Wind wave model

Last updated
NOAA WAVEWATCH III (R) 120-hour Forecast for the North Atlantic NOAA Wavewatch III Sample Forecast.gif
NOAA WAVEWATCH III (R) 120-hour Forecast for the North Atlantic

In fluid dynamics, wind wave modeling describes the effort to depict the sea state and predict the evolution of the energy of wind waves using numerical techniques. These simulations consider atmospheric wind forcing, nonlinear wave interactions, and frictional dissipation, and they output statistics describing wave heights, periods, and propagation directions for regional seas or global oceans. Such wave hindcasts and wave forecasts are extremely important for commercial interests on the high seas. [1] For example, the shipping industry requires guidance for operational planning and tactical seakeeping purposes. [1]

Contents

For the specific case of predicting wind wave statistics on the ocean, the term ocean surface wave model is used.

Other applications, in particular coastal engineering, have led to the developments of wind wave models specifically designed for coastal applications.

Historical overview

Early forecasts of the sea state were created manually based upon empirical relationships between the present state of the sea, the expected wind conditions, the fetch/duration, and the direction of the wave propagation. [2] Alternatively, the swell part of the state has been forecasted as early as 1920 using remote observations. [3]

During the 1950s and 1960s, much of the theoretical groundwork necessary for numerical descriptions of wave evolution was laid. For forecasting purposes, it was realized that the random nature of the sea state was best described by a spectral decomposition in which the energy of the waves was attributed to as many wave trains as necessary, each with a specific direction and period. This approach allowed to make combined forecasts of wind seas and swells. The first numerical model based on the spectral decomposition of the sea state was operated in 1956 by the French Weather Service, and focused on the North Atlantic. [4] The 1970s saw the first operational, hemispheric wave model: the spectral wave ocean model (SWOM) at the Fleet Numerical Oceanography Center. [5]

First generation wave models did not consider nonlinear wave interactions. Second generation models, available by the early 1980s, parameterized these interactions. They included the “coupled hybrid” and “coupled discrete” formulations. [6] Third generation models explicitly represent all the physics relevant for the development of the sea state in two dimensions. The wave modeling project (WAM), an international effort, led to the refinement of modern wave modeling techniques during the decade 1984-1994. [7] Improvements included two-way coupling between wind and waves, assimilation of satellite wave data, and medium-range operational forecasting.

Wind wave models are used in the context of a forecasting or hindcasting system. Differences in model results arise (with decreasing order of importance) from: differences in wind and sea ice forcing, differences in parameterizations of physical processes, the use of data assimilation and associated methods, and the numerical techniques used to solve the wave energy evolution equation.

In the aftermath of World War II, the study of wave growth garnered significant attention. The global nature of the war, encompassing battles in the Pacific, Atlantic, and Mediterranean seas, necessitated the execution of landing operations on enemy-held coasts. Safe landing was paramount, given that choppy waters posed the danger of capsizing landing craft. Consequently, the precise forecasting of weather and wave conditions became essential, prompting the recruitment of meteorologists and oceanographers by the warring nations. [8] [9]

During this period, both Japan and the United States embarked on wave prediction research. In the U.S., comprehensive studies were carried out at the Scripps Institution of Oceanography affiliated with the University of California. Under the guidance of Harald Svedrup, Walter Munk devised an avant-garde wave calculation methodology for the United States Navy and later refined this approach for the Office of Naval Research.

This pioneering effort led to the creation of the significant wave method, which underwent subsequent refinements and data integrations. The method, in due course, came to be popularly referred to as the SMB method, an acronym derived from its founders Sverdrup, Munk, and Charles L. Bretschneider. [10] [11]

Between 1950 and 1980, various formulae were proposed. Given that two-dimensional field models had not been formulated during that time, studies were initiated in the Netherlands by Rijkswaterstaat and the Technische Adviescommissie voor de Waterkeringen (TAW - Technical Advisory Committee for Flood Defences) to discern the most appropriate formula to compute wave height at the base of a dike. [12] This work concluded that the 1973 Bretschneider formula was the most suitable. However, subsequent studies by Young and Verhagen in 1997 suggested that adjusting certain coefficients enhanced the formula's efficacy in shallow water regions.

General strategy

Input

A wave model requires as initial conditions information describing the state of the sea. An analysis of the sea or ocean can be created through data assimilation, where observations such as buoy or satellite altimeter measurements are combined with a background guess from a previous forecast or climatology to create the best estimate of the ongoing conditions. In practice, many forecasting system rely only on the previous forecast, without any assimilation of observations. [13]

A more critical input is the "forcing" by wind fields: a time-varying map of wind speed and directions. The most common sources of errors in wave model results are the errors in the wind field. Ocean currents can also be important, in particular in western boundary currents such as the Gulf Stream, Kuroshio or Agulhas current, or in coastal areas where tidal currents are strong. Waves are also affected by sea ice and icebergs, and all operational global wave models take at least the sea ice into account.

This figures show an example of the effects of currents on the wave heights. This example is adapted from scientific paper published in the Journal of Physical Oceanography (vol. 42, December 2012). The top panels show the tidal currents at 3 AM and 11 AM on 28 October 2008, off the West coast of France, around the island of Ouessant, which lies 20 km from the mainland. The bottom panel show the heights and directions of waves, computed with the numerical model WAVEWATCH III (R), using a triangular mesh with variable resolution. The strong currents south of Ouessant deflect the waves away from the measuring buoy at low tide. Wave model current induced refraction.png
This figures show an example of the effects of currents on the wave heights. This example is adapted from scientific paper published in the Journal of Physical Oceanography (vol. 42, December 2012). The top panels show the tidal currents at 3 AM and 11 AM on 28 October 2008, off the West coast of France, around the island of Ouessant, which lies 20 km from the mainland. The bottom panel show the heights and directions of waves, computed with the numerical model WAVEWATCH III (R), using a triangular mesh with variable resolution. The strong currents south of Ouessant deflect the waves away from the measuring buoy at low tide.

Representation

The sea state is described as a spectrum; the sea surface can be decomposed into waves of varying frequencies using the principle of superposition. The waves are also separated by their direction of propagation. The model domain size can range from regional to the global ocean. Smaller domains can be nested within a global domain to provide higher resolution in a region of interest. The sea state evolves according to physical equations – based on a spectral representation of the conservation of wave action – which include: wave propagation / advection, refraction (by bathymetry and currents), shoaling, and a source function which allows for wave energy to be augmented or diminished. The source function has at least three terms: wind forcing, nonlinear transfer, and dissipation by whitecapping. [6] Wind data are typically provided from a separate atmospheric model from an operational weather forecasting center.

For intermediate water depths the effect of bottom friction should also be added. [14] At ocean scales, the dissipation of swells - without breaking - is a very important term. [15]

Output

The output of a wind wave model is a description of the wave spectra, with amplitudes associated with each frequency and propagation direction. Results are typically summarized by the significant wave height, which is the average height of the one-third largest waves, and the period and propagation direction of the dominant wave.

Coupled models

Wind waves also act to modify atmospheric properties through frictional drag of near-surface winds and heat fluxes. [16] Two-way coupled models allow the wave activity to feed back upon the atmosphere. The European Centre for Medium-Range Weather Forecasts (ECMWF) coupled atmosphere-wave forecast system described below facilitates this through exchange of the Charnock parameter which controls the sea surface roughness. This allows the atmosphere to respond to changes in the surface roughness as the wind sea builds up or decays.

Examples

WAVEWATCH

The operational wave forecasting systems at NOAA are based on the WAVEWATCH III (R) model. [17] This system has a global domain of approximately 50 km resolution, with nested regional domains for the northern hemisphere oceanic basins at approximately 18 km and approximately 7 km resolution. Physics includes wave field refraction, nonlinear resonant interactions, sub-grid representations of unresolved islands, and dynamically updated ice coverage. Wind data is provided from the GDAS data assimilation system for the GFS weather model. Up to 2008, the model was limited to regions outside the surf zone where the waves are not strongly impacted by shallow depths. [18]

The model can incorporate the effects of currents on waves from its early design by Hendrik Tolman in the 1990s, and is now extended for near shore applications.

WAM

The wave model WAM was the first so-called third generation prognostic wave model where the two-dimensional wave spectrum was allowed to evolve freely (up to a cut-off frequency) with no constraints on the spectral shape. [19] The model underwent a series of software updates from its inception in the late 1980s. [20] The last official release is Cycle 4.5, maintained by the German Helmholtz Zentrum, Geesthacht. [21]

ECMWF has incorporated WAM into its deterministic and ensemble forecasting system., [22] known as the Integrated Forecast System (IFS). The model currently comprises 36 frequency bins and 36 propagation directions at an average spatial resolution of 25 km. The model has been coupled to the atmospheric component of IFS since 1998. [23] [24]

Other models

Wind wave forecasts are issued regionally by Environment Canada. [25]

Regional wave predictions are also produced by universities, such as Texas A&M University’s use of the SWAN model (developed by Delft University of Technology) to forecast waves in the Gulf of Mexico. [26]

Another model, CCHE2D-COAST is a processes-based integrated model which is capable of simulating coastal processes in different coasts with complex shorelines such as irregular wave deformation from offshore to onshore, nearshore currents induced by radiation stresses, wave set-up, wave set-down, sediment transport, and seabed morphological changes. [27]

Other wind wave models include the U.S. Navy Standard Surf Model (NSSM). [28]

The formulae of Bretschneider, Wilson, and Young & Verhagen

For determining wave growth in deep waters subjected to prolonged fetch, the basic formula set is:

Where:

= gravitational acceleration (m/s2)
= significant wave height (s)
= significant wave period (s)
= wind speed (m/s)

The constants in these formulas are deduced from empirical data. Factoring in water depth, wind fetch, and storm duration complicates the equations considerably. However, the application of dimensionless values facilitates the identification of patterns for all these variables. The dimensionless parameters employed are:

Where:

= water depth (m)
= wind fetch (m)
= storm duration (s)

When plotted against the dimensionless wind fetch, both dimensionless wave height and wave period tend to align linearly. However, this trend becomes notably more flattened for more extended dimensionless wind fetches. Various researchers have endeavoured to formulate equations capturing this observed behaviour.

Dimensionless wave height and period against the backdrop of the dimensionless fetch (data courtesy of Wilson, 1965) DataWilson-Golfgroei.jpg
Dimensionless wave height and period against the backdrop of the dimensionless fetch (data courtesy of Wilson, 1965)

Common Formulas for Deep Water

Bretschneider (1952, 1977):

Wilson (1965): [30]

Wave growth chart based on the formulas by Groen & Dorrestein Wave forecasting nomogram.png
Wave growth chart based on the formulas by Groen & Dorrestein
Graph depicting the variation of significant wave height with dimensionless fetch based on the Young & Verhagen wave growth formula, set against a specific water depth and wind speed. Youngverhagen.png
Graph depicting the variation of significant wave height with dimensionless fetch based on the Young & Verhagen wave growth formula, set against a specific water depth and wind speed.

In the Netherlands, a formula devised by Groen & Dorrestein (1976) is also in common use: [31]

for
for
for

During periods when programmable computers weren't commonly utilised, these formulas were cumbersome to use. Consequently, for practical applications, nomograms were developed which did away with dimensionless units, instead presenting wave heights in metres, storm duration in hours, and the wind fetch in km.

Integrating the water depth into the same chart was problematic as it introduced too many input parameters. Therefore, during the primary usage of nomograms, separate nomograms were crafted for distinct depths. The use of computers has resulted in reduced reliance on nomograms.

For deep water, the distinctions between the various formulas are subtle. However, for shallow water, the formula modified by Young & Verhagen [32] proves more suitable. It's defined as:

and

and

and

Research by Bart demonstrated that, under Dutch conditions (for example, in the IJsselmeer), this formula is reliable. [33]

Example: Lake Garda

Lake Garda in Italy is a deep, elongated lake, measuring about 350 m in depth and spanning 45 km in length. With a wind speed of 25 m/s from the SSW, the Bretschneider and Wilson formulas suggest an Hs of 3.5 m and a period of roughly 7 s (assuming the storm persists for at least 4 hours). The Young and Verhagen formula, however, predicts a lower wave height of 2.6 m. This diminished result is attributed to the formula's calibration for shallow waters, whilst Lake Garda is notably deep.

Bretschneider Formula: Lake Garda

Based on Bretschneider's formula:

  • Predicted wave height: 3.54 meters
  • Predicted wave period: 7.02 seconds
Wilson Formula: Lake Garda

Utilizing Wilson's formula, the predictions are:

  • Predicted wave height: 3.56 meters
  • Predicted wave period: 7.01 seconds
Young & Verhagen Formula: Lake Garda

Young & Verhagen's formula, which typically applies to shallow waters, yields:

  • Predicted wave height: 2.63 meters
  • Predicted wave period: 6.89 seconds

Shallow and coastal waters

Global wind wave models such as WAVEWATCH and WAM are not reliable in shallow water areas near the coast. To address this issue, the SWAN (Simulating WAves Nearshore) program was developed in 1993 by Delft University of Technology, in collaboration with Rijkswaterstaat and the Office of Naval Research in the United States. [34] [35] Initially, the main focus of this development was on wave changes due to the effects of breaking, refraction, and the like. The program was subsequently developed to include analysis of wave growth. [36]

SWAN essentially calculates the energy of a wave field (in the form of a wave spectrum) and derives the significant wave height from this spectrum. SWAN lacks a user interface for easily creating input files and presenting the output. The program is open-source, and many institutions and companies have since developed their own user environments for SWAN. The program has become a global standard for such calculations, and can be used in both one-dimensional and two-dimensional modes. [37]

One-dimensional approach

Wave growth in the Western Scheldt Golfprofiel-Westerschelde-en.jpg
Wave growth in the Western Scheldt

The computation time for a calculation with SWAN is in the order of seconds. In one-dimensional mode, results are available from the input of a cross-sectional profile and wind information. In many cases, this can yield a sufficiently reliable value for the local wave spectrum, particularly when the wind path crosses shallow areas.

Example: wave growth calculation in The Netherlands

As an example, a calculation of the wave growth in the Westerschelde has been made. For this example, the one-dimensional version of SWAN and the open-source user interface SwanOne were used. [38] The wave height at the base of the sea dike near Goudorpe on South Beveland, just west of the Westerscheldetunnel, was calculated, with the wind coming from the SW at a speed of 25m/s (force 9 to 10). In the graph, this is from left to right. The dike is quite far from deep water, with a salt marsh in front of it.

The calculation was made for low water, average water level, and high water. At high tide, the salt marsh is under water; at low tide, only the salt marsh is submerged (the tidal difference here is about 5 metres). At high tide, there is a constant increase in wave height, which is faster in deep water than in shallow water. At low tide, some plates are dry, and wave growth has to start all over again. Close to the shore (beyond the Gat van Borssele), there's a tall salt marsh; at low tide, there are no waves there, at average tide, the wave height decreases to almost nothing at the dike, and at high tide, there's still a wave height of 1 m present. The measure of period shown in these graphs is the spectral period (Tm-1,0).

Two-dimensional approach

Calculation of wave growth (and decline) on Lake Garda due to strong wind (25 m/s) from the SSW (210deg). Lake Garda waves-Favaretto.png
Calculation of wave growth (and decline) on Lake Garda due to strong wind (25 m/s) from the SSW (210°).

In situations where significant refraction occurs, or where the coastline is irregular, the one-dimensional method falls short, necessitating the use of a field model. Even in a relatively rectangular lake like Lake Garda, a two-dimensional calculation provides considerably more information, especially in its southern regions. The figure below demonstrates the results of such a calculation.

This case highlights another limitation of the one-dimensional approach: at certain points, the actual wave growth is less than predicted by the one-dimensional model. This discrepancy arises because the model assumes a broad wave field, which isn't the case for narrow lakes. [40]

Validation

Comparison of the wave model forecasts with observations is essential for characterizing model deficiencies and identifying areas for improvement. In-situ observations are obtained from buoys, ships and oil platforms. Altimetry data from satellites, such as GEOSAT and TOPEX, can also be used to infer the characteristics of wind waves.

Hindcasts of wave models during extreme conditions also serves as a useful test bed for the models. [41]

Reanalyses

A retrospective analysis, or reanalysis, combines all available observations with a physical model to describe the state of a system over a time period of decades. Wind waves are a part of both the NCEP Reanalysis [42] and the ERA-40 from the ECMWF. [43] Such resources permit the creation of monthly wave climatologies, and can track the variation of wave activity on interannual and multi-decadal time scales. During the northern hemisphere winter, the most intense wave activity is located in the central North Pacific south of the Aleutians, and in the central North Atlantic south of Iceland. During the southern hemisphere winter, intense wave activity circumscribes the pole at around 50°S, with 5 m significant wave heights typical in the southern Indian Ocean. [43]

Related Research Articles

<span class="mw-page-title-main">Wind wave</span> Surface waves generated by wind on open water

In fluid dynamics, a wind wave, or wind-generated water wave, is a surface wave that occurs on the free surface of bodies of water as a result of the wind blowing over the water's surface. The contact distance in the direction of the wind is known as the fetch. Waves in the oceans can travel thousands of kilometers before reaching land. Wind waves on Earth range in size from small ripples to waves over 30 m (100 ft) high, being limited by wind speed, duration, fetch, and water depth.

<span class="mw-page-title-main">Thomas Stevenson</span> Scottish civil engineer, lighthouse designer and meteorologist (1818–1887)

Thomas StevensonPRSE MInstCE FRSSA FSAScot was a pioneering Scottish civil engineer, lighthouse designer and meteorologist, who designed over thirty lighthouses in and around Scotland, as well as the Stevenson screen used in meteorology. His designs, celebrated as ground breaking, ushered in a new era of lighthouse creation.

<span class="mw-page-title-main">Pacific decadal oscillation</span> Recurring pattern of climate variability

The Pacific decadal oscillation (PDO) is a robust, recurring pattern of ocean-atmosphere climate variability centered over the mid-latitude Pacific basin. The PDO is detected as warm or cool surface waters in the Pacific Ocean, north of 20°N. Over the past century, the amplitude of this climate pattern has varied irregularly at interannual-to-interdecadal time scales. There is evidence of reversals in the prevailing polarity of the oscillation occurring around 1925, 1947, and 1977; the last two reversals corresponded with dramatic shifts in salmon production regimes in the North Pacific Ocean. This climate pattern also affects coastal sea and continental surface air temperatures from Alaska to California.

<span class="mw-page-title-main">Swell (ocean)</span> Series of waves generated by distant weather systems

A swell, also sometimes referred to as ground swell, in the context of an ocean, sea or lake, is a series of mechanical waves that propagate along the interface between water and air under the predominating influence of gravity, and thus are often referred to as surface gravity waves. These surface gravity waves have their origin as wind waves, but are the consequence of dispersion of wind waves from distant weather systems, where wind blows for a duration of time over a fetch of water, and these waves move out from the source area at speeds that are a function of wave period and length. More generally, a swell consists of wind-generated waves that are not greatly affected by the local wind at that time. Swell waves often have a relatively long wavelength, as short wavelength waves carry less energy and dissipate faster, but this varies due to the size, strength, and duration of the weather system responsible for the swell and the size of the water body, and varies from event to event, and from the same event, over time. Occasionally, swells that are longer than 700m occur as a result of the most severe storms.

In physical oceanography, the significant wave height (SWH, HTSGW or Hs) is defined traditionally as the mean wave height (trough to crest) of the highest third of the waves (H1/3). It is usually defined as four times the standard deviation of the surface elevation – or equivalently as four times the square root of the zeroth-order moment (area) of the wave spectrum. The symbol Hm0 is usually used for that latter definition. The significant wave height (Hs) may thus refer to Hm0 or H1/3; the difference in magnitude between the two definitions is only a few percent. SWH is used to characterize sea state, including winds and swell.

Backtesting is a term used in modeling to refer to testing a predictive model on historical data. Backtesting is a type of retrodiction, and a special type of cross-validation applied to previous time period(s).

In physical oceanography and fluid dynamics, the wind stress is the shear stress exerted by the wind on the surface of large bodies of water – such as oceans, seas, estuaries and lakes. When wind is blowing over a water surface, the wind applies a wind force on the water surface. The wind stress is the component of this wind force that is parallel to the surface per unit area. Also, the wind stress can be described as the flux of horizontal momentum applied by the wind on the water surface. The wind stress causes a deformation of the water body whereby wind waves are generated. Also, the wind stress drives ocean currents and is therefore an important driver of the large-scale ocean circulation. The wind stress is affected by the wind speed, the shape of the wind waves and the atmospheric stratification. It is one of the components of the air–sea interaction, with others being the atmospheric pressure on the water surface, as well as the exchange of energy and mass between the water and the atmosphere.

<span class="mw-page-title-main">Wave shoaling</span> Effect by which surface waves entering shallower water change in wave height

In fluid dynamics, wave shoaling is the effect by which surface waves, entering shallower water, change in wave height. It is caused by the fact that the group velocity, which is also the wave-energy transport velocity, changes with water depth. Under stationary conditions, a decrease in transport speed must be compensated by an increase in energy density in order to maintain a constant energy flux. Shoaling waves will also exhibit a reduction in wavelength while the frequency remains constant.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

<span class="mw-page-title-main">Mild-slope equation</span> Physics phenomenon and formula

In fluid dynamics, the mild-slope equation describes the combined effects of diffraction and refraction for water waves propagating over bathymetry and due to lateral boundaries—like breakwaters and coastlines. It is an approximate model, deriving its name from being originally developed for wave propagation over mild slopes of the sea floor. The mild-slope equation is often used in coastal engineering to compute the wave-field changes near harbours and coasts.

Equatorial waves are oceanic and atmospheric waves trapped close to the equator, meaning that they decay rapidly away from the equator, but can propagate in the longitudinal and vertical directions. Wave trapping is the result of the Earth's rotation and its spherical shape which combine to cause the magnitude of the Coriolis force to increase rapidly away from the equator. Equatorial waves are present in both the tropical atmosphere and ocean and play an important role in the evolution of many climate phenomena such as El Niño. Many physical processes may excite equatorial waves including, in the case of the atmosphere, diabatic heat release associated with cloud formation, and in the case of the ocean, anomalous changes in the strength or direction of the trade winds.

<span class="mw-page-title-main">Iribarren number</span> Dimensionless parameter

In fluid dynamics, the Iribarren number or Iribarren parameter – also known as the surf similarity parameter and breaker parameter – is a dimensionless parameter used to model several effects of (breaking) surface gravity waves on beaches and coastal structures. The parameter is named after the Spanish engineer Ramón Iribarren Cavanilles (1900–1967), who introduced it to describe the occurrence of wave breaking on sloping beaches. The parameter used to describe breaking wave types on beaches; or wave run-up on – and reflection by – beaches, breakwaters and dikes.

Wind-wave dissipation or "swell dissipation" is process in which a wave generated via a weather system loses its mechanical energy transferred from the atmosphere via wind. Wind waves, as their name suggests, are generated by wind transferring energy from the atmosphere to the ocean's surface, capillary gravity waves play an essential role in this effect, "wind waves" or "swell" are also known as surface gravity waves.

<span class="mw-page-title-main">Ian Young (academic)</span> Australian academic

Ian Robert Young, AO is an Australian academic. He is the Kernot Professor of Engineering at the University of Melbourne. He previously held the senior administrative roles of Vice-Chancellor of Swinburne University of Technology (2003–2011) and Vice-Chancellor of the Australian National University (2011–2016).

The nonlinearity of surface gravity waves refers to their deviations from a sinusoidal shape. In the fields of physical oceanography and coastal engineering, the two categories of nonlinearity are skewness and asymmetry. Wave skewness and asymmetry occur when waves encounter an opposing current or a shallow area. As waves shoal in the nearshore zone, in addition to their wavelength and height changing, their asymmetry and skewness also change. Wave skewness and asymmetry are often implicated in ocean engineering and coastal engineering for the modelling of random sea states, in particular regarding the distribution of wave height, wavelength and crest length. For practical engineering purposes, it is important to know the probability of these wave characteristics in seas and oceans at a given place and time. This knowledge is crucial for the prediction of extreme waves, which are a danger for ships and offshore structures. Satellite altimeter Envisat RA-2 data shows geographically coherent skewness fields in the ocean and from the data has been concluded that large values of skewness occur primarily in regions of large significant wave height.

In physical oceanography and fluid mechanics, the Miles-Phillips mechanism describes the generation of wind waves from a flat sea surface by two distinct mechanisms. Wind blowing over the surface generates tiny wavelets. These wavelets develop over time and become ocean surface waves by absorbing the energy transferred from the wind. The Miles-Phillips mechanism is a physical interpretation of these wind-generated surface waves.
Both mechanisms are applied to gravity-capillary waves and have in common that waves are generated by a resonance phenomenon. The Miles mechanism is based on the hypothesis that waves arise as an instability of the sea-atmosphere system. The Phillips mechanism assumes that turbulent eddies in the atmospheric boundary layer induce pressure fluctuations at the sea surface. The Phillips mechanism is generally assumed to be important in the first stages of wave growth, whereas the Miles mechanism is important in later stages where the wave growth becomes exponential in time.

<span class="mw-page-title-main">Wave overtopping</span> Transmission of water waves over a coastal structure

Wave overtopping is the time-averaged amount of water that is discharged per linear metre by waves over a structure such as a breakwater, revetment or dike which has a crest height above still water level.

<span class="mw-page-title-main">Wave run-up</span> Height that waves reach on a slope

Wave run-up is the height to which waves run up the slope of a revetment, bank or dike, regardless of whether the waves are breaking or not. Conversely, wave run-down is the height to which waves recede. These heights are always measured vertically. The wave run-up height, denoted by , , or , is a very important parameter in coastal engineering as, together with the design highest still water level, it determines the required crest height of a dike or revetment.

<span class="mw-page-title-main">Wind setup</span> Rise of water due to wind blowing over the water surface

Wind setup, also known as wind effect or storm effect, refers to the rise in water level in seas or lakes caused by winds pushing the water in a specific direction. As the wind moves across the water's surface, it applies a shear stress to the water, prompting the formation of a wind-driven current. When this current encounters a shoreline, the water level along the shore increases, generating a hydrostatic counterforce in equilibrium with the shear force.

Basil Wrigley Wilson was an oceanographic engineer and researcher in the field of coastal engineering who made significant contributions to the study of ocean waves, ship motion, and mooring technology.

References

  1. 1 2 Cox, Andrew T. & Vincent J. Cardone (2002). "20 Years Of Operational Forecasting At Oceanweather" (PDF). 7th International Workshop on Wave Hindcasting and Forecasting October 21–25, 2002, Banff, Alberta, Canada. Retrieved 2008-11-21.
  2. Wittmann, Paul and Mike Clancy, "Thirty Years of Operational Ocean Wave Forecasting at Fleet Numerical Meteorology and Oceanography Center", Symposium on the 50th Anniversary of Operational Numerical Weather Prediction, 14–17 June 2004, University of Maryland
  3. Robert Montagne, The swell forecasting service in Morocco (In French), 1922, Annales Hydrographiques, pp. 157-186. This paper describes the use of the method published by Gain in the same journal (1918) which combines a classification of North Atlantic Storms with the use of observations in Azores and Portugal to forecast the swells in Morocco.
  4. Gelci, R., H. Cazalé, J. Vassal (1957) Sea state forecasting. The spectral method (In French), Bulletin d'information du Comité d'Océanographie et d'Etude des Côtes, Vol. 9 (1957), pp. 416-435.
  5. "Wave Modeling" Archived 2001-11-21 at the Wayback Machine , Oceanweather Inc
  6. 1 2 Komen, Gerbrand, "The Wave Modeling Group, a historical perspective"
  7. G.J. Komen, L. Cavaleri, M. Donelan, K. Hasselmann, S. Hasselmann and P.A.E.M. Janssen, 1994. Dynamics and Modelling of Ocean Waves. Cambridge University Press, 532p.
  8. Goda, Y. (1999). 「波動問題」の歴史的変遷 [On the Historical Development of the Mathematical Theory of Water Waves](PDF) (in Japanese). Tokyo: Japan Society of Civil Engineers . Retrieved 27 August 2023.
  9. Goda, Y. (1999). "On the Historical Development of the Mathematical Theory of Water Waves". 35th Summer Training Course Lecture Collection, Japan Society of Civil Engineers Coastal Engineering Committee. Retrieved 27 August 2023 via TU Delft Repository.
  10. Saville, T. (1954). North Atlantic Coast Wave Statistics: Hindcast by Bretschneider-Revised Sverdrup-Munk Method. U.S. Beach Erosion Board.
  11. Bretschneider, C.L. (2011). "Revisions in Wave Forecasting: Deep and Shallow Water". Coastal Engineering Proceedings. 1 (6): 3. doi: 10.9753/icce.v6.3 . ISSN   2156-1028 . Retrieved 11 August 2023.
  12. Holthuijsen, L.H. (1980). "Methods for wave prediction". P80-01. Rijkswaterstaat - TAW.
  13. Tolman, Hendrik. "User manual and system documentation of WAVEWATCH III TM version 3.14 †" (PDF). National Oceanographic and Atmospheric Administration. Retrieved 22 March 2022.
  14. Ardhuin, F.; O'Reilly, W. C.; Herbers, T. H. C.; Jessen, P. F. (2003). "Swell transformation across the continental shelf. part I: Attenuation and directional broadening". J. Phys. Oceanogr. 33 (9): 1921–1939. Bibcode:2003JPO....33.1921A. doi: 10.1175/1520-0485(2003)033<1921:statcs>2.0.co;2 .
  15. Ardhuin, F.; Chapron, B.; Collard, F. (2009). "Observation of swell dissipation across oceans". Geophys. Res. Lett. 36 (6): L06607. arXiv: 0809.2497 . Bibcode:2009GeoRL..36.6607A. doi:10.1029/2008GL037030. S2CID   6470677.
  16. Bender, L.C. (1996). "Modification of the Physics and Numerics in a Third-Generation Ocean Wave Model". Journal of Atmospheric and Oceanic Technology. 13 (3): 726–750. Bibcode:1996JAtOT..13..726B. doi: 10.1175/1520-0426(1996)013<0726:motpan>2.0.co;2 .
  17. Tolman, H. L., "WAVEWATCH III Model Description"
  18. Tolman, 2002g: User manual and system documentation of WAVEWATCH-III version 2.22. NOAA / NWS / NCEP / MMAB Technical Note 222, 133 pp.
  19. Komen, GJ and Cavaleri, L. and Donelan, M. and Hasselmann, K. and Hasselmann, S. and Janssen, P. et al, 1994: "Dynamics and Modelling of Ocean Waves", Cambridge, 534 pp
  20. Hasselmann, S; Hasselmann, K; Janssen, P A E M; et al. (1988). "The WAM model - A third generation ocean wave prediction model". Journal of Physical Oceanography. 18 (12): 1775–1810. Bibcode:1988JPO....18.1775W. doi: 10.1175/1520-0485(1988)018<1775:twmtgo>2.0.co;2 .
  21. "Entwicklungen KSD WAM Cycle 4.5". Archived from the original on 2013-08-23. Retrieved 2012-03-22.
  22. "The Ocean Wave Model" Archived 2008-06-03 at the Wayback Machine , European Centre for Medium-Range Weather Forecasts
  23. Janssen, P. A. E. M., J. D. Doyle, J. Bidlot, B. Hansen, L. Isaksen and P. Viterbo, 2002: "Impact and feedback of ocean waves on the atmosphere", in Advances in Fluid Mechanics, Atmosphere-Ocean Interactions, Vol. I, WITpress, Ed. W.Perrie., pp 155-197
  24. Janssen, P. A. E. M., 2004: The interaction of ocean waves and wind, Cambridge, 300 pages
  25. "Operational Model Forecasts", Environment Canada
  26. "Surf's Up: Professor Using Models To Predict Huge Waves", ScienceDaily, Feb. 23, 2005
  27. "CCHE2D-Coast | the National Center for Computational Hydroscience and Engineering". Archived from the original on 2016-03-04. Retrieved 2015-06-01.
  28. "Validation Test Report for the Navy Standard Surf Model", US Naval Research Lab
  29. Battjes, J.A. (1982). Wind Waves. TU Delft. Retrieved 10 August 2023.
  30. Wilson, B.W. (1965). "Numerical Prediction of Ocean Waves in the North Atlantic for December 1959". German Hydrographic Journal. 18 (3): 114–130. Bibcode:1965DeHyZ..18..114W. doi:10.1007/BF02333333 . Retrieved 10 August 2023.
  31. 1 2 Groen, P.; Dorrestein, R. (1976). "Sea Waves". Knmi Publicatie 11. Royal Netherlands Meteorological Institute . Retrieved 10 August 2023.
  32. Young, I.R.; Verhagen, L.A. (1996). "The Growth of Fetch Limited Waves in Water of Finite Depth. Part 1. Total Energy and Peak Frequency". Coastal Engineering. 29 (1–2): 47–78. doi:10.1016/S0378-3839(96)00006-3 . Retrieved 10 August 2023.
  33. Bart, L. (2013). The Accuracy of the Young and Verhagen Formula for Waves in Water of Finite Depth. TU Delft. Retrieved 10 August 2023.
  34. "SWAN homepage". TU Delft, fluid mechanics section.
  35. Holthuijsen, L.H.; Booij, N.; Ris, R.C. (1993). "A spectral wave model for the coastal zone". Proceedings International Symposium on Ocean Wave Measurement and Analysis. 2: 630–641. Retrieved 11 August 2023.
  36. "SWAN". TU Delft. Retrieved 2023-08-27.
  37. Guisado-Pintado, E. (2020), "Shallow water wave modelling in the nearshore (SWAN)", Sandy Beach Morphodynamics, Elsevier, pp. 391–419, doi:10.1016/b978-0-08-102927-5.00017-5, ISBN   978-0-08-102927-5, S2CID   219883066 , retrieved 27 August 2023
  38. Verhagen, H.J.; van Vledder, G.P.; Eslami Arab, S. (2008). "A practical method for design of coastal structures in shallow water". International Conference on Coastal Engineering. 31. Retrieved 11 August 2023.
  39. Favaretto, C.; Martinelli, L.; Philipine Vigneron, E.; Ruol, P. (2022). "Wave Hindcast in Enclosed Basins: Comparison among SWAN, STWAVE and CMS-Wave Models". Water. 14 (7): 1087. doi: 10.3390/w14071087 . hdl: 11577/3439620 .
  40. Breugem, A. (2003). A reanalysis of the wave observations in Lake George. TU Delft. Retrieved 11 August 2023.
  41. Cardone, V.; Jensen, R.; Resio, D.; Swail, V.; Cox, A. (1996). "Evaluation of Contemporary Ocean Wave Models in Rare Extreme Events: The "Halloween Storm" of October 1991 and the "Storm of the Century" of March 1993". J. Atmos. Oceanic Technol. 13 (1): 198–230. Bibcode:1996JAtOT..13..198C. doi: 10.1175/1520-0426(1996)013<0198:eocowm>2.0.co;2 .
  42. Cox, A., V. Cardone, and V. Swail, "Evaluation Of NCEP-NCAR Reanalysis Project Marine Surface Wind Products For A Long Term North Atlantic Wave Hindcast"
  43. 1 2 Caires, S., A. Sterl, G. Burgers, and G. Komen, ERA-40, "Forty-year European Re-Analysis of the Global Atmosphere; Ocean wave product validation and analysis" Archived 2007-02-07 at the Wayback Machine