Skip to main content
Log in

Fully nonlinear Boussinesq equations for long wave propagation and run-up in sloping channels with parabolic cross sections

  • Original Paper
  • Published:
Natural Hazards Aims and scope Submit manuscript

Abstract

A general framework for derivation of long wave equations in narrow channels, and their transformation to Lagrangian coordinates is briefly established. Then, fully nonlinear Boussinesq equations are derived for channels of parabolic cross sections. The simplified version with normal nonlinearity is compared with corresponding models from the literature, and propagation properties are discussed. A Lagrangian run-up model is adapted to the fully nonlinear set. This model is tested by means of controlled residues and by a well-controlled comparison to exact analytic solutions from the literature. Then, run-up of solitary waves in simple geometries is simulated and compared to a semi-analytic solution that is derived for propagation and run-up in a composite channel. The dispersive model retains the higher run-up height in a parabolic channels, as reported in the recent literature for NLSW solutions, as compared to a rectangular channel.

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

Similar content being viewed by others

References

  • Antuono M, Brocchini M (2007) The boundary value problem for the nonlinear shallow water equations. Stud Appl Math 119:71–91

    Article  Google Scholar 

  • Antuono M, Brocchini M (2008) Maximum run-up, breaking conditions and dynamical forces in the swash zone: a boundary value approach. Coast Eng 55:732–740

    Article  Google Scholar 

  • Brocchini M (2013) A reasoned overview on Boussinesq-type models: the interplay between physics, mathematics and numerics. Proc R Soc 469:20130,496

    Article  Google Scholar 

  • Carrier GF, Greenspan HP (1958) Water waves of finite amplitude on a sloping beach. J Fluid Mech 4:97–109

    Article  Google Scholar 

  • Chang P, Melville WK, Miles JW (1979) On the evolution of a solitary wave in a gradually varying channel. J Fluid Mech 95:401–414

    Article  Google Scholar 

  • Choi B, Pelinovsky E, Kim D, Didenkulova I (2008) Two- and three- dimensional computation of solitary wave runup on non-plane beach. Nonlinear Process Geophys 15:489–502

    Article  Google Scholar 

  • Choi BH, Pelinovsky E, Kim DC, Didenkulova I, Woo SB (2008) Two- and three-dimensional computation of solitary wave runup on non-plane beach. Nonlinear Process Geophys 15:489–502

    Article  Google Scholar 

  • Didenkulova I, Pelinovsky E (2009) Non-dispersive traveling waves in inclined shallow water channels. Phys Lett A 373:3883–3887

    Article  Google Scholar 

  • Didenkulova I, Pelinovsky E (2011) Nonlinear wave evolution and runup in an inclined channel of a parabolic cross-section. Phys Fluids 23:086,602

    Article  Google Scholar 

  • Didenkulova I, Pelinovsky E (2011) Runup of tsunami waves in u-shaped bays. Pure Appl Geophys 168:1239–1249

    Article  Google Scholar 

  • Didenkulova I, Pelinovsky E, Soomere T (2009) Long surface wave dynamics along a convex bottom. J Geophys Res 114(C7):2156–2202

    Article  Google Scholar 

  • Fenton JD (1973) Cnoidal waves and bores in uniform channels of arbitrary cross-section. J Fluid Mech 58:417–434

    Article  Google Scholar 

  • Grimshaw R (1971) The solitary wave in water of variable depth. Part 2. J Fluid Mech 46:611–622

    Article  Google Scholar 

  • Jensen A, Pedersen G, Wood DJ (2003) An experimental study of wave run-up at a steep beach. J Fluid Mech 486:161–188

    Article  Google Scholar 

  • Liu PLF, Yeh H, Synolakis CE (eds) (2008) Advances in coastal and ocean engineering, vol 10. World Scientific Publishing, Singapore

    Google Scholar 

  • Løvholt F, Lynett P, Pedersen GK (2013) Simulating run-up on steep slopes with operational Boussinesq models; capabilities, spurious effects and instabilities. Nonlinear Process Geophys 20:379–395

    Article  Google Scholar 

  • Madsen P, Sørensen O (1992) A new form of the Boussinesq equations with improved linear dispersion characteristics. Part 2. A slowly-varying bathymetry. Coast Eng 18:183–204

    Article  Google Scholar 

  • Madsen PA, Schäffer HA (1999) A review of Boussinesq-type equations for surface gravity waves. World Scientific Publishing Co., Singapore

    Book  Google Scholar 

  • Nachbin A, Simoes VS (2015) Solitary waves in forked channel regions. J Fluid Mech 777:544–568

    Article  Google Scholar 

  • Pedersen G (2008) Modeling run-up with depth integrated equation models. In: Liu PLF, Yeh H, Synolakis CE (eds) Advanced numerical models for simulating tsunami waves and runup, advances in coastal and ocean engineering, vol 10. World Scientific Publishing, Singapore, pp 3–42

    Chapter  Google Scholar 

  • Pedersen G, Gjevik B (1983) Run-up of solitary waves. J Fluid Mech 135:283–299

    Article  Google Scholar 

  • Peregrine DH (1968) Long waves in a uniform channel of arbitrary cross-section. J Fluid Mech 32:353–365

    Article  Google Scholar 

  • Peregrine DH (1969) Solitary waves in trapezoidal channels. J Fluid Mech 35:1–6

    Article  Google Scholar 

  • Peters AS (1966) Rotational and irrotational solitary waves in a channel with arbitrary cross. Commun Pure Appl Math 19(4):445–471

    Article  Google Scholar 

  • Synolakis CE (1987) The run-up of solitary waves. J Fluid Mech 185:523–545

    Article  Google Scholar 

  • Teng MH, Wu TY (1992) Nonlinear water waves in channels of arbitrary shape. J Fluid Mech 242:211–233

    Article  Google Scholar 

Download references

Acknowledgments

The subject presented was first suggested after a talk by Efim Pelinovsky back in 2011. Unfortunately, it took a while before the present study materialized. The author is also grateful to Patricio Winckler Grez for helpful communication. This work has been supported by the Norwegian Research Council under project no. 205184.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to G. Pedersen.

Appendices

Appendix 1: The standard Boussinesq and the KdV equation

After deletion of the nonlinear parts in E (4) reads

$$\begin{aligned} p=p_s+{\tilde{\eta }}+\mu ^2\left( \widetilde{\frac{\partial \phi }{\partial t} } -\frac{\partial \phi }{\partial t} \right) +O(\epsilon \mu ^2,\mu ^4). \end{aligned}$$

Insertion of this expression in the momentum equation followed by averaging then yields

$$\begin{aligned} {{\overline{D}}}{\overline{u}}=-\frac{\partial {\tilde{\eta }}}{\partial x}+\mu ^2\left( \overline{\left( \frac{\partial ^2\phi }{\partial t\partial x}\right) }-\frac{\partial \ }{\partial x}\left( \widetilde{\frac{\partial \phi }{\partial t} }\right) \right) +O(\epsilon \mu ^2,\mu ^4), \end{aligned}$$
(23)

and

$$\begin{aligned} \displaystyle {{\overline{D}}}{\overline{u}}=-\frac{\partial {\tilde{\eta }}}{\partial x} +\mu ^2\left( \left\{ \frac{4h_m^2}{21}+\frac{h_m}{45\beta }\right\} \frac{\partial ^3 {\overline{u}}}{\partial x^2\partial t} \displaystyle +h_m'\left\{ \frac{2h_m}{3}+\frac{1}{18\beta }\right\} \frac{\partial ^2 {\overline{u}}}{\partial x\partial t} +\frac{2h_mh_m''}{5}\frac{\partial {\overline{u}}}{\partial t} \right) +O(\epsilon \mu ^2,\mu ^4). \end{aligned}$$
(24)

When compared to equation 69 of Teng and Wu (1992), there are differences in the reciprocal \(\beta \) terms. This is apparently due to a switching of the outer spatial derivative and the surface (\(\,\widetilde{\ }\) ) average in the equation equivalent to (23) herein (see equation 38 in the reference).

To derive a KdV equation for unidirectional waves, it is most convenient to rescale according to \({\tilde{\eta }}=\epsilon {\hat{\eta }}\) and \({\overline{u}}=\epsilon {\hat{u}}\) and assume that \(\epsilon =\mu ^2\) is appropriate. For \(h_m=h_0=\hbox {const}\). the momentum and continuity equation become

$$\begin{aligned} \begin{array}{l} \displaystyle \frac{\partial {\hat{u}}}{\partial t}+\epsilon {\hat{u}} \frac{\partial {\hat{u}}}{\partial x}=-\frac{\partial {\hat{\eta }}}{\partial x}+\epsilon M \frac{\partial ^3 {\hat{u}}}{\partial x^2\partial t}+O(\epsilon ^2),\\ \displaystyle \frac{\partial {\hat{\eta }}}{\partial t}+\epsilon {\hat{u}} \frac{\partial {\hat{\eta }}}{\partial x}=-\nu (h_0+\epsilon {\hat{\eta }})\frac{\partial {\hat{u}}}{\partial x}+O(\epsilon ^2), \end{array} \end{aligned}$$

where M and \(\nu \) are as in (18). Next, these equations are transformed by introducing a translating spatial variable, \(\xi \), and a slow temporal variable \(\tau \)

$$\begin{aligned} \xi =x-c_0 t,\quad \tau =\epsilon t, \end{aligned}$$

where \(c_0\) is as in (18). The equations become

$$\begin{aligned}&\,&\frac{\partial {\hat{u}}}{\partial \tau }+(\epsilon {\hat{u}} -c_0)\frac{\partial {\hat{u}}}{\partial \xi }=-\frac{\partial {\hat{\eta }}}{\partial \xi }+\epsilon M \frac{\partial ^3 {\hat{u}}}{\partial \xi ^3}+O(\epsilon ^2), \end{aligned}$$
(25)
$$\begin{aligned}&\,&\frac{\partial {\hat{\eta }}}{\partial \tau }+(\epsilon {\hat{u}} -c_0)\frac{\partial {\hat{\eta }}}{\partial \xi }=-\nu (h_0+\epsilon {\hat{\eta }})\frac{\partial {\hat{u}}}{\partial \xi }+O(\epsilon ^2). \end{aligned}$$
(26)

To leading order, these imply \(c_0{\hat{u}}={\hat{\eta }}+O(\epsilon )\) which is used to eliminate \({\hat{u}}\) from all \(O(\epsilon )\) terms in (25) and (26). From the resulting two equations \(\frac{\partial {\hat{u}}}{\partial \xi }\) is then eliminated and a single equation for \({\hat{\eta }}\) is obtained. The KdV equation on the form (18) then follows when \({\tilde{\eta }}, x\) and t are restored.

Appendix 2: A linear solution for the wave tank geometry

The incident wave in a geometry like the one in Fig. 4 will be modified when entering the slope. To assess this effect, we employ the availability of explicit local solutions to find a LSW solution with a sharp apex between the beach and the flat bottom at \(x=\ell =\alpha ^{-1}\).

In the linear approximation, the incident wave of (11) and (12) may be expressed, for \(x\le \ell \), according to

$$\begin{aligned} \eta =\nu x^{-\frac{1}{2}}I'(\tau _s),\quad u=\nu \left( \frac{1}{2}x^{-\frac{3}{2}}I(\tau _s)-\sqrt{\frac{3}{2\alpha }}x^{-1}I'(\tau _s)\right) , \quad \tau _s=2\left( \sqrt{\frac{3x}{2\alpha }}-\sqrt{\frac{3\ell }{2\alpha }}\right) +t+\frac{L}{c_0}-t_a \end{aligned}$$
(27)

where I is a potential related to \({\varPhi }, \nu \) is a constant and the last part of \(\tau _s\) is included to make the expressions compatible with the incident wave as specified below. We skip the details, but note that (27) may also be readily verified by direct substitution into the LSW equations. If we assume that \(I'\) defines a pulse of finite extent and omit the first term in the expression for u we obtain the optical approximation [see Didenkulova and Pelinovsky (2011b)] for which the energy is conserved and for which kinetic and potential energies are equal. On the other hand, this solution does not conserve volume. Instead the volume is decreasing as the pulse approaches the shore. Hence, in the full u in (27), we have the first term which gives a reduced kinetic energy, thereby also a deviation from energy equipartition, and a constant offshore cross-integrated flux in the wake of the incident pulse to account for the volume loss. Since this residual current is stationary, it is consistent with a surface elevation of finite extent. However, u does decrease toward deeper water since the cross-sectional area increases. Thus, there is a convective retardation and this is consistent with the weak trailing surface depression inherent in the nonlinear solution, as given by the last term of expression for \(\eta \) in (11). While we could reproduce the trailing current by the wave paddle in the test case of Sect. 2.7, it is difficult to conceive how a surface pulse with a wake current could constitute the whole solution on the slope in a composite geometry as shown in Fig. 4. Even if the current could extend beyond the apex it would have to be finite in extent, since the incident wave from the constant depth region is confined. A finite current implies a surface gradient and a propagating wave. Hence, we must expect both modifications in the shape of the wave transmitted to the slope and a reflected wave.

If we combine (27) with the odd shoreline reflection, we have the full LSW solution on the slope. For \(x\ge \ell \), the solution is

$$\begin{aligned} \eta =A_0 (Y(\tau _i)+P(\tau _r)),\quad u=\frac{A_0}{c_0} \left( -Y(\tau _i)+P(\tau _r)\right) ,\quad \tau _i=\frac{x}{c_0}+t-t_a,\quad \tau _r=-\frac{\ell -x}{c_0}+t-t_a, \end{aligned}$$
(28)

where \(c_0\) is the shallow water speed over the flat bottom and Y and P represents the incident wave and the one reflected from the apex, respectively. If we restrict ourselves to the time period before the reflected wave from the beach reaches \(x=\ell \), we may patch (27) and (28) by requiring continuous \(\eta \) and u. Choosing \(\nu =\sqrt{\ell }A_0\) and employing some elementary manipulation we obtain

$$\begin{aligned} I'-\frac{c_0}{4\ell }I=Y,\quad P=\frac{c_0}{4\ell }I. \end{aligned}$$
(29)

When the incident wave, \(A_0 Y\), is specified this is a first order, linear ODE for I. Some Y shapes yield I as simple elementary function. However, from (19), we obtain (ignoring \(c_1\)) \(Y(\tau )=\mathrm {sech}^2(kc_0\tau )\) and an I that may expressed by hypergeometric functions. It is then simpler to solve (29) with high accuracy by a Runge–Kutta method. Derivatives of any order may then be obtained, without quality loss, by repeated differentiation of (29). Combining the incident and reflected waves on the beach and invoking the limit \(x\rightarrow 0^+\) we find the shoreline elevation

$$\begin{aligned} \eta _s=\frac{4A_0\ell }{c_0}I'', \end{aligned}$$
(30)

from which the maximum value, R, can be readily and accurately obtained by representing \(I''\) by splines and differentiate. When the wavelength of the incident wave is small compared to \(\ell \), the leading order approximation in (29) is \(I'=Y\) and corrections may then be obtained by perturbation, or by manipulating the hypergeometric integral in the full solution. This yields

$$\begin{aligned} I'(\tau )\approx \mathrm {sech}^2(kc_0\tau )+\frac{1}{4k\ell }\left( 1+ \mathrm {tanh}(kc_0\tau )\right) ,\quad \frac{A_\ell }{A_0}\approx 1+\frac{1}{4k\ell },\quad \frac{R}{A_0}\approx \frac{8k}{3\sqrt{3}\alpha }\left( 1+\frac{\sqrt{3}}{8k\ell }\right) . \end{aligned}$$
(31)

where relative errors of order \((k\ell )^{-2}\) are implicit in all three relations and \(A_\ell \) is the maximum surface elevation at \(x=\ell \). We observe (see Fig. 8) that the bell shape of the solitary wave is transformed into a bell shape plus a long shelf when it passes the vertex. The height of the surface elevation is then also increased, from \(A_0\) to \(A_\ell \). Correspondingly, the maximum run-up height is increased as compared to the formula (21). Since linear and nonlinear theory yield the same R from initial conditions on the slope, the expression in (31) should apply when \(A_0\) small, while the wave may be strongly nonlinear near the shore. The reflection from the apex is a long elevation. In principle, the incident wave as obtained above could be inserted into the theory outlined in Sect. 2.5.2. However, this is rather cumbersome and hardly worthwhile when the nonlinearities will not alter the maximum run-up height. If other properties are wanted, we would anyhow be better served by the numerical solutions of the Lagrangian NLSW.

An asymptotic solution is also readily obtained for long waves (solitary waves of low amplitudes) incident on steep beaches. The details are omitted, but the maximum run-up height becomes

$$\begin{aligned} \frac{R}{A_0}\approx 2+\frac{16}{9}(k\ell )^2, \end{aligned}$$
(32)

demonstrating that the run-up height approaches twice the amplitude in this limit, just as for the case with a plane beach.

If we instead of adding a horizontal channel section specified a paddle velocity, \(A_0U/c_0\), on the slope, the ODE in (29) would be replaced by \(I'-\frac{c_0}{2\ell }I=U\). If U was a pulse of finite duration, we would still have tails in the waves on the slope, as in the solution from (29).

Fig. 8
figure 8

The linear surface elevation for the transmitted and reflected waves for \(A_0=0.1\) and \(\theta =3^\circ \). The curve marked “full” is computed with the Runge--Kutta solution for I, the one marked “pert.” corresponds to equation (31), while the curve market “opt.” displays the optical approximations in the sense that only the first term in the expression for \(\eta \) from (31) is retained. The gray, vertical bar shows the position of the bottom apex

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pedersen, G. Fully nonlinear Boussinesq equations for long wave propagation and run-up in sloping channels with parabolic cross sections. Nat Hazards 84 (Suppl 2), 599–619 (2016). https://doi.org/10.1007/s11069-016-2448-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11069-016-2448-0

Keywords

Navigation