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 10354 for NEMO/trunk/doc/latex/NEMO/subfiles/chap_model_basics.tex – NEMO

Ignore:
Timestamp:
2018-11-21T17:59:55+01:00 (5 years ago)
Author:
nicolasmartin
Message:

Vast edition of LaTeX subfiles to improve the readability by cutting sentences in a more suitable way
Every sentence begins in a new line and if necessary is splitted around 110 characters lenght for side-by-side visualisation,
this setting may not be adequate for everyone but something has to be set.
The punctuation was the primer trigger for the cutting process, otherwise subordinators and coordinators, in order to mostly keep a meaning for each line

File:
1 edited

Legend:

Unmodified
Added
Removed
  • NEMO/trunk/doc/latex/NEMO/subfiles/chap_model_basics.tex

    r9414 r10354  
    2626 
    2727 
    28 The ocean is a fluid that can be described to a good approximation by the primitive  
    29 equations, $i.e.$ the Navier-Stokes equations along with a nonlinear equation of  
    30 state which couples the two active tracers (temperature and salinity) to the fluid  
    31 velocity, plus the following additional assumptions made from scale considerations: 
    32  
    33 \textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to  
    34 be spheres so that gravity (local vertical) is parallel to the earth's radius 
    35  
    36 \textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius 
    37  
    38 \textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect  
    39 of small scale processes on the large-scale) are expressed in terms of large-scale features 
    40  
    41 \textit{(4) Boussinesq hypothesis:} density variations are neglected except in their  
    42 contribution to the buoyancy force 
    43  
    44 \textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a  
    45 balance between the vertical pressure gradient and the buoyancy force (this removes  
    46 convective processes from the initial Navier-Stokes equations and so convective processes  
    47 must be parameterized instead) 
    48  
    49 \textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity  
    50 vector is assumed to be zero. 
    51  
    52 Because the gravitational force is so dominant in the equations of large-scale motions,  
    53 it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked  
    54 to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two  
    55 vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define  
    56 the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$  
     28The ocean is a fluid that can be described to a good approximation by the primitive equations, 
     29$i.e.$ the Navier-Stokes equations along with a nonlinear equation of state which 
     30couples the two active tracers (temperature and salinity) to the fluid velocity, 
     31plus the following additional assumptions made from scale considerations: 
     32 
     33\textit{(1) spherical earth approximation:} the geopotential surfaces are assumed to be spheres so that 
     34gravity (local vertical) is parallel to the earth's radius 
     35 
     36\textit{(2) thin-shell approximation:} the ocean depth is neglected compared to the earth's radius 
     37 
     38\textit{(3) turbulent closure hypothesis:} the turbulent fluxes 
     39(which represent the effect of small scale processes on the large-scale) are expressed in terms of 
     40large-scale features 
     41 
     42\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their contribution to 
     43the buoyancy force 
     44 
     45\textit{(5) Hydrostatic hypothesis:} the vertical momentum equation is reduced to a balance between 
     46the vertical pressure gradient and the buoyancy force 
     47(this removes convective processes from the initial Navier-Stokes equations and so 
     48convective processes must be parameterized instead) 
     49 
     50\textit{(6) Incompressibility hypothesis:} the three dimensional divergence of the velocity vector is assumed to 
     51be zero. 
     52 
     53Because the gravitational force is so dominant in the equations of large-scale motions, 
     54it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked to 
     55the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to 
     56\textbf{k}, $i.e.$ tangent to the geopotential surfaces. 
     57Let us define the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$  
    5758(the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane),  
    58 $T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density.  
    59 The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k})  
    60 vector system provides the following six equations (namely the momentum balance, the  
    61 hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation  
    62 equations and an equation of state): 
     59$T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density. 
     60The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k}) vector system 
     61provides the following six equations 
     62(namely the momentum balance, the hydrostatic equilibrium, the incompressibility equation, 
     63the heat and salt conservation equations and an equation of state): 
    6364\begin{subequations} \label{eq:PE} 
    6465  \begin{equation}     \label{eq:PE_dyn} 
     
    8586  \end{equation} 
    8687\end{subequations} 
    87 where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions,  
    88 $t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by  
    89 the equation of state (\autoref{eq:PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,  
    90 $f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's  
    91 angular velocity vector), and $g$ is the gravitational acceleration.  
    92 ${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterisations of small-scale  
    93 physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$  
    94 and $F^S$ surface forcing terms. Their nature and formulation are discussed in  
    95 \autoref{sec:PE_zdf_ldf} and \autoref{subsec:PE_boundary_condition}. 
    96  
    97 . 
     88where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions, $t$ is the time, 
     89$z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by the equation of state 
     90(\autoref{eq:PE_eos}), $\rho_o$ is a reference density, $p$ the pressure, 
     91$f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration 
     92(where $\bf \Omega$ is the Earth's angular velocity vector), and $g$ is the gravitational acceleration. 
     93${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterisations of small-scale physics for momentum, 
     94temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$ surface forcing terms. 
     95Their nature and formulation are discussed in \autoref{sec:PE_zdf_ldf} and \autoref{subsec:PE_boundary_condition}. 
     96 
     97 
    9898 
    9999% ------------------------------------------------------------------------------------------------------------- 
     
    103103\label{subsec:PE_boundary_condition} 
    104104 
    105 An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea  
    106 or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$  
    107 and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height  
    108 of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$,  
    109 chosen as a mean sea surface (\autoref{fig:ocean_bc}). Through these two boundaries,  
    110 the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth,  
    111 the continental margins, the sea ice and the atmosphere. However, some of these fluxes are  
    112 so weak that even on climatic time scales of thousands of years they can be neglected.  
    113 In the following, we briefly review the fluxes exchanged at the interfaces between the ocean  
    114 and the other components of the earth system. 
     105An ocean is bounded by complex coastlines, bottom topography at its base and 
     106an air-sea or ice-sea interface at its top. 
     107These boundaries can be defined by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, 
     108where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface. 
     109Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$, chosen as a mean sea surface 
     110(\autoref{fig:ocean_bc}). 
     111Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt, and momentum with 
     112the solid earth, the continental margins, the sea ice and the atmosphere. 
     113However, some of these fluxes are so weak that even on climatic time scales of thousands of years 
     114they can be neglected. 
     115In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and 
     116the other components of the earth system. 
    115117 
    116118%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    117 \begin{figure}[!ht]   \begin{center} 
    118 \includegraphics[width=0.90\textwidth]{Fig_I_ocean_bc} 
    119 \caption{    \protect\label{fig:ocean_bc}  
    120 The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,t)$, where $H$  
    121 is the depth of the sea floor and $\eta$ the height of the sea surface.  
    122 Both $H$ and $\eta$ are referenced to $z=0$.} 
    123 \end{center}   \end{figure} 
     119\begin{figure}[!ht] 
     120  \begin{center} 
     121    \includegraphics[width=0.90\textwidth]{Fig_I_ocean_bc} 
     122    \caption{   \protect\label{fig:ocean_bc} 
     123      The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,t)$, 
     124      where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface. 
     125      Both $H$ and $\eta$ are referenced to $z=0$. 
     126    } 
     127  \end{center} 
     128\end{figure} 
    124129%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    125130 
    126131 
    127132\begin{description} 
    128 \item[Land - ocean interface:] the major flux between continental margins and the ocean is  
    129 a mass exchange of fresh water through river runoff. Such an exchange modifies the sea  
    130 surface salinity especially in the vicinity of major river mouths. It can be neglected for short  
    131 range integrations but has to be taken into account for long term integrations as it influences  
    132 the characteristics of water masses formed (especially at high latitudes). It is required in order  
    133 to close the water cycle of the climate system. It is usually specified as a fresh water flux at  
    134 the air-sea interface in the vicinity of river mouths. 
    135 \item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small,  
    136 except in special areas of little extent. They are usually neglected in the model  
    137 \footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling  
    138 ($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world  
    139 ocean (see \autoref{subsec:TRA_bbc}).}.  
    140 The boundary condition is thus set to no flux of heat and salt across solid boundaries.  
    141 For momentum, the situation is different. There is no flow across solid boundaries,  
    142 $i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words,  
    143 the bottom velocity is parallel to solid boundaries). This kinematic boundary condition  
    144 can be expressed as: 
    145 \begin{equation} \label{eq:PE_w_bbc} 
    146 w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right) 
    147 \end{equation} 
    148 In addition, the ocean exchanges momentum with the earth through frictional processes.  
    149 Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized  
    150 in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification  
    151 depends on the nature of the physical parameterisation used for ${\rm {\bf D}}^{\rm {\bf U}}$  
    152 in \autoref{eq:PE_dyn}. It is discussed in \autoref{eq:PE_zdf}.% and Chap. III.6 to 9. 
    153 \item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux  
    154 of fresh water PE  (the precipitation minus evaporation budget) leads to:  
    155 \begin{equation} \label{eq:PE_w_sbc} 
    156 w = \frac{\partial \eta }{\partial t}  
    157     + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right)  
     133\item[Land - ocean interface:] 
     134  the major flux between continental margins and the ocean is a mass exchange of fresh water through river runoff. 
     135  Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths. 
     136  It can be neglected for short range integrations but has to be taken into account for long term integrations as 
     137  it influences the characteristics of water masses formed (especially at high latitudes). 
     138  It is required in order to close the water cycle of the climate system. 
     139  It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths. 
     140\item[Solid earth - ocean interface:] 
     141  heat and salt fluxes through the sea floor are small, except in special areas of little extent. 
     142  They are usually neglected in the model \footnote{ 
     143    In fact, it has been shown that the heat flux associated with the solid Earth cooling 
     144    ($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world ocean 
     145    (see \autoref{subsec:TRA_bbc}). 
     146  }. 
     147  The boundary condition is thus set to no flux of heat and salt across solid boundaries. 
     148  For momentum, the situation is different. There is no flow across solid boundaries, 
     149  $i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words, 
     150  the bottom velocity is parallel to solid boundaries). This kinematic boundary condition 
     151  can be expressed as: 
     152  \begin{equation} \label{eq:PE_w_bbc} 
     153    w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right) 
     154  \end{equation} 
     155  In addition, the ocean exchanges momentum with the earth through frictional processes. 
     156  Such momentum transfer occurs at small scales in a boundary layer. 
     157  It must be parameterized in terms of turbulent fluxes using bottom and/or lateral boundary conditions. 
     158  Its specification depends on the nature of the physical parameterisation used for 
     159  ${\rm {\bf D}}^{\rm {\bf U}}$ in \autoref{eq:PE_dyn}. 
     160  It is discussed in \autoref{eq:PE_zdf}.% and Chap. III.6 to 9. 
     161\item[Atmosphere - ocean interface:] 
     162  the kinematic surface condition plus the mass flux of fresh water PE (the precipitation minus evaporation budget) 
     163  leads to: 
     164  \begin{equation} \label{eq:PE_w_sbc} 
     165    w = \frac{\partial \eta }{\partial t} 
     166    + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right) 
    158167    + P-E 
    159 \end{equation} 
    160 The dynamic boundary condition, neglecting the surface tension (which removes capillary  
    161 waves from the system) leads to the continuity of pressure across the interface $z=\eta$.  
    162 The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat. 
    163 \item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water  
    164 and momentum. The sea surface temperature is constrained to be at the freezing point  
    165 at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the  
    166 ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and  
    167 salt fluxes that cannot be neglected. 
     168  \end{equation} 
     169  The dynamic boundary condition, neglecting the surface tension (which removes capillary waves from the system) 
     170  leads to the continuity of pressure across the interface $z=\eta$. 
     171  The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat. 
     172\item[Sea ice - ocean interface:] 
     173  the ocean and sea ice exchange heat, salt, fresh water and momentum. 
     174  The sea surface temperature is constrained to be at the freezing point at the interface. 
     175  Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the ocean ($\sim34 \,psu$). 
     176  The cycle of freezing/melting is associated with fresh water and salt fluxes that cannot be neglected. 
    168177\end{description} 
    169178 
     
    184193\label{subsec:PE_p_formulation} 
    185194 
    186 The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a  
    187 reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that:  
    188 $p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\autoref{eq:PE_hydrostatic}),  
    189 assuming that pressure in decibars can be approximated by depth in meters in (\autoref{eq:PE_eos}).  
     195The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at 
     196a reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that: 
     197$p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. 
     198The latter is computed by integrating (\autoref{eq:PE_hydrostatic}), 
     199assuming that pressure in decibars can be approximated by depth in meters in (\autoref{eq:PE_eos}). 
    190200The hydrostatic pressure is then given by: 
    191201\begin{equation} \label{eq:PE_pressure} 
     
    193203 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,\varsigma} \right)\;d\varsigma }  
    194204\end{equation} 
    195  Two strategies can be considered for the surface pressure term: $(a)$ introduce of a  
    196  new variable $\eta$, the free-surface elevation, for which a prognostic equation can be  
    197  established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the  
    198  pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used,  
    199  one solution of the free-surface elevation consists of the excitation of external gravity waves.  
    200  The flow is barotropic and the surface moves up and down with gravity as the restoring force.  
    201  The phase speed of such waves is high (some hundreds of metres per second) so that  
    202  the time step would have to be very short if they were present in the model. The latter  
    203  strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$  
    204  the sea surface is the surface $z=0$. This well known approximation increases the surface  
    205  wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic  
    206  Rossby or planetary waves). The rigid-lid hypothesis is an obsolescent feature in modern  
    207  OGCMs. It has been available until the release 3.1 of  \NEMO, and it has been removed 
    208  in release 3.2 and followings. Only the free surface formulation is now described in the 
    209  this document (see the next sub-section). 
     205Two strategies can be considered for the surface pressure term: 
     206$(a)$ introduce of a  new variable $\eta$, the free-surface elevation, 
     207for which a prognostic equation can be established and solved; 
     208$(b)$ assume that the ocean surface is a rigid lid, 
     209on which the pressure (or its horizontal gradient) can be diagnosed. 
     210When the former strategy is used, one solution of the free-surface elevation consists of 
     211the excitation of external gravity waves. 
     212The flow is barotropic and the surface moves up and down with gravity as the restoring force. 
     213The phase speed of such waves is high (some hundreds of metres per second) so that 
     214the time step would have to be very short if they were present in the model. 
     215The latter strategy filters out these waves since the rigid lid approximation implies $\eta=0$, 
     216$i.e.$ the sea surface is the surface $z=0$. 
     217This well known approximation increases the surface wave speed to infinity and 
     218modifies certain other longwave dynamics ($e.g.$ barotropic Rossby or planetary waves). 
     219The rigid-lid hypothesis is an obsolescent feature in modern OGCMs. 
     220It has been available until the release 3.1 of  \NEMO, and it has been removed in release 3.2 and followings. 
     221Only the free surface formulation is now described in the this document (see the next sub-section). 
    210222 
    211223% ------------------------------------------------------------------------------------------------------------- 
     
    215227\label{subsec:PE_free_surface} 
    216228 
    217 In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced  
    218 which describes the shape of the air-sea interface. This variable is solution of a  
    219 prognostic equation which is established by forming the vertical average of the kinematic  
    220 surface condition (\autoref{eq:PE_w_bbc}): 
     229In the free surface formulation, a variable $\eta$, the sea-surface height, 
     230is introduced which describes the shape of the air-sea interface. 
     231This variable is solution of a prognostic equation which is established by forming the vertical average of 
     232the kinematic surface condition (\autoref{eq:PE_w_bbc}): 
    221233\begin{equation} \label{eq:PE_ssh} 
    222234\frac{\partial \eta }{\partial t}=-D+P-E 
     
    226238and using (\autoref{eq:PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$. 
    227239 
    228 Allowing the air-sea interface to move introduces the external gravity waves (EGWs)  
    229 as a class of solution of the primitive equations. These waves are barotropic because  
    230 of hydrostatic assumption, and their phase speed is quite high. Their time scale is  
    231 short with respect to the other processes described by the primitive equations. 
    232  
    233 Two choices can be made regarding the implementation of the free surface in the model,  
     240Allowing the air-sea interface to move introduces the external gravity waves (EGWs) as 
     241a class of solution of the primitive equations. 
     242These waves are barotropic because of hydrostatic assumption, and their phase speed is quite high. 
     243Their time scale is short with respect to the other processes described by the primitive equations. 
     244 
     245Two choices can be made regarding the implementation of the free surface in the model, 
    234246depending on the physical processes of interest.  
    235247 
    236 $\bullet$ If one is interested in EGWs, in particular the tides and their interaction  
    237 with the baroclinic structure of the ocean (internal waves) possibly in shallow seas,  
    238 then a non linear free surface is the most appropriate. This means that no  
    239 approximation is made in (\autoref{eq:PE_ssh}) and that the variation of the ocean  
    240 volume is fully taken into account. Note that in order to study the fast time scales  
    241 associated with EGWs it is necessary to minimize time filtering effects (use an  
    242 explicit time scheme with very small time step, or a split-explicit scheme with  
    243 reasonably small time step, see \autoref{subsec:DYN_spg_exp} or \autoref{subsec:DYN_spg_ts}. 
    244  
    245 $\bullet$ If one is not interested in EGW but rather sees them as high frequency  
    246 noise, it is possible to apply an explicit filter to slow down the fastest waves while  
    247 not altering the slow barotropic Rossby waves. If further, an approximative conservation  
    248 of heat and salt contents is sufficient for the problem solved, then it is  
    249 sufficient to solve a linearized version of (\autoref{eq:PE_ssh}), which still allows  
    250 to take into account freshwater fluxes applied at the ocean surface \citep{Roullet_Madec_JGR00}. 
     248$\bullet$ If one is interested in EGWs, in particular the tides and their interaction with 
     249the baroclinic structure of the ocean (internal waves) possibly in shallow seas, 
     250then a non linear free surface is the most appropriate. 
     251This means that no approximation is made in (\autoref{eq:PE_ssh}) and that 
     252the variation of the ocean volume is fully taken into account. 
     253Note that in order to study the fast time scales associated with EGWs it is necessary to 
     254minimize time filtering effects 
     255(use an explicit time scheme with very small time step, or a split-explicit scheme with reasonably small time step, 
     256see \autoref{subsec:DYN_spg_exp} or \autoref{subsec:DYN_spg_ts}). 
     257 
     258$\bullet$ If one is not interested in EGW but rather sees them as high frequency noise, 
     259it is possible to apply an explicit filter to slow down the fastest waves while 
     260not altering the slow barotropic Rossby waves. 
     261If further, an approximative conservation of heat and salt contents is sufficient for the problem solved, 
     262then it is sufficient to solve a linearized version of (\autoref{eq:PE_ssh}), 
     263which still allows to take into account freshwater fluxes applied at the ocean surface \citep{Roullet_Madec_JGR00}. 
    251264Nevertheless, with the linearization, an exact conservation of heat and salt contents is lost. 
    252265 
    253 The filtering of EGWs in models with a free surface is usually a matter of discretisation  
    254 of the temporal derivatives, using a split-explicit method \citep{Killworth_al_JPO91, Zhang_Endoh_JGR92}  
    255 or the implicit scheme \citep{Dukowicz1994} or the addition of a filtering force in the momentum equation  
    256 \citep{Roullet_Madec_JGR00}. With the present release, \NEMO offers the choice between  
    257 an explicit free surface (see \autoref{subsec:DYN_spg_exp}) or a split-explicit scheme strongly  
    258 inspired the one proposed by \citet{Shchepetkin_McWilliams_OM05} (see \autoref{subsec:DYN_spg_ts}). 
     266The filtering of EGWs in models with a free surface is usually a matter of discretisation of 
     267the temporal derivatives, 
     268using a split-explicit method \citep{Killworth_al_JPO91, Zhang_Endoh_JGR92} or 
     269the implicit scheme \citep{Dukowicz1994} or 
     270the addition of a filtering force in the momentum equation \citep{Roullet_Madec_JGR00}. 
     271With the present release, \NEMO offers the choice between 
     272an explicit free surface (see \autoref{subsec:DYN_spg_exp}) or 
     273a split-explicit scheme strongly inspired the one proposed by \citet{Shchepetkin_McWilliams_OM05} 
     274(see \autoref{subsec:DYN_spg_ts}). 
    259275 
    260276%\newpage 
     
    274290\label{subsec:PE_tensorial} 
    275291 
    276 In many ocean circulation problems, the flow field has regions of enhanced dynamics  
    277 ($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts).  
    278 The representation of such dynamical processes can be improved by specifically increasing  
    279 the model resolution in these regions. As well, it may be convenient to use a lateral  
    280 boundary-following coordinate system to better represent coastal dynamics. Moreover,  
    281 the common geographical coordinate system has a singular point at the North Pole that  
    282 cannot be easily treated in a global model without filtering. A solution consists of introducing  
    283 an appropriate coordinate transformation that shifts the singular point onto land  
    284 \citep{Madec_Imbard_CD96, Murray_JCP96}. As a consequence, it is important to solve the primitive  
    285 equations in various curvilinear coordinate systems. An efficient way of introducing an  
    286 appropriate coordinate transform can be found when using a tensorial formalism.  
    287 This formalism is suited to any multidimensional curvilinear coordinate system. Ocean  
    288 modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth  
    289 approximation), with preservation of the local vertical. Here we give the simplified equations  
    290 for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey  
    291 of the conservation laws of fluid dynamics. 
    292  
    293 Let (\textit{i},\textit{j},\textit{k}) be a set of orthogonal curvilinear coordinates on the sphere  
    294 associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k})  
    295 linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are  
    296 two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (\autoref{fig:referential}).  
    297 Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined  
    298 by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of  
    299 the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea  
    300 level (\autoref{fig:referential}). The local deformation of the curvilinear coordinate system is  
    301 given by $e_1$, $e_2$ and $e_3$, the three scale factors: 
     292In many ocean circulation problems, the flow field has regions of enhanced dynamics 
     293($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts). 
     294The representation of such dynamical processes can be improved by 
     295specifically increasing the model resolution in these regions. 
     296As well, it may be convenient to use a lateral boundary-following coordinate system to 
     297better represent coastal dynamics. 
     298Moreover, the common geographical coordinate system has a singular point at the North Pole that 
     299cannot be easily treated in a global model without filtering. 
     300A solution consists of introducing an appropriate coordinate transformation that 
     301shifts the singular point onto land \citep{Madec_Imbard_CD96, Murray_JCP96}. 
     302As a consequence, it is important to solve the primitive equations in various curvilinear coordinate systems. 
     303An efficient way of introducing an appropriate coordinate transform can be found when using a tensorial formalism. 
     304This formalism is suited to any multidimensional curvilinear coordinate system. 
     305Ocean modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth approximation), 
     306with preservation of the local vertical. Here we give the simplified equations for this particular case. 
     307The general case is detailed by \citet{Eiseman1980} in their survey of the conservation laws of fluid dynamics. 
     308 
     309Let (\textit{i},\textit{j},\textit{k}) be a set of orthogonal curvilinear coordinates on 
     310the sphere associated with the positively oriented orthogonal set of unit vectors 
     311(\textbf{i},\textbf{j},\textbf{k}) linked to the earth such that 
     312\textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to \textbf{k}, 
     313$i.e.$ along geopotential surfaces (\autoref{fig:referential}). 
     314Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by 
     315the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and 
     316the distance from the centre of the earth $a+z(k)$ where $a$ is the earth's radius and 
     317$z$ the altitude above a reference sea level (\autoref{fig:referential}). 
     318The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$, 
     319the three scale factors: 
    302320\begin{equation} \label{eq:scale_factors} 
    303 \begin{aligned} 
    304  e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda  
    305 }{\partial i}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi  
    306 }{\partial i}} \right)^2} \right]^{1/2} \\  
    307  e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda  
    308 }{\partial j}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi  
    309 }{\partial j}} \right)^2} \right]^{1/2} \\  
    310  e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\  
    311  \end{aligned} 
    312  \end{equation} 
    313  
     321  \begin{aligned} 
     322    e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda}{\partial i}\cos \varphi } \right)^2 
     323        +\left( {\frac{\partial \varphi }{\partial i}} \right)^2} \right]^{1/2} \\ 
     324    e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda }{\partial j}\cos \varphi } \right)^2+ 
     325        \left( {\frac{\partial \varphi }{\partial j}} \right)^2} \right]^{1/2} \\ 
     326    e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\ 
     327  \end{aligned} 
     328\end{equation} 
     329 
     330% >>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     331\begin{figure}[!tb] 
     332  \begin{center} 
     333    \includegraphics[width=0.60\textwidth]{Fig_I_earth_referential} 
     334    \caption{  \protect\label{fig:referential} 
     335      the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear 
     336      coordinate system (\textbf{i},\textbf{j},\textbf{k}). } 
     337  \end{center} 
     338\end{figure} 
    314339%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    315 \begin{figure}[!tb]   \begin{center} 
    316 \includegraphics[width=0.60\textwidth]{Fig_I_earth_referential} 
    317 \caption{   \protect\label{fig:referential}  
    318 the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear  
    319 coordinate system (\textbf{i},\textbf{j},\textbf{k}). } 
    320 \end{center}   \end{figure} 
    321 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    322  
    323 Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by  
    324 $a$ in (\autoref{eq:scale_factors}) (thin-shell approximation). The resulting horizontal scale  
    325 factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single  
    326 function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that  
    327 appear in the primitive equations (\autoref{eq:PE_dyn} to \autoref{eq:PE_eos}) can  
    328 be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate  
    329 system transformation: 
     340 
     341Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by $a$ in 
     342(\autoref{eq:scale_factors}) (thin-shell approximation). 
     343The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while 
     344the vertical scale factor is a single function of $k$ as \textbf{k} is parallel to \textbf{z}. 
     345The scalar and vector operators that appear in the primitive equations 
     346(\autoref{eq:PE_dyn} to \autoref{eq:PE_eos}) can be written in the tensorial form, 
     347invariant in any orthogonal horizontal curvilinear coordinate system transformation: 
    330348\begin{subequations} \label{eq:PE_discrete_operators} 
    331349\begin{equation} \label{eq:PE_grad} 
     
    369387\label{subsec:PE_zco_Eq} 
    370388 
    371 In order to express the Primitive Equations in tensorial formalism, it is necessary to compute  
    372 the horizontal component of the non-linear and viscous terms of the equation using  
    373 \autoref{eq:PE_grad}) to \autoref{eq:PE_lap_vector}.  
    374 Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate  
    375 system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity  
    376 field $\chi$, by: 
     389In order to express the Primitive Equations in tensorial formalism, 
     390it is necessary to compute the horizontal component of the non-linear and viscous terms of the equation using 
     391\autoref{eq:PE_grad}) to \autoref{eq:PE_lap_vector}. 
     392Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate system and 
     393define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$, by: 
    377394\begin{equation} \label{eq:PE_curl_Uh} 
    378395\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v}  
     
    386403\end{equation} 
    387404 
    388 Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$  
    389 and that $e_3$  is a function of the single variable $k$, the nonlinear term of  
    390 \autoref{eq:PE_dyn} can be transformed as follows: 
     405Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that 
     406$e_3$  is a function of the single variable $k$, 
     407the nonlinear term of \autoref{eq:PE_dyn} can be transformed as follows: 
    391408\begin{flalign*} 
    392409&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}} 
     
    426443\end{flalign*} 
    427444 
    428 The last term of the right hand side is obviously zero, and thus the nonlinear term of  
     445The last term of the right hand side is obviously zero, and thus the nonlinear term of 
    429446\autoref{eq:PE_dyn} is written in the $(i,j,k)$ coordinate system: 
    430447\begin{equation} \label{eq:PE_vector_form} 
     
    437454\end{equation} 
    438455 
    439 This is the so-called \textit{vector invariant form} of the momentum advection term.  
    440 For some purposes, it can be advantageous to write this term in the so-called flux form,  
    441 $i.e.$ to write it as the divergence of fluxes. For example, the first component of  
    442 \autoref{eq:PE_vector_form} (the $i$-component) is transformed as follows: 
     456This is the so-called \textit{vector invariant form} of the momentum advection term. 
     457For some purposes, it can be advantageous to write this term in the so-called flux form, 
     458$i.e.$ to write it as the divergence of fluxes. 
     459For example, the first component of \autoref{eq:PE_vector_form} (the $i$-component) is transformed as follows: 
    443460\begin{flalign*} 
    444461&{ \begin{array}{*{20}l} 
     
    525542\end{multline} 
    526543 
    527 The flux form has two terms, the first one is expressed as the divergence of momentum  
    528 fluxes (hence the flux form name given to this formulation) and the second one is due to  
    529 the curvilinear nature of the coordinate system used. The latter is called the \emph{metric}  
    530 term and can be viewed as a modification of the Coriolis parameter:  
     544The flux form has two terms, 
     545the first one is expressed as the divergence of momentum fluxes (hence the flux form name given to this formulation) 
     546and the second one is due to the curvilinear nature of the coordinate system used. 
     547The latter is called the \emph{metric} term and can be viewed as a modification of the Coriolis parameter:  
    531548\begin{equation} \label{eq:PE_cor+metric} 
    532549f \to f + \frac{1}{e_1\;e_2}  \left(  v \frac{\partial e_2}{\partial i} 
     
    534551\end{equation} 
    535552 
    536 Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$  
    537 and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of  
    538 the Coriolis parameter $f \to f+(u/a) \tan \varphi$. 
     553Note that in the case of geographical coordinate, 
     554$i.e.$ when $(i,j) \to (\lambda ,\varphi )$ and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, 
     555we recover the commonly used modification of the Coriolis parameter $f \to f+(u/a) \tan \varphi$. 
    539556 
    540557 
    541558$\ $\newline    % force a new ligne 
    542559 
    543 To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be  
    544 written in the following tensorial formalism: 
     560To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be written in 
     561the following tensorial formalism: 
    545562 
    546563\vspace{+10pt} 
     
    594611\end{multline} 
    595612\end{subequations} 
    596 where $\zeta$, the relative vorticity, is given by \autoref{eq:PE_curl_Uh} and $p_s $,  
    597 the surface pressure, is given by: 
     613where $\zeta$, the relative vorticity, is given by \autoref{eq:PE_curl_Uh} and 
     614$p_s $, the surface pressure, is given by: 
    598615\begin{equation} \label{eq:PE_spg} 
    599616p_s =  \rho \,g \,\eta  
    600617\end{equation} 
    601 with $\eta$ is solution of \autoref{eq:PE_ssh} 
     618with $\eta$ is solution of \autoref{eq:PE_ssh}. 
    602619 
    603620The vertical velocity and the hydrostatic pressure are diagnosed from the following equations: 
     
    628645\end{equation} 
    629646 
    630 The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale  
    631 parameterisation used. It will be defined in \autoref{eq:PE_zdf}. The nature and formulation of  
    632 ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed  
    633 in \autoref{chap:SBC}. 
     647The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale parameterisation used. 
     648It will be defined in \autoref{eq:PE_zdf}. 
     649The nature and formulation of ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, 
     650are discussed in \autoref{chap:SBC}. 
    634651 
    635652 
     
    642659\label{sec:PE_gco} 
    643660 
    644 The ocean domain presents a huge diversity of situation in the vertical. First the ocean surface is a time dependent surface (moving surface). Second the ocean floor depends on the geographical position, varying from more than 6,000 meters in abyssal trenches to zero at the coast. Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.  
    645 Therefore, in order to represent the ocean with respect to the first point a space and time dependent vertical coordinate that follows the variation of the sea surface height $e.g.$ an $z$*-coordinate; for the second point, a space variation to fit the change of bottom topography $e.g.$ a terrain-following or $\sigma$-coordinate; and for the third point, one will be tempted to use a space and time dependent coordinate that follows the isopycnal surfaces, $e.g.$ an isopycnic coordinate. 
    646  
    647 In order to satisfy two or more constrains one can even be tempted to mixed these coordinate systems, as in HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and $\sigma$ at the ocean bottom) \citep{Chassignet_al_JPO03}  or OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and $\sigma$-coordinate elsewhere) \citep{Madec_al_JPO96} among others. 
    648  
    649 In fact one is totally free to choose any space and time vertical coordinate by introducing an arbitrary vertical coordinate : 
     661The ocean domain presents a huge diversity of situation in the vertical. 
     662First the ocean surface is a time dependent surface (moving surface). 
     663Second the ocean floor depends on the geographical position, 
     664varying from more than 6,000 meters in abyssal trenches to zero at the coast. 
     665Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.  
     666Therefore, in order to represent the ocean with respect to 
     667the first point a space and time dependent vertical coordinate that follows the variation of the sea surface height 
     668$e.g.$ an $z$*-coordinate; 
     669for the second point, a space variation to fit the change of bottom topography 
     670$e.g.$ a terrain-following or $\sigma$-coordinate; 
     671and for the third point, one will be tempted to use a space and time dependent coordinate that 
     672follows the isopycnal surfaces, $e.g.$ an isopycnic coordinate. 
     673 
     674In order to satisfy two or more constrains one can even be tempted to mixed these coordinate systems, as in 
     675HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and $\sigma$ at 
     676the ocean bottom) \citep{Chassignet_al_JPO03} or 
     677OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and $\sigma$-coordinate elsewhere) 
     678\citep{Madec_al_JPO96} among others. 
     679 
     680In fact one is totally free to choose any space and time vertical coordinate by 
     681introducing an arbitrary vertical coordinate : 
    650682\begin{equation} \label{eq:PE_s} 
    651683s=s(i,j,k,t) 
    652684\end{equation} 
    653 with the restriction that the above equation gives a single-valued monotonic relationship between $s$ and $k$, when $i$, $j$ and $t$ are held fixed. \autoref{eq:PE_s} is a transformation from the $(i,j,k,t)$ coordinate system with independent variables into the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through \autoref{eq:PE_s}. 
    654 This so-called \textit{generalised vertical coordinate} \citep{Kasahara_MWR74} is in fact an Arbitrary Lagrangian--Eulerian (ALE) coordinate. Indeed, choosing an expression for $s$ is an arbitrary choice that determines which part of the vertical velocity (defined from a fixed referential) will cross the levels (Eulerian part) and which part will be used to move them (Lagrangian part). 
    655 The coordinate is also sometime referenced as an adaptive coordinate \citep{Hofmeister_al_OM09}, since the coordinate system is adapted in the course of the simulation. Its most often used implementation is via an ALE algorithm, in which a pure lagrangian step is followed by regridding and remapping steps, the later step implicitly embedding the vertical advection \citep{Hirt_al_JCP74, Chassignet_al_JPO03, White_al_JCP09}. Here we follow the \citep{Kasahara_MWR74} strategy : a regridding step (an update of the vertical coordinate) followed by an eulerian step with an explicit computation of vertical advection relative to the moving s-surfaces. 
     685with the restriction that the above equation gives a single-valued monotonic relationship between $s$ and $k$, 
     686when $i$, $j$ and $t$ are held fixed. 
     687\autoref{eq:PE_s} is a transformation from the $(i,j,k,t)$ coordinate system with independent variables into 
     688the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through 
     689\autoref{eq:PE_s}. 
     690This so-called \textit{generalised vertical coordinate} \citep{Kasahara_MWR74} is in fact 
     691an Arbitrary Lagrangian--Eulerian (ALE) coordinate. 
     692Indeed, choosing an expression for $s$ is an arbitrary choice that determines 
     693which part of the vertical velocity (defined from a fixed referential) will cross the levels (Eulerian part) and 
     694which part will be used to move them (Lagrangian part). 
     695The coordinate is also sometime referenced as an adaptive coordinate \citep{Hofmeister_al_OM09}, 
     696since the coordinate system is adapted in the course of the simulation. 
     697Its most often used implementation is via an ALE algorithm, 
     698in which a pure lagrangian step is followed by regridding and remapping steps, 
     699the later step implicitly embedding the vertical advection 
     700\citep{Hirt_al_JCP74, Chassignet_al_JPO03, White_al_JCP09}. 
     701Here we follow the \citep{Kasahara_MWR74} strategy: 
     702a regridding step (an update of the vertical coordinate) followed by an eulerian step with 
     703an explicit computation of vertical advection relative to the moving s-surfaces. 
    656704 
    657705%\gmcomment{ 
     
    659707%A key point here is that the $s$-coordinate depends on $(i,j)$ ==> horizontal pressure gradient... 
    660708 
    661 the generalized vertical coordinates used in ocean modelling are not orthogonal,  
    662 which contrasts with many other applications in mathematical physics.  
    663 Hence, it is useful to keep in mind the following properties that may seem  
    664 odd on initial encounter. 
    665  
    666 The horizontal velocity in ocean models measures motions in the horizontal plane,  
    667 perpendicular to the local gravitational field. That is, horizontal velocity is mathematically  
    668 the same regardless the vertical coordinate, be it geopotential, isopycnal, pressure,  
    669 or terrain following. The key motivation for maintaining the same horizontal velocity  
    670 component is that the hydrostatic and geostrophic balances are dominant in the large-scale ocean.  
    671 Use of an alternative quasi-horizontal velocity, for example one oriented parallel  
    672 to the generalized surface, would lead to unacceptable numerical errors.  
    673 Correspondingly, the vertical direction is anti-parallel to the gravitational force in all  
    674 of the coordinate systems. We do not choose the alternative of a quasi-vertical  
    675 direction oriented normal to the surface of a constant generalized vertical coordinate.  
    676  
    677 It is the method used to measure transport across the generalized vertical coordinate  
    678 surfaces which differs between the vertical coordinate choices. That is, computation  
    679 of the dia-surface velocity component represents the fundamental distinction between  
    680 the various coordinates. In some models, such as geopotential, pressure, and  
    681 terrain following, this transport is typically diagnosed from volume or mass conservation.  
    682 In other models, such as isopycnal layered models, this transport is prescribed based  
    683 on assumptions about the physical processes producing a flux across the layer interfaces.  
    684  
    685  
    686 In this section we first establish the PE in the generalised vertical $s$-coordinate,  
     709the generalized vertical coordinates used in ocean modelling are not orthogonal, 
     710which contrasts with many other applications in mathematical physics. 
     711Hence, it is useful to keep in mind the following properties that may seem odd on initial encounter. 
     712 
     713The horizontal velocity in ocean models measures motions in the horizontal plane, 
     714perpendicular to the local gravitational field. 
     715That is, horizontal velocity is mathematically the same regardless the vertical coordinate, be it geopotential, 
     716isopycnal, pressure, or terrain following. 
     717The key motivation for maintaining the same horizontal velocity component is that 
     718the hydrostatic and geostrophic balances are dominant in the large-scale ocean. 
     719Use of an alternative quasi-horizontal velocity, for example one oriented parallel to the generalized surface, 
     720would lead to unacceptable numerical errors. 
     721Correspondingly, the vertical direction is anti-parallel to the gravitational force in 
     722all of the coordinate systems. 
     723We do not choose the alternative of a quasi-vertical direction oriented normal to 
     724the surface of a constant generalized vertical coordinate.  
     725 
     726It is the method used to measure transport across the generalized vertical coordinate surfaces which differs between 
     727the vertical coordinate choices. 
     728That is, computation of the dia-surface velocity component represents the fundamental distinction between 
     729the various coordinates. 
     730In some models, such as geopotential, pressure, and terrain following, this transport is typically diagnosed from 
     731volume or mass conservation. 
     732In other models, such as isopycnal layered models, this transport is prescribed based on assumptions about 
     733the physical processes producing a flux across the layer interfaces.  
     734 
     735 
     736In this section we first establish the PE in the generalised vertical $s$-coordinate, 
    687737then we discuss the particular cases available in \NEMO, namely $z$, $z$*, $s$, and $\tilde z$.   
    688738%} 
     
    693743\subsection{\textit{S-}coordinate formulation} 
    694744 
    695 Starting from the set of equations established in \autoref{sec:PE_zco} for the special case $k=z$  
    696 and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes  
    697 $z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and  
    698 $s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed  
    699 equations is given in \autoref{apdx:A}. Let us define the vertical scale factor by  
    700 $e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the  
    701 (\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by : 
     745Starting from the set of equations established in \autoref{sec:PE_zco} for the special case $k=z$ and thus $e_3=1$, 
     746we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, 
     747which includes $z$-, \textit{z*}- and $\sigma-$coordinates as special cases 
     748($s=z$, $s=\textit{z*}$, and $s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). 
     749A formal derivation of the transformed equations is given in \autoref{apdx:A}. 
     750Let us define the vertical scale factor by $e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), 
     751and the slopes in the (\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by: 
    702752\begin{equation} \label{eq:PE_sco_slope} 
    703753\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s  
     
    711761\end{equation} 
    712762 
    713 The equations solved by the ocean model \autoref{eq:PE} in $s-$coordinate can be written as follows (see \autoref{sec:A_momentum}): 
     763The equations solved by the ocean model \autoref{eq:PE} in $s-$coordinate can be written as follows 
     764(see \autoref{sec:A_momentum}): 
    714765 
    715766 \vspace{0.5cm} 
     
    763814\end{multline} 
    764815 
    765 where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic  
    766 pressure have the same expressions as in $z$-coordinates although they do not represent  
    767 exactly the same quantities. $\omega$ is provided by the continuity equation  
    768 (see \autoref{apdx:A}): 
     816where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, 
     817and the hydrostatic pressure have the same expressions as in $z$-coordinates although 
     818they do not represent exactly the same quantities. 
     819$\omega$ is provided by the continuity equation (see \autoref{apdx:A}): 
    769820\begin{equation} \label{eq:PE_sco_continuity} 
    770821\frac{\partial e_3}{\partial t} + e_3 \; \chi + \frac{\partial \omega }{\partial s} = 0    
     
    791842\end{multline} 
    792843 
    793 The equation of state has the same expression as in $z$-coordinate, and similar expressions  
    794 are used for mixing and forcing terms. 
     844The equation of state has the same expression as in $z$-coordinate, 
     845and similar expressions are used for mixing and forcing terms. 
    795846 
    796847\gmcomment{ 
     
    808859 
    809860%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    810 \begin{figure}[!b]    \begin{center} 
    811 \includegraphics[width=1.0\textwidth]{Fig_z_zstar} 
    812 \caption{   \protect\label{fig:z_zstar}  
    813 (a) $z$-coordinate in linear free-surface case ;  
    814 (b) $z-$coordinate in non-linear free surface case ;  
    815 (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate  
    816 \citep{Adcroft_Campin_OM04} ).} 
    817 \end{center}   \end{figure} 
     861\begin{figure}[!b] 
     862  \begin{center} 
     863    \includegraphics[width=1.0\textwidth]{Fig_z_zstar} 
     864    \caption{  \protect\label{fig:z_zstar} 
     865      (a) $z$-coordinate in linear free-surface case ; 
     866      (b) $z-$coordinate in non-linear free surface case ; 
     867      (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate 
     868      \citep{Adcroft_Campin_OM04} ). 
     869    } 
     870  \end{center} 
     871\end{figure} 
    818872%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    819873 
    820874 
    821 In that case, the free surface equation is nonlinear, and the variations of volume are fully  
    822 taken into account. These coordinates systems is presented in a report \citep{Levier2007}  
    823 available on the \NEMO web site.  
     875In that case, the free surface equation is nonlinear, and the variations of volume are fully taken into account. 
     876These coordinates systems is presented in a report \citep{Levier2007} available on the \NEMO web site.  
    824877 
    825878%\gmcomment{ 
    826 The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation  
    827 which allows one to deal with large amplitude free-surface 
    828 variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In 
    829 the  \textit{z*} formulation, the variation of the column thickness due to sea-surface 
    830 undulations is not concentrated in the surface level, as in the $z$-coordinate formulation, 
    831 but is equally distributed over the full water column. Thus vertical 
    832 levels naturally follow sea-surface variations, with a linear attenuation with 
    833 depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in 
    834 fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent. 
    835 The definition and modified oceanic equations for the rescaled vertical coordinate 
    836  \textit{z*}, including the treatment of fresh-water flux at the surface, are 
    837 detailed in Adcroft and Campin (2004). The major points are summarized 
    838 here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as: 
     879The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation which allows one to 
     880deal with large amplitude free-surface variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. 
     881In the \textit{z*} formulation, 
     882the variation of the column thickness due to sea-surface undulations is not concentrated in the surface level, 
     883as in the $z$-coordinate formulation, but is equally distributed over the full water column. 
     884Thus vertical levels naturally follow sea-surface variations, with a linear attenuation with depth, 
     885as illustrated by figure fig.1c. 
     886Note that with a flat bottom, such as in fig.1c, the bottom-following $z$ coordinate and \textit{z*} are equivalent. 
     887The definition and modified oceanic equations for the rescaled vertical coordinate  \textit{z*}, 
     888including the treatment of fresh-water flux at the surface, are detailed in Adcroft and Campin (2004). 
     889The major points are summarized here. 
     890The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as: 
    839891\begin{equation} \label{eq:z-star} 
    840892H +  \textit{z*} = (H + z) / r \quad \text{and} \ \delta \textit{z*} = \delta z / r \quad \text{with} \ r = \frac{H+\eta} {H} 
    841893\end{equation}  
    842 Since the vertical displacement of the free surface is incorporated in the vertical 
    843 coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,   
    844 $\textit{z*} = 0$ and  $\textit{z*} = -H$ respectively. Also the divergence of the flow field  
    845 is no longer zero as shown by the continuity equation: 
     894Since the vertical displacement of the free surface is incorporated in the vertical coordinate \textit{z*}, 
     895the upper and lower boundaries are at fixed  \textit{z*} position, 
     896$\textit{z*} = 0$ and  $\textit{z*} = -H$ respectively. 
     897Also the divergence of the flow field is no longer zero as shown by the continuity equation: 
    846898\begin{equation*}  
    847899\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right) 
     
    853905% from MOM4p1 documentation 
    854906 
    855 To overcome problems with vanishing surface and/or bottom cells, we consider the  
    856 zstar coordinate  
     907To overcome problems with vanishing surface and/or bottom cells, we consider the zstar coordinate  
    857908\begin{equation} \label{eq:PE_} 
    858909   z^\star = H \left( \frac{z-\eta}{H+\eta} \right) 
    859910\end{equation} 
    860911 
    861 This coordinate is closely related to the "eta" coordinate used in many atmospheric  
    862 models (see Black (1994) for a review of eta coordinate atmospheric models). It  
    863 was originally used in ocean models by Stacey et al. (1995) for studies of tides  
    864 next to shelves, and it has been recently promoted by Adcroft and Campin (2004)  
    865 for global climate modelling. 
    866  
    867 The surfaces of constant $z^\star$ are quasi-horizontal. Indeed, the $z^\star$ coordinate reduces to $z$ when $\eta$ is zero. In general, when noting the large differences between  
    868 undulations of the bottom topography versus undulations in the surface height, it  
    869 is clear that surfaces constant $z^\star$ are very similar to the depth surfaces. These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to terrain following sigma models discussed in \autoref{subsec:PE_sco}.  
    870 Additionally, since $z^\star$ when $\eta = 0$, no flow is spontaneously generated in an  
    871 unforced ocean starting from rest, regardless the bottom topography. This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in  
    872 the presence of nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state, depending on the sophistication of the pressure  
    873 gradient solver. The quasi-horizontal nature of the coordinate surfaces also facilitates the implementation of neutral physics parameterizations in $z^\star$ models using  
    874 the same techniques as in $z$-models (see Chapters 13-16 of \cite{Griffies_Bk04}) for a  
    875 discussion of neutral physics in $z$-models, as well as \autoref{sec:LDF_slp}  
    876 in this document for treatment in \NEMO).  
    877  
    878 The range over which $z^\star$ varies is time independent $-H \leq z^\star \leq 0$. Hence, all  
    879 cells remain nonvanishing, so long as the surface height maintains $\eta > ?H$. This  
    880 is a minor constraint relative to that encountered on the surface height when using  
    881 $s = z$ or $s = z - \eta$.  
    882  
    883 Because $z^\star$ has a time independent range, all grid cells have static increments  
    884 ds, and the sum of the ver tical increments yields the time independent ocean  
    885 depth %·k ds = H.  
    886 The $z^\star$ coordinate is therefore invisible to undulations of the  
    887 free surface, since it moves along with the free surface. This proper ty means that  
    888 no spurious ver tical transpor t is induced across surfaces of constant $z^\star$ by the  
    889 motion of external gravity waves. Such spurious transpor t can be a problem in  
    890 z-models, especially those with tidal forcing. Quite generally, the time independent  
    891 range for the $z^\star$ coordinate is a very convenient proper ty that allows for a nearly  
    892 arbitrary ver tical resolution even in the presence of large amplitude fluctuations of  
    893 the surface height, again so long as $\eta > -H$.  
     912This coordinate is closely related to the "eta" coordinate used in many atmospheric models 
     913(see Black (1994) for a review of eta coordinate atmospheric models). 
     914It was originally used in ocean models by Stacey et al. (1995) for studies of tides next to shelves, 
     915and it has been recently promoted by Adcroft and Campin (2004) for global climate modelling. 
     916 
     917The surfaces of constant $z^\star$ are quasi-horizontal. 
     918Indeed, the $z^\star$ coordinate reduces to $z$ when $\eta$ is zero. 
     919In general, when noting the large differences between 
     920undulations of the bottom topography versus undulations in the surface height, 
     921it is clear that surfaces constant $z^\star$ are very similar to the depth surfaces. 
     922These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to 
     923terrain following sigma models discussed in \autoref{subsec:PE_sco}. 
     924Additionally, since $z^\star$ when $\eta = 0$, 
     925no flow is spontaneously generated in an unforced ocean starting from rest, regardless the bottom topography. 
     926This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in the presence of 
     927nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state, 
     928depending on the sophistication of the pressure gradient solver. 
     929The quasi-horizontal nature of the coordinate surfaces also facilitates the implementation of 
     930neutral physics parameterizations in $z^\star$ models using the same techniques as in $z$-models 
     931(see Chapters 13-16 of \cite{Griffies_Bk04}) for a discussion of neutral physics in $z$-models, 
     932as well as \autoref{sec:LDF_slp} in this document for treatment in \NEMO). 
     933 
     934The range over which $z^\star$ varies is time independent $-H \leq z^\star \leq 0$. 
     935Hence, all cells remain nonvanishing, so long as the surface height maintains $\eta > ?H$. 
     936This is a minor constraint relative to that encountered on the surface height when using $s = z$ or $s = z - \eta$.  
     937 
     938Because $z^\star$ has a time independent range, all grid cells have static increments ds, 
     939and the sum of the ver tical increments yields the time independent ocean depth. %·k ds = H.  
     940The $z^\star$ coordinate is therefore invisible to undulations of the free surface, 
     941since it moves along with the free surface. 
     942This proper ty means that no spurious vertical transport is induced across surfaces of constant $z^\star$ by 
     943the motion of external gravity waves. 
     944Such spurious transpor t can be a problem in z-models, especially those with tidal forcing. 
     945Quite generally, the time independent range for the $z^\star$ coordinate is a very convenient property that 
     946allows for a nearly arbitrary ver tical resolution even in the presence of large amplitude fluctuations of 
     947the surface height, again so long as $\eta > -H$. 
    894948 
    895949%end MOM doc %%% 
     
    909963\subsubsection{Introduction} 
    910964 
    911 Several important aspects of the ocean circulation are influenced by bottom topography.  
    912 Of course, the most important is that bottom topography determines deep ocean sub-basins,  
    913 barriers, sills and channels that strongly constrain the path of water masses, but more subtle  
    914 effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary  
    915 one along continental slopes. Topographic Rossby waves can be excited and can interact  
    916 with the mean current. In the $z-$coordinate system presented in the previous section  
    917 (\autoref{sec:PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is  
    918 discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom  
    919 and to large localized depth gradients associated with large localized vertical velocities.  
    920 The response to such a velocity field often leads to numerical dispersion effects.  
    921 One solution to strongly reduce this error is to use a partial step representation of bottom  
    922 topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.  
    923 Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)  
    924  
    925 The $s$-coordinate avoids the discretisation error in the depth field since the layers of  
    926 computation are gradually adjusted with depth to the ocean bottom. Relatively small  
    927 topographic features as well as  gentle, large-scale slopes of the sea floor in the deep  
    928 ocean, which would be ignored in typical $z$-model applications with the largest grid  
    929 spacing at greatest depths, can easily be represented (with relatively low vertical resolution).  
    930 A terrain-following model (hereafter $s-$model) also facilitates the modelling of the  
    931 boundary layer flows over a large depth range, which in the framework of the $z$-model  
    932 would require high vertical resolution over the whole depth range. Moreover, with a  
    933 $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface  
    934 as the only boundaries of the domain (nomore lateral boundary condition to specify).  
    935 Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a  
    936 homogeneous ocean, it has strong limitations as soon as stratification is introduced.  
    937 The main two problems come from the truncation error in the horizontal pressure  
    938 gradient and a possibly increased diapycnal diffusion. The horizontal pressure force  
    939 in $s$-coordinate consists of two terms (see \autoref{apdx:A}), 
     965Several important aspects of the ocean circulation are influenced by bottom topography. 
     966Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers, sills and 
     967channels that strongly constrain the path of water masses, but more subtle effects exist. 
     968For example, the topographic $\beta$-effect is usually larger than the planetary one along continental slopes. 
     969Topographic Rossby waves can be excited and can interact with the mean current. 
     970In the $z-$coordinate system presented in the previous section (\autoref{sec:PE_zco}), 
     971$z-$surfaces are geopotential surfaces. 
     972The bottom topography is discretised by steps. 
     973This often leads to a misrepresentation of a gradually sloping bottom and to 
     974large localized depth gradients associated with large localized vertical velocities. 
     975The response to such a velocity field often leads to numerical dispersion effects. 
     976One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of 
     977a full step one \cite{Pacanowski_Gnanadesikan_MWR98}. 
     978Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate). 
     979 
     980The $s$-coordinate avoids the discretisation error in the depth field since the layers of 
     981computation are gradually adjusted with depth to the ocean bottom. 
     982Relatively small topographic features as well as  gentle, large-scale slopes of the sea floor in the deep ocean, 
     983which would be ignored in typical $z$-model applications with the largest grid spacing at greatest depths, 
     984can easily be represented (with relatively low vertical resolution). 
     985A terrain-following model (hereafter $s-$model) also facilitates the modelling of the boundary layer flows over 
     986a large depth range, which in the framework of the $z$-model would require high vertical resolution over 
     987the whole depth range. 
     988Moreover, with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as 
     989the only boundaries of the domain (no more lateral boundary condition to specify). 
     990Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean, 
     991it has strong limitations as soon as stratification is introduced. 
     992The main two problems come from the truncation error in the horizontal pressure gradient and 
     993a possibly increased diapycnal diffusion. 
     994The horizontal pressure force in $s$-coordinate consists of two terms (see \autoref{apdx:A}), 
    940995 
    941996\begin{equation} \label{eq:PE_p_sco} 
     
    944999\end{equation} 
    9451000 
    946 The second term in \autoref{eq:PE_p_sco} depends on the tilt of the coordinate surface  
    947 and introduces a truncation error that is not present in a $z$-model. In the special case  
    948 of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),  
    949 \citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude  
    950 of this truncation error. It depends on topographic slope, stratification, horizontal and  
    951 vertical resolution, the equation of state, and the finite difference scheme. This error  
    952 limits the possible topographic slopes that a model can handle at a given horizontal  
    953 and vertical resolution. This is a severe restriction for large-scale applications using  
    954 realistic bottom topography. The large-scale slopes require high horizontal resolution,  
    955 and the computational cost becomes prohibitive. This problem can be at least partially  
    956 overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec_al_JPO96}. However, the definition of the model  
    957 domain vertical coordinate becomes then a non-trivial thing for a realistic bottom  
    958 topography: a envelope topography is defined in $s$-coordinate on which a full or  
    959 partial step bottom topography is then applied in order to adjust the model depth to  
    960 the observed one (see \autoref{sec:DOM_zgr}. 
    961  
    962 For numerical reasons a minimum of diffusion is required along the coordinate surfaces  
    963 of any finite difference model. It causes spurious diapycnal mixing when coordinate  
    964 surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as  
    965 well as for a $s$-model. However, density varies more strongly on $s-$surfaces than  
    966 on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal  
    967 diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a  
    968 $z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal  
    969 circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.  
    970 For example, imagine an isolated bump of topography in an ocean at rest with a horizontally  
    971 uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral  
    972 surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,  
    973 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  
    974 ($i.e.$ the main thermocline) \citep{Madec_al_JPO96}. An alternate solution consists of rotating  
    975 the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \autoref{subsec:PE_ldf}.  
    976 Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,  
     1001The second term in \autoref{eq:PE_p_sco} depends on the tilt of the coordinate surface and 
     1002introduces a truncation error that is not present in a $z$-model. 
     1003In the special case of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$), 
     1004\citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude of this truncation error. 
     1005It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state, 
     1006and the finite difference scheme. 
     1007This error limits the possible topographic slopes that a model can handle at 
     1008a given horizontal and vertical resolution. 
     1009This is a severe restriction for large-scale applications using realistic bottom topography. 
     1010The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive. 
     1011This problem can be at least partially overcome by mixing $s$-coordinate and 
     1012step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec_al_JPO96}. 
     1013However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for 
     1014a realistic bottom topography: 
     1015a envelope topography is defined in $s$-coordinate on which a full or 
     1016partial step bottom topography is then applied in order to adjust the model depth to the observed one 
     1017(see \autoref{sec:DOM_zgr}. 
     1018 
     1019For numerical reasons a minimum of diffusion is required along the coordinate surfaces of any finite difference model. 
     1020It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces. 
     1021This is the case for a $z$-model as well as for a $s$-model. 
     1022However, density varies more strongly on $s-$surfaces than on horizontal surfaces in regions of 
     1023large topographic slopes, implying larger diapycnal diffusion in a $s$-model than in a $z$-model. 
     1024Whereas such a diapycnal diffusion in a $z$-model tends to weaken horizontal density (pressure) gradients and thus 
     1025the horizontal circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation. 
     1026For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification. 
     1027Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography, 
     1028and thus will generate there a baroclinic eddy. 
     1029In contrast, the ocean will stay at rest in a $z$-model. 
     1030As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below 
     1031the strongly stratified portion of the water column ($i.e.$ the main thermocline) \citep{Madec_al_JPO96}. 
     1032An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces 
     1033(see \autoref{subsec:PE_ldf}). 
     1034Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large, 
    9771035strongly exceeding the stability limit of such a operator when it is discretized (see \autoref{chap:LDF}).  
    9781036 
    979 The $s-$coordinates introduced here \citep{Lott_al_OM90,Madec_al_JPO96} differ mainly in two  
    980 aspects from similar models:  it allows  a representation of bottom topography with mixed  
    981 full or partial step-like/terrain following topography ; It also offers a completely general  
    982 transformation, $s=s(i,j,z)$ for the vertical coordinate. 
     1037The $s-$coordinates introduced here \citep{Lott_al_OM90,Madec_al_JPO96} differ mainly in two aspects from 
     1038similar models: 
     1039it allows a representation of bottom topography with mixed full or partial step-like/terrain following topography; 
     1040It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate. 
    9831041 
    9841042 
     
    9911049 
    9921050The $\tilde{z}$-coordinate has been developed by \citet{Leclair_Madec_OM11}. 
    993 It is available in \NEMO since the version 3.4. Nevertheless, it is currently not robust enough  
    994 to be used in all possible configurations. Its use is therefore not recommended. 
     1051It is available in \NEMO since the version 3.4. 
     1052Nevertheless, it is currently not robust enough to be used in all possible configurations. 
     1053Its use is therefore not recommended. 
    9951054 
    9961055 
     
    10021061\label{sec:PE_zdf_ldf} 
    10031062 
    1004 The primitive equations describe the behaviour of a geophysical fluid at  
    1005 space and time scales larger than a few kilometres in the horizontal, a few  
    1006 meters in the vertical and a few minutes. They are usually solved at larger  
    1007 scales: the specified grid spacing and time step of the numerical model. The  
    1008 effects of smaller scale motions (coming from the advective terms in the  
    1009 Navier-Stokes equations) must be represented entirely in terms of  
    1010 large-scale patterns to close the equations. These effects appear in the  
    1011 equations as the divergence of turbulent fluxes ($i.e.$ fluxes associated with  
    1012 the mean correlation of small scale perturbations). Assuming a turbulent  
    1013 closure hypothesis is equivalent to choose a formulation for these fluxes.  
    1014 It is usually called the subgrid scale physics. It must be emphasized that  
    1015 this is the weakest part of the primitive equations, but also one of the  
    1016 most important for long-term simulations as small scale processes \textit{in fine}  
    1017 balance the surface input of kinetic energy and heat. 
    1018  
    1019 The control exerted by gravity on the flow induces a strong anisotropy  
    1020 between the lateral and vertical motions. Therefore subgrid-scale physics   
    1021 \textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \autoref{eq:PE_dyn},  
    1022 \autoref{eq:PE_tra_T} and \autoref{eq:PE_tra_S} are divided into a lateral part   
    1023 \textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part   
    1024 \textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms  
    1025 and their underlying physics are briefly discussed in the next two subsections. 
     1063The primitive equations describe the behaviour of a geophysical fluid at space and time scales larger than 
     1064a few kilometres in the horizontal, a few meters in the vertical and a few minutes. 
     1065They are usually solved at larger scales: the specified grid spacing and time step of the numerical model. 
     1066The effects of smaller scale motions (coming from the advective terms in the Navier-Stokes equations) must be represented entirely in terms of large-scale patterns to close the equations. 
     1067These effects appear in the equations as the divergence of turbulent fluxes 
     1068($i.e.$ fluxes associated with the mean correlation of small scale perturbations). 
     1069Assuming a turbulent closure hypothesis is equivalent to choose a formulation for these fluxes. 
     1070It is usually called the subgrid scale physics. 
     1071It must be emphasized that this is the weakest part of the primitive equations, 
     1072but also one of the most important for long-term simulations as 
     1073small scale processes \textit{in fine} balance the surface input of kinetic energy and heat. 
     1074 
     1075The control exerted by gravity on the flow induces a strong anisotropy between the lateral and vertical motions. 
     1076Therefore subgrid-scale physics \textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in 
     1077\autoref{eq:PE_dyn}, \autoref{eq:PE_tra_T} and \autoref{eq:PE_tra_S} are divided into 
     1078a lateral part  \textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and 
     1079a vertical part  \textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. 
     1080The formulation of these terms and their underlying physics are briefly discussed in the next two subsections. 
    10261081 
    10271082% ------------------------------------------------------------------------------------------------------------- 
     
    10311086\label{subsec:PE_zdf} 
    10321087 
    1033 The model resolution is always larger than the scale at which the major  
    1034 sources of vertical turbulence occur (shear instability, internal wave  
    1035 breaking...). Turbulent motions are thus never explicitly solved, even  
    1036 partially, but always parameterized. The vertical turbulent fluxes are  
    1037 assumed to depend linearly on the gradients of large-scale quantities (for  
    1038 example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$,  
    1039 where $A^{vT}$ is an eddy coefficient). This formulation is  
    1040 analogous to that of molecular diffusion and dissipation. This is quite  
    1041 clearly a necessary compromise: considering only the molecular viscosity  
    1042 acting on large scale severely underestimates the role of turbulent  
    1043 diffusion and dissipation, while an accurate consideration of the details of  
    1044 turbulent motions is simply impractical. The resulting vertical momentum and  
    1045 tracer diffusive operators are of second order: 
     1088The model resolution is always larger than the scale at which the major sources of vertical turbulence occur 
     1089(shear instability, internal wave breaking...). 
     1090Turbulent motions are thus never explicitly solved, even partially, but always parameterized. 
     1091The vertical turbulent fluxes are assumed to depend linearly on the gradients of large-scale quantities 
     1092(for example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$, 
     1093where $A^{vT}$ is an eddy coefficient). 
     1094This formulation is analogous to that of molecular diffusion and dissipation. 
     1095This is quite clearly a necessary compromise: considering only the molecular viscosity acting on 
     1096large scale severely underestimates the role of turbulent diffusion and dissipation, 
     1097while an accurate consideration of the details of turbulent motions is simply impractical. 
     1098The resulting vertical momentum and tracer diffusive operators are of second order: 
    10461099\begin{equation} \label{eq:PE_zdf} 
    10471100   \begin{split} 
     
    10521105   \end{split} 
    10531106\end{equation} 
    1054 where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients,  
    1055 respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat  
    1056 and salt must be specified (see \autoref{chap:SBC} and \autoref{chap:ZDF} and \autoref{sec:TRA_bbl}).  
    1057 All the vertical physics is embedded in the specification of the eddy coefficients.  
    1058 They can be assumed to be either constant, or function of the local fluid properties  
    1059 ($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a  
    1060 turbulent closure model. The choices available in \NEMO are discussed in \autoref{chap:ZDF}). 
     1107where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients, respectively. 
     1108At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified 
     1109(see \autoref{chap:SBC} and \autoref{chap:ZDF} and \autoref{sec:TRA_bbl}). 
     1110All the vertical physics is embedded in the specification of the eddy coefficients. 
     1111They can be assumed to be either constant, or function of the local fluid properties 
     1112($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), 
     1113or computed from a turbulent closure model. 
     1114The choices available in \NEMO are discussed in \autoref{chap:ZDF}). 
    10611115 
    10621116% ------------------------------------------------------------------------------------------------------------- 
     
    10661120\label{subsec:PE_ldf} 
    10671121 
    1068 Lateral turbulence can be roughly divided into a mesoscale turbulence  
    1069 associated with eddies (which can be solved explicitly if the resolution is  
    1070 sufficient since their underlying physics are included in the primitive  
    1071 equations), and a sub mesoscale turbulence which is never explicitly solved  
    1072 even partially, but always parameterized. The formulation of lateral eddy  
    1073 fluxes depends on whether the mesoscale is below or above the grid-spacing  
     1122Lateral turbulence can be roughly divided into a mesoscale turbulence associated with eddies 
     1123(which can be solved explicitly if the resolution is sufficient since 
     1124their underlying physics are included in the primitive equations), 
     1125and a sub mesoscale turbulence which is never explicitly solved even partially, but always parameterized. 
     1126The formulation of lateral eddy fluxes depends on whether the mesoscale is below or above the grid-spacing 
    10741127($i.e.$ the model is eddy-resolving or not). 
    10751128 
    1076 In non-eddy-resolving configurations, the closure is similar to that used  
    1077 for the vertical physics. The lateral turbulent fluxes are assumed to depend  
    1078 linearly on the lateral gradients of large-scale quantities. The resulting  
    1079 lateral diffusive and dissipative operators are of second order.  
    1080 Observations show that lateral mixing induced by mesoscale turbulence tends  
    1081 to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987})  
    1082 rather than across them.  
    1083 As the slope of neutral surfaces is small in the ocean, a common  
    1084 approximation is to assume that the `lateral' direction is the horizontal,  
    1085 $i.e.$ the lateral mixing is performed along geopotential surfaces. This leads  
    1086 to a geopotential second order operator for lateral subgrid scale physics.  
    1087 This assumption can be relaxed: the eddy-induced turbulent fluxes can be  
    1088 better approached by assuming that they depend linearly on the gradients of  
    1089 large-scale quantities computed along neutral surfaces. In such a case,  
    1090 the diffusive operator is an isoneutral second order operator and it has  
    1091 components in the three space directions. However, both horizontal and  
    1092 isoneutral operators have no effect on mean ($i.e.$ large scale) potential  
    1093 energy whereas potential energy is a main source of turbulence (through  
    1094 baroclinic instabilities). \citet{Gent1990} have proposed a  
    1095 parameterisation of mesoscale eddy-induced turbulence which associates an  
    1096 eddy-induced velocity to the isoneutral diffusion. Its mean effect is to  
    1097 reduce the mean potential energy of the ocean. This leads to a formulation  
    1098 of lateral subgrid-scale physics made up of an isoneutral second order  
    1099 operator and an eddy induced advective part. In all these lateral diffusive  
    1100 formulations, the specification of the lateral eddy coefficients remains the  
    1101 problematic point as there is no really satisfactory formulation of these  
    1102 coefficients as a function of large-scale features. 
    1103  
    1104 In eddy-resolving configurations, a second order operator can be used, but  
    1105 usually the more scale selective biharmonic operator is preferred as the  
    1106 grid-spacing is usually not small enough compared to the scale of the  
    1107 eddies. The role devoted to the subgrid-scale physics is to dissipate the  
    1108 energy that cascades toward the grid scale and thus to ensure the stability of  
    1109 the model while not interfering with the resolved mesoscale activity. Another approach  
    1110 is becoming more and more popular: instead of specifying explicitly a sub-grid scale  
    1111 term in the momentum and tracer time evolution equations, one uses a advective  
    1112 scheme which is diffusive enough to maintain the model stability. It must be emphasised 
    1113 that then, all the sub-grid scale physics is included in the formulation of the 
    1114 advection scheme.  
    1115  
    1116 All these parameterisations of subgrid scale physics have advantages and  
    1117 drawbacks. There are not all available in \NEMO. For active tracers (temperature and  
    1118 salinity) the main ones are: Laplacian and bilaplacian operators acting along  
    1119 geopotential or iso-neutral surfaces, \citet{Gent1990} parameterisation,  
    1120 and various slightly diffusive advection schemes.  
    1121 For momentum, the main ones are: Laplacian and bilaplacian operators acting along  
    1122 geopotential surfaces, and UBS advection schemes when flux form is chosen for the momentum advection. 
     1129In non-eddy-resolving configurations, the closure is similar to that used for the vertical physics. 
     1130The lateral turbulent fluxes are assumed to depend linearly on the lateral gradients of large-scale quantities. 
     1131The resulting lateral diffusive and dissipative operators are of second order. 
     1132Observations show that lateral mixing induced by mesoscale turbulence tends to be along isopycnal surfaces 
     1133(or more precisely neutral surfaces \cite{McDougall1987}) rather than across them. 
     1134As the slope of neutral surfaces is small in the ocean, a common approximation is to assume that 
     1135the `lateral' direction is the horizontal, $i.e.$ the lateral mixing is performed along geopotential surfaces. 
     1136This leads to a geopotential second order operator for lateral subgrid scale physics. 
     1137This assumption can be relaxed: the eddy-induced turbulent fluxes can be better approached by assuming that 
     1138they depend linearly on the gradients of large-scale quantities computed along neutral surfaces. 
     1139In such a case, the diffusive operator is an isoneutral second order operator and 
     1140it has components in the three space directions. 
     1141However, 
     1142both horizontal and isoneutral operators have no effect on mean ($i.e.$ large scale) potential energy whereas 
     1143potential energy is a main source of turbulence (through baroclinic instabilities). 
     1144\citet{Gent1990} have proposed a parameterisation of mesoscale eddy-induced turbulence which 
     1145associates an eddy-induced velocity to the isoneutral diffusion. 
     1146Its mean effect is to reduce the mean potential energy of the ocean. 
     1147This leads to a formulation of lateral subgrid-scale physics made up of an isoneutral second order operator and 
     1148an eddy induced advective part. 
     1149In all these lateral diffusive formulations, 
     1150the specification of the lateral eddy coefficients remains the problematic point as 
     1151there is no really satisfactory formulation of these coefficients as a function of large-scale features. 
     1152 
     1153In eddy-resolving configurations, a second order operator can be used, 
     1154but usually the more scale selective biharmonic operator is preferred as 
     1155the grid-spacing is usually not small enough compared to the scale of the eddies. 
     1156The role devoted to the subgrid-scale physics is to dissipate the energy that 
     1157cascades toward the grid scale and thus to ensure the stability of the model while 
     1158not interfering with the resolved mesoscale activity. 
     1159Another approach is becoming more and more popular: 
     1160instead of specifying explicitly a sub-grid scale term in the momentum and tracer time evolution equations, 
     1161one uses a advective scheme which is diffusive enough to maintain the model stability. 
     1162It must be emphasised that then, all the sub-grid scale physics is included in the formulation of 
     1163the advection scheme.  
     1164 
     1165All these parameterisations of subgrid scale physics have advantages and drawbacks. 
     1166There are not all available in \NEMO. For active tracers (temperature and salinity) the main ones are: 
     1167Laplacian and bilaplacian operators acting along geopotential or iso-neutral surfaces, 
     1168\citet{Gent1990} parameterisation, and various slightly diffusive advection schemes. 
     1169For momentum, the main ones are: Laplacian and bilaplacian operators acting along geopotential surfaces, 
     1170and UBS advection schemes when flux form is chosen for the momentum advection. 
    11231171 
    11241172\subsubsection{Lateral laplacian tracer diffusive operator} 
     
    11331181\end{array} }} \right) 
    11341182\end{equation} 
    1135 where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along  
    1136 which the diffusive operator acts and the model level ($e. g.$ $z$- or  
    1137 $s$-surfaces). Note that the formulation \autoref{eq:PE_iso_tensor} is exact for the  
    1138 rotation between geopotential and $s$-surfaces, while it is only an approximation  
    1139 for the rotation between isoneutral and $z$- or $s$-surfaces. Indeed, in the latter  
    1140 case, two assumptions are made to simplify \autoref{eq:PE_iso_tensor} \citep{Cox1987}.  
    1141 First, the horizontal contribution of the dianeutral mixing is neglected since the ratio  
    1142 between iso and dia-neutral diffusive coefficients is known to be several orders of  
    1143 magnitude smaller than unity. Second, the two isoneutral directions of diffusion are  
    1144 assumed to be independent since the slopes are generally less than $10^{-2}$ in the  
    1145 ocean (see \autoref{apdx:B}). 
    1146  
    1147 For \textit{iso-level} diffusion, $r_1$ and $r_2 $ are zero. $\Re $ reduces to the identity  
    1148 in the horizontal direction, no rotation is applied.  
    1149  
    1150 For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the  
    1151 geopotential and computational surfaces: they are equal to $\sigma _1$ and $\sigma _2$,  
    1152 respectively (see \autoref{eq:PE_sco_slope}). 
    1153  
    1154 For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral  
    1155 and computational surfaces. Therefore, they are different quantities, 
    1156 but have similar expressions in $z$- and $s$-coordinates. In $z$-coordinates: 
     1183where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along which the diffusive operator acts and 
     1184the model level ($e. g.$ $z$- or $s$-surfaces). 
     1185Note that the formulation \autoref{eq:PE_iso_tensor} is exact for 
     1186the rotation between geopotential and $s$-surfaces, 
     1187while it is only an approximation for the rotation between isoneutral and $z$- or $s$-surfaces. 
     1188Indeed, in the latter case, two assumptions are made to simplify \autoref{eq:PE_iso_tensor} \citep{Cox1987}. 
     1189First, the horizontal contribution of the dianeutral mixing is neglected since the ratio between iso and 
     1190dia-neutral diffusive coefficients is known to be several orders of magnitude smaller than unity. 
     1191Second, the two isoneutral directions of diffusion are assumed to be independent since 
     1192the slopes are generally less than $10^{-2}$ in the ocean (see \autoref{apdx:B}). 
     1193 
     1194For \textit{iso-level} diffusion, $r_1$ and $r_2 $ are zero. 
     1195$\Re $ reduces to the identity in the horizontal direction, no rotation is applied.  
     1196 
     1197For \textit{geopotential} diffusion, 
     1198$r_1$ and $r_2 $ are the slopes between the geopotential and computational surfaces: 
     1199they are equal to $\sigma _1$ and $\sigma _2$, respectively (see \autoref{eq:PE_sco_slope}). 
     1200 
     1201For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral and computational surfaces. 
     1202Therefore, they are different quantities, but have similar expressions in $z$- and $s$-coordinates. 
     1203In $z$-coordinates: 
    11571204\begin{equation} \label{eq:PE_iso_slopes} 
    11581205r_1 =\frac{e_3 }{e_1 }  \left( \pd[\rho]{i} \right) \left( \pd[\rho]{k} \right)^{-1} \, \quad 
     
    11621209 
    11631210\subsubsection{Eddy induced velocity} 
    1164  When the \textit{eddy induced velocity} parametrisation (eiv) \citep{Gent1990} is used,  
     1211When the \textit{eddy induced velocity} parametrisation (eiv) \citep{Gent1990} is used, 
    11651212an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers: 
    11661213\begin{equation} \label{eq:PE_iso+eiv} 
     
    11681215           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right) 
    11691216\end{equation} 
    1170 where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent,  
     1217where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent, 
    11711218eddy-induced transport velocity. This velocity field is defined by: 
    11721219\begin{equation} \label{eq:PE_eiv} 
     
    11791226   \end{split} 
    11801227\end{equation} 
    1181 where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral  
    1182 thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes  
    1183 between isoneutral and \emph{geopotential} surfaces. Their values are 
    1184 thus independent of the vertical coordinate, but their expression depends on the coordinate:  
     1228where $A^{eiv}$ is the eddy induced velocity coefficient 
     1229(or equivalently the isoneutral thickness diffusivity coefficient), 
     1230and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes between isoneutral and \emph{geopotential} surfaces. 
     1231Their values are thus independent of the vertical coordinate, but their expression depends on the coordinate:  
    11851232\begin{align} \label{eq:PE_slopes_eiv} 
    11861233\tilde{r}_n = \begin{cases} 
     
    11911238\end{align} 
    11921239 
    1193 The normal component of the eddy induced velocity is zero at all the boundaries.  
    1194 This can be achieved in a model by tapering either the eddy coefficient or the slopes  
    1195 to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. \autoref{chap:LDF}). 
     1240The normal component of the eddy induced velocity is zero at all the boundaries. 
     1241This can be achieved in a model by tapering either the eddy coefficient or 
     1242the slopes to zero in the vicinity of the boundaries. 
     1243The latter strategy is used in \NEMO (cf. \autoref{chap:LDF}). 
    11961244 
    11971245\subsubsection{Lateral bilaplacian tracer diffusive operator} 
     
    12011249D^{lT}= - \Delta \left( \;\Delta T \right)  
    12021250\qquad \text{where} \;\; \Delta \bullet = \nabla \left( {\sqrt{B^{lT}\,}\;\Re \;\nabla \bullet} \right) 
    1203  \end{equation} 
    1204 It is the Laplacian operator given by \autoref{eq:PE_iso_tensor} applied twice with  
     1251\end{equation} 
     1252It is the Laplacian operator given by \autoref{eq:PE_iso_tensor} applied twice with 
    12051253the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.  
    12061254 
     
    12081256\subsubsection{Lateral Laplacian momentum diffusive operator} 
    12091257 
    1210 The Laplacian momentum diffusive operator along $z$- or $s$-surfaces is found by  
     1258The Laplacian momentum diffusive operator along $z$- or $s$-surfaces is found by 
    12111259applying \autoref{eq:PE_lap_vector} to the horizontal velocity vector (see \autoref{apdx:B}): 
    12121260\begin{equation} \label{eq:PE_lapU} 
     
    12241272\end{equation} 
    12251273 
    1226 Such a formulation ensures a complete separation between the vorticity and  
    1227 horizontal divergence fields (see \autoref{apdx:C}).  
    1228 Unfortunately, it is only available in \textit{iso-level} direction.  
    1229 When a rotation is required ($i.e.$ geopotential diffusion in $s-$coordinates  
    1230 or isoneutral diffusion in both $z$- and $s$-coordinates), the $u$ and $v-$fields  
    1231 are considered as independent scalar fields, so that the diffusive operator is given by: 
     1274Such a formulation ensures a complete separation between the vorticity and horizontal divergence fields 
     1275(see \autoref{apdx:C}). 
     1276Unfortunately, it is only available in \textit{iso-level} direction. 
     1277When a rotation is required 
     1278($i.e.$ geopotential diffusion in $s-$coordinates or isoneutral diffusion in both $z$- and $s$-coordinates), 
     1279the $u$ and $v-$fields are considered as independent scalar fields, so that the diffusive operator is given by: 
    12321280\begin{equation} \label{eq:PE_lapU_iso} 
    12331281\begin{split} 
    12341282 D_u^{l{\rm {\bf U}}} &= \nabla .\left( {A^{lm} \;\Re \;\nabla u} \right) \\  
    12351283 D_v^{l{\rm {\bf U}}} &= \nabla .\left( {A^{lm} \;\Re \;\nabla v} \right) 
    1236  \end{split} 
    1237  \end{equation} 
    1238 where $\Re$ is given by \autoref{eq:PE_iso_tensor}. It is the same expression as  
    1239 those used for diffusive operator on tracers. It must be emphasised that such a  
    1240 formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or  
    1241 $\beta-$plane, not on the sphere. It is also a very good approximation in vicinity  
    1242 of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}. 
     1284\end{split} 
     1285\end{equation} 
     1286where $\Re$ is given by \autoref{eq:PE_iso_tensor}. 
     1287It is the same expression as those used for diffusive operator on tracers. 
     1288It must be emphasised that such a formulation is only exact in a Cartesian coordinate system, 
     1289$i.e.$ on a $f-$ or $\beta-$plane, not on the sphere. 
     1290It is also a very good approximation in vicinity of the Equator in 
     1291a geographical coordinate system \citep{Lengaigne_al_JGR03}. 
    12431292 
    12441293\subsubsection{lateral bilaplacian momentum diffusive operator} 
    12451294 
    1246 As for tracers, the bilaplacian order momentum diffusive operator is a  
    1247 re-entering Laplacian operator with the harmonic eddy diffusion coefficient  
    1248 set to the square root of the biharmonic one. Nevertheless it is currently  
    1249 not available in the iso-neutral case. 
     1295As for tracers, the bilaplacian order momentum diffusive operator is a re-entering Laplacian operator with 
     1296the harmonic eddy diffusion coefficient set to the square root of the biharmonic one. 
     1297Nevertheless it is currently not available in the iso-neutral case. 
    12501298 
    12511299\end{document} 
Note: See TracChangeset for help on using the changeset viewer.