rotational surf zone modeling for boussinesq–greenâ

11
Rotational surf zone modeling for Oðl 4 Þ Boussinesq–Green–Naghdi systems Yao Zhang a , Andrew B. Kennedy a,, Aaron S. Donahue a , Joannes J. Westerink a , Nishant Panda b , Clint Dawson b a Department of Civil & Environmental Engineering & Earth Sciences, University of Notre Dame, Notre Dame, IN 46556, USA b Institute for Computational Engineering and Sciences, University of Texas at Austin, Austin, TX 78712-1229, USA article info Article history: Received 21 November 2013 Received in revised form 22 February 2014 Accepted 4 April 2014 Available online 26 April 2014 Keywords: Boussinesq modeling Nearshore hydrodynamics Numerical methods Wave breaking abstract A surf zone model is developed and tested based on the Oðl 4 Þ Boussinesq–Green–Naghdi system of Zhang et al. (2013). Because the model is fundamentally rotational, it uses fewer ad hoc assumptions than are found in many Boussinesq breaking wave systems. Eddy viscosity is used to describe both breaking dissipation and bottom friction, with breaking viscosities derived from the turbulent kinetic energy equa- tion coupled with an Oðl 4 Þ rotational wave model. In contrast, bottom friction is included by imposing the frictional coefficient-derived boundary stress as an equivalent eddy viscosity. Numerical tests for one horizontal dimension show good agreement with regular and irregular wave breaking tests, and for solitary wave runup. Surface elevation decay, setup, runup, interior orbital velocities, and depth-vary- ing undertow velocities can all be modeled reasonably. Comparison with an Oðl 2 Þ system shows similar performance for water surface elevations in the surf zone, demonstrating that the dissipation model is the major controlling factor. However, the Oðl 4 Þ model shows significantly improved representations of the velocity profile in the surf zone, as expected. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction Boussinesq-scaled water wave models give a phase resolving approximation of nearshore wave dynamics with good computa- tional efficiency and application range. Classic depth-averaged Boussinesq theory for varying depth was introduced by Peregrine (1967), but has seen strong improvements in accuracy and range of application over the past 2 decades (Madsen and Sørensen, 1992; Nwogu, 1993; Wei et al., 1995; Kennedy et al., 2001; Madsen and Schäffer, 1998; Gobbi and Kirby, 1999; Gobbi et al., 2000; Lynett and Liu, 2002). Extensions to include wave breaking and moving shorelines were introduced into previously inviscid Boussinesq models that allowed them to simulate surf zone dynamics such as wave evolution, wave setup, and large scale wave-induced currents such as rip currents, and longshore cur- rents. These have created systems that have provided good results for a variety of studies (e.g. Schäffer and Madsen, 1993; Sørensen et al., 1998; Chen et al., 2000; Kennedy et al., 2000a,b; Lynett et al., 2002; Musumeci et al., 2005; Nwogu and Demirbilek, 2010; Bonneton et al., 2011; Kim and Lynett, 2013). However, most Boussinesq models feature partial or full irrota- tionality assumptions for orbital velocities. While reasonable for nonbreaking waves, these assumptions are strongly violated under wave breaking in the surf zone. Rotational scaling has been intro- duced to some Boussinesq models with additional formulations for vorticity evolution (Veeramony and Svendsen, 2000; Musumeci et al., 2005; Kim and Lynett, 2013). These can give good results, but none of these approaches has been widely adopted, which is significant as surf zone simulations require the inclusion of vortic- ity to provide accurate reconstructions of internal velocities. For most quasi-irrotational Boussinesq models, there are two common breaking techniques. The first is the ad hoc eddy viscosity formula- tion originally developed by Zelt (1991), and extended by Kennedy et al. (2000a) that yields extra viscous terms in momentum equa- tions leading to wave dissipation. Dissipative terms have in some instances also been introduced into mass equations (Cienfuegos et al., 2010; Klonaris et al., 2013). The second technique is the sur- face roller approach attributed to Schäffer and Madsen (1993) based on the flux version of Boussinesq equations. This approach evolved from the roller concept introduced by Svendsen (1984) and was further developed by Madsen et al. (1997). Both eddy viscosity and surface roller techniques are of comparable accuracy. A third, more recent, approach turns off dispersive terms in the http://dx.doi.org/10.1016/j.ocemod.2014.04.001 1463-5003/Ó 2014 Elsevier Ltd. All rights reserved. Corresponding author. Tel.: +1 574 631 6686. E-mail address: [email protected] (A.B. Kennedy). Ocean Modelling 79 (2014) 43–53 Contents lists available at ScienceDirect Ocean Modelling journal homepage: www.elsevier.com/locate/ocemod

Upload: others

Post on 27-Nov-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Rotational surf zone modeling for Boussinesq–Greenâ

Ocean Modelling 79 (2014) 43–53

Contents lists available at ScienceDirect

Ocean Modelling

journal homepage: www.elsevier .com/locate /ocemod

Rotational surf zone modeling for Oðl4Þ Boussinesq–Green–Naghdisystems

http://dx.doi.org/10.1016/j.ocemod.2014.04.0011463-5003/� 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author. Tel.: +1 574 631 6686.E-mail address: [email protected] (A.B. Kennedy).

Yao Zhang a, Andrew B. Kennedy a,⇑, Aaron S. Donahue a, Joannes J. Westerink a, Nishant Panda b,Clint Dawson b

a Department of Civil & Environmental Engineering & Earth Sciences, University of Notre Dame, Notre Dame, IN 46556, USAb Institute for Computational Engineering and Sciences, University of Texas at Austin, Austin, TX 78712-1229, USA

a r t i c l e i n f o

Article history:Received 21 November 2013Received in revised form 22 February 2014Accepted 4 April 2014Available online 26 April 2014

Keywords:Boussinesq modelingNearshore hydrodynamicsNumerical methodsWave breaking

a b s t r a c t

A surf zone model is developed and tested based on the Oðl4Þ Boussinesq–Green–Naghdi system ofZhang et al. (2013). Because the model is fundamentally rotational, it uses fewer ad hoc assumptions thanare found in many Boussinesq breaking wave systems. Eddy viscosity is used to describe both breakingdissipation and bottom friction, with breaking viscosities derived from the turbulent kinetic energy equa-tion coupled with an Oðl4Þ rotational wave model. In contrast, bottom friction is included by imposingthe frictional coefficient-derived boundary stress as an equivalent eddy viscosity. Numerical tests forone horizontal dimension show good agreement with regular and irregular wave breaking tests, andfor solitary wave runup. Surface elevation decay, setup, runup, interior orbital velocities, and depth-vary-ing undertow velocities can all be modeled reasonably. Comparison with an Oðl2Þ system shows similarperformance for water surface elevations in the surf zone, demonstrating that the dissipation model isthe major controlling factor. However, the Oðl4Þ model shows significantly improved representationsof the velocity profile in the surf zone, as expected.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Boussinesq-scaled water wave models give a phase resolvingapproximation of nearshore wave dynamics with good computa-tional efficiency and application range. Classic depth-averagedBoussinesq theory for varying depth was introduced by Peregrine(1967), but has seen strong improvements in accuracy and rangeof application over the past 2 decades (Madsen and Sørensen,1992; Nwogu, 1993; Wei et al., 1995; Kennedy et al., 2001;Madsen and Schäffer, 1998; Gobbi and Kirby, 1999; Gobbi et al.,2000; Lynett and Liu, 2002). Extensions to include wave breakingand moving shorelines were introduced into previously inviscidBoussinesq models that allowed them to simulate surf zonedynamics such as wave evolution, wave setup, and large scalewave-induced currents such as rip currents, and longshore cur-rents. These have created systems that have provided good resultsfor a variety of studies (e.g. Schäffer and Madsen, 1993; Sørensenet al., 1998; Chen et al., 2000; Kennedy et al., 2000a,b; Lynettet al., 2002; Musumeci et al., 2005; Nwogu and Demirbilek,2010; Bonneton et al., 2011; Kim and Lynett, 2013).

However, most Boussinesq models feature partial or full irrota-tionality assumptions for orbital velocities. While reasonable fornonbreaking waves, these assumptions are strongly violated underwave breaking in the surf zone. Rotational scaling has been intro-duced to some Boussinesq models with additional formulations forvorticity evolution (Veeramony and Svendsen, 2000; Musumeciet al., 2005; Kim and Lynett, 2013). These can give good results,but none of these approaches has been widely adopted, which issignificant as surf zone simulations require the inclusion of vortic-ity to provide accurate reconstructions of internal velocities. Formost quasi-irrotational Boussinesq models, there are two commonbreaking techniques. The first is the ad hoc eddy viscosity formula-tion originally developed by Zelt (1991), and extended by Kennedyet al. (2000a) that yields extra viscous terms in momentum equa-tions leading to wave dissipation. Dissipative terms have in someinstances also been introduced into mass equations (Cienfuegoset al., 2010; Klonaris et al., 2013). The second technique is the sur-face roller approach attributed to Schäffer and Madsen (1993)based on the flux version of Boussinesq equations. This approachevolved from the roller concept introduced by Svendsen (1984)and was further developed by Madsen et al. (1997). Both eddyviscosity and surface roller techniques are of comparable accuracy.A third, more recent, approach turns off dispersive terms in the

Page 2: Rotational surf zone modeling for Boussinesq–Greenâ

44 Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53

vicinity of the breaking roller and allows the dissipative nature ofshallow water bores to remove energy from waves while conserv-ing momentum (Shi et al., 2012; Tissier et al., 2012; Gallerano atal., 2014). This shallow water bore approach requires shock-cap-turing numerical schemes to ensure stability at the bore frontbut is relatively simple to implement, as it has few adjustableparameters other than criteria for switching to shallow waterequations. None of these breaking schemes applied to quasi-irrota-tional models has well-defined velocities in the vicinity of thebreaker: all assume partially irrotational velocities which is knownto be untrue, and shallow water shock-capturing methods have anear-discontinuity in velocities at the bore front which leads tosecond order velocities being poorly defined.

An alternate but related shallow water wave theory was intro-duced by Green and Naghdi (1976). Using a polynomial structureof the velocity profile, the mass and momentum equations weresolved in a weighted residual sense with no irrotationality con-straint. Rotational Green–Naghdi equations have shown excellentnonlinear properties and very fast convergence with increasingnumbers of terms in the series (Shields and Webster, 1988; Zhaoet al., 2014b). However, their extreme complexity has generallyconfined research efforts to low levels of approximation (e.g.Ertekin et al., 1986). Although irrotational characteristics and scal-ing have also been introduced into the Green–Naghdi theory (Kimet al., 2003), these irrotational formulations hinder their extensionto the surf zone.

In contrast, Zhang et al. (2013) developed a Boussinesq–Green–Naghdi model that shows a resemblance to both Boussinesq andGreen–Naghdi systems. Polynomial expansions (Shields andWebster, 1988) and Boussinesq scaling were both applied. The sys-tems may be extended to higher order and show excellent conver-gence towards exact solutions for dispersion, shoaling, and orbitalvelocities. Most of the asymptotic rearrangement techniques usedfor Boussinesq models may also be employed to improve accuracyfor given levels of approximation. Importantly, the partial or com-plete irrotationality assumption of most Boussinesq systems wasremoved so that rotational surf zone flows may be modelednaturally.

In this paper, we extend the rotational Boussinesq–Green–Nag-hdi model (Zhang et al., 2013) to the surf zone including wavebreaking and moving shorelines. Due to the rotational nature ofthe system, fewer assumptions are needed for vorticity and viscousterms (also called dissipative/breaking terms). Viscous terms in theNavier Stokes equation represented as eddy viscosity are kept byproper scaling to reproduce the energy dissipation under thebreaking wave crest, while the eddy viscosity is modeled by thedepth-integrated turbulent-kinetic-energy equation. The spa-tially-and-temporally-varying eddy viscosity model is coupledwith the wave model to model rotational flow naturally in the surfzone. Numerous breaking tests for plunging, spilling and irregularwave breaking are shown in this paper, with computed results ofsurface elevation decay, setup, and depth-varying velocities show-ing good matches to data. A thin-layer moving shoreline techniquewas derived based on Benque’s method (1982) and the breakingsolitary wave runup simulation showed good results. Numericaltests in this paper were performed with an Oðl4Þ system – this isquite unusual as most breaking wave systems have used Oðl2Þequations. Additional Oðl2Þ runs for comparison purposes demon-strates the effect of the additional higher order terms on the overallsystem accuracy.

Because the Boussinesq class of models will never be able tosimulate the complex free surfaces found in, for example, plungingbreakers, there is an upper limit on accuracy, particularly in theroller region and near the wave crest. The weakly nonlinearOðl4Þ system used here also has its own accuracy limitations. Forthese reasons the breaking model will be intermediate in accuracy

and expense between comprehensive Navier–Stokes computa-tional fluid dynamics models (e.g. Ma et al., 2012; Higuera et al.,2013), and the much more efficient existing Boussinesq breakingmodels that have more parameterized representations of breakingprocesses. In particular, the present methods will be able to predictinternal velocities in the surf zone with more accuracy than exist-ing Boussinesq models, but with lower cost than Navier–Stokessolvers.

2. Breaking model

2.1. Boussinesq–Green–Naghdi rotational water wave system

It is convenient to put the governing equations and the bound-ary conditions into dimensionless form. We define the Boussinesq-shallow water scaling for non-dimensional variables as

ðx;yÞ¼k0ðx�;y�Þ; z¼h�10 z�; t¼k0ðg0h0Þ

12t�; h¼h�1

0 h�;

g¼ðh0Þ�1g� P¼ðq�g0h0Þ�1P�; g¼g�10 g�; ðu;vÞ¼ðg0h0Þ�1=2ðu�;v�Þ;

w¼ðk0h0Þ�1ðg0h0Þ�1=2w� mt¼h�10 ðg0h0Þ�1=2ðk0h0Þ�1m�t

sxx¼g�10 k�2

0 h�30 s�xx; szx¼g�1

0 k�10 h�2

0 s�zx ð2:1Þ

where the superscript � indicates dimensional variables, and g� isgravitational acceleration. Horizontal spatial coordinates arex� � ðx�; y�Þ, while the vertical coordinate ðz�Þ is oriented positiveupward. Time t� is scaled by typical long wave speed ðg0h0Þ1=2 andwavenumber k0, while depth h� and surface elevation g� scale withtypical water depth h0. The pressure P� is hydrostatically scaled.Horizontal and vertical velocities ðu�;v�;w�Þ all scale with waveorbital velocities taken from shallow water theory. This scalingassumes that the wave may be strongly nonlinear, although ofcourse the system is also valid for small amplitude waves. Eddy vis-cosity m�t is assumed to scale with depth and gravity, and turbulentstresses use eddy viscosity scaling combined with wavenumber anddepth, and velocity scales.

In the present work we use a Boussinesq–Green–Naghdi rota-tional water wave system that assumes a polynomial expansionfor the horizontal velocity,

u ¼XN

n¼0

lbn unðx; y; tÞfnðqÞ; where; q ¼ zþ hðx; yÞgðx; y; tÞ þ hðx; yÞ : ð2:2Þ

and N is the approximation order, which must be even. The dimen-sionless wavenumber, l ¼ k0h0, describes how far the system isfrom the shallow water condition. Scaling exponent bn gives theorder of each velocity component, where bn ¼ n when n is even;bn ¼ nþ 1 when n is odd – this is the typical Boussinesq scalingfor the order of l. The polyonomial basis functions fnðqÞ have theform fn ¼

Pnm¼0anmqm, with arbitrary real constants for polynomial

coefficients, anm, and ann – 0. It is assumed without loss of general-ity that f0 ¼ 1. The sigma-like coordinate q varies between zero atthe bed and one at the free surface. All horizontal velocities compo-nents u0;u1; . . . ;uN are independent so that higher order velocitycomponents are not defined by lower order components as in stan-dard irrotational Boussinesq theories (e.g. Peregrine, 1967). Specifi-cation of both the order of approximation, OðlNÞ, and thepolynomial coefficients, anm, will define the specific systems oncesubstituted into the mass and momentum equations. In particular,different choices of anm will yield different wave properties throughasymptotic rearrangement which is similar to changes in propertiesby using different reference velocities in Nwogu (1993). All systemswill converge with an increasing order of approximation, but havedifferent properties with different choice of basis functions.

Page 3: Rotational surf zone modeling for Boussinesq–Greenâ

Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53 45

The vertical velocity, w, is then uniquely specified from inte-grating the continuity equation with bottom boundary conditionas

w¼XN

n¼0

lbn �ðr�unÞðhþgÞgnþ un �rðhþgÞð Þrn�ðun �rhÞfn½ � ð2:3Þ

where gn and rn are integral functions of fn, e.g. gn �R q

0 fnðq0Þdq0,with many other functions defined (Appendix B). All integrals haveconstant of integration defined such that gnjq¼0 ¼ 0 and thusgnj

q¼10 ¼ gnjq¼1.To arrive at a final system of equations, the velocity expansions,

and viscous/turbulent stress approximations are inserted into themass and momentum equations and terms are kept or discardedaccording to the assumed order of approximation, OðlNÞ. This isidentical to the procedure in Zhang et al. (2013), but retaining tur-bulent stresses, which must be specified through a turbulenceclosure.

The vertically integrated mass conservation equation is

@g@tþr �

Z g

�hudz ¼ 0: ð2:4Þ

which, after substitution of specific fn, gives Eqs. (A.1), (A.3) inAppendix A for Oðl2Þ and Oðl4Þ, respectively.

The momentum equations are then integrated in a weightedresidual sense using the N þ 1 basis functions from the bed tothe free surface:Z g

�hfm

@u@tþ u � ruþw

@u@zþrP � l2r � sxx �

@

@zszx

� �dz ¼ 0;

m ¼ 0;1; . . . ;N: ð2:5Þ

The nonhydrostatic pressure is found by integrating the dimen-sionless vertical momentum equation from z to g assuming a zerogauge pressure at the free surface

PðzÞ ¼ l2Z g

z

@w@t

dzþ l2Z g

zu � rwdzþ l2

Z g

zw@w@z

dz

� l2Z g

zr � szx þ

@

@zszz

� �dzþ gðg� zÞ ð2:6Þ

After substitution and integration, these give the coupledmomentum Eqs. (A.2) and (A.4) in Appendix A for Oðl2Þ andOðl4Þ, respectively. The systems resemble coupled lower orderBoussinesq equations. Unlike standard Boussinesq expansions,rotational processes are included and evolve naturally once turbu-lent stresses sij are specified. This involves determination of thebreaking terms and bottom friction through models for turbulentviscosity and its evolution, subject to scaling and complexitylimitations.

2.2. Breaking terms and bottom friction

The breaking model developed here aims at a state intermediatebetween Boussinesq models with eddy viscosity breaking termsinserted after the model derivation and calibrated to give reason-able wave heights in the surf zone (Kennedy et al., 2000a; Lynettet al., 2002), and more sophisticated computational fluid dynamicsmodels with complex turbulence closures. The goals of the presentmodel are to provide accurate wave heights, water levels, andinternal water velocities in the surf zone while retaining a reason-able computational cost and implementational complexity withinthe Boussinesq–Green–Naghdi framework. The system will notattempt to simulate detailed turbulence properties, or to representthe bottom boundary layer. The basic framework of the presentmodel will use eddy viscosity with a k� l turbulence model, withmixing length l a function of water depth.

Significant simplifications must be imposed to arrive at a trac-table model:

� Turbulent terms in the pressure Eq. (2.6) are neglected,� Bottom stresses and wave breaking will be treated as separate

terms, with different evolution equations, and acting on differ-ent portions of the system,� Only leading order vertical velocity terms are retained.

This division between breaking and bottom friction has a physicalbasis, as bottom stresses are bed-generated and diffuse upwardswhile wave breaking arises from surface processes and diffusesdownwards (e.g. Richard and Gavrilyuk, 2012). In the wave roller,breaking stresses will dwarf the much smaller bottom-generatedshear stresses, while breaking stresses are small outside the rollerwhere the bottom stresses dominate. The neglect of turbulentstresses in the pressure equation, while making the system moretractable, also has a physical basis: turbulent (or any) pressuresgenerate only normal stresses and thus will not contribute to thevorticity generation leading to depth-varying currents. Together,these simplifications will reduce the generality of the system andmean that details of pressures and turbulent stresses will be moresimplified than reality; however, the single-valued free surface inthis and many other models means that representation of plungingjets, other free surface complexities such as surface turbulence, andtheir effects will always have an upper limit on accuracy in the surfzone. The additional approximations will allow the ability to modelimportant aspects of both current-related shear and roller-relateddissipation in the surf zone, which are the two most importantaspects desired.

Because breaking dissipation and bottom stress effects are sep-arated, they will be modeled with separate eddy viscosities,mt1ðx; tÞ, and mt2ðx; z; tÞ, respectively. Applying these assumptionsto the eddy viscosity approximation, and assuming that eddy vis-cosity scaling is OðmtÞ ¼ Oðlh0ðgh0Þ

1=2Þ, the respective turbulentstress terms become

l2r � sxx ¼ l2r � mt1 ruþ ðruÞT� �h i

ð2:7Þ

@

@zsxz ¼

@

@zmt2 uz þ l2rw� ��

ð2:8Þ

In order to include a specified bottom friction, the bottomstress-generated term in Eq. (2.5) is integrated by parts,Z g

�hfm@sxz

@zdz ¼

Z g

�h

@

@zðfmsxzÞdz�

Z g

�h

@fm

@zsxzdz

¼ ðfmsxzÞjg�h �Z g

�h

@fm

@zmt2ðx; zÞ uz þ l2rw

� �dz: ð2:9Þ

If there is no air–water shear stress, then sxzðgÞ ¼ 0. For bottom fric-tional dissipation, some sort of information is required based onbottom conditions, whether from bed roughness or vegetation type.These are often placed into a drag framework such assxzð�hÞ ¼ Cf ubjubj, where ub is the near-bottom velocity, and Cf isa bottom friction coefficient normally in the order of 10�3, andthis approach will be adopted. A bottom stress coefficient andgradient also imply a vertical profile for the eddy viscosity,mt2ðx; zÞ ¼ ~mt2ðxÞfmðqÞ. Combining eddy viscosity and bottomfriction, the depth-varying eddy viscosity will then be mt2ðx; zÞ ¼�fmCf ðgþ hÞjubj, where � is a constant applied to control the velocitydifference between surface and bed, and the influence of the bedvelocity (or alternatively mean velocity) on eddy viscosity arisesfrom scaling considerations. The simplest application will be aconstant vertical profile fmðqÞ ¼ 1 (e.g. Kim and Lynett, 2013),which will use � � 0:4, but other vertical profiles may also beemployed.

Page 4: Rotational surf zone modeling for Boussinesq–Greenâ

46 Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53

Using these vertical variations for mt1 and mt2 and performingintegrations over depth, the turbulent stresses retained are then

Z g

�hfm@sxz

@zdz¼�fmjq¼0Cf ubjubj�

Z 1

0f 0m mt2u;q=ðgþhÞþl2mt2rw� �

dq

¼�fmjq¼0Cf ubjubj��Cf jubjXN

n¼0

lbn unwmnmj10

þl2�Cf jubjðgþhÞ rðr�u0ÞðgþhÞemmj10h

þ r�u0rhþrðu0 �rhÞð Þfmmj10i

ð2:10Þ

andZ g

�hl2fmr�sxxdz¼

Z g

�hl2fmr� mt ruþðruÞT

� �� �dz

¼l2r� ðgþhÞZ 1

0fmmt1ðruþðruÞT Þdq

�l2ðgþhÞZ 1

0mt1ðrfmÞ � ðruþðruÞTÞdq

�l2fmmt1ðrgÞ � ðruþðruÞTÞjq¼1

�l2fmmt1ðrhÞ � ðruþðruÞTÞjq¼0

¼XN

n¼0

lbnþ2r� mt1ðgþhÞðrunþðrunÞTÞ/mnj10

h

þmt1ðuTnrTþðuT

nrTÞT Þ hfmnj10�ðgþhÞemnj10� �i

�XN

n¼0

lbnþ2mt1 ðfnmj10rh�enmj10rðgþhÞÞh

�ðrunþðrunÞTÞþðEmnj10rh� Jmnj10rðgþhÞÞ

� ðuTnrT hþðuT

nrT ÞT hÞ=ðgþhÞ�ðJmnj10rh�Kmnj10rðgþhÞÞ

�ðuTnrT ðgþhÞþðuT

nrTÞT ðgþhÞÞ=ðgþhÞi

þXN

n¼0

lbnþ2mt1ðfmrqÞj10 � ðrunþðrunÞTÞðgþhÞfnj10h

þðuTnrTþðuT

nrTÞT Þðh�ðgþhÞqj10Þf 0nj10

ið2:11Þ

2.3. k� l Turbulent-kinetic-energy model for mt1

The final stage of the breaking model is the evolution equationfor eddy viscosity, mt1. This arises directly from representations forturbulent kinetic energy evolution combined with assumptionsand simplifications relevant to the structure of the Boussinesq–Green–Naghdi system and eddy viscosity. The k� l model is argu-ably the simplest incomplete turbulence model, and hence it has abroad range of applicability with a low computational cost (e.g.Pope, 2003). Similarly to k� � model, k� l is a favorable numericalcoupling between the flow and turbulence equations with the

term, mt@<Ui>

@xjþ @<Uj>

@xi

� �. This actually allows us to get better velocity

profile under breaking and comparable computational efficiencycompared to most other Boussinesq-type model. The process ofproduction, transport and dissipation equation for the turbulentkinetic energy k can be expressed as Pope (2003)

DkDt� �r � T0 þ P � e: ð2:12Þ

Here, the turbulent energy flux T0 is modeled with a gradient-diffu-sion hypothesis as

T0 ¼ � mt1

rkrk: ð2:13Þ

where rk is the turbulent Prandtl number for kinetic energy, and isgenerally taken to be 1:0. According the turbulent-viscosity hypoth-

esis introduced by Boussinesq (1877), the production of turbulentkinetic energy is then

P ¼ mt1½ru � ðruþ ðruÞTÞ þ 2u;z � rwþ 1l2 u;z � u;z þ 2w2

;z�:

ð2:14Þ

where Oðl2Þ terms are neglected. The turbulent viscosity is definedby Pope (2003)

mt1 ¼ ck1=2�lm: ð2:15Þ

where �lm is the vertically averaged mixing length, lm. At highReynolds number the dissipation rate is modeled ase ¼ Cek

3=2=�lm: ð2:16Þ

where Ce ¼ c3 in k� l model. Nezu and Rodi (1986), based on exper-imental measurements made with stationary currents, concludedthat the length scale tends to zero near the water surface. Theincrease of the mixing length and its reduction near the free surfaceis given by

lm ¼ jqffiffiffiffiffiffiffiffiffiffiffiffi1� q

pðgþ hÞ ð2:17Þ

�lm ¼Z 1

0lmdq ¼ 4=15jðgþ hÞ ð2:18Þ

where j ¼ 0:412 (von Karman constant).After inserting expressions for T 0; P; � into (2.12) and substi-

tuting mt1 for k, the dimensionless TKE approximation equationtransforms to

Dmt1

Dt¼ l

mt1rkr � ðm2

t1rmt1Þ þ l c2�l2m

2mt1P � 1

lc2

2�l2m

m2t1 ð2:19Þ

A value of c ¼ 0:55 yields the correct behavior for shear flow in k� lmodels, and this is used here. As assumed above, kinematic viscos-ity is depth-uniform. Integrating (2.19) from bottom to the surfacethen gives

@mt1

@tþrmt1 �

XN

n¼0

ungnj10 �

lmt1rk

r � ðm2t1rmt1Þ � l c2�l2

m

2mt1

Z 1

0Pdq

þ 1l

c2

2�l2m

m2t1 ¼ 0: ð2:20Þ

whereZ 1

0Pdq¼mt1

XN

n¼0

XN

m¼0

lbmþbn rum � ðrunþðrunÞTÞ/mnj10

h(

þrum � ðunþðunÞTÞðrhfmn�ðgþhÞemnÞj10� �

þumðrhfmn�ðgþhÞemnÞj10 � ðrunþðrunÞTÞ

þZ 1

0umrfm � ðunrfnþðunrfnÞTÞdq

��XN

n¼0

lbn 2un � ½rðr�u0Þrnj10

þ r�u0rhþrðu0 �rhÞð Þ=ðgþhÞfnj10�þXN

n¼0

XN

m¼0

lbmþbn�2um

�un=ðgþhÞ2Emnj10þ2 ðrðr�u0ÞðgþhÞÞ2=3 r�u0rhð�

þrðu0 �rhÞÞ2þ2rðr�u0ÞðgþhÞ r�u0rhþrðu0 �rhÞð ÞÞg:ð2:21Þ

This one equation model for the evolution of eddy viscosity is usedto find depth-averaged values to use in the velocity calculations forthe Boussinesq–Green–Naghdi rotational wave model.

3. Moving shoreline

For computations of wave runup, a moving shoreline method isdeveloped here. Benque et al. (1982) proposed a way to treat the

Page 5: Rotational surf zone modeling for Boussinesq–Greenâ

0 5 10 15 20 25 30−0.4

−0.2

0

0.2

z(m

)x(m)

abcdef g

wavemaker

Fig. 1. Experimental setup for regular wave breaking test. Gauges (a–g) are at still water depths [0.169, 0.156, 0.142, 0.128, 0.113, 0.096, 0.079] m, respectively.

Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53 47

moving boundary by assuming a thin water layer over the dry landregion so that all grids in the computational domain are alwayswet so that there is no need to update the wet computationalboundary during the computation. The method we are using issimilar but a bit different in the governing equations for the ‘‘thinlayer’’. Like most moving shoreline methods, a threshold, d, todetermine the dividing point between wet and dry area needs tobe set. If the total water depth hþ g > d, the model Eqs. (2.4)and (2.5) will be applied; otherwise, a new set of equations willbe adopted as following,

g ¼ �hþ d; ð3:1Þ@u@t¼ xð0� uÞ: ð3:2Þ

These equations set a thin layer over the ‘‘dry area’’ (hþ g <¼ d)allowing the water to propagate up to the shoreline but preventingthe thin layer water to flow back down by setting a damping coef-ficient x for the horizontal velocity, where x ¼ Cðx� xðdÞÞ. Sounlike Benque’s method, we do need to update the threshold posi-tion the at each computing time step and the total mass is betterconserved.

4. Numerical verification

Here we present three numerical tests to validate the perfor-mance of the breaking model, all for one horizontal dimension.

15 20 25 30−0.1

00.10.2

η(m

)

h=0.169 m

15 20 25 30−0.1

00.10.2

η(m

)

h=0.142 m

15 20 25 30−0.1

00.10.2

time(s)

η(m

)

h=.096 m

Fig. 2. Surface elevation decay for Plunging Breaking. ( ) Oðl

The first test examines regular wave breaking on plane slopes asreported by Ting and Kirby (1995, 1996) for plunging and spillingcases, and compares both surface elevations and interior fluidvelocities. The second test compares computed and measuredirregular wave evolution over the conditions of Mase and Kirby(1992). The final test examines solitary wave runup on a beach(Synolakis, 1987).

All numerical tests used a weakly nonlinear Oðl4Þ Boussinesq–Green–Naghdi rotational system (Zhang et al., 2013) with secondorder spatial central differencing scheme and fourth orderRunge–Kutta time differencing. For all quantities in the Ting andKirby experiments, results were also computed using a fullynonlinear Oðl2Þ system to compare with the Oðl4Þ data. A spatialresolution of x = 0.025 m and t = 0.02 s was used for all tests shownhere.

4.1. Wave breaking

Fig. 1 shows the experimental setup where regular waves weregenerated in a water depth of 0.4 m using second order generation(Zhang et al., 2014a) and propagated to the 1:35 sloping beach.Two cases were studied.

1. Spilling breaker, H = 12.5 cm; T = 2 s, kh ¼ 0:680,2. Plunging breaker, H = 12.8 cm; T = 5 s, kh ¼ 0:257.

15 20 25 30−0.1

00.10.2

η(m

)

h=0.156 m

15 20 25 30−0.1

00.10.2

η(m

)

h=0.113 m

15 20 25 30−0.1

00.10.2

time(s)

η(m

)

h=0.079 m

4Þ; ( ) Oðl2Þ; (–) measured. Some lines are obscured.

Page 6: Rotational surf zone modeling for Boussinesq–Greenâ

0.1 0.2 0.3

0

10

20x 10−3

h (m)

Setu

p (m

)

0.1 0.2 0.30.05

0.10.15

0.2

h (m)

Hei

ght (

m)

Fig. 3. Wave heights and setup for Plunging Breaking. ( ) Oðl4Þ; ( ) Oðl2Þ; (o)measured.

0.1 0.2 0.3 0.4

0

10

20x 10−3

h (m)

Setu

p (m

)

0.1 0.2 0.3 0.4

0.1

0.2

h (m)

Hei

ght (

m)

Fig. 4. Wave heights and setup for Spilling Breaking. ( ) Oðl4Þ; ( ) Oðl2Þcomputed; (o) measured.

0 0.5 1−0.4

−0.2

0

0.2

0.4

z=−0.04m

0 0.5 1−0.4

−0.2

0

0.2

0.4

time (t/T)

z=−0.06m

Fig. 5. Time-varying horizontal velocity for plunging breaker at different points in the wmeasured.

48 Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53

The wave height-depth ratio in the constant-depth region waslarge (H=h > 0:3) and the breaking was intense. Fig. 2 showsmeasured and computed time series of surface elevation at differ-ent depths for the plunging breaker case, with good agreementseen.

Figs. 3 and 4 show numerical and measured results for waveheight and setup from outside the breaking region through to theshoreline, with good agreement for both quantities. Wave heightsincreased slowly before breaking and then rapidly decayedthrough the surf zone. Setup increased strongly after breaking asexpected, with good agreement between computed and measuredvalues. For both the spilling and plunging breakers, the presentsystem does not model the very sharp decrease in wave height atinitial breaking, and shows a more gradual decline. This is likelybecause of the time needed to build up turbulent kinetic energyusing the present breaking model, which acts on velocity gradientsand does not model the initial overturning. However, overall agree-ment is good and is comparable to other systems (Tissier et al.,2012). In somewhat of a surprise, wave heights and setups arealmost identical for Oðl2Þ and Oðl4Þ systems. This suggeststhat the breaking model dominates evolution in the surf zone,and the specific equations used are of much less importance,at least for the cases considered. Because both test cases haverelatively small wavenumbers even in the initial depth(kh ¼ 0:680; kh ¼ 0:257), both Oðl2Þ and Oðl4Þ systems shouldhave similar properties for these cases, at least to second order(Zhang et al., 2013). Thus it seems that Oðl4Þ systems may notalways provide much of an improvement in the computation ofsurface elevations in the surf zone. However, as will be seen verysoon, interior velocities and undertow benefit significantly fromthe increase from Oðl2Þ to Oðl4Þ systems.

One of the major benefits of Boussinesq–Green–Naghdi systemsis the removal of the irrotationality assumption - thus we expect tosimulate well the interior water velocities well under breaking.Here we choose two numerical results for orbital velocities:time-varying velocities at different levels, and time-averagedundertow. Fig. 5 compares the time series of horizontal velocityto the data reported by Ting and Kirby (1994) for the plunging case.The location was at water depth h ¼ 0:079 m where the breaking

0 0.5 1−0.4

−0.2

0

0.2

0.4

z=−0.05m

0 0.5 1−0.4

−0.2

0

0.2

0.4

time (t/T)

z=−0.0665m

ater column. h=hb ¼ 0:584; h ¼ 0:079m. ( ) computed ul4 ; ( ) computed ul2 ; (–)

Page 7: Rotational surf zone modeling for Boussinesq–Greenâ

−0.1 0.1−0.15

−0.1

−0.05

0 a

−0.1 0.1−0.15

−0.1

−0.05

0 b

−0.1 0.1−0.15

−0.1

−0.05

0 c

−0.1 0.1−0.15

−0.1

−0.05

0 d

−0.1 0.1−0.15

−0.1

−0.05

0 e

−0.1 0.1−0.15

−0.1

−0.05

0 f

−0.1 0.1−0.15

−0.1

−0.05

0 g

z(m)

u

Fig. 6. Variation of dimensionless undertow velocity u=ffiffiffiffiffiffigh

pwith depth for plunging breaking. ( ) Oðl4Þ; ( ) Oðl2Þ computed; (o) measured.

Table 1Dimensionless, depth-averaged, root-mean-square error in undertow profiles

ðuln �uÞ2

gh

� �1=2

for plunging breaker using Oðl2Þ and Oðl4Þ systems.

System Still Water Depth

0.169 m 0.156 m 0.142 m 0.128 m 0.113 m 0.096 m 0.079 m

Oðl2Þ 0.0172 0.0157 0.0152 0.0279 0.0213 0.0289 0.0294

Oðl4Þ 0.0093 0.0104 0.0133 0.0153 0.0167 0.0130 0.0123

Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53 49

was very strong. Four measurement points were vertically distrib-uted and numerical results were compared to measured values.The magnitude of horizontal velocity decreased with increasingsubmergence, as expected and showed a sharp increase on thewave front and more gentle decrease on the back side. HigherOðl4Þ results showed good agreement with measured values atall depths for both magnitude and phase. Small differences werenoticeable near the times of maximum velocity, where details ofthe plunging jet would likely have been important and are notmodeled in the present system. In contrast, the Oðl2Þ systemtended to overpredict velocities at both the crest and troughalthough overall agreement was still reasonable. A likely reasonfor this is that the quadratic velocity profile was unable to resolvethe complexity of velocities in the breaking region.

Fig. 6 shows the variation of time-averaged horizontal velocity(undertow) with depth at all measurement locations for the plung-ing breaker. Negative undertow velocities near the bottomregion can be seen for both sites while the mean surface velocity

0 0.5 10

0.01

0.02

0.03

0.04

k/(g

h)

z=−0.04m

0 0.5 10

0.01

0.02

0.03

0.04

time (t/T)

k/(g

h)

z=−0.06m

Fig. 7. Time-varying turbulent kinetic energy for plunging breaker at different pointscomputed; (–) measured.

is positive. The Oðl4Þ system shows reasonable agreement,demonstrating that surf zone undertow can be modeled with thisrelatively simple breaking model, and giving good confidence mov-ing forward for more complex three dimensional topographies.However, there are some differences for near-bed velocities in dee-per depths, where a small secondary recurvature is observed incomputations, but not in the data. This is likely a function of theq4 vertical variation in velocity profiles, which is less complex thanshown by the actual data. When decreasing order of approximationto Oðl2Þ, agreement with data declines. Again, this is likely a func-tion of the limited z2 velocity profile for the Oðl2Þ model. Table 1shows depth-averaged, dimensionless undertow error for boththe Oðl2Þ and Oðl4Þ systems, demonstrating the significantincrease in accuracy for the higher order system. When combinedwith the surface elevation and orbital velocity results, it appearsthat a higher order model will provide significant improvementsto interior velocities, but is not as important in improving surfaceelevation results in the surf zone.

0 0.5 10

0.01

0.02

0.03

0.04

k/(g

h)

z=−0.05m

0 0.5 10

0.01

0.02

0.03

0.04

time (t/T)

k/(g

h)

z=−0.0665m

in the water column. h=hb ¼ 0:584; h ¼ 0:079m. ( ) Oðl4Þ computed; ( ) Oðl2Þ

Page 8: Rotational surf zone modeling for Boussinesq–Greenâ

0 5 10 15 20−0.47

−0.4−0.3−0.2−0.1

00.10.2

z(m

)

x(m)

a b c d e f g h i j k

still water level

wavemaker

Fig. 8. Definition sketch for Mase and Kirby irregular wave tests. Depths for (a–k) are [0.4 0.35 0.3 0.2 0.175 0.15, 0.125, 0.075, 0.025] m.

50 Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53

Because the breaking system developed and implementedhere provides direct estimates of the turbulent kinetic energy,k, it becomes possible to compare with experimental values atdifferent points in the water column as shown in Fig. 7. Theseresults show predictions that match well the general behaviorof the experimental data, with both Oðl2Þ and Oðl4Þ systems giv-ing similar results. However, there appear non-negligible over-predictions of k at the times of maximum breaking, and overallagreement is not as good as with surface elevations or interiorvelocities. This increased error is to be expected to some degree,since higher order quantities like turbulent kinetic energy andshear stress are more difficult to model than interior velocitiesor surface elevations (Pope, 2003). Given the relatively simplemixing length model implemented here, agreement seems

15 20−0.04−0.02

00.020.04

η(m

)

h

15 20−0.04−0.02

00.020.04

η(m

)

h

15 20−0.04−0.02

00.020.04

η(m

)

h=

15 20−0.04−0.02

00.020.04

η(m

)

h=

15 20−0.04−0.02

00.020.04

η(m

)

h=

15 20−0.04−0.02

00.020.04

ti

η(m

)

h=

Fig. 9. Time series of surface elevation at different

reasonable but it remains likely that a more complete turbulencemodel would be able to provide non-negligible improvements inhigher order turbulent quantities. (See Fig. 7)

4.2. Irregular wave breaking

For a test of irregular waves the experiments of Mase and Kirby(1992) were modeled. The irregular waves were generated in47 cm-depth water transitioning to a 1:20 planar slope to theshoreline as described in Fig. 8. Fig. 9 shows a truncated time seriesof computed and measured surface elevations at 6 gauges. Thecomputed result almost overtops the data for the 5 deepest gauges.However, as waves propagate near the shore, the computedelevation goes off a bit but the agreement is still good, and

25 30

=.4 m

25 30

=.2 m

25 30

.125 m

25 30

.075 m

25 30

.05 m

25 30me(s)

.025 m

gauge locations. (- -) computed; (–) measured.

Page 9: Rotational surf zone modeling for Boussinesq–Greenâ

0.05 0.1 0.15 0.2 0.25 0.30

0.5

1

1.5

h (m)

−Asy

mm

etry

0.05 0.1 0.15 0.2 0.25 0.30

0.5

1

1.5

h (m)

Skew

ness

Fig. 11. Asymmetry and skewness for Mase–Kirby irregular wave tests. (o)measured; (-⁄-) computed.

−20 −10 0 10−0.2

00.20.4 t=10

η(m

)

−20 −10 0 10−0.2

00.20.4 t=20

η(m

)

−20 −10 0 10−0.2

00.20.4 t=45

x(m)

η(m

)

Fig. 12. Wave runup on a sloping bea

0.05 0.1 0.15 0.2 0.25 0.3 0.35−2

0

2x 10−3

h (m)

Setu

p (m

)

0.05 0.1 0.15 0.2 0.25 0.3 0.350

0.01

0.02

h (m)

Hei

ght (

m)

Fig. 10. Setup and wave height for Mase–Kirby irregular wave tests. (o) measured;(-) computed).

Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53 51

Fig. 10 shows a good match for wave setup and wave height. Otherstatistical parameters can tell how well the model performs non-linearly. Wave asymmetry is a measure of left–right differencesin a wave. Skewness is a measure of crest-trough shape. They areboth computed from time series of surface elevation at chosenlocations (Kennedy et al., 2000a). The computed results are com-pared with the data in Fig. 11. The trend is caught well from mod-erate depth through breaking to the shore. The wave becomesmore asymmetric as propagating to the shore. The skewnessincreases as the wave shoals and breaks, and then decreases nearthe shoreline.

4.3. Solitary wave runup

As a check of the moving boundary algorithm, a runup test for abreaking solitary wave was simulated for a beach slope of 1=19:85.The computed water surface runup for an initial solitary wave withwave height to still water depth ratio of H=h ¼ 0:3 was comparedto the experimental data by Synolakis (1987). Fig. 12 shows thewater surface compared to the data during shoaling, breaking,runup and rundown. Panels 1, 2 and 3 show the wave becomingasymmetric followed by breaking in the surf zone. Panel 4(t ¼ 25) shows the breaking wave run up onto the beach andreaching near the highest point in Panel 5 (t ¼ 45), while Panel 6(t ¼ 50) shows the rundown. Numerical results show good agree-ment with the data, providing a verification of the overall systemfor breaking and runup.

5. Discussion and conclusions

This paper is an extension and partial verification of the Bous-sinesq–Green–Naghdi rotational water wave model (Zhang et al.,2013). Due to the removal of the irrotationality assumption andthe use of standard turbulence models, fewer ad hoc breakingassumptions have been made than many other eddy viscosityBoussinesq-type breaking models. Bottom friction and even sur-face wind shear stress can be added into the model by applyingthe boundary conditions when integrating the vertical force termin the momentum equation. Using simplifications to the well-known k� l turbulence model, surf zone breaking phenomenasuch as wave height decay, setup, undertow velocities, and wave

−20 −10 0 10−0.2

00.20.4 t=15

η(m

)

−20 −10 0 10−0.2

00.20.4 t=25

η(m

)

−20 −10 0 10−0.2

00.20.4 t=50

x(m)

η(m

)

ch. (�) measured; (–) computed.

Page 10: Rotational surf zone modeling for Boussinesq–Greenâ

52 Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53

runup are shown to be modeled well. One area of significant pro-gress compared to some other Boussinesq surf zone models isthe good agreement with the undertow velocity data which furthervalidates the rotational feature of our model.

Both Oðl2Þ and Oðl4Þ systems were tested against each otherand against experimental data. Results showed that the lowerorder systems performed very similarly to the higher order sys-tems in computing wave heights and mean setup. In contrast,the more complex velocity profile found in the Oðl4Þ systemallowed for a significantly improved representation of both time-varying and time-averaged interior velocities. However there isstill room for improvement. The mixing length profile used forbreaking is highly approximate and might not represent the actualmixing length with high accuracy. This affects many quantitiesincluding the turbulent kinetic energy, k, which was not predictedas well as either surface elevations or interior velocities. A morecomplete turbulence model may be warranted for the future. Thesingle-value surface assumption will additionally impose an upperlimit on surf zone accuracy. However, the systems describe theoverall breaking water surface, wave shape, rotational orbitalvelocities, and moving shoreline for wave runup reasonably well.

Although the tests herein provided a glimpse of the improvedrotational capabilities of the model, further verification for twohorizontal dimensions is necessary. This work is ongoing and willexamine performance for breaking-driven flows where the lackof any irrotationality condition is expected to be important in mod-eling the interior velocities, and applications to geophysical andengineering problems.

Acknowledgments

This work was funded under the National Science FoundationGrant 1025519, and by the Office of Naval Research under theawards N00014-11-1-0045 and N00014-13-1-0123. Their supportis gratefully acknowledged.

Appendix A. Expressions for depth- integrated Oðl2Þ; Oðl4Þsystems

The vertically integrated mass conservation equation is, toOðl2Þ

@g@tþr �

Z g

�hudz ¼ g;t þr �

X2

n¼1

lbn unðgþ hÞgnjq¼1

!¼ 0: ðA:1Þ

where gn is defined in Appendix B.Weighted momentum equation keeping all terms to Oðl2Þ are

u0;tðgþ hÞgmjq¼1 þ u0 � ru0ðgþ hÞgmjq¼1 þ grgðgþ hÞgmjq¼1

þ l2X2

n¼1

un;tðgþ hÞ/mn � ung;temn

� �jq¼1 � l2

12rðr � u0;tÞðgþ hÞ3

ðgm � mmÞ þ ðr � u0;tÞðgþ hÞ2rðgþ hÞgm þrðu0;t � rhÞðgþ hÞ2

ðgm � SmÞ þ u0;t � rhrgðgþ hÞgm � ðr � u0;tÞðgþ hÞ2rhSm�jq¼1

þ l2X2

n¼1

ðun � ru0 þ u0 � runÞðgþ hÞ/mn � unr � ðu0ðgþ hÞÞemn½ �jq¼1

þ l2ðgþ hÞ2½ðr � u0Þ2 � u0 � rðr � u0Þ�ðrggm þrhðgm � SmÞÞjq¼1

þ l2

2ðgþ hÞ3r½ðr � u0Þ2 � u0 � rðr � u0Þ

�ðgm � mmÞjq¼1 � l2ðgþ hÞ

rgu0 � rðu0 � rhÞgmjq¼1 � l2ðgþ hÞ2rðu0 � rðu0 � rhÞÞðgm � SmÞjq¼1

�Z g

�hfm@sxz

@zdz�

Z g

�hl2fmr � sxxdz ¼ 0; m ¼ 0;1;2: ðA:2Þ

The vertically integrated mass conservation equation is, to Oðl4Þ

@g@tþr �

Z g

�hudz ¼ g;t þr �

X4

n¼1

lbn unðgþ hÞgnjq¼1

!¼ 0: ðA:3Þ

The integrated conservation of momentum equations for Oðl4Þ levelare

u0;tðgþ hÞ þ u0 � ru0ðgþ hÞ þ grgðgþ hÞð Þgmjq¼1 þ l2 . . .ð Þ

þXN

n¼3

lbn hun;t/mnjq¼1 �XN�2

n¼1

lbn ½hrðh2r � un;tÞðGnjq¼1gmjq¼1

�Cmnjq¼1Þ � h2r � un;trhðcmnjq¼1 � hmnjq¼1Þ þ hrðhðun;t � rhÞÞ ððgn � RnÞjq¼1gmjq¼1 � cmnjq¼1 þHmnjq¼1Þ þ hðun;t � rhÞrh

ðqmnjq¼1 � Fmnjq¼1 � /mnjq¼1 þWmnjq¼1Þ� �Z g

�hfm@sxz

@zdz

�Z g

�hl2fmr � sxxdz ¼ 0; m ¼ 0;1;2;3;4: ðA:4Þ

For both Oðl2Þ and Oðl4Þ systems, depth-integrated stress terms aregiven by (2.10) and (2.11).

Appendix B. Integral functions

All indefinite integrals will be assumed to have integration con-stants defined to give values of 0 at q ¼ 0. Thus, for example,gnj

q¼10 ¼ gnjq¼1.

gn¼R

fndq rn¼R

f 0nqdq Gn¼R

gndq wmnm¼R

f 0mf 0nfmdq

Rn¼R

rndq /mn¼R

fmfndq cmn¼R

fmgndq fmn¼R

fmf 0ndq

qmn¼R

fmrndq Cmn¼R

fmGndq Hmn¼R

fmRndq Emn¼R

f 0mf 0ndq

hmn¼R

fmgnqdq mm¼R

q2fmdq Sm¼R

qfmdq Kmn¼R

f 0mf 0nq2dq

emn¼R

fmf 0nqdq Wmn¼R

fmfnqdq Fmn¼R

fmrnqdq Jmn¼R

f 0mf 0nqdq

References

Benque, J.P., Cunge, J.A., Hauguel, J.F., Holly, F.M.A., 1982. New method for tidalcurrent circulation. J. Waterway Port Coastal Ocean Div.-ASCE 108, 396–417.

Bonneton, P., Barthelemy, E., Chazel, F., Cienfuegos, R., Lannes, D., Marche, F., Tissier,M., 2011. Recent advances in Serre–Green Naghdi modelling for wavetransformation, breaking, and runup. Eur. J. Mech. B-Fluids 30, 589–597.

Boussinesq, J., 1877. Essai sur la théorie des eaux courantes. Mémoires présentéspar divers savants à l’Académie des Sciences XXIII, 1–680.

Chen, Q., Kirby, J.T., Dalrymple, R.A., Kennedy, A.B., Chawla, A., 2000. Boussinesqmodeling of wave transformation, breaking, and runup. II: 2D. J. Waterway PortCoastal Ocean Eng. 126, 48–56.

Cienfuegos, R., Barthelemy, E., Bonneton, P., 2010. Wave-breaking model forBoussinesq-type equations including roller effects in the mass conservationequation. J. Waterway Port Coastal Ocean Eng. 136, 10–26.

Ertekin, R.C., Webster, W.C., Wehausen, J.V., 1986. Waves caused by a movingdisturbance in a shallow channel of finite width. J. Fluid Mech 169, 275–292.

Gallerano, F., Cannata, G., Villani, M., 2014. An integral contravariant formulation ofthe fully non-linear Boussinesq equations. Coastal Eng. 83, 119–136.

Gobbi, M.F., Kirby, J.T., 1999. Wave evolution over submerged sills: tests of a high-order Boussinesq model. Coastal Eng. 37, 57–96, and erratum, Coastal Eng. 40,277.

Gobbi, M.F.G., Kirby, J.T., Wei, G., 2000. A fully nonlinear Boussinesq model forsurface waves. II. Extension to Oðl4Þ. J. Fluid Mech. 405, 181–210.

Green, A.E., Naghdi, P.M., 1976. A derivation of equations for wave propagation inwater of variable depth. J. Fluid Mech. 78, 237–246.

Higuera, P., Lara, J.L., Losada, I.J., 2013. Simulating coastal engineering processeswith OpenFOAM. Coastal Eng. 71, 119–134.

Kennedy, A.B., Chen, Q., Kirby, J.T., Dalrymple, R.A., 2000a. Boussinesq modeling ofwave transformation, breaking, and runup. I: 1D. J. Waterway Port CoastalOcean Eng. 126, 39–47.

Kennedy, A.B., Dalrymple, R.A., Kirby, J.T., Chen, Q., 2000b. Determination of inversedepths using direct Boussinesq modeling. J. Waterway Port Coastal Ocean Eng.126, 206–214.

Kennedy, A.B., Kirby, J.T., Chen, Q., Dalrymple, R.A., 2001. Boussinesq-type equationswith improved nonlinear performance. Wave Motion 33, 225–243.

Kim, D.-H., Lynett, P.J., 2013. A sigma-coordinate transport model coupled withrotational Boussinesq-type equations. Environ. Fluid Mech. 13, 51–72.

Kim, J.W., Bai, K.J., Ertekin, R.C., Webster, W.C., 2003. A strongly-nonlinear model forwater waves in water of variable depth – the irrotational Green–Naghdi model.J. Offshore Mech. Arctic Eng. 125, 25–32.

Page 11: Rotational surf zone modeling for Boussinesq–Greenâ

Y. Zhang et al. / Ocean Modelling 79 (2014) 43–53 53

Klonaris, G.Th., Memos, C.D., Karambas, Th.V., 2013. A Boussinesq-type modelincluding wave-breaking terms in both continuity and momentum equations.Ocean Modell. 57, 128–140.

Lynett, P.J., Liu, P.L.F., 2002. A two-dimensional, depth-integrated model for internalwave propagation over variable bathymetry. Wave Motion 36, 221–240.

Lynett, P.J., Wu, T.R., Liu, P.L.F., 2002. Modeling wave runup with depth-integratedequations. Coastal Eng. 46, 89–107.

Madsen, P.A., Schäffer, H.A., 1998. Higher order Boussinesq-type equations forsurface gravity waves- derivation and analysis. Philos. Trans. R. Soc. A 356,1–59.

Madsen, P.A., Srensen, O.R., 1992. A new form of the Boussinesq equationswith improved linear dispersion characteristics. Part 2. A slowly varyingbathymetry. Coastal Eng. 18, 183–204.

Madsen, P., Srensen, O., Schäffer, H., 1997. Surf zone dynamics simulated by aBoussinesq-type model: Part I. Model description and cross-shore motion ofregular waves. Coastal Eng. 32, 255–288.

Ma, G., Shi, F.Y., Kirby, J.T., 2012. Schock-capturing non-hydrostatic model for fullydispersive surface wave processes. Ocean Modell. 43–44, 22–35.

Mase, H., Kirby, J.T., 1992. Hybrid frequency-domain KdV equation forrandom wave transformation. In: Proceedings of 23rd InternationalCoastal Engineering Conference,. ASCE, New York, pp. 474–487.

Musumeci, R.E., Svendsen, I.A., Veeramony, J., 2005. The flow in the surf zone: a fullynonlinear Boussinesq-type of approach. Coastal Eng. 52, 565–598.

Nezu, I., Rodi, W., 1986. Open-channel flow measurements with a laser doppleranemometer. J. Hydraulic Eng. 112 (5), 335–355.

Nwogu, O., 1993. Alternative form of Boussinesq equations for nearshore wavepropagation. J. Waterway Port Coastal Ocean Eng. 119, 618–638.

Nwogu, O., Demirbilek, Z., 2010. Infragravity wave motions and runup over shallowfringing reefs. J. Waterway, Port, Coastal and Ocean Eng. 136 (6),295–305.

Peregrine, D.H., 1967. Long waves on a beach. J. Fluid Mech. 27, 815827.Pope, S.B., 2003. Turbulent Flows. Cambridge University Press, Cambridge.Richard, G.L., Gavrilyuk, S.L., 2012. A new model of roll waves: comparison with

Brock’s experiments. J. Fluid Mech. 698, 374–405.

Schäffer, H.A., Madsen, P.A., 1993. A Boussinesq model for waves breaking inshallow water. Coastal Eng. 20, 185–202.

Shields, J.J., Webster, W.C., 1988. On direct methods in water-wave theory. J. FluidMech. 197, 171–199.

Shi, F.Y., Kirby, J.T., Harris, J.C., Geiman, J.D., Grilli, S.T., 2012. A high-order adaptivetime-stepping TVD solver for Boussinesq modeling of breaking waves andcoastal inundation. Ocean Modell. 43–44, 36–51.

Srensen, O.R., Schäffer, H.A., Madsen, P.A., 1998. Surf zone dynamics simulated by aBoussinesq type model. III Wave-induced horizontal circulations. Coastal Eng.33, 155–176.

Svendsen, I., 1984. Mass flux and undertow in a surf zone. Coastal Eng. 8, 347365.Synolakis, C.E., 1987. The run-up of solitary waves. J. Fluid Mech. 185, 523–545.Ting, F.C.K., Kirby, J.T., 1994. Observation of undertow and turbulence in a

laboratory surf zone. Coastal Eng. 24, 5180.Ting, F.C.K., Kirby, J.T., 1995. Dynamics of surf-zone turbulence in a strong plunging

breaker. Coastal Eng. 24, 177204.Ting, F.C.K., Kirby, J.T., 1996. Dynamics of surf-zone turbulence in a spilling breaker.

Coastal Eng. 27, 131160.Tissier, M., Bonneton, P., Marche, F., Chazel, F., Lannes, D., 2012. A new approach to

handle wave breaking in fully non-linear Boussinesq models. Coastal Eng. 67,54–66.

Veeramony, J., Svendsen, I.A., 2000. The flow in surf-zone waves. Coastal Eng. 39,93–122.

Wei, G., Kirby, J.T., Grilli, S.T., Subramanya, R., 1995. A fully nonlinear Boussinesqmodel for surface waves. I. Highly nonlinear, unsteady waves. J. Fluid Mech.294, 71–92.

Zelt, J.A., 1991. The runup of nonbreaking and breaking solitary waves. Coastal Eng.15, 205246.

Zhang, Y., Kennedy, A.B., Panda, N., Dawson, C., Westerink, J.J., 2013. Boussinesq–Green–Naghdi rotational water wave theory. Coastal Eng. 73, 13–27.

Zhang, Y., Kennedy, A.B., Panda, N., Dawson, C., Westerink, J.J., 2014a. Generating-absorbing sponge layers for phase-resolving wave models. Coastal Eng. 84, 1–9.

Zhao, B.B., Duan, W.Y., Ertekin, R.C., 2014b. Application of higher-level GN theory tosome wave transformation problems. Coastal Eng. 83, 177–189.