New URL for NEMO forge!   http://forge.nemo-ocean.eu

Since March 2022 along with NEMO 4.2 release, the code development moved to a self-hosted GitLab.
This present forge is now archived and remained online for history.
Changeset 1224 for trunk/DOC/TexFiles/Chapters/Chap_Model_Basics.tex – NEMO

Ignore:
Timestamp:
2008-11-26T14:52:28+01:00 (15 years ago)
Author:
gm
Message:

minor corrections in the Chapters from Steven + gm see ticket #283

File:
1 edited

Legend:

Unmodified
Added
Removed
  • trunk/DOC/TexFiles/Chapters/Chap_Model_Basics.tex

    r998 r1224  
    2727velocity, plus the following additional assumptions made from scale considerations: 
    2828 
    29 \textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to be spheres so that gravity (local vertical) is parallel to the earth's radius 
     29\textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to  
     30be spheres so that gravity (local vertical) is parallel to the earth's radius 
    3031 
    3132\textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius 
    3233 
    33 \textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect of small scale processes on the large-scale) are expressed in terms of large-scale features 
    34  
    35 \textit{(4) Boussinesq hypothesis:} density variations are neglected except in their contribution to the buoyancy force 
    36  
    37 \textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a balance between the vertical pressure gradient and the buoyancy force (this removes convective processes from  
    38 the initial Navier-Stokes equations and so convective processes must be parameterized instead) 
    39  
    40 \textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity vector is assumed to be zero. 
    41  
    42 Because the gravitational force is so dominant in the equations of large-scale motions, it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$ (the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane), $T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density. The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k}) vector system provides the following six equations (namely the momentum balance, the hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation equations and an equation of state): 
     34\textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect  
     35of small scale processes on the large-scale) are expressed in terms of large-scale features 
     36 
     37\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their  
     38contribution to the buoyancy force 
     39 
     40\textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a  
     41balance between the vertical pressure gradient and the buoyancy force (this removes  
     42convective processes from the initial Navier-Stokes equations and so convective processes  
     43must be parameterized instead) 
     44 
     45\textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity  
     46vector is assumed to be zero. 
     47 
     48Because the gravitational force is so dominant in the equations of large-scale motions,  
     49it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked  
     50to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two  
     51vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define  
     52the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$  
     53(the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane),  
     54$T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density.  
     55The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k})  
     56vector system provides the following six equations (namely the momentum balance, the  
     57hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation  
     58equations and an equation of state): 
    4359\begin{subequations} \label{Eq_PE} 
    4460  \begin{equation}     \label{Eq_PE_dyn} 
     
    6581  \end{equation} 
    6682\end{subequations} 
    67 where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions, $t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by the equation of state (\ref{Eq_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure, $f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's angular velocity vector), and $g$ is the gravitational acceleration.  
    68 ${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterizations of small-scale physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$ surface forcing terms. Their nature and formulation are discussed in \S\ref{PE_zdf_ldf} and page \S\ref{PE_boundary_condition}. 
     83where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions,  
     84$t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by  
     85the equation of state (\ref{Eq_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,  
     86$f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's  
     87angular velocity vector), and $g$ is the gravitational acceleration.  
     88${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterisations of small-scale  
     89physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$  
     90and $F^S$ surface forcing terms. Their nature and formulation are discussed in  
     91\S\ref{PE_zdf_ldf} and page \S\ref{PE_boundary_condition}. 
    6992 
    7093. 
     
    7699\label{PE_boundary_condition} 
    77100 
    78 An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$, chosen as a mean sea surface (Fig.~\ref{Fig_ocean_bc}). Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth, the continental margins, the sea ice and the atmosphere. However, some of these fluxes are so weak that even on climatic time scales of thousands of years they can be neglected. In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and the other components of the earth system. 
     101An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea  
     102or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$  
     103and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height  
     104of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$,  
     105chosen as a mean sea surface (Fig.~\ref{Fig_ocean_bc}). Through these two boundaries,  
     106the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth,  
     107the continental margins, the sea ice and the atmosphere. However, some of these fluxes are  
     108so weak that even on climatic time scales of thousands of years they can be neglected.  
     109In the following, we briefly review the fluxes exchanged at the interfaces between the ocean  
     110and the other components of the earth system. 
    79111 
    80112%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    81113\begin{figure}[!ht] \label{Fig_ocean_bc}  \begin{center} 
    82114\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_I_ocean_bc.pdf} 
    83 \caption{The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface. Both $H$ and $\eta $ are referenced to $z=0$.} 
     115\caption{The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$  
     116is the depth of the sea floor and $\eta$ the height of the sea surface. Both $H$ and $\eta $  
     117are referenced to $z=0$.} 
    84118\end{center}   \end{figure} 
    85119%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     
    87121 
    88122\begin{description} 
    89 \item[Land - ocean interface:] the major flux between continental margins and the ocean is a mass exchange of fresh water through river runoff. Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths. It can be neglected for short range integrations but has to be taken into account for long term integrations as it influences the characteristics of water masses formed (especially at high latitudes). It is required in order to close the water cycle of the climate system. It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths. 
    90 \item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small, except in special areas of little extent. They are usually neglected in the model \footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling ($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world ocean (see \ref{TRA_bbc}).}. The boundary condition is thus set to no flux of heat and salt across solid boundaries. For momentum, the situation is different. There is no flow across solid boundaries, $i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words, the bottom velocity is parallel to solid boundaries). This kinematic boundary condition can be expressed as: 
     123\item[Land - ocean interface:] the major flux between continental margins and the ocean is  
     124a mass exchange of fresh water through river runoff. Such an exchange modifies the sea  
     125surface salinity especially in the vicinity of major river mouths. It can be neglected for short  
     126range integrations but has to be taken into account for long term integrations as it influences  
     127the characteristics of water masses formed (especially at high latitudes). It is required in order  
     128to close the water cycle of the climate system. It is usually specified as a fresh water flux at  
     129the air-sea interface in the vicinity of river mouths. 
     130\item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small,  
     131except in special areas of little extent. They are usually neglected in the model  
     132\footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling  
     133($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world  
     134ocean (see \ref{TRA_bbc}).}.  
     135The boundary condition is thus set to no flux of heat and salt across solid boundaries.  
     136For momentum, the situation is different. There is no flow across solid boundaries,  
     137$i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words,  
     138the bottom velocity is parallel to solid boundaries). This kinematic boundary condition  
     139can be expressed as: 
    91140\begin{equation} \label{Eq_PE_w_bbc} 
    92141w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right) 
    93142\end{equation} 
    94 In addition, the ocean exchanges momentum with the earth through frictional processes. Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification depends on the nature of the physical parameterization used for ${\rm {\bf D}}^{\rm {\bf U}}$ in \eqref{Eq_PE_dyn}. It is discussed in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9. 
    95 \item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux of fresh water PE  (the precipitation minus evaporation budget) leads to:  
     143In addition, the ocean exchanges momentum with the earth through frictional processes.  
     144Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized  
     145in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification  
     146depends on the nature of the physical parameterisation used for ${\rm {\bf D}}^{\rm {\bf U}}$  
     147in \eqref{Eq_PE_dyn}. It is discussed in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9. 
     148\item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux  
     149of fresh water PE  (the precipitation minus evaporation budget) leads to:  
    96150\begin{equation} \label{Eq_PE_w_sbc} 
    97151w = \frac{\partial \eta }{\partial t}  
     
    99153    + P-E 
    100154\end{equation} 
    101 The dynamic boundary condition, neglecting the surface tension (which removes capillary waves from the system) leads to the continuity of pressure across the interface $z=\eta$. The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat. 
    102 \item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water and momentum. The sea surface temperature is constrained to be at the freezing point at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and salt fluxes that cannot be neglected. 
     155The dynamic boundary condition, neglecting the surface tension (which removes capillary  
     156waves from the system) leads to the continuity of pressure across the interface $z=\eta$.  
     157The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat. 
     158\item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water  
     159and momentum. The sea surface temperature is constrained to be at the freezing point  
     160at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the  
     161ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and  
     162salt fluxes that cannot be neglected. 
    103163\end{description} 
    104164 
     
    116176\label{PE_p_formulation} 
    117177 
    118 The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that: $p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\ref{Eq_PE_hydrostatic}), assuming that pressure in decibars can be approximated by depth in meters in (\ref{Eq_PE_eos}). The hydrostatic pressure is then given by: 
     178The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a  
     179reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that:  
     180$p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\ref{Eq_PE_hydrostatic}),  
     181assuming that pressure in decibars can be approximated by depth in meters in (\ref{Eq_PE_eos}).  
     182The hydrostatic pressure is then given by: 
    119183\begin{equation} \label{Eq_PE_pressure} 
    120184p_h \left( {i,j,z,t} \right) 
    121  = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,z} \right)\;d\varsigma }  
    122 \end{equation} 
    123  Two strategies can be considered for the surface pressure term: $(a)$ introduce of a new variable $\eta$, the free-surface elevation, for which a prognostic equation can be established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used, one solution of the free-surface elevation consists of the excitation of external gravity waves. The flow is barotropic and the surface moves up and down with gravity as the restoring force. The phase speed of such waves is high (some hundreds of metres per second) so that the time step would have to be very short if they were present in the model. The latter strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$ the sea surface is the surface $z=0$. This well known approximation increases the surface wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic Rossby or planetary waves). In the present release of \NEMO, both strategies are still available. They are further described in the next two sub-sections. 
     185 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,\varsigma} \right)\;d\varsigma }  
     186\end{equation} 
     187 Two strategies can be considered for the surface pressure term: $(a)$ introduce of a  
     188 new variable $\eta$, the free-surface elevation, for which a prognostic equation can be  
     189 established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the  
     190 pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used,  
     191 one solution of the free-surface elevation consists of the excitation of external gravity waves.  
     192 The flow is barotropic and the surface moves up and down with gravity as the restoring force.  
     193 The phase speed of such waves is high (some hundreds of metres per second) so that  
     194 the time step would have to be very short if they were present in the model. The latter  
     195 strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$  
     196 the sea surface is the surface $z=0$. This well known approximation increases the surface  
     197 wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic  
     198 Rossby or planetary waves). In the present release of \NEMO, both strategies are still available.  
     199 They are further described in the next two sub-sections. 
    124200 
    125201% ------------------------------------------------------------------------------------------------------------- 
     
    129205\label{PE_free_surface} 
    130206 
    131 In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced which describes the shape of the air-sea interface. This variable is solution of a prognostic equation which is established by forming the vertical average of the kinematic surface condition (\ref{Eq_PE_w_bbc}): 
     207In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced  
     208which describes the shape of the air-sea interface. This variable is solution of a  
     209prognostic equation which is established by forming the vertical average of the kinematic  
     210surface condition (\ref{Eq_PE_w_bbc}): 
    132211\begin{equation} \label{Eq_PE_ssh} 
    133212\frac{\partial \eta }{\partial t}=-D+P-E 
     
    137216and using (\ref{Eq_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$. 
    138217 
    139 Allowing the air-sea interface to move introduces the external gravity waves (EGWs) as a class of solution of the primitive equations. These waves are barotropic because of hydrostatic assumption, and their phase speed is quite high. Their time scale is short with respect to the other processes described by the primitive equations. 
    140  
    141 Three choices can be made regarding the implementation of the free surface in the model, depending on the physical processes of interest.  
     218Allowing the air-sea interface to move introduces the external gravity waves (EGWs)  
     219as a class of solution of the primitive equations. These waves are barotropic because  
     220of hydrostatic assumption, and their phase speed is quite high. Their time scale is  
     221short with respect to the other processes described by the primitive equations. 
     222 
     223Three choices can be made regarding the implementation of the free surface in the model,  
     224depending on the physical processes of interest.  
    142225 
    143226$\bullet$ If one is interested in EGWs, in particular the tides and their interaction  
    144 with the baroclinic structure of the ocean (internal waves) possibly in  
    145 shallow seas, then a non linear free surface is the most appropriate. This  
    146 means that no approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the  
    147 ocean volume is fully taken into account. Note that in order to study the  
    148 fast time scales associated with EGWs it is necessary to minimize time  
    149 filtering effects (use an explicit time scheme with very small time step, or  
    150 a split-explicit scheme with reasonably small time step, see \S\ref{DYN_spg_exp} or 
    151 \S\ref{DYN_spg_ts}. 
     227with the baroclinic structure of the ocean (internal waves) possibly in shallow seas,  
     228then a non linear free surface is the most appropriate. This means that no  
     229approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the ocean  
     230volume is fully taken into account. Note that in order to study the fast time scales  
     231associated with EGWs it is necessary to minimize time filtering effects (use an  
     232explicit time scheme with very small time step, or a split-explicit scheme with  
     233reasonably small time step, see \S\ref{DYN_spg_exp} or \S\ref{DYN_spg_ts}. 
    152234 
    153235$\bullet$ If one is not interested in EGW but rather sees them as high frequency  
     
    163245external waves are removed from the system.  
    164246 
    165 The filtering of EGWs in models with a free surface is usually a matter of discretisation of the temporal derivatives, using the time splitting method \citep{Killworth1991, Zhang1992} or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach developed by \citet{Roullet2000}: the damping of EGWs is ensured by introducing an additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:  
     247The filtering of EGWs in models with a free surface is usually a matter of discretisation  
     248of the temporal derivatives, using the time splitting method \citep{Killworth1991, Zhang1992}  
     249or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach  
     250developed by \citet{Roullet2000}: the damping of EGWs is ensured by introducing an  
     251additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:  
    166252\begin{equation} \label{Eq_PE_flt} 
    167253\frac{\partial {\rm {\bf U}}_h }{\partial t}= {\rm {\bf M}} 
     
    169255- g \ T_c \nabla \left( \widetilde{\rho} \ \partial_t \eta \right)  
    170256\end{equation} 
    171 where $T_c$, is a parameter with dimensions of time which characterizes the force, $\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and $\rm {\bf M}$ represents the collected contributions of the Coriolis, hydrostatic pressure gradient, non-linear and viscous terms in \eqref{Eq_PE_dyn}. 
    172  
    173 The new force can be interpreted as a diffusion of vertically integrated volume flux divergence. The time evolution of $D$ is thus governed by a balance of two terms, $-g$ \textbf{A} $\eta$ and $g \, T_c \,$ \textbf{A} $D$, associated with a propagative regime and a diffusive regime in the temporal spectrum, respectively. In the diffusive regime, the EGWs no longer propagate, $i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than $T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs can be damped by choosing $T_c > \Delta t$. \citet{Roullet2000} demonstrate that (\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which has to be computed implicitly. This is not surprising since the use of a large time step has a necessarily numerical cost. Two gains arise in comparison with the previous formulations. Firstly, the damping of EGWs can be quantified through the magnitude of the additional term. Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as soon as $T_c > \Delta t$. 
    174  
    175 When the variations of free surface elevation are small compared to the thickness of the first model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized by \citet{Roullet2000} the linearization of (\ref{Eq_PE_ssh}) has consequences on the conservation of salt in the model. With the nonlinear free surface equation, the time evolution of the total salt content is  
     257where $T_c$, is a parameter with dimensions of time which characterizes the force,  
     258$\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and $\rm {\bf M}$  
     259represents the collected contributions of the Coriolis, hydrostatic pressure gradient,  
     260non-linear and viscous terms in \eqref{Eq_PE_dyn}. 
     261 
     262The new force can be interpreted as a diffusion of vertically integrated volume flux divergence.  
     263The time evolution of $D$ is thus governed by a balance of two terms, $-g$ \textbf{A} $\eta$  
     264and $g \, T_c \,$ \textbf{A} $D$, associated with a propagative regime and a diffusive regime  
     265in the temporal spectrum, respectively. In the diffusive regime, the EGWs no longer propagate,  
     266$i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than  
     267$T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs  
     268can be damped by choosing $T_c > \Delta t$. \citet{Roullet2000} demonstrate that  
     269(\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which  
     270has to be computed implicitly. This is not surprising since the use of a large time step has a  
     271necessarily numerical cost. Two gains arise in comparison with the previous formulations.  
     272Firstly, the damping of EGWs can be quantified through the magnitude of the additional term.  
     273Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as  
     274soon as $T_c > \Delta t$. 
     275 
     276When the variations of free surface elevation are small compared to the thickness of the first  
     277model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized  
     278by \citet{Roullet2000} the linearization of (\ref{Eq_PE_ssh}) has consequences on the  
     279conservation of salt in the model. With the nonlinear free surface equation, the time evolution  
     280of the total salt content is  
    176281\begin{equation} \label{Eq_PE_salt_content} 
    177 \frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} =\int\limits_S  
    178 {S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds} 
    179 \end{equation} 
    180 where $S$ is the salinity, and the total salt is integrated over the whole ocean volume $D_\eta$ bounded by the time-dependent free surface. The right hand side (which is an integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh}) is satisfied, so that the salt is perfectly conserved. When the free surface equation is linearized, \citet{Roullet2000} show that the total salt content integrated in the fixed volume $D$ (bounded by the surface $z=0$) is no longer conserved: 
     282    \frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv}  
     283                        =\int\limits_S {S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds} 
     284\end{equation} 
     285where $S$ is the salinity, and the total salt is integrated over the whole ocean volume  
     286$D_\eta$ bounded by the time-dependent free surface. The right hand side (which is an  
     287integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh})  
     288is satisfied, so that the salt is perfectly conserved. When the free surface equation is  
     289linearized, \citet{Roullet2000} show that the total salt content integrated in the fixed  
     290volume $D$ (bounded by the surface $z=0$) is no longer conserved: 
    181291\begin{equation} \label{Eq_PE_salt_content_linear} 
    182 \frac{\partial }{\partial t}\int\limits_D {S\;dv} =-\int\limits_S  
    183 {S\;\frac{\partial \eta }{\partial t}ds}  
    184 \end{equation} 
    185  
    186 The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions \citep{Roullet2000}. It can be significant when the freshwater forcing is not balanced and the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $} results in a decrease of the salinity in the fixed volume $D$. Even in that case though, the total salt integrated in the variable volume $D_{\eta}$ varies much less, since (\ref{Eq_PE_salt_content_linear}) can be rewritten as  
     292         \frac{\partial }{\partial t}\int\limits_D {S\;dv}  
     293               = - \int\limits_S {S\;\frac{\partial \eta }{\partial t}ds}  
     294\end{equation} 
     295 
     296The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions  
     297\citep{Roullet2000}. It can be significant when the freshwater forcing is not balanced and  
     298the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $}  
     299results in a decrease of the salinity in the fixed volume $D$. Even in that case though,  
     300the total salt integrated in the variable volume $D_{\eta}$ varies much less, since  
     301(\ref{Eq_PE_salt_content_linear}) can be rewritten as  
    187302\begin{equation} \label{Eq_PE_salt_content_corrected} 
    188303\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv}  
     
    191306\end{equation} 
    192307 
    193 Although the total salt content is not exactly conserved with the linearized free surface, its variations are driven by correlations of the time variation of surface salinity with the sea surface height, which is a negligible term. This situation contrasts with  
    194 the case of the rigid lid approximation (following section) in which case freshwater forcing is represented by a virtual salt flux, leading to a spurious source of salt at the ocean surface \citep{Roullet2000}. 
     308Although the total salt content is not exactly conserved with the linearized free surface,  
     309its variations are driven by correlations of the time variation of surface salinity with the  
     310sea surface height, which is a negligible term. This situation contrasts with the case of  
     311the rigid lid approximation (following section) in which case freshwater forcing is  
     312represented by a virtual salt flux, leading to a spurious source of salt at the ocean  
     313surface \citep{Roullet2000}. 
    195314 
    196315% ------------------------------------------------------------------------------------------------------------- 
     
    200319\label{PE_rigid_lid} 
    201320 
    202 With the rigid lid approximation, we assume that the ocean surface ($z=0$) is a rigid lid on which a pressure $p_s$ is exerted. This implies that the vertical velocity at the surface is equal to zero. From the continuity equation \eqref{Eq_PE_continuity} and the kinematic condition at the bottom \eqref{Eq_PE_w_bbc} (no flux across the bottom), it can be shown that the vertically integrated flow $H{\rm {\bf \bar {U}}}_h$ is non-divergent (where the overbar indicates a vertical average over the whole water column, i.e. from $z=-H$, the ocean bottom, to $z=0$ , the rigid-lid). Thus, $\rm {\bf \bar {U}}_h$ can be derived from a volume transport streamfunction $\psi$: 
     321With the rigid lid approximation, we assume that the ocean surface ($z=0$) is a rigid lid  
     322on which a pressure $p_s$ is exerted. This implies that the vertical velocity at the surface  
     323is equal to zero. From the continuity equation \eqref{Eq_PE_continuity} and the kinematic  
     324condition at the bottom \eqref{Eq_PE_w_bbc} (no flux across the bottom), it can be shown  
     325that the vertically integrated flow $H{\rm {\bf \bar {U}}}_h$ is non-divergent (where the  
     326overbar indicates a vertical average over the whole water column, i.e. from $z=-H$,  
     327the ocean bottom, to $z=0$ , the rigid-lid). Thus, $\rm {\bf \bar {U}}_h$ can be derived  
     328from a volume transport streamfunction $\psi$: 
    203329\begin{equation} \label{Eq_PE_u_psi} 
    204330\overline{\vect{U}}_h =\frac{1}{H}\left(   \vect{k} \times \nabla \psi   \right) 
    205331\end{equation} 
    206332 
    207 As $p_s$ does not depend on depth, its horizontal gradient is obtained by forming the vertical average of \eqref{Eq_PE_dyn} and using \eqref{Eq_PE_u_psi}: 
     333As $p_s$ does not depend on depth, its horizontal gradient is obtained by forming the  
     334vertical average of \eqref{Eq_PE_dyn} and using \eqref{Eq_PE_u_psi}: 
    208335 
    209336\begin{equation} \label{Eq_PE_u_barotrope} 
     
    214341\end{equation} 
    215342 
    216 Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ represents the collected contributions of the Coriolis, hydrostatic pressure gradient, nonlinear and viscous terms in \eqref{Eq_PE_dyn}. The time derivative of $\psi $ is the solution of an elliptic equation which is obtained from the vertical component of the curl of (\ref{Eq_PE_u_barotrope}): 
     343Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ represents the collected contributions of the  
     344Coriolis, hydrostatic pressure gradient, nonlinear and viscous terms in \eqref{Eq_PE_dyn}.  
     345The time derivative of $\psi $ is the solution of an elliptic equation which is obtained from  
     346the vertical component of the curl of (\ref{Eq_PE_u_barotrope}): 
    217347\begin{equation} \label{Eq_PE_psi} 
    218348\left[   {\nabla \times \left[ {\frac{1}{H} \vect{\bf k}  
     
    221351\end{equation} 
    222352 
    223 Using the proper boundary conditions, \eqref{Eq_PE_psi} can be solved to find $\partial_t \psi$ and thus using \eqref{Eq_PE_u_barotrope} the horizontal surface pressure gradient. It should be noted that $p_s$ can be computed by taking the divergence of \eqref{Eq_PE_u_barotrope} and solving the resulting elliptic equation. Thus the surface pressure is a diagnostic quantity that can be recovered for analysis purposes. 
    224  
    225 A difficulty lies in the determination of the boundary condition on $\partial_t \psi$. The boundary condition on velocity is that there is no flow normal to a solid wall, $i.e.$ the coastlines are streamlines. Therefore \eqref{Eq_PE_psi} is solved with the following Dirichlet boundary condition: $\partial_t \psi$ is constant along each coastline of the same continent or of the same island. When all the coastlines are connected (there are no islands), the constant value of $\partial_t \psi$ along the coast can be arbitrarily chosen to be zero. When islands are present in the domain, the value of the barotropic streamfunction will generally be different for each island and for the continent, and will vary with respect to time. So the boundary condition is: $\psi=0$ along the continent and $\psi=\mu_n$ along island $n$ ($1 \leq n \leq Q$), where $Q$ is the number of islands present in the domain and $\mu_n$ is a time dependent variable. A time evolution equation of the unknown $\mu_n$ can be found by evaluating the circulation of the time derivative of the vertical average (barotropic) velocity field along a closed contour around each island. Since the circulation of a gradient field along a closed contour is zero, from \eqref{Eq_PE_u_barotrope} we have: 
     353Using the proper boundary conditions, \eqref{Eq_PE_psi} can be solved to find $\partial_t \psi$  
     354and thus using \eqref{Eq_PE_u_barotrope} the horizontal surface pressure gradient.  
     355It should be noted that $p_s$ can be computed by taking the divergence of  
     356\eqref{Eq_PE_u_barotrope} and solving the resulting elliptic equation. Thus the surface  
     357pressure is a diagnostic quantity that can be recovered for analysis purposes. 
     358 
     359A difficulty lies in the determination of the boundary condition on $\partial_t \psi$.  
     360The boundary condition on velocity is that there is no flow normal to a solid wall,  
     361$i.e.$ the coastlines are streamlines. Therefore \eqref{Eq_PE_psi} is solved with  
     362the following Dirichlet boundary condition: $\partial_t \psi$ is constant along each  
     363coastline of the same continent or of the same island. When all the coastlines are  
     364connected (there are no islands), the constant value of $\partial_t \psi$ along the  
     365coast can be arbitrarily chosen to be zero. When islands are present in the domain,  
     366the value of the barotropic streamfunction will generally be different for each island  
     367and for the continent, and will vary with respect to time. So the boundary condition is:  
     368$\psi=0$ along the continent and $\psi=\mu_n$ along island $n$ ($1 \leq n \leq Q$),  
     369where $Q$ is the number of islands present in the domain and $\mu_n$ is a time  
     370dependent variable. A time evolution equation of the unknown $\mu_n$ can be found  
     371by evaluating the circulation of the time derivative of the vertical average (barotropic)  
     372velocity field along a closed contour around each island. Since the circulation of a  
     373gradient field along a closed contour is zero, from \eqref{Eq_PE_u_barotrope} we have: 
    226374\begin{equation} \label{Eq_PE_isl_circulation} 
    227375\oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(  
     
    256404\right)_{1\leqslant n\leqslant Q} ={\rm {\bf B}} 
    257405\end{equation} 
    258 where \textbf{A} is a $Q  \times Q$ matrix and \textbf{B} is a time dependent vector. As \textbf{A} is independent of time, it can be calculated and inverted once. The time derivative of the streamfunction when islands are present is thus given by: 
     406where \textbf{A} is a $Q  \times Q$ matrix and \textbf{B} is a time dependent vector.  
     407As \textbf{A} is independent of time, it can be calculated and inverted once. The time  
     408derivative of the streamfunction when islands are present is thus given by: 
    259409\begin{equation} \label{Eq_PE_psi_isl_dt} 
    260410\frac{\partial \psi }{\partial t}=\frac{\partial \psi _o }{\partial  
     
    277427\label{PE_tensorial} 
    278428 
    279 In many ocean circulation problems, the flow field has regions of enhanced dynamics ($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts). The representation of such dynamical processes can be improved by specifically increasing the model resolution in these regions. As well, it may be convenient to use a lateral boundary-following coordinate system to better represent coastal dynamics. Moreover, the common geographical coordinate system has a singular point at the North Pole that cannot be easily treated in a global model without filtering. A solution consists of introducing an appropriate coordinate transformation that shifts the singular point onto land \citep{MadecImb1996, Murray1996}. As a consequence, it is important to solve the primitive equations in various curvilinear coordinate systems. An efficient way of introducing an appropriate coordinate transform can be found when using a tensorial formalism. This formalism is suited to any multidimensional curvilinear coordinate system. Ocean modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth approximation), with preservation of the local vertical. Here we give the simplified equations for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey of the conservation laws of fluid dynamics. 
    280  
    281 Let (\textbf{i},\textbf{j},\textbf{k}) be a set of orthogonal curvilinear coordinates on the sphere associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (Fig.\ref{Fig_referential}). Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea level (Fig.\ref{Fig_referential}). The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$, the three scale factors: 
     429In many ocean circulation problems, the flow field has regions of enhanced dynamics  
     430($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts).  
     431The representation of such dynamical processes can be improved by specifically increasing  
     432the model resolution in these regions. As well, it may be convenient to use a lateral  
     433boundary-following coordinate system to better represent coastal dynamics. Moreover,  
     434the common geographical coordinate system has a singular point at the North Pole that  
     435cannot be easily treated in a global model without filtering. A solution consists of introducing  
     436an appropriate coordinate transformation that shifts the singular point onto land  
     437\citep{MadecImb1996, Murray1996}. As a consequence, it is important to solve the primitive  
     438equations in various curvilinear coordinate systems. An efficient way of introducing an  
     439appropriate coordinate transform can be found when using a tensorial formalism.  
     440This formalism is suited to any multidimensional curvilinear coordinate system. Ocean  
     441modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth  
     442approximation), with preservation of the local vertical. Here we give the simplified equations  
     443for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey  
     444of the conservation laws of fluid dynamics. 
     445 
     446Let (\textbf{i},\textbf{j},\textbf{k}) be a set of orthogonal curvilinear coordinates on the sphere  
     447associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k})  
     448linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are  
     449two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (Fig.\ref{Fig_referential}).  
     450Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined  
     451by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of  
     452the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea  
     453level (Fig.\ref{Fig_referential}). The local deformation of the curvilinear coordinate system is  
     454given by $e_1$, $e_2$ and $e_3$, the three scale factors: 
    282455\begin{equation} \label{Eq_scale_factors} 
    283456\begin{aligned} 
     
    295468\begin{figure}[!tb] \label{Fig_referential}  \begin{center} 
    296469\includegraphics[width=0.60\textwidth]{./TexFiles/Figures/Fig_I_earth_referential.pdf} 
    297 \caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear coordinate system (\textbf{i},\textbf{j},\textbf{k}). } 
     470\caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear  
     471coordinate system (\textbf{i},\textbf{j},\textbf{k}). } 
    298472\end{center}   \end{figure} 
    299473%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    300474 
    301 Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by $a$ in (\ref{Eq_scale_factors}) (thin-shell approximation). The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that appear in the primitive equations (Eqs. \eqref{Eq_PE_dyn} to \eqref{Eq_PE_eos}) can be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate system transformation: 
     475Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by  
     476$a$ in (\ref{Eq_scale_factors}) (thin-shell approximation). The resulting horizontal scale  
     477factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single  
     478function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that  
     479appear in the primitive equations (Eqs. \eqref{Eq_PE_dyn} to \eqref{Eq_PE_eos}) can  
     480be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate  
     481system transformation: 
    302482\begin{subequations} \label{Eq_PE_discrete_operators} 
    303483\begin{equation} \label{Eq_PE_grad} 
     
    341521\label{PE_zco_Eq} 
    342522 
    343 In order to express the Primitive Equations in tensorial formalism, it is necessary to compute the horizontal component of the non-linear and viscous terms of the equation using \eqref{Eq_PE_grad}) to \eqref{Eq_PE_lap_vector}. Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$, by: 
     523In order to express the Primitive Equations in tensorial formalism, it is necessary to compute  
     524the horizontal component of the non-linear and viscous terms of the equation using  
     525\eqref{Eq_PE_grad}) to \eqref{Eq_PE_lap_vector}.  
     526Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate  
     527system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity  
     528field $\chi$, by: 
    344529\begin{equation} \label{Eq_PE_curl_Uh} 
    345530\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v}  
     
    353538\end{equation} 
    354539 
    355 Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that $e_3$  is a function of the single variable $k$, the nonlinear term of \eqref{Eq_PE_dyn} can be transformed as follows: 
     540Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$  
     541and that $e_3$  is a function of the single variable $k$, the nonlinear term of  
     542\eqref{Eq_PE_dyn} can be transformed as follows: 
    356543\begin{flalign*} 
    357544&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}} 
     
    391578\end{flalign*} 
    392579 
    393 The last term of the right hand side is obviously zero, and thus the nonlinear term of \eqref{Eq_PE_dyn} is written in the $(i,j,k)$ coordinate system: 
     580The last term of the right hand side is obviously zero, and thus the nonlinear term of  
     581\eqref{Eq_PE_dyn} is written in the $(i,j,k)$ coordinate system: 
    394582\begin{equation} \label{Eq_PE_vector_form} 
    395583\left[ {\left( {  \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}} 
     
    401589\end{equation} 
    402590 
    403 This is the so-called \textit{vector invariant form} of the momentum advection term. For some purposes, it can be advantageous to write this term in the so-called flux form, $i.e.$ to write it as the divergence of fluxes. For example, the first component of \eqref{Eq_PE_vector_form} (the $i$-component) is transformed as follows: 
     591This is the so-called \textit{vector invariant form} of the momentum advection term.  
     592For some purposes, it can be advantageous to write this term in the so-called flux form,  
     593$i.e.$ to write it as the divergence of fluxes. For example, the first component of  
     594\eqref{Eq_PE_vector_form} (the $i$-component) is transformed as follows: 
    404595\begin{flalign*} 
    405596&{ \begin{array}{*{20}l} 
     
    486677\end{multline} 
    487678 
    488 The flux form has two terms, the first one is expressed as the divergence of momentum fluxes (hence the flux form name given to this formulation) and the second one is due to the curvilinear nature of the coordinate system used. The latter is called the \emph{metric} term and can be viewed as a modification of the Coriolis parameter:  
     679The flux form has two terms, the first one is expressed as the divergence of momentum  
     680fluxes (hence the flux form name given to this formulation) and the second one is due to  
     681the curvilinear nature of the coordinate system used. The latter is called the \emph{metric}  
     682term and can be viewed as a modification of the Coriolis parameter:  
    489683\begin{equation} \label{Eq_PE_cor+metric} 
    490684f \to f + \frac{1}{e_1 \; e_2}   \left(    v \frac{\partial e_2}{\partial i} 
     
    492686\end{equation} 
    493687 
    494 Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$ and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of the Coriolis parameter $f \to f+(u/a) \tan \varphi$. 
     688Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$  
     689and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of  
     690the Coriolis parameter $f \to f+(u/a) \tan \varphi$. 
    495691 
    496692To sum up, the equations solved by the ocean model can be written in the following tensorial formalism: 
     
    545741\end{multline} 
    546742\end{subequations} 
    547 where $\zeta$ is given by \eqref{Eq_PE_curl_Uh} and the surface pressure gradient formulation depends on the one of the free surface: 
     743where $\zeta$ is given by \eqref{Eq_PE_curl_Uh} and the surface pressure gradient formulation  
     744depends on the one of the free surface: 
    548745 
    549746$*$ free surface formulation 
     
    566763\end{equation} 
    567764where ${\vect{M}}= \left( M_u,M_v \right)$ represents the collected contributions of nonlinear,  
    568 viscosity and hydrostatic pressure gradient terms in \eqref{Eq_PE_dyn_vect} and the overbar indicates a vertical average over the whole water column ($i.e.$ from $z=-H$, the ocean bottom, to $z=0$, the rigid-lid), and where the time derivative of $\psi$ is the solution of an elliptic equation: 
     765viscosity and hydrostatic pressure gradient terms in \eqref{Eq_PE_dyn_vect} and the overbar  
     766indicates a vertical average over the whole water column ($i.e.$ from $z=-H$, the ocean bottom,  
     767to $z=0$, the rigid-lid), and where the time derivative of $\psi$ is the solution of an elliptic equation: 
    569768\begin{multline} \label{Eq_psi_total} 
    570769  \frac{\partial }{\partial i}\left[ {\frac{e_2 }{H\,e_1}\frac{\partial}{\partial i} 
     
    605804\end{equation} 
    606805 
    607 The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid  
    608 scale parameterization used. It will be defined in \S\ref{PE_zdf}. The nature and formulation of ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed in Chapter~\ref{SBC}. 
     806The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale  
     807parameterisation used. It will be defined in \S\ref{PE_zdf}. The nature and formulation of  
     808${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed  
     809in Chapter~\ref{SBC}. 
    609810 
    610811\newpage  
     
    617818\begin{figure}[!b] \label{Fig_z_zstar}  \begin{center} 
    618819\includegraphics[width=1.0\textwidth]{./TexFiles/Figures/Fig_z_zstar.pdf} 
    619 \caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate \citep{Adcroft_Campin_OM04} ).} 
     820\caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear  
     821free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate  
     822\citep{Adcroft_Campin_OM04} ).} 
    620823\end{center}   \end{figure} 
    621824%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    622825 
    623826 
    624 In that case, the free surface equation is nonlinear, and the variations of volume are fully taken into account. These coordinates systems is presented in a report  
    625 \citep{Levier2007} available on the \NEMO web site.  
     827In that case, the free surface equation is nonlinear, and the variations of volume are fully  
     828taken into account. These coordinates systems is presented in a report \citep{Levier2007}  
     829available on the \NEMO web site.  
    626830 
    627831\gmcomment{ 
    628 The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation which allows one to deal with large amplitude free-surface 
     832The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation  
     833which allows one to deal with large amplitude free-surface 
    629834variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In 
    630835the  \textit{z*} formulation, the variation of the column thickness due to sea-surface 
     
    642847 
    643848Since the vertical displacement of the free surface is incorporated in the vertical 
    644 coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,  $\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field is no longer zero as shown by the continuity equation: 
     849coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,   
     850$\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field  
     851is no longer zero as shown by the continuity equation: 
    645852 
    646853$\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right) 
     
    662869\subsection{Introduction} 
    663870 
    664 Several important aspects of the ocean circulation are influenced by bottom topography. Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers, sills and channels that strongly constrain the path of water masses, but more subtle effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary one along continental slopes. Topographic Rossby waves can be excited and can interact with the mean current. In the $z-$coordinate system presented in the previous section (\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom and to large localized depth gradients associated with large localized vertical velocities. The response to such a velocity field often leads to numerical dispersion effects. One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}. Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)  
    665  
    666 The $s$-coordinate avoids the discretisation error in the depth field since the layers of computation are gradually adjusted with depth to the ocean bottom. Relatively small topographic features as well as  gentle, large-scale slopes of the sea floor in the deep ocean, which would be ignored in typical $z$-model applications with the largest grid spacing at greatest depths, can easily be represented (with relatively low vertical resolution). A terrain-following model (hereafter $s-$model) also facilitates the modelling of the boundary layer flows over a large depth range, which in the framework of the $z$-model would require high vertical resolution over the whole depth range. Moreover, with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as the only boundaries of the domain (nomore lateral boundary condition to specify). Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean, it has strong limitations as soon as stratification is introduced. The main two problems come from the truncation error in the horizontal pressure gradient and a possibly increased diapycnal diffusion. The horizontal pressure force in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}), 
     871Several important aspects of the ocean circulation are influenced by bottom topography.  
     872Of course, the most important is that bottom topography determines deep ocean sub-basins,  
     873barriers, sills and channels that strongly constrain the path of water masses, but more subtle  
     874effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary  
     875one along continental slopes. Topographic Rossby waves can be excited and can interact  
     876with the mean current. In the $z-$coordinate system presented in the previous section  
     877(\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is  
     878discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom  
     879and to large localized depth gradients associated with large localized vertical velocities.  
     880The response to such a velocity field often leads to numerical dispersion effects.  
     881One solution to strongly reduce this error is to use a partial step representation of bottom  
     882topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.  
     883Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)  
     884 
     885The $s$-coordinate avoids the discretisation error in the depth field since the layers of  
     886computation are gradually adjusted with depth to the ocean bottom. Relatively small  
     887topographic features as well as  gentle, large-scale slopes of the sea floor in the deep  
     888ocean, which would be ignored in typical $z$-model applications with the largest grid  
     889spacing at greatest depths, can easily be represented (with relatively low vertical resolution).  
     890A terrain-following model (hereafter $s-$model) also facilitates the modelling of the  
     891boundary layer flows over a large depth range, which in the framework of the $z$-model  
     892would require high vertical resolution over the whole depth range. Moreover, with a  
     893$s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface  
     894as the only boundaries of the domain (nomore lateral boundary condition to specify).  
     895Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a  
     896homogeneous ocean, it has strong limitations as soon as stratification is introduced.  
     897The main two problems come from the truncation error in the horizontal pressure  
     898gradient and a possibly increased diapycnal diffusion. The horizontal pressure force  
     899in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}), 
    667900 
    668901\begin{equation} \label{Eq_PE_p_sco} 
     
    671904\end{equation} 
    672905 
    673 The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface and introduces a truncation error that is not present in a $z$-model. In the special case of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$), \citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude of this truncation error. It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state, and the finite difference scheme. This error limits the possible topographic slopes that a model can handle at a given horizontal and vertical resolution. This is a severe restriction for large-scale applications using realistic bottom topography. The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive. This problem can be at least partially overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec1996}. However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for a realistic bottom topography: a envelope topography is defined in $s$-coordinate on which a full or partial step bottom topography is then applied in order to adjust the model depth to the observed one (see \S\ref{DOM_zgr}. 
    674  
    675 For numerical reasons a minimum of diffusion is required along the coordinate surfaces of any finite difference model. It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as well as for a $s$-model.  
    676 However, density varies more strongly on $s-$surfaces than on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a $z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation. For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast, the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column (i.e. the main thermocline) \citep{Madec1996}. An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}. Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large, strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).  
    677  
    678 The $s-$coordinates introduced here \citep{Lott1990,Madec1996} differ mainly in two aspects from similar models:  it allows  a representation of bottom topography with mixed full or partial step-like/terrain following topography ; It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate. 
     906The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface  
     907and introduces a truncation error that is not present in a $z$-model. In the special case  
     908of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),  
     909\citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude  
     910of this truncation error. It depends on topographic slope, stratification, horizontal and  
     911vertical resolution, the equation of state, and the finite difference scheme. This error  
     912limits the possible topographic slopes that a model can handle at a given horizontal  
     913and vertical resolution. This is a severe restriction for large-scale applications using  
     914realistic bottom topography. The large-scale slopes require high horizontal resolution,  
     915and the computational cost becomes prohibitive. This problem can be at least partially  
     916overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec1996}. However, the definition of the model  
     917domain vertical coordinate becomes then a non-trivial thing for a realistic bottom  
     918topography: a envelope topography is defined in $s$-coordinate on which a full or  
     919partial step bottom topography is then applied in order to adjust the model depth to  
     920the observed one (see \S\ref{DOM_zgr}. 
     921 
     922For numerical reasons a minimum of diffusion is required along the coordinate surfaces  
     923of any finite difference model. It causes spurious diapycnal mixing when coordinate  
     924surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as  
     925well as for a $s$-model. However, density varies more strongly on $s-$surfaces than  
     926on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal  
     927diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a  
     928$z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal  
     929circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.  
     930For example, imagine an isolated bump of topography in an ocean at rest with a horizontally  
     931uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral  
     932surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,  
     933the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column  
     934($i.e.$ the main thermocline) \citep{Madec1996}. An alternate solution consists of rotating  
     935the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}.  
     936Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,  
     937strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).  
     938 
     939The $s-$coordinates introduced here \citep{Lott1990,Madec1996} differ mainly in two  
     940aspects from similar models:  it allows  a representation of bottom topography with mixed  
     941full or partial step-like/terrain following topography ; It also offers a completely general  
     942transformation, $s=s(i,j,z)$ for the vertical coordinate. 
    679943 
    680944% ------------------------------------------------------------------------------------------------------------- 
     
    683947\subsection{The \textit{s-}coordinate Formulation} 
    684948 
    685 Starting from the set of equations established in \S\ref{PE_zco} for the special case $k=z$ and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes $z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and $s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed equations is given in Appendix~\ref{Apdx_A}. Let us define the vertical scale factor by $e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the (\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by : 
     949Starting from the set of equations established in \S\ref{PE_zco} for the special case $k=z$  
     950and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes  
     951$z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and  
     952$s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed  
     953equations is given in Appendix~\ref{Apdx_A}. Let us define the vertical scale factor by  
     954$e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the  
     955(\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by : 
    686956\begin{equation} \label{Eq_PE_sco_slope} 
    687957\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s  
     
    689959\sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right|_s 
    690960\end{equation} 
    691 We also introduce  $\omega $, a dia-surface velocity component, defined as the velocity relative to the moving $s$-surfaces and normal to them: 
     961We also introduce  $\omega $, a dia-surface velocity component, defined as the velocity  
     962relative to the moving $s$-surfaces and normal to them: 
    692963\begin{equation} \label{Eq_PE_sco_w} 
    693964\omega  = w - e_3 \, \frac{\partial s}{\partial t} - \sigma _1 \,u - \sigma _2 \,v    \\ 
     
    716987   +  D_v^{\vect{U}}  +   F_v^{\vect{U}} \quad 
    717988\end{multline} 
    718 where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic pressure have the same expressions as in $z$-coordinates although they do not represent exactly the same quantities. $\omega$ is provided by the continuity equation (see Appendix~\ref{Apdx_A}): 
     989where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic  
     990pressure have the same expressions as in $z$-coordinates although they do not represent  
     991exactly the same quantities. $\omega$ is provided by the continuity equation  
     992(see Appendix~\ref{Apdx_A}): 
    719993 
    720994\begin{equation} \label{Eq_PE_sco_continuity} 
     
    7421016\end{multline} 
    7431017 
    744 The equation of state has the same expression as in $z$-coordinate, and similar expressions are used for mixing and forcing terms. 
     1018The equation of state has the same expression as in $z$-coordinate, and similar expressions  
     1019are used for mixing and forcing terms. 
    7451020 
    7461021\gmcomment{ 
     
    7681043It is usually called the subgrid scale physics. It must be emphasized that  
    7691044this is the weakest part of the primitive equations, but also one of the  
    770 most important for long-term simulations as small scale processes \textit{in fine} balance  
    771 the surface input of kinetic energy and heat. 
     1045most important for long-term simulations as small scale processes \textit{in fine}  
     1046balance the surface input of kinetic energy and heat. 
    7721047 
    7731048The control exerted by gravity on the flow induces a strong anisotropy  
    774 between the lateral and vertical motions. Therefore subgrid-scale physics  \textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \eqref{Eq_PE_dyn}, \eqref{Eq_PE_tra_T} and \eqref{Eq_PE_tra_S} are divided into a lateral part  \textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part  \textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms and their underlying physics are briefly discussed in the next two subsections. 
     1049between the lateral and vertical motions. Therefore subgrid-scale physics   
     1050\textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \eqref{Eq_PE_dyn},  
     1051\eqref{Eq_PE_tra_T} and \eqref{Eq_PE_tra_S} are divided into a lateral part   
     1052\textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part   
     1053\textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms  
     1054and their underlying physics are briefly discussed in the next two subsections. 
    7751055 
    7761056% ------------------------------------------------------------------------------------------------------------- 
     
    7851065partially, but always parameterized. The vertical turbulent fluxes are  
    7861066assumed to depend linearly on the gradients of large-scale quantities (for  
    787 example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$, where $A^{vT}$ is an eddy coefficient). This formulation is  
     1067example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$,  
     1068where $A^{vT}$ is an eddy coefficient). This formulation is  
    7881069analogous to that of molecular diffusion and dissipation. This is quite  
    7891070clearly a necessary compromise: considering only the molecular viscosity  
     
    7941075\begin{equation} \label{Eq_PE_zdf} 
    7951076   \begin{split} 
    796 {\vect{D}}^{v \vect{U}} 
    797 &=\frac{\partial }{\partial z}\left( {A^{vm}\frac{\partial {\vect{U}}_h }{\partial z}} \right) \ , \\         D^{vT} &= \frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial T}{\partial z}} \right) \ , 
     1077{\vect{D}}^{v \vect{U}} &=\frac{\partial }{\partial z}\left( {A^{vm}\frac{\partial {\vect{U}}_h }{\partial z}} \right) \ , \\          
     1078D^{vT}                        &= \frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial T}{\partial z}} \right) \ , 
    7981079\quad 
    7991080D^{vS}=\frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial S}{\partial z}} \right) 
    8001081   \end{split} 
    8011082\end{equation} 
    802 where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients, respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified (see Chap.~\ref{SBC} and \ref{ZDF} and \S\ref{TRA_bbl}). All the vertical physics is embedded in the specification of the eddy coefficients. They can be assumed to be either constant, or function of the local fluid properties ($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a turbulent closure model. The choices available in \NEMO are discussed in \S\ref{ZDF}). 
     1083where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients,  
     1084respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat  
     1085and salt must be specified (see Chap.~\ref{SBC} and \ref{ZDF} and \S\ref{TRA_bbl}).  
     1086All the vertical physics is embedded in the specification of the eddy coefficients.  
     1087They can be assumed to be either constant, or function of the local fluid properties  
     1088($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a  
     1089turbulent closure model. The choices available in \NEMO are discussed in \S\ref{ZDF}). 
    8031090 
    8041091% ------------------------------------------------------------------------------------------------------------- 
     
    8211108lateral diffusive and dissipative operators are of second order.  
    8221109Observations show that lateral mixing induced by mesoscale turbulence tends  
    823 to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987}) rather than across them.  
     1110to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987})  
     1111rather than across them.  
    8241112As the slope of neutral surfaces is small in the ocean, a common  
    8251113approximation is to assume that the `lateral' direction is the horizontal,  
     
    8341122energy whereas potential energy is a main source of turbulence (through  
    8351123baroclinic instabilities). \citet{Gent1990} have proposed a  
    836 parameterization of mesoscale eddy-induced turbulence which associates an  
     1124parameterisation of mesoscale eddy-induced turbulence which associates an  
    8371125eddy-induced velocity to the isoneutral diffusion. Its mean effect is to  
    8381126reduce the mean potential energy of the ocean. This leads to a formulation  
     
    8501138the model while not interfering with the solved mesoscale activity. Another approach  
    8511139is becoming more and more popular: instead of specifying explicitly a sub-grid scale  
    852 term in the momentum and tracer time evolution equations, one uses a advective scheme which is diffusive enough to maintain the model stability. It must be emphasised 
     1140term in the momentum and tracer time evolution equations, one uses a advective  
     1141scheme which is diffusive enough to maintain the model stability. It must be emphasised 
    8531142that then, all the sub-grid scale physics is in this case include in the formulation of the 
    8541143advection scheme.  
    8551144 
    856 All these parameterizations of subgrid scale physics present advantages and  
     1145All these parameterisations of subgrid scale physics present advantages and  
    8571146drawbacks. There are not all available in \NEMO. In the $z$-coordinate  
    8581147formulation, five options are offered for active tracers (temperature and  
    8591148salinity): second order geopotential operator, second order isoneutral  
    860 operator, \citet{Gent1990} parameterization, fourth order  
    861 geopotential operator, and various slightly diffusive advection schemes. The same options are available for momentum, except  
    862 \citet{Gent1990} parameterization which only involves tracers. In the 
     1149operator, \citet{Gent1990} parameterisation, fourth order  
     1150geopotential operator, and various slightly diffusive advection schemes.  
     1151The same options are available for momentum, except  
     1152\citet{Gent1990} parameterisation which only involves tracers. In the 
    8631153$s$-coordinate formulation, additional options are offered for tracers: second  
    8641154order operator acting along $s-$surfaces, and for momentum: fourth order  
     
    8811171rotation between geopotential and $s$-surfaces, while it is only an approximation  
    8821172for the rotation between isoneutral and $z$- or $s$-surfaces. Indeed, in the latter  
    883 case, two assumptions are made to simplify  \eqref{Eq_PE_iso_tensor} \citep{Cox1987}. First, the horizontal contribution of the dianeutral mixing is neglected since the ratio between iso and dia-neutral diffusive coefficients is known to be several orders of magnitude smaller than unity. Second, the two isoneutral directions of diffusion are assumed to be independent since the slopes are generally less than $10^{-2}$ in the ocean (see Appendix~\ref{Apdx_B}). 
     1173case, two assumptions are made to simplify  \eqref{Eq_PE_iso_tensor} \citep{Cox1987}.  
     1174First, the horizontal contribution of the dianeutral mixing is neglected since the ratio  
     1175between iso and dia-neutral diffusive coefficients is known to be several orders of  
     1176magnitude smaller than unity. Second, the two isoneutral directions of diffusion are  
     1177assumed to be independent since the slopes are generally less than $10^{-2}$ in the  
     1178ocean (see Appendix~\ref{Apdx_B}). 
    8841179 
    8851180For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the  
    886 geopotential and computational surfaces: in $z$-coordinates they are zero ($r_1 = r_2 = 0$) while in $s$-coordinate (including $\textit{z*}$ case) they are equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ). 
    887  
    888 For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral and computational surfaces. Therefore, they have a same expression in $z$- and $s$-coordinates: 
     1181geopotential and computational surfaces: in $z$-coordinates they are zero  
     1182($r_1 = r_2 = 0$) while in $s$-coordinate (including $\textit{z*}$ case) they are  
     1183equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ). 
     1184 
     1185For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral  
     1186and computational surfaces. Therefore, they have a same expression in $z$- and $s$-coordinates: 
    8891187\begin{equation} \label{Eq_PE_iso_slopes} 
    8901188r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right) 
     
    8941192\end{equation} 
    8951193 
    896 When the \textit{Eddy Induced Velocity} parametrisation (eiv) \citep{Gent1990} is used, an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers: 
     1194When the \textit{Eddy Induced Velocity} parametrisation (eiv) \citep{Gent1990} is used,  
     1195an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers: 
    8971196\begin{equation} \label{Eq_PE_iso+eiv} 
    8981197D^{lT}=\nabla \cdot \left( {A^{lT}\;\Re \;\nabla T} \right) 
    8991198           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right) 
    9001199\end{equation} 
    901 where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent, eddy-induced transport velocity. This velocity field is defined by: 
     1200where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent,  
     1201eddy-induced transport velocity. This velocity field is defined by: 
    9021202\begin{equation} \label{Eq_PE_eiv} 
    9031203   \begin{split} 
     
    9091209   \end{split} 
    9101210\end{equation} 
    911 where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes between isoneutral and \emph{geopotential} surfaces and thus depends on the coordinate considered:  
     1211where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral  
     1212thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes  
     1213between isoneutral and \emph{geopotential} surfaces and thus depends on the coordinate  
     1214considered:  
    9121215\begin{align} \label{Eq_PE_slopes_eiv} 
    9131216\tilde{r}_n = \begin{cases} 
     
    9181221\end{align} 
    9191222 
    920 The normal component of the eddy induced velocity is zero at all the boundaries. this can be achieved in a model by tapering either the eddy coefficient or the slopes to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. Chap.~\ref{LDF}). 
     1223The normal component of the eddy induced velocity is zero at all the boundaries.  
     1224This can be achieved in a model by tapering either the eddy coefficient or the slopes  
     1225to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. Chap.~\ref{LDF}). 
    9211226 
    9221227\subsubsection{lateral fourth order tracer diffusive operator} 
     
    9281233 \end{equation} 
    9291234 
    930 It is the second order operator given by \eqref{Eq_PE_iso_tensor} applied twice with the eddy diffusion coefficient correctly placed.  
     1235It is the second order operator given by \eqref{Eq_PE_iso_tensor} applied twice with  
     1236the eddy diffusion coefficient correctly placed.  
    9311237 
    9321238 
     
    9521258horizontal divergence fields (see Appendix~\ref{Apdx_C}). Unfortunately, it is not  
    9531259available for geopotential diffusion in $s-$coordinates and for isoneutral  
    954 diffusion in both $z$- and $s$-coordinates ($i.e.$ when a rotation is required). In these two cases, the $u$ and $v-$fields are considered as independent scalar fields, so that the diffusive operator is given by: 
     1260diffusion in both $z$- and $s$-coordinates ($i.e.$ when a rotation is required).  
     1261In these two cases, the $u$ and $v-$fields are considered as independent scalar  
     1262fields, so that the diffusive operator is given by: 
    9551263\begin{equation} \label{Eq_PE_lapU_iso} 
    9561264\begin{split} 
     
    9591267 \end{split} 
    9601268 \end{equation} 
    961 where $\Re$ is given by  \eqref{Eq_PE_iso_tensor}. It is the same expression as those used for diffusive operator on tracers. It must be emphasised that such a formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or $\beta-$plane, not on the sphere. It is also a very good approximation in vicinity of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}. 
     1269where $\Re$ is given by  \eqref{Eq_PE_iso_tensor}. It is the same expression as  
     1270those used for diffusive operator on tracers. It must be emphasised that such a  
     1271formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or  
     1272$\beta-$plane, not on the sphere. It is also a very good approximation in vicinity  
     1273of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}. 
    9621274 
    9631275\subsubsection{lateral fourth order momentum diffusive operator} 
    9641276 
    965 As for tracers, the fourth order momentum diffusive operator along $z$ or $s$-surfaces is a re-entering second order operator \eqref{Eq_PE_lapU} or \eqref{Eq_PE_lapU} with the eddy viscosity coefficient correctly placed: 
     1277As for tracers, the fourth order momentum diffusive operator along $z$ or $s$-surfaces  
     1278is a re-entering second order operator \eqref{Eq_PE_lapU} or \eqref{Eq_PE_lapU}  
     1279with the eddy viscosity coefficient correctly placed: 
    9661280 
    9671281geopotential diffusion in $z$-coordinate: 
Note: See TracChangeset for help on using the changeset viewer.