Mild-slope equation

Source: Wikipedia, the free encyclopedia.
CGWAVE
(which solves the mild-slope equation).

In

harbours and coasts
.

The mild-slope equation models the propagation and transformation of water waves, as they travel through waters of varying depth and interact with lateral boundaries such as

bathymetric changes of the sea bed and coastline, mean flow fields and mass transfer of dissolved and floating materials. Most often, the mild-slope equation is solved by computer using methods from numerical analysis
.

A first form of the mild-slope equation was developed by

approximations to the mild-slope equation are often used, in order to reduce the computational cost.

In case of a constant depth, the mild-slope equation reduces to the Helmholtz equation for wave diffraction.

Formulation for monochromatic wave motion

For

monochromatic waves according to linear theory—with the free surface
elevation given as and the waves propagating on a fluid layer of mean water depth —the mild-slope equation is:[4]
where:

  • is the
    complex-valued amplitude
    of the free-surface elevation
  • is the horizontal position;
  • is the angular frequency of the monochromatic wave motion;
  • is the imaginary unit;
  • means taking the
    real part
    of the quantity between braces;
  • is the horizontal gradient operator;
  • is the divergence operator;
  • is the wavenumber;
  • is the
    phase speed
    of the waves and
  • is the
    group speed
    of the waves.

The phase and group speed depend on the dispersion relation, and are derived from Airy wave theory as:[5]

where

  • is
    Earth's gravity
    and
  • is the
    hyperbolic tangent
    .

For a given angular frequency , the wavenumber has to be solved from the dispersion equation, which relates these two quantities to the water depth .

Transformation to an inhomogeneous Helmholtz equation

Through the transformation

the mild slope equation can be cast in the form of an
inhomogeneous Helmholtz equation:[4][6]
where is the
Laplace operator.

Propagating waves

In spatially

complex amplitude
in its amplitude and phase, both real valued:[7]
where

  • is the amplitude or absolute value of and
  • is the wave phase, which is the
    argument
    of

This transforms the mild-slope equation in the following set of equations (apart from locations for which is singular):[7]

where

  • is the average wave-energy density per unit horizontal area (the sum of the kinetic and potential energy densities),
  • is the effective wavenumber vector, with components
  • is the effective group velocity vector,
  • is the fluid density, and
  • is the acceleration by the
    Earth's gravity
    .

The last equation shows that wave energy is conserved in the mild-slope equation, and that the wave energy is transported in the -direction normal to the wave

crests (in this case of pure wave motion without mean currents).[7]
The effective group speed is different from the group speed

The first equation states that the effective wavenumber is

irrotational
, a direct consequence of the fact it is the derivative of the wave phase , a
scalar field. The second equation is the eikonal equation. It shows the effects of diffraction on the effective wavenumber: only for more-or-less progressive waves, with the splitting into amplitude and phase leads to consistent-varying and meaningful fields of and . Otherwise, κ2 can even become negative. When the diffraction effects are totally neglected, the effective wavenumber κ is equal to , and the
geometric optics approximation for wave refraction can be used.[7]

Details of the derivation of the above equations

When is used in the mild-slope equation, the result is, apart from a factor :

Now both the real part and the imaginary part of this equation have to be equal to zero:

The effective wavenumber vector is defined as the gradient of the wave phase:

and its
vector length
is

Note that is an

is zero:

Now the real and imaginary parts of the transformed mild-slope equation become, first multiplying the imaginary part by :

The first equation directly leads to the eikonal equation above for , while the second gives:

which—by noting that in which the angular frequency is a constant for time-harmonic motion—leads to the wave-energy conservation equation.

Derivation of the mild-slope equation

The mild-slope equation can be derived by the use of several methods. Here, we will use a

Stokes boundary layers (for the oscillatory part of the flow). Because the flow is irrotational, the wave motion can be described using potential flow
theory.

Details of the derivation of the mild-slope equation

Luke's variational principle

Luke's

non-linear surface gravity waves.[9]
For the case of a horizontally unbounded domain with a constant density , a free fluid surface at and a fixed sea bed at Luke's variational principle uses the Lagrangian
where is the horizontal
Lagrangian density
, given by:

where is the velocity potential, with the flow velocity components being and in the , and directions, respectively. Luke's Lagrangian formulation can also be recast into a Hamiltonian formulation in terms of the surface elevation and velocity potential at the free surface.[10] Taking the variations of with respect to the potential and surface elevation leads to the

Laplace equation
for in the fluid interior, as well as all the boundary conditions both on the free surface as at the bed at

Linear wave theory

In case of linear wave theory, the vertical integral in the Lagrangian density is split into a part from the bed to the mean surface at and a second part from to the free surface . Using a Taylor series expansion for the second integral around the mean free-surface elevation and only retaining quadratic terms in and the Lagrangian density for linear wave motion becomes

The term in the vertical integral is dropped since it has become dynamically uninteresting: it gives a zero contribution to the Euler–Lagrange equations, with the upper integration limit now fixed. The same is true for the neglected bottom term proportional to in the potential energy.

The waves propagate in the horizontal plane, while the structure of the potential is not wave-like in the vertical -direction. This suggests the use of the following assumption on the form of the potential

with normalisation
at the mean free-surface elevation

Here is the velocity potential at the mean free-surface level Next, the mild-slope assumption is made, in that the vertical shape function changes slowly in the -plane, and horizontal derivatives of can be neglected in the flow velocity. So:

As a result:

with

The Euler–Lagrange equations for this Lagrangian density are, with representing either or

Now is first taken equal to and then to As a result, the evolution equations for the wave motion become:[4]

with the horizontal gradient operator: ∇ ≡ (∂/∂x, ∂/∂y)T where superscript T denotes the transpose.

The next step is to choose the shape function and to determine and

Vertical shape function from Airy wave theory

Since the objective is the description of waves over mildly sloping beds, the shape function is chosen according to Airy wave theory. This is the linear theory of waves propagating in constant depth The form of the shape function is:[4]

with now in general not a constant, but chosen to vary with and according to the local depth and the linear dispersion relation:[4]

Here a constant angular frequency, chosen in accordance with the characteristics of the wave field under study. Consequently, the integrals and become:[4]

The following time-dependent equations give the evolution of the free-surface elevation and free-surface potential [4]

From the two evolution equations, one of the variables or can be eliminated, to obtain the time-dependent form of the mild-slope equation:[4]

and the corresponding equation for the free-surface potential is identical, with replaced by The time-dependent mild-slope equation can be used to model waves in a narrow band of frequencies around

Monochromatic waves

Consider monochromatic waves with complex amplitude and angular frequency :

with and chosen equal to each other, Using this in the time-dependent form of the mild-slope equation, recovers the classical mild-slope equation for time-harmonic wave motion:[4]

Applicability and validity of the mild-slope equation

The standard mild slope equation, without extra terms for bed slope and bed curvature, provides accurate results for the wave field over bed slopes ranging from 0 to about 1/3.[11] However, some subtle aspects, like the amplitude of reflected waves, can be completely wrong, even for slopes going to zero. This mathematical curiosity has little practical importance in general since this reflection becomes vanishingly small for small bottom slopes.

Notes

  1. ^ Berkhoff, J. C. W. (1972), "Computation of combined refraction–diffraction", Proceedings 13th International Conference on Coastal Engineering, Vancouver, pp. 471–490{{citation}}: CS1 maint: location missing publisher (link)
  2. ^ Berkhoff, J. C. W. (1976), Mathematical models for simple harmonic linear water wave models; wave refraction and diffraction (PDF) (PhD. Thesis), Delft University of Technology
  3. ^ a b c d e f g h i j Dingemans (1997, pp. 248–256 & 378–379)
  4. ^ Dingemans (1997, p. 49)
  5. ^ Mei (1994, pp. 86–89)
  6. ^ a b c d Dingemans (1997, pp. 259–262)

References