Skip to main content

Advertisement

Log in

Statistics of long-crested extreme waves in single and mixed sea states

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

Most of the studies on extreme waves are focused on the systems with single-peak wave spectra. However, according to the statistics of occurrence, the bimodal spectral system is also frequent in real sea conditions. In order to summarize the statistics of extreme waves, irregular wave trains under single-peak and bimodal spectra for long durations are simulated in this paper, based on a two-dimensional High Order Spectral (HOS) numerical wave tank. A large number of configurations have been tested under unimodal and bimodal spectra. The investigation on the wave trains under single-peak spectrum indicates that although in conditions often referred as deep water (kph > π), the relative water depth has a significant influence on the probabilities of occurrence of extreme waves. A detailed analysis of the combined effect of Benjamin-Feir Index (BFI) and relative water depth is provided. However, the situation is more complex in real sea conditions, which may exhibit multimodal spectra. We focus in this study on long-crested bimodal spectra characterized by the same significant wave height Hs and mean zero-crossing period Tz of the sea states as the single-peak spectrum. The wave conditions under bimodal spectrum present milder extreme wave statistics than those under single-peak spectrum. In addition, mixed ocean systems with equivalent energy distribution (i.e., Sea-Swell Energy Ratio (SSER) is close to 1.0) and larger separation between partitions (i.e., Intermodal Distance (ID) > 0.10) are the less prominent to extreme waves appearance. The comparison of the mixed sea states and the corresponding single independent systems demonstrates that the complexity of the underlying physics of a given sea state (for instance the presence of modulational instability or other nonlinear process) cannot be deduced by an analysis limited to the statistical content of the combined sea state. The wave energy being distributed among frequencies plays a major role. Additionally, Gram-Charlier distribution can accurately predict the probability of large waves (1.5 < H/Hs < 2.0) compared to the MER distribution, but it underestimates the statistics of the wave height distribution when H/Hs is larger than 2.0 for both single-peak and bimodal states.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

References

  • Boccotti P (2000) Wave mechanics for ocean engineering. Elsevier Science, Oxford

    Google Scholar 

  • Bonnefoy F, Ducrozet G, Le Touzé D, Ferrant P (2010) Time domain simulation of nonlinear water waves using spectral methods. Adv Numer Simul Nonlinear Water Waves:129–164

  • Dommermuth DG, Yue DKP (1987) A high-order spectral method for the study of nonlinear gravity waves. J Fluid Mech 184:267–288

    Article  Google Scholar 

  • Ducrozet G, Gouin M (2017) Influence of varying bathymetry in rogue wave occurrence within unidirectional and directional sea-states. J Ocean Eng Mar Energy 3(4):309–324

    Article  Google Scholar 

  • Ducrozet G, Bonnefoy F, Le Touzé D, Ferrant P (2012) A modified high-order spectral method for wavemaker modeling in a numerical wave tank. Eur J Mech -B/Fluids 34:19–34

    Article  Google Scholar 

  • Fedele F (2015) On the kurtosis of deep-water gravity waves. J Fluid Mech 782(7):25–36

    Article  Google Scholar 

  • Fedele F, Brennan J, Ponce de León S, Dudley J, Dias F (2016) Real world ocean rogue waves explained without the modulational instability. Sci Rep 6:27715

    Article  Google Scholar 

  • Fernandez L, Onorato M, Monbaliu J, Toffoli A (2014) Modulational instability and wave amplification in finite water depth. Nat Hazards Earth Syst Sci 14(3):705–711

    Article  Google Scholar 

  • Fernandez L, Onorato M, Monbaliu J, Toffoli A (2016) Occurrence of extreme waves in finite water depth. Extreme Ocean Waves. Springer, Cham, pp 45–62

    Google Scholar 

  • Forristall GZ (1984) The distribution of measured and simulated wave heights as a function of spectral shape. J Geophys Res 89(C6):547–552

    Google Scholar 

  • Forristall GZ (2000) Wave crest distributions: observations and second-order theory. J Phys Oceanogr 30(8):1931–1943

    Article  Google Scholar 

  • Gramstad O, Trulsen K (2010) Can swell increase the number of freak waves in a wind-sea? J Fluid Mech 650:57–79

    Article  Google Scholar 

  • Guedes Soares C (1984) Representation of double-peaked sea wave spectra. Ocean Eng 11(2):185–207

    Article  Google Scholar 

  • Guedes Soares C (1991) On the occurrence of double peaked wave spectra. Ocean Eng 18(1–2):167–171

    Article  Google Scholar 

  • Janssen PAEM (2003) Nonlinear four-wave interactions and freak waves. J Phys Oceanogr 33(4):863–884

    Article  Google Scholar 

  • Janssen PAEM, Bidlot J-R (2009) On the extension of the freak wave warning system and its verification. Technical Memorandum, 588. ECMWF, Reading, pp 42

  • Janssen PAEM, Onorato M (2007) The intermediate water depth limit of the Zakharov equation and consequences for wave prediction. J Phys Oceanogr 37(10):2389–2400

    Article  Google Scholar 

  • Klinting P, Sand SE (1987) Analysis of prototype freak waves. Coastal Hydrodynamics, ASCE, 618–632

  • Li JX, Liu SX (2015) Focused wave properties based on a high order spectral method with a non-periodic boundary. China Ocean Eng 29(1):1–16

    Article  Google Scholar 

  • Li JX, Li PF, Liu SX (2013) Observations of freak waves in random wave field in 2D experimental wave flume. China Ocean Eng 27(5):659–670

    Article  Google Scholar 

  • Longuet-Higgins MS (1952) On the statistical distribution of the heights of sea waves. J Mar Res 11:245–266

    Google Scholar 

  • Mori N, Janssen PAEM (2006) On kurtosis and occurrence probability of freak waves. J Phys Oceanogr 36(7):1471–1483

    Article  Google Scholar 

  • Mori N, Yasuda T (2002) A weakly non-Gaussian model of wave height distribution for random wave train. Ocean Eng 29(10):1219–1231

    Article  Google Scholar 

  • Mori N, Onorato M, Janssen PAEM, Osborne AR, Serio M (2007) On the extreme statistics of long-crested deep water waves: theory and experiments. J Geophys Res 112(C9):C09011

    Google Scholar 

  • Naess A (1985) On the statistical distribution of crest to trough wave heights. Ocean Eng 12:221–234

    Article  Google Scholar 

  • Ochi MK, Hubble EN (1976) Six-parameter wave spectra. Proceeding of 15th Coastal Engineering Conference, 301-328

  • Onorato M, Osborne AR, Serio M, Cavaleri L, Brandini C, Stansberg CT (2004) Observation of strongly non-Gaussian statistics for random sea surface gravity waves in wave flume experiments. Phys Rev E 70(6):067302

    Article  Google Scholar 

  • Onorato M, Osborne AR, Serio M, Cavaleri L (2005) Modulation instability and non Gaussian statistics in experimental random water-wave trains. Phys Fluids 17(7):1–4

    Article  Google Scholar 

  • Onorato M, Osborne AR, Serio M (2006) Modulational instability in crossing sea states: a possible mechanism for the formation of freak waves. Phys Rev Lett 96(1):014503

    Article  Google Scholar 

  • Onorato M, Cavaleri L, Fouques S, Gramstad O, Janssen PAEM, Monbaliu J, Osborne AR, Pakozdi C, Serio M, Stansberg C, Toffoli A, Trulsen K (2009) Statistical properties of mechanically generated surface gravity waves: a laboratory experiment in a three-dimensional wave basin. J Fluid Mech 627:235–257

    Article  Google Scholar 

  • Petrova PG, Guedes Soares C (2009) Probability distributions of wave heights in bimodal seas in an offshore basin. Appl Ocean Res 31(2):90–100

    Article  Google Scholar 

  • Petrova PG, Guedes Soares C (2011) Wave height distributions in bimodal sea states from offshore basins. Ocean Eng 38(4):658–672

    Article  Google Scholar 

  • Regev A, Agnon Y, Stiassnie M, Gramstad O (2008) Sea-swell interaction as a mechanism for the generation of freak waves. Phys Fluids 20(11):112102

    Article  Google Scholar 

  • Rice SO (1944) Mathematical analysis of random noise. Bell Syst Tech J 23(3):282–332

    Article  Google Scholar 

  • Rodriguez G, Guedes Soares C, Pacheco M, Perez-Martell E (2002) Wave height distribution in mixed sea states. J Offshore Mech Arct Eng 124:34–40

  • Shemer L, Sergeeva A (2009) An experimental study of spatial evolution of statistical parameters in a unidirectional narrow-banded random wavefield. J Geophys Res 114(C1)

  • Slunyaev AV (2005) A high-order nonlinear envelope equation for gravity waves in finite-depth water. J Exp Theor Phys 101(5):926–941

    Article  Google Scholar 

  • Stansell P (2004) Distributions of freak wave heights measured in the North Sea. Appl Ocean Res 26(1–2):35–48

    Article  Google Scholar 

  • Støle-Hentschel S, Trulsen K, Nieto Borge JC, Olluri S (2020) Extreme wave statistics in combined and partitioned windsea and swell. Water Waves

  • Tayfun MA (1981) Distribution of crest-to-trough wave heights. J Waterw Port Coast Ocean Eng 107(3):149

    Google Scholar 

  • Tayfun MA (1990) Distribution of large wave heights. J Waterw Port Coast Ocean Eng 116(6):686–707

    Article  Google Scholar 

  • Tayfun MA (2006) Statistics of nonlinear wave crests and groups. Ocean Eng 33(11–12):1589–1622

    Article  Google Scholar 

  • Tayfun MA, Fedele F (2007) Wave-height distributions and nonlinear effects. Ocean Eng 34(11–12):1631–1649

    Article  Google Scholar 

  • Toffoli A, Benoit M, Onorato M, Bitner-Gregersen EM (2009) The effect of third-order nonlinearity on statistical properties of random directional waves in finite depth. Nonlinear Process Geophys 16:131–139

    Article  Google Scholar 

  • Toffoli A, Bitner-Gregersen EM, Osborne AR, Serio M, Monbaliu J, Onorato M (2011) Extreme waves in random crossing seas: laboratory experiments and numerical simulations. Geophys Res Lett 38:L06605

    Article  Google Scholar 

  • West BJ, Brueckner KA, Janda RS, Milder DM, Milton RL (1987) A new numerical method for surface hydrodynamics. J Geophys Res 92(C11):11803–11824

    Article  Google Scholar 

Download references

Funding

This research was supported by the National Natural Science Foundation of China (51739010, 51879037), the National Key Research and Development Plan (2016YFC1401405, 2016YFE0200100), the Fundamental Research Funds for the Central Universities of China, and the National Scholarship for Building High Level Universities, China Scholarship Council (201706060082).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jinxuan Li.

Additional information

Responsible Editor: Sandro Carniel

Appendices

Appendix A: Generation of irregular waves

The wave-maker motion is determined by the target free surface elevation and the linear transfer function, which depends on its geometry. According to the linear superposition theory, the two-dimensional irregular free surface elevation can be represented as a superposition of regular wave components with different frequencies:

$$ \eta \left(x,t\right)=\sum \limits_{i=1}^{N_f}{a}_i\cos \left({k}_ix-{\omega}_it+{\varphi}_i\right) $$
(17)

where the subscript i stands for the ith wave component; Nf is the total number of wave components; ai, ki, and ωi are the wave amplitude, wave number, and wave frequency of each component; and φi is the initial random phase, which is uniformly distributed from 0 to 2π. The value of ki can be obtained through the dispersion equation:

$$ {\omega_i}^2={gk}_i\mathrm{th}{k}_ih $$
(18)

with g the gravity acceleration. The wave amplitude of each component ai is defined from the wave spectrum:

$$ {a}_i=\sqrt{2S\left({\omega}_i\right)\Delta {\omega}_i} $$
(19)

where Δωi is the division widths of the frequencies, Δωi = 2π(fHfL)/(Nf−1). (fL, fH) is the considered frequency range. To avoid the periodic time repetition, the actual frequency of the ith wave component ωi is selected randomly in the frequency spin as

$$ {\omega}_i=\left(i-1+\mathit{\operatorname{rand}}\right)\Delta {\omega}_i $$
(20)

where rand is a random number distributed from 0 to 1.

Appendix B: Convergence analysis

The data from the irregular wave experiment under single-peak spectrum by Li et al. (2013) are used to conduct the convergence analysis of the established numerical model. The experiment was carried out in the State Key Laboratory of Coastal and Offshore Engineering in Dalian University of Technology, China. The wave flume is 69.0 m long, 2.0 m wide, and 1.8 m deep. The water depth for the experiment is set at 1.2 m. Eleven resistive wave gauges are arranged along the length of the wave flume to record the free surface elevations. The experimental setup is displayed in Fig. 15.

Fig. 15
figure 15

The setup of the experiment

Irregular waves are characterized by the JONSWAP spectrum with random initial phases. Details of the experimental parameters are listed in Table 5. Two cases with different BFI values of 0.51 and 0.87, which represent different stabilities of the wave group, have been chosen by varying the peak period Tp and the peak enhancement factor γ in the JONSWAP spectrum. BFI is defined as deep-water BFI\( =\frac{\sqrt{2}\varepsilon }{2\Delta f/{f}_p} \), with ε = kpHs/2 the wave steepness and Δf/fp the relative spectral bandwidth. For the two cases, the significant wave height Hs is the same. However, case II has a deeper relative water depth kph, larger wave steepness ε, and larger BFI value, which represents a stronger nonlinear wave condition than case I. Thus, the convergence analysis is performed with the irregular wave trains for case II.

Table 5 Parameters of the physical experiment

The numerical wave tank replicas the experimental setup. Its length is enlarged to 80 m to ensure correct wave absorption. The error is measured thanks to the free surface elevation recorded at the middle of the wave tank on a time window fixed to 100Tp and calculated as:

$$ \mathrm{error}=\frac{\int_t\left|{\eta}_{\mathrm{numerical}}-{\eta}_{\exp \mathrm{eriment}}\right|}{\int_t\left|{\eta}_{\exp \mathrm{eriment}}\right|} $$
(21)

The convergence analysis is carried out with the parameters Nz = Nx/4, M = 5, and the result of the convergence analysis with respect to the number of points per peak wavelength NLp is represented in Fig. 16. The figure displays two lines representing the convergence rates of 2 and 3 as a reference. The numerical convergence rate is slightly larger than 2nd order, which is the theoretical expected value (Ducrozet et al. 2012). Considering the numerical simulation of the unidirectional irregular wave trains, the overall error is about 5% with 30 points per corresponding peak wavelength. This value of NLp = 30 is chosen as converged parameter for the rest of the study. It ensures the accuracy of the numerical simulation as well as a fast solution. Similar convergence analysis has been conducted for the discretization in time (time-step Δt) as well as the order of nonlinearity of the method M. The final numerical configuration is NLp = 30, Tpt = 100, and M = 5.

Fig. 16
figure 16

Convergence of the numerical simulation with respect to the number of points per corresponding peak wavelength

The comparison of the input target spectrum and numerically generated one at x = 5 m is exhibited in Fig. 17. The comparison ensures the correct wave generation procedure. In addition, the results obtained with a linear solver (M = 1) is added as a reference. It can be observed that the generated spectrum with M = 1 is exactly the target spectrum, which demonstrates the effectiveness of the established numerical model in dealing with random wave trains. Consequently, the discrepancies between the input target spectrum and the generated one in a nonlinear context (M = 5) are only due to the nonlinear effects in the process of the wave propagation, the generation of waves being strictly equivalent.

Fig. 17
figure 17

Comparison of amplitude spectra input target and numerically generated at x = 5 m

Appendix C: Numerical validation

Complementary to the previous convergence analysis, which demonstrated the accuracy of the HOS numerical wave tank compared to the experiments on the free surface elevation at a single location, this paragraph assesses its relevance for the systematic study presented in Sections 3 and 4. The experimental data listed in Table 5 are further used to validate the accuracy of the established numerical model for long-time simulation and statistical characteristics.

The total simulation duration is 2000 s, i.e., more than 1000 waves are involved in the simulated wave trains. 2D HOS numerical model is adopted to reproduce the two wave trains. Figure 18 compares the free surface elevations between the experimental data and numerical results at some pre-setting locations along the wave tank. The horizontal axis represents the time corrected from the group velocity at the peak of the spectrum cg. This allows to follow the wave groups in their evolution and possibly identify the large wave events in each case. It can be noted that all the free surface elevations of the numerical simulation remain consistent with the experimental data along the wave tank, irrespective of the value of BFI.

Fig. 18
figure 18

The comparisons of the free surface elevations between experimental data and numerical results along the wave tank (the black solid lines represent the experimental data, and red solid lines represent the numerical results). a Case I. b Case II

The evolution of the kurtosis along the wave flume for both cases with different BFI values is presented in Fig. 19, where experiments and numerical results are compared. The horizontal axis represents the distance away from the wave-maker normalized by the corresponding peak wavelength. Complementary, the probability distribution of the wave height at different positions of the numerical waves is compared to the experimental data in Fig. 20, in which Naess distribution are plotted as a reference. It can be found that the kurtosis and the wave height distribution of the numerical results have good agreement with the experimental data, even for case II that exhibits the largest waves. These comparisons in Figs. 18, 19, and 20 demonstrate that the established HOS numerical model is able to accurately simulate the irregular wave trains for a long time, even in the presence of significant modulation instability. Furthermore, in Fig. 20, the wave height distributions show different behaviors for the two cases at different locations. This is an awaited behavior from literature: lower kph and lower BFI induce weaker modulation instability (Janssen 2003; Janssen and Bidlot 2009; Fedele 2015) and consequently smaller occurrence of extreme waves (characterized by a smaller kurtosis). We are consequently confident in the accuracy of the numerical model to study this phenomenon.

Fig. 19
figure 19

The comparisons of kurtosis between experimental data and numerical results along the wave tank (Lp is the wavelength corresponding to the peak frequency). a Case I. b Case II

Fig. 20
figure 20

The comparisons of the probability distribution of the wave height between experimental data and numerical results at different locations (Lp is the wavelength corresponding to the peak frequency). a Case I. b Case II

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, L., Li, J., Liu, S. et al. Statistics of long-crested extreme waves in single and mixed sea states. Ocean Dynamics 71, 21–42 (2021). https://doi.org/10.1007/s10236-020-01418-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-020-01418-9

Keywords

Navigation