Tides in marginal seas

Last updated

Tides in marginal seas are tides affected by their location in semi-enclosed areas along the margins of continents and differ from tides in the open oceans. Tides are water level variations caused by the gravitational interaction between the Moon, the Sun and the Earth. The resulting tidal force is a secondary effect of gravity: it is the difference between the actual gravitational force and the centrifugal force. While the centrifugal force is constant across the Earth, the gravitational force is dependent on the distance between the two bodies and is therefore not constant across the Earth. The tidal force is thus the difference between these two forces on each location on the Earth. [1]

Contents

In an idealized situation, assuming a planet with no landmasses (an aqua planet), the tidal force would result in two tidal bulges on opposite sides of the earth. This is called the equilibrium tide. However, due to global and local ocean responses different tidal patterns are generated. The complicated ocean responses are the result of the continental barriers, resonance due to the shape of the ocean basin, the tidal waves impossibility to keep up with the Moons tracking, the Coriolis acceleration and the elastic response of the solid earth. [2]

In addition, when the tide arrives in the shallow seas it interacts with the sea floor which leads to the deformation of the tidal wave. As a results, tides in shallow waters tend to be larger, of shorter wavelength, and possibly nonlinear relative to tides in the deep ocean. [3]

Tides on the continental shelf

The transition from the deep ocean to the continental shelf, known as the continental slope, is characterized by a sudden decrease in water depth. In order to apply to the conservation of energy, the tidal wave has to deform as a result of the decrease in water depth. The total energy of a linear progressive wave per wavelength is the sum of the potential energy (PE) and the kinetic energy (KE). The potential and kinetic energy integrated over a complete wavelength are the same, under the assumption that the water level variations are small compared to the water depth (). [4]

Figure 1: Tidal wave transmission on the continental shelf. At the transition from the open ocean to the continental shelf, the water depth decreases abruptly. As a result the tidal wave speed decreases as its phase and group speed are dependent on depth. In order to conserve the energy flux, the amplitude of the tidal wave has to increase on the continental shelf. Tidal wave transmission at the continental shelf.svg
Figure 1: Tidal wave transmission on the continental shelf. At the transition from the open ocean to the continental shelf, the water depth decreases abruptly. As a result the tidal wave speed decreases as its phase and group speed are dependent on depth. In order to conserve the energy flux, the amplitude of the tidal wave has to increase on the continental shelf.

where is the density, the gravitation acceleration and the vertical tidal elevation. The total wave energy becomes:

If we now solve for a harmonic wave , where is the wave number and the amplitude, the total energy per unit area of surface becomes: [5]

A tidal wave has a wavelength that is much larger than the water depth. And thus according to the dispersion of gravity waves, they travel with the phase and group velocity of a shallow water wave: . The wave energy is transmitted by the group velocity of a wave [6] and thus the energy flux () is given by:

The energy flux needs to be conserved and with and constant, this leads to:

where and thus .

When the tidal wave propagates onto the continental shelf, the water depth decreases. In order to conserve the energy flux, the amplitude of the wave needs to increase (see figure 1).

Transmission coefficient

The above explanation is a simplification as not all tidal wave energy is transmitted, but it is partly reflected at the continental slope. The transmission coefficient of the tidal wave is given by:

[4]

This equation indicates that when the transmitted tidal wave has the same amplitude as the original wave. Furthermore, the transmitted wave will be larger than the original wave when as is the case for the transition to the continental shelf.

The reflected wave amplitude () is determined by the reflection coefficient of the tidal wave:

[4]

This equation indicates that when there is no reflected wave and if the reflected tidal wave will be smaller than the original tidal wave.

Internal tide and mixing

At the continental shelf the reflection and transmission of the tidal wave can lead to the generation of internal tides on the pycnocline. The surface (i.e. barotropic) tide generates these internal tides where stratified waters are forced upwards over a sloping bottom topography. [7] The internal tide extracts energy from the surface tide and propagates both in shoreward and seaward direction. [8] The shoreward propagating internal waves shoals when reaching shallower water where the wave energy is dissipated by wave breaking. The shoaling of the internal tide drives mixing across the pycnocline, high levels carbon sequestration and sediment resuspension. [9] [10] Furthermore, through nutrient mixing the shoaling of the internal tide has a fundamental control on the functioning of ecosystems on the continental margin. [11]

Tidal propagation along coasts

After entering the continental shelf, a tidal wave quickly faces a boundary in the form of a landmass. When the tidal wave reaches a continental margin, it continues as a boundary trapped Kelvin wave. Along the coast, a boundary trapped Kelvin is also known as a coastal Kelvin wave or Edge wave. A Kelvin wave is a special type of gravity wave that can exist when there is (1) gravity and stable stratification, (2) sufficient Coriolis force and (3) the presence of a vertical boundary. [12] Kelvin waves are important in the ocean and shelf seas, they form a balance between inertia, the Coriolis force and the pressure gradient force. The simplest equations that describe the dynamics of Kelvin waves are the linearized shallow water equations for homogeneous, in-viscid flows. These equations can be linearized for a small Rossby number, no frictional forces and under the assumption that the wave height is small compared to the water depth (). The linearized depth-averaged shallow water equations become:

u momentum equation:

v momentum equation:

the continuity equation:

where is the zonal velocity ( direction), the meridional velocity ( direction), is time and is the Coriolis frequency.

Kelvin waves are named after Lord Kelvin, who first described them after finding solutions to the linearized shallow water equations with the boundary condition . [13] When this assumption is made the linearized depth-averaged shallow water equations that can describe a Kelvin wave become:

u momentum equation:

v momentum equation:

the continuity equation:

Now it is possible to get an expression for , by taking the time derivative of the continuity equation and substituting the momentum equation:

The same can be done for , by taking the time derivative of the v momentum equation and substituting the continuity equation

Both of these equations take the form of the classical wave equation, where . Which is the same velocity as the tidal wave and thus of a shallow water wave. These preceding equations govern the dynamics of a one-dimensional non-dispersive wave, for which the following general solution exist:

where length is the Rossby radius of deformation and is an arbitrary function describing the wave motion. In the most simple form is a cosine or sine function which describes a wave motion in the positive and negative direction. The Rossby radius of deformation is a typical length scale in the ocean and atmosphere that indicates when rotational effects become important. The Rossby radius of deformation is a measure for the trapping distance of a coastal Kelvin wave. [14] The exponential term results in an amplitude that decays away from the coast.

The expression of tides as a bounded Kelvin wave is well observable in enclosed shelf seas around the world (e.g. the English channel, the North Sea or the Yellow sea). Animation 1 shows the behaviour of a simplified case of a Kelvin wave in an enclosed shelf sea for the case with (lower panel) and without friction (upper panel). The shape of an enclosed shelf sea is represented as a simple rectangular domain in the Northern Hemisphere which is open on the left hand side and closed on the right hand side. The tidal wave, a Kelvin wave, enters the domain in the lower left corner and travels to the right with the coast on its right. The sea surface height (SSH, left panels of animation 1), the tidal elevation, is maximum at the coast and decreases towards the centre of the domain. The tidal currents (right panels of animation 1) are in the direction of wave propagation under the crest and in the opposite direction under the through. They are both maximum under the crest and the trough of the waves and decrease towards the centre. This was expected as the equations for and are in phase as they both depend on the same arbitrary function describing the wave motion and exponential decay term. Therefore this set of equations describes a wave that travels along the coast with a maximum amplitude at the coast which declines towards the ocean. These solutions also indicate that a Kelvin wave always travels with the coast on their right hand side in the Northern Hemisphere and with the coast at their left hand side in the Southern Hemisphere. In the limit of no rotation where , the exponential term increase without a bound and the wave will become a simple gravity wave orientated perpendicular to the coast. [14] In the next section, it will be shown how these Kelvin waves behaves when traveling along a coast, in an enclosed shelf seas or in estuaries and basins.

Tides in enclosed shelf seas

Animation 1: A traveling Kelvin wave is shown for the Northern hemisphere with (lower panel) and without (upper panel) friction, the coast is always at the right of the direction of motion. The domain is closed on the right hand side, to mimic an enclosed shelf sea. The wave enters the domain on the lower left hand side and travels towards the right. On the right hand side the wave is reflected and travels back towards the left. On the closed side the reflection happens through the creation of Poincare waves which are not modelled here. The panels on the left shows the sea surface height and the panels on the right the velocity. The colours are indicated in the colour bar, the arrows are scaled velocity vectors Kelvin Wave shelf.gif
Animation 1: A traveling Kelvin wave is shown for the Northern hemisphere with (lower panel) and without (upper panel) friction, the coast is always at the right of the direction of motion. The domain is closed on the right hand side, to mimic an enclosed shelf sea. The wave enters the domain on the lower left hand side and travels towards the right. On the right hand side the wave is reflected and travels back towards the left. On the closed side the reflection happens through the creation of Poincare waves which are not modelled here. The panels on the left shows the sea surface height and the panels on the right the velocity. The colours are indicated in the colour bar, the arrows are scaled velocity vectors
Figure 2: The time averaged sea surface height is shown in red shading and the dashed lines are cotidal lines at intervals of roughly an hour. The Amphidromic points are located at the intersect of all lines. The upper panel shows the case for without friction and the lower panel for the case with friction Amphidromic points.png
Figure 2: The time averaged sea surface height is shown in red shading and the dashed lines are cotidal lines at intervals of roughly an hour. The Amphidromic points are located at the intersect of all lines. The upper panel shows the case for without friction and the lower panel for the case with friction

The expression of tides as a bounded Kelvin wave is well observable in enclosed shelf seas around the world (e.g. the English channel, the North Sea or the Yellow sea). Animation 1 shows the behaviour of a simplified case of a Kelvin wave in an enclosed shelf sea for the case with (lower panel) and without friction (upper panel). The shape of an enclosed shelf sea is represented as a simple rectangular domain in the Northern Hemisphere which is open on the left hand side and closed on the right hand side. The tidal wave, a Kelvin wave, enters the domain in the lower left corner and travels to the right with the coast on its right. The sea surface height (SSH, left panels of animation 1), the tidal elevation, is maximum at the coast and decreases towards the centre of the domain. The tidal currents (right panels of animation 1) are in the direction of wave propagation under the crest and in the opposite direction under the through. They are both maximum under the crest and the trough of the waves and decrease towards the centre. This was expected as the equations for and are in phase as they both depend on the same arbitrary function describing the wave motion and exponential decay term.

On the enclosed right hand side, the Kelvin wave is reflected and because it always travels with the coast on its right, it will now travel in the opposite direction. The energy of the incoming Kelvin wave is transferred through Poincare waves along the enclosed side of the domain to the outgoing Kelvin wave. The final pattern of the SSH and the tidal currents is made up of the sum of the two Kelvin waves. These two can amplify each other and this amplification is maximum when the length of the shelf sea is a quarter wavelength of the tidal wave. [2] Next to that, the sum of the two Kelvin waves result in several static minima's in the centre of the domain which hardly experience any tidal motion, these are called Amphidromic points. In the upper panel of figure 2, the absolute time averaged SSH is shown in red shading and the dotted lines show the zero tidal elevation level at roughly hourly intervals, also known as cotidal lines. Where these lines intersect the tidal elevation is zero during a full tidal period and thus this is the location of the Amphidromic points.

In the real world, the reflected Kelvin wave has a lower amplitude due to energy loss as a result of friction and through the transfer via Poincare waves (lower left panel of animation 1). The tidal currents are proportional to the wave amplitude and therefore also decrease on the side of the reflected wave (lower right panel of animation 1). Finally, the static minima's are no longer in the centre of the domain as wave amplitude is no longer symmetric. Therefore, the Amphidromic points shift towards the side of the reflected wave (lower panel figure 2).

The dynamics of a tidal Kelvin wave in enclosed shelf sea is well manifested and studied in the North Sea. [15]

Tides in estuaries and basins

When tides enter estuaries or basins, the boundary conditions change as the geometry changes drastically. The water depth becomes shallower and the width decreases, next to that the depth and width become significantly variable over the length and width of the estuary or basin. As a result the tidal wave deforms which affects the tidal amplitude, phase speed and the relative phase between tidal velocity and elevation. The deformation of the tide is largely controlled by the competition between bottom friction and channel convergence. [16] Channel convergence increases the tidal amplitude and phase speed as the energy of the tidal wave is traveling through a smaller area while bottom friction decrease the amplitude through energy loss. [17] The modification of the tide leads to the creation of overtides (e.g. tidal constituents) or higher harmonics. These overtides are multiples, sums or differences of the astronomical tidal constituents and as a result the tidal wave can become asymmetric. [18] A tidal asymmetry is a difference between the duration of the rise and the fall of the tidal water elevation and this can manifest itself as a difference in flood/ebb tidal currents. [19] The tidal asymmetry and the resulting currents are important for the sediment transport and turbidity in estuaries and tidal basins. [20] Each estuary and basin has its own distinct geometry and these can be subdivided in several groups of similar geometries with its own tidal dynamics. [16]

See also

Related Research Articles

<span class="mw-page-title-main">Wave equation</span> Differential equation important in physics

The (two-way) wave equation is a second-order linear partial differential equation for the description of waves or standing wave fields – as they occur in classical physics – such as mechanical waves or electromagnetic waves. It arises in fields like acoustics, electromagnetism, and fluid dynamics. Single mechanical or electromagnetic waves propagating in a pre-defined direction can also be described with the first-order one-way wave equation, which is much easier to solve and also valid for inhomogeneous media.

<span class="mw-page-title-main">Gravity wave</span> Wave in or at the interface between fluids where gravity is the main equilibrium force

In fluid dynamics, gravity waves are waves generated in a fluid medium or at the interface between two media when the force of gravity or buoyancy tries to restore equilibrium. An example of such an interface is that between the atmosphere and the ocean, which gives rise to wind waves.

The primitive equations are a set of nonlinear partial differential equations that are used to approximate global atmospheric flow and are used in most atmospheric models. They consist of three main sets of balance equations:

  1. A continuity equation: Representing the conservation of mass.
  2. Conservation of momentum: Consisting of a form of the Navier–Stokes equations that describe hydrodynamical flow on the surface of a sphere under the assumption that vertical motion is much smaller than horizontal motion (hydrostasis) and that the fluid layer depth is small compared to the radius of the sphere
  3. A thermal energy equation: Relating the overall temperature of the system to heat sources and sinks
<span class="mw-page-title-main">Rossby wave</span> Inertial wave occurring in rotating fluids

Rossby waves, also known as planetary waves, are a type of inertial wave naturally occurring in rotating fluids. They were first identified by Sweden-born American meteorologist Carl-Gustaf Arvid Rossby in the Earth's atmosphere in 1939. They are observed in the atmospheres and oceans of Earth and other planets, owing to the rotation of Earth or of the planet involved. Atmospheric Rossby waves on Earth are giant meanders in high-altitude winds that have a major influence on weather. These waves are associated with pressure systems and the jet stream. Oceanic Rossby waves move along the thermocline: the boundary between the warm upper layer and the cold deeper part of the ocean.

<span class="mw-page-title-main">Wave power</span> Transport of energy by wind waves, and the capture of that energy to do useful work

Wave power is the capture of energy of wind waves to do useful work – for example, electricity generation, water desalination, or pumping water. A machine that exploits wave power is a wave energy converter (WEC).

A Kelvin wave is a wave in the ocean or atmosphere that balances the Earth's Coriolis force against a topographic boundary such as a coastline, or a waveguide such as the equator. A feature of a Kelvin wave is that it is non-dispersive, i.e., the phase speed of the wave crests is equal to the group speed of the wave energy for all frequencies. This means that it retains its shape as it moves in the alongshore direction over time.

<span class="mw-page-title-main">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

Atmospheric tides are global-scale periodic oscillations of the atmosphere. In many ways they are analogous to ocean tides. Atmospheric tides can be excited by:

<span class="mw-page-title-main">Boussinesq approximation (water waves)</span> Approximation valid for weakly non-linear and fairly long waves

In fluid dynamics, the Boussinesq approximation for water waves is an approximation valid for weakly non-linear and fairly long waves. The approximation is named after Joseph Boussinesq, who first derived them in response to the observation by John Scott Russell of the wave of translation. The 1872 paper of Boussinesq introduces the equations now known as the Boussinesq equations.

<span class="mw-page-title-main">Stokes wave</span> Nonlinear and periodic surface wave on an inviscid fluid layer of constant mean depth

In fluid dynamics, a Stokes wave is a nonlinear and periodic surface wave on an inviscid fluid layer of constant mean depth. This type of modelling has its origins in the mid 19th century when Sir George Stokes – using a perturbation series approach, now known as the Stokes expansion – obtained approximate solutions for nonlinear wave motion.

In fluid dynamics, Luke's variational principle is a Lagrangian variational description of the motion of surface waves on a fluid with a free surface, under the action of gravity. This principle is named after J.C. Luke, who published it in 1967. This variational principle is for incompressible and inviscid potential flows, and is used to derive approximate wave models like the mild-slope equation, or using the averaged Lagrangian approach for wave propagation in inhomogeneous media.

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">Cnoidal wave</span> Nonlinear and exact periodic wave solution of the Korteweg–de Vries equation

In fluid dynamics, a cnoidal wave is a nonlinear and exact periodic wave solution of the Korteweg–de Vries equation. These solutions are in terms of the Jacobi elliptic function cn, which is why they are coined cnoidal waves. They are used to describe surface gravity waves of fairly long wavelength, as compared to the water depth.

<span class="mw-page-title-main">Radiation stress</span> Term in physical oceanography

In fluid dynamics, the radiation stress is the depth-integrated – and thereafter phase-averaged – excess momentum flux caused by the presence of the surface gravity waves, which is exerted on the mean flow. The radiation stresses behave as a second-order tensor.

<span class="mw-page-title-main">Green's law</span> Equation describing evolution of waves in shallow water

In fluid dynamics, Green's law, named for 19th-century British mathematician George Green, is a conservation law describing the evolution of non-breaking, surface gravity waves propagating in shallow water of gradually varying depth and width. In its simplest form, for wavefronts and depth contours parallel to each other, it states:

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.

Nonlinear tides are generated by hydrodynamic distortions of tides. A tidal wave is said to be nonlinear when its shape deviates from a pure sinusoidal wave. In mathematical terms, the wave owes its nonlinearity due to the nonlinear advection and frictional terms in the governing equations. These become more important in shallow-water regions such as in estuaries. Nonlinear tides are studied in the fields of coastal morphodynamics, coastal engineering and physical oceanography. The nonlinearity of tides has important implications for the transport of sediment.

<span class="mw-page-title-main">Topographic Rossby waves</span> Description of waves in the ocean and atmosphere created by bottom irregularities

Topographic Rossby waves are geophysical waves that form due to bottom irregularities. For ocean dynamics, the bottom irregularities are on the ocean floor such as the mid-ocean ridge. For atmospheric dynamics, the other primary branch of geophysical fluid dynamics, the bottom irregularities are found on land, for example in the form of mountains. Topographic Rossby waves are one of two types of geophysical waves named after the meteorologist Carl-Gustaf Rossby. The other type of Rossby waves are called planetary Rossby waves and have a different physical origin. Planetary Rossby waves form due to the changing Coriolis parameter over the earth. Rossby waves are quasi-geostrophic, dispersive waves. This means that not only the Coriolis force and the pressure-gradient force influence the flow, as in geostrophic flow, but also inertia.

References

  1. https://beltoforion.de/en/tides/simulation.php Tidal forces explained
  2. 1 2 Pugh, David; Woodworth, Philip (2014). Sea-Level Science: Understanding Tides, Surges, Tsunamis and Mean Sea-Level Changes. Cambridge: Cambridge University Press. pp. 97–132. doi:10.1017/cbo9781139235778. ISBN   978-1-107-02819-7.
  3. Parker, Bruce B.; Davies, Alan M.; Xing, Jiuxing (1999), "Tidal height and current prediction", Coastal and Estuarine Studies, Washington, D. C.: American Geophysical Union, pp. 277–327, doi:10.1029/ce056p0277, ISBN   0-87590-270-7 , retrieved 2021-05-13
  4. 1 2 3 Pugh, David; Woodworth, Philip (2014). Sea-Level Science: Understanding Tides, Surges, Tsunamis and Mean Sea-Level Changes. Cambridge: Cambridge University Press. pp. 370–375. doi:10.1017/cbo9781139235778. ISBN   978-1-107-02819-7.
  5. "Wave Energy - Energy associated with wave motion". web.mit.edu. Retrieved 2021-05-13.
  6. Cushman-Roisin, Benoit; Beckers, Jean-Marie (2011). Introduction to Geophysical Fluid Dynamics: Physical and Numerical Aspects. Academic Press. pp. 778–780. doi:10.1016/c2009-0-00052-x. ISBN   978-0-12-088759-0.
  7. Garrett, Chris; Kunze, Eric (2007-01-21). "Internal Tide Generation in the Deep Ocean". Annual Review of Fluid Mechanics. 39 (1): 57–87. Bibcode:2007AnRFM..39...57G. doi:10.1146/annurev.fluid.39.050905.110227. ISSN   0066-4189.
  8. Holloway, Peter E.; Chatwin, Paul G.; Craig, Peter (2001-06-06). "Internal Tide Observations from the Australian North West Shelf in Summer 1995". Journal of Physical Oceanography. 31 (5): 1182–1199. Bibcode:2001JPO....31.1182H. doi: 10.1175/1520-0485(2001)031<1182:itofta>2.0.co;2 . ISSN   0022-3670.
  9. Palmer, M.R.; Stephenson, G.R.; Inall, M.E.; Balfour, C.; Düsterhus, A.; Green, J.A.M. (2015-04-01). "Turbulence and mixing by internal waves in the Celtic Sea determined from ocean glider microstructure measurements". Journal of Marine Systems. 144: 57–69. Bibcode:2015JMS...144...57P. doi: 10.1016/j.jmarsys.2014.11.005 . ISSN   0924-7963. S2CID   128841660.
  10. Kelly, S. M.; Nash, J. D. (2010-12-10). "Internal-tide generation and destruction by shoaling internal tides". Geophysical Research Letters. 37 (23): n/a. Bibcode:2010GeoRL..3723611K. doi: 10.1029/2010gl045598 . ISSN   0094-8276.
  11. Sharples, Jonathan; Moore, C. Mark; Hickman, Anna E.; Holligan, Patrick M.; Tweddle, Jacqueline F.; Palmer, Matthew R.; Simpson, John H. (2009-12-02). "Internal tidal mixing as a control on continental margin ecosystems". Geophysical Research Letters. 36 (23). Bibcode:2009GeoRL..3623603S. doi: 10.1029/2009gl040683 . ISSN   0094-8276. S2CID   129002095.
  12. Wang, B. (2003), "Kelvin Waves", Encyclopedia of Atmospheric Sciences, Elsevier, pp. 1062–1068, doi:10.1016/b0-12-227090-8/00191-3, ISBN   978-0-12-227090-1 , retrieved 2021-05-17
  13. "Midlatitude Dynamics and Quasi-Geostrophic Models". International Geophysics. 66: 291–336. 2000-01-01. doi:10.1016/S0074-6142(00)80009-1. ISBN   9780124340688. ISSN   0074-6142.
  14. 1 2 Cushman-Roisin, Benoit; Beckers, Jean-Marie (2011). Introduction to Geophysical Fluid Dynamics: Physical and Numerical Aspects. Academic Press. pp. 273–276. doi:10.1016/c2009-0-00052-x. ISBN   978-0-12-088759-0.
  15. Roos, Pieter C.; Velema, Jorick J.; Hulscher, Suzanne J. M. H.; Stolk, Ad (2011-07-19). "An idealized model of tidal dynamics in the North Sea: resonance properties and response to large-scale changes". Ocean Dynamics. 61 (12): 2019–2035. Bibcode:2011OcDyn..61.2019R. doi: 10.1007/s10236-011-0456-x . ISSN   1616-7341. S2CID   19852281.
  16. 1 2 Friedrichs, Carl T. (2010), Valle-Levinson, Arnoldo (ed.), "Barotropic tides in channelized estuaries", Contemporary Issues in Estuarine Physics, Cambridge: Cambridge University Press, pp. 27–61, doi:10.1017/cbo9780511676567.004, ISBN   978-0-511-67656-7 , retrieved 2021-05-18
  17. Friedrichs, Carl T.; Aubrey, David G. (1994). "Tidal propagation in strongly convergent channels". Journal of Geophysical Research. 99 (C2): 3321. Bibcode:1994JGR....99.3321F. doi:10.1029/93jc03219. ISSN   0148-0227. S2CID   56037856.
  18. Pugh, David; Woodworth, Philip (2014). Sea-Level Science: Understanding Tides, Surges, Tsunamis and Mean Sea-Level Changes. Cambridge: Cambridge University Press. pp. 133–154. doi:10.1017/cbo9781139235778. ISBN   978-1-107-02819-7.
  19. Cavalcante, Geórgenes H. (2016), "Tidal Asymmetry", in Kennish, Michael J. (ed.), Encyclopedia of Estuaries, Encyclopedia of Earth Sciences Series, Dordrecht: Springer Netherlands, p. 664, doi:10.1007/978-94-017-8801-4_183, ISBN   978-94-017-8800-7 , retrieved 2021-05-18
  20. Gallo, Marcos Nicolás; Vinzon, Susana Beatriz (2005-08-30). "Generation of overtides and compound tides in Amazon estuary". Ocean Dynamics. 55 (5–6): 441–448. Bibcode:2005OcDyn..55..441G. doi:10.1007/s10236-005-0003-8. ISSN   1616-7341. S2CID   129123300.