45 m �gauge 8� fromthe wave maker, reproduced from the work of Cherneva et al. �Ref. 21�. Inthis case, the theoretical distributions QGC for wave heights, QGC

+ for wavecrests, and QGC

− for trough amplitudes follow from the expressions given inEqs. �8�, �11�, and �12� of the same reference.

10-6

10-4

10-2

100

Q

QhQR

(a)

QGCQGC±

10-6

10-4

10-2

100

Q

QhQR

(b)

QGCQGC±

0 2 4 6 8 10 1210-6

10-4

10-2

100

Q

QhQR

(c)

z

QGCQGC±

FIG. 6. Exceedance distributions observed at �a� x=10 m �gauge 1�, �b� x=30 m �gauge 5�, and �c� x=45 m �gauge 8� from the wave maker, describ-ing wave heights �points�, crests �hollow triangles� and trough amplitudes�solid triangles� compared with the predictions from QGC

� of Eq. �18� forwave crest and trough amplitudes, and QGC of Eq. �22� for wave heights.Gaussian limits �dashed curves� are represented by QR of Eq. �20� for linearcrest and trough amplitudes and Qh of Eq. �21� for wave heights.

036601-7 NLS invariants and wave statistics Phys. Fluids 22, 036601 �2010�

�1 = � −�

2��2 − �̂2� +

�2

8��3 − 3��̂2� + O��3� . �A2�

The requirement that ��13�=0 leads to the expression

A2�2 − A1� + A0 = 0, �A3�

correct to O��2�, where

A0 = ��3�, A1 = 3��4� − ��2�̂2�

2, �A4�

and

A2 =9��5� − 21��3�̂2� + 6���̂4�

8. �A5�

The physically meaningful solution of Eq. �A3� is

� =A1 − �A1

2 − 4A0A2

2A2. �A6�

APPENDIX B: EVALUATION OF THE INTEGRAL J

The integrand of J��� is not singular at �=0. So, fromEq. �16�

dJ

d�=

1

�2��3/2

�� � � exp−z1

2 + z22 + z3

2

2�sin��Z�dz1dz2dz3.

�B1�

Using complex notation, this expression can be rewritten as

dJ

d�=

1

�2��3/2 Im

�� � � exp−z1

2 + z22 + z3

2 − 2i�Z

2�dz1dz2dz3,

�B2�

or more simply,

dJ

d�=

1

�2��3/2 Im� � � exp−uT�u

2�du , �B3�

where

� = � 1 − i� i�

− i� 1 i�

i� i� 1 − 2i��, u = �z1

z2

z3� . �B4�

From the Gaussian identity

1

�2��3/2��−1� � � exp−

uT�u

2�du = 1, �B5�

it immediately follows that

dJ

d�= ��−1 =

1�1 − 2i� + 3�2

. �B6�

Integrating the preceding expression with J�0�=0 will give

J��� = Im�0

� dx�1 − 2ix + 3x2

=1�3��

6− Im�i sin−11 + 2i�

2��� . �B7�

1G. Lindgren, “Local maxima of Gaussian fields,” Ark. Mat. 10, 195�1972�.

2P. Boccotti, “Some new results on statistical properties of wind waves,”Appl. Ocean Res. 5, 134 �1983�.

3P. Boccotti, Wave Mechanics for Ocean Engineering �Elsevier, New York,2000�.

4F. Fedele, “Rogue waves in oceanic turbulence,” Physica D 237, 2127�2008�.

5F. Fedele and M. A. Tayfun, “On nonlinear wave groups and crest statis-tics,” J. Fluid Mech. 620, 221 �2009�.

6M. S. Longuet-Higgins, “The effect of nonlinearities on statistical distri-butions in the theory of sea waves,” J. Fluid Mech. 17, 459 �1963�.

7M. A. Tayfun and F. Fedele, “Wave-height distributions and nonlineareffects,” Ocean Eng. 34, 1631 �2007�.

8M. A. Tayfun, “Distributions of envelope and phase in wind waves,” J.Phys. Oceanogr. 38, 2784 �2008�.

9V. E. Zakharov, “Stability of periodic waves of finite amplitude on thesurface of deep fluid,” J. Appl. Mech. Tech. Phys. 2, 190 �1968� �Englishtranslation�.

10V. E. Zakharov, “Statistical theory of gravity and capillary waves on thesurface of a finite-depth fluid,” Eur. J. Mech. B 18, 327 �1999�.

11M. Onorato, A. R. Osborne, M. Serio, and S. Bertone, “Freak waves inrandom oceanic sea states,” Phys. Rev. Lett. 86, 5831 �2001�.

12C. Guedes Soares, Z. Cherneva, and E. Antao, “Characteristics of abnor-mal waves in North Sea storm sea states,” Appl. Ocean Res. 25, 337�2003�.

13C. Guedes Soares, Z. Cherneva, and E. Antao, “Abnormal waves duringhurricane Camille,” J. Geophys. Res. 109, C08008, doi:10.1029/2003JC002244 �2004�.

14H. Socquet-Juglard, K. Dysthe, K. Trulsen, H. E. Krogstad, and J. D. Liu,“Probability distributions of surface gravity waves during spectralchanges,” J. Fluid Mech. 542, 195 �2005�.

15P. Petrova, Z. Cherneva, and C. Guedes Soares, “On the adequacy ofsecond-order models to predict abnormal waves,” Ocean Eng. 34, 956�2007�.

16K. B. Dysthe, H. E. Krogstad, and P. Müller, “Oceanic rogue waves,”Annu. Rev. Fluid Mech. 40, 287 �2008�.

17M. Onorato, A. R. Osborne, M. Serio, L. Cavaleri, C. Brandini, and C. T.Stansberg, “Extreme waves, modulational instability and second ordertheory: Wave flume experiments on irregular waves,” Eur. J. Mech. B 25,586 �2006�.

18M. Onorato, L. Cavaleri, S. Fouques, O. Gramstad, P. A. E. M. Janssen, J.Monbaliu, A. R. Osborne, C. Pakozdi, M. Serio, C. T. Stansberg, A. Tof-foli, and K. Trulsen, “Statistical properties of mechanically generated sur-face gravity waves: A laboratory experiment in a three dimensional wavebasin,” J. Fluid Mech. 627, 235 �2009�.

19P. Petrova and C. Guedes Soares, “Maximum wave crest and height sta-tistics of irregular and abnormal waves in an offshore basin,” Appl. OceanRes. 30, 144 �2008�.

20L. Shemer and A. Sergeeva, “An experimental study of spatial evolutionof statistical parameters in a unidirectional narrow-banded random wave-field,” J. Geophys. Res. 114, C01015, doi:10.1029/2008JC005077 �2009�.

21Z. I. Cherneva, M. A. Tayfun, and C. Guedes Soares, “Statistics of non-linear waves generated in an offshore wave basin,” J. Geophys. Res. 114,C08005, doi:10.1029/2009JC005332 �2009�.

22T. B. Benjamin and J. E. Feir, “The disintegration of wave trains in deepwater: Part 1, theory,” J. Fluid Mech. 27, 417 �1967�.

23P. A. E. M. Janssen, “Nonlinear four-wave interactions and freak waves,”J. Phys. Oceanogr. 33, 863 �2003�.

24K. B. Dysthe, “Note on a modification to the nonlinear Schrödinger equa-tion for application to deep water waves,” Proc. R. Soc. London Ser. A369, 105 �1979�.

25L. Shemer, E. Kit, and H. Jiao, “An experimental and numerical study ofthe spatial evolution of unidirectional nonlinear water-wave groups,”Phys. Fluids 14, 3380 �2002�.

036601-8 Fedele et al. Phys. Fluids 22, 036601 �2010�

26C. C. Mei, The Applied Dynamics of Ocean Surface Waves �World Scien-tific, Singapore, 1989�, pp. 1–760.

27E. Kit and L. Shemer, “On dissipation coefficients in a rectangular wavetank,” Acta Mech. 77, 171 �1989�.

28L. Shemer, K. Goulitski, and E. Kit, “Evolution of wide-spectrum unidi-rectional wave groups in a tank: an experimental and numerical study,”Eur. J. Mech. B/Fluids 26, 193 �2007�.

29M. J. Ablowitz and H. Segur, “On the evolution of packets of water,” J.Fluid Mech. 92, 691 �1979�.

30A. Thyagaraja, “Recurrent motions in certain continuum dynamical sys-tems,” Phys. Fluids 22, 2093 �1979�.

31P. Petrova, C. Guedes Soares, and Z. Cherneva, “Influence of the thirdorder nonlinearity on the distribution of wave height maxima in an off-shore basin,” Proceedings of the 27th International Conference on Off-shore Mechanics and Arctic Engineering �OMAE2008�, Estoril, Portugal,15–20 June 2008, ASME Paper No. OMAE2008-58049.

32N. Mori and P. A. E. M. Janssen, “On kurtosis and occurrence probabilityof freak waves,” J. Phys. Oceanogr. 36, 1471 �2006�.

33S. Y. Annenkov and V. I. Shrira, “Evolution of kurtosis for wind waves,”Geophys. Res. Lett. 36, L13603, doi:10.1029/2009GL038613 �2009�.

34M. A. Tayfun, “Distribution of large wave heights,” J. Waterway, Port,Coastal and Ocean Eng. 116, 686 �1990�.

036601-9 NLS invariants and wave statistics Phys. Fluids 22, 036601 �2010�