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.
Chap_Model_Basics.tex in trunk/DOC/TexFiles/Chapters – NEMO

source: trunk/DOC/TexFiles/Chapters/Chap_Model_Basics.tex @ 996

Last change on this file since 996 was 996, checked in by gm, 16 years ago

trunk - DOC update to correct error in previous update

  • Property svn:executable set to *
File size: 72.4 KB
Line 
1% ================================================================
2% Chapter 1 Ñ Model Basics
3% ================================================================
4
5\chapter{Model basics}
6\label{PE}
7\minitoc
8
9
10% ================================================================
11% Primitive Equations
12% ================================================================
13\section{Primitive Equations}
14\label{PE_PE}
15
16% -------------------------------------------------------------------------------------------------------------
17%        Vector Invariant Formulation
18% -------------------------------------------------------------------------------------------------------------
19
20\subsection{Vector Invariant Formulation}
21\label{PE_Vector}
22
23
24The ocean is a fluid that can be described to a good approximation by the primitive
25equations, $i.e.$ the Navier-Stokes equations along with a nonlinear equation of
26state which couples the two active tracers (temperature and salinity) to the fluid
27velocity, plus the following additional assumptions made from scale considerations:
28
29\textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to be spheres so that gravity (local vertical) is parallel to the earth's radius
30
31\textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius
32
33\textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect of small scale processes on the large-scale) are expressed in terms of large-scale features
34
35\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their contribution to the buoyancy force
36
37\textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a balance between the vertical pressure gradient and the buoyancy force (this removes convective processes from
38the initial Navier-Stokes equations and so convective processes must be parameterized instead)
39
40\textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity vector is assumed to be zero.
41
42Because the gravitational force is so dominant in the equations of large-scale motions, it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$ (the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane), $T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density. The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k}) vector system provides the following six equations (namely the momentum balance, the hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation equations and an equation of state):
43\begin{subequations} \label{Eq_PE}
44  \begin{equation}     \label{Eq_PE_dyn}
45\frac{\partial {\rm {\bf U}}_h }{\partial t}=
46-\left[    {\left( {\nabla \times {\rm {\bf U}}} \right)\times {\rm {\bf U}}
47            +\frac{1}{2}\nabla \left( {{\rm {\bf U}}^2} \right)}    \right]_h
48 -f\;{\rm {\bf k}}\times {\rm {\bf U}}_h
49-\frac{1}{\rho _o }\nabla _h p + {\rm {\bf D}}^{\rm {\bf U}} + {\rm {\bf F}}^{\rm {\bf U}}
50  \end{equation}
51  \begin{equation}     \label{Eq_PE_hydrostatic}
52\frac{\partial p }{\partial z} = - \rho \ g
53  \end{equation}
54  \begin{equation}     \label{Eq_PE_continuity}
55\nabla \cdot {\bf U}=  0
56  \end{equation}
57\begin{equation} \label{Eq_PE_tra_T}
58\frac{\partial T}{\partial t} = - \nabla \cdot  \left( T \ \rm{\bf U} \right) + D^T + F^T
59  \end{equation}
60  \begin{equation}     \label{Eq_PE_tra_S}
61\frac{\partial S}{\partial t} = - \nabla \cdot  \left( S \ \rm{\bf U} \right) + D^S + F^S
62  \end{equation}
63  \begin{equation}     \label{Eq_PE_eos}
64\rho = \rho \left( T,S,p \right)
65  \end{equation}
66\end{subequations}
67where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions, $t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by the equation of state (\ref{Eq_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure, $f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's angular velocity vector), and $g$ is the gravitational acceleration.
68${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterizations of small-scale physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$ surface forcing terms. Their nature and formulation are discussed in \S\ref{PE_zdf_ldf} and page \S\ref{PE_boundary_condition}.
69
70.
71
72% -------------------------------------------------------------------------------------------------------------
73% Boundary condition
74% -------------------------------------------------------------------------------------------------------------
75\subsection{Boundary Conditions}
76\label{PE_boundary_condition}
77
78An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$, chosen as a mean sea surface (Fig.~\ref{Fig_ocean_bc}). Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth, the continental margins, the sea ice and the atmosphere. However, some of these fluxes are so weak that even on climatic time scales of thousands of years they can be neglected. In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and the other components of the earth system.
79
80%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
81\begin{figure}[!ht] \label{Fig_ocean_bc}  \begin{center}
82\includegraphics[width=0.90\textwidth]{./Figures/Fig_I_ocean_bc.pdf}
83\caption{The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface. Both $H$ and $\eta $ are referenced to $z=0$.}
84\end{center}   \end{figure}
85%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
86
87
88\begin{description}
89\item[Land - ocean interface:] the major flux between continental margins and the ocean is a mass exchange of fresh water through river runoff. Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths. It can be neglected for short range integrations but has to be taken into account for long term integrations as it influences the characteristics of water masses formed (especially at high latitudes). It is required in order to close the water cycle of the climate system. It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths.
90\item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small, except in special areas of little extent. They are usually neglected in the model \footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling ($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world ocean (see \ref{TRA_bbc}).}. The boundary condition is thus set to no flux of heat and salt across solid boundaries. For momentum, the situation is different. There is no flow across solid boundaries, $i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words, the bottom velocity is parallel to solid boundaries). This kinematic boundary condition can be expressed as:
91\begin{equation} \label{Eq_PE_w_bbc}
92w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right)
93\end{equation}
94In addition, the ocean exchanges momentum with the earth through frictional processes. Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification depends on the nature of the physical parameterization used for ${\rm {\bf D}}^{\rm {\bf U}}$ in \eqref{Eq_PE_dyn}. It is discussed in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9.
95\item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux of fresh water PE  (the precipitation minus evaporation budget) leads to:
96\begin{equation} \label{Eq_PE_w_sbc}
97w = \frac{\partial \eta }{\partial t} 
98    + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right)
99    + P-E
100\end{equation}
101The dynamic boundary condition, neglecting the surface tension (which removes capillary waves from the system) leads to the continuity of pressure across the interface $z=\eta$. The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
102\item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water and momentum. The sea surface temperature is constrained to be at the freezing point at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and salt fluxes that cannot be neglected.
103\end{description}
104
105
106% ================================================================
107% The Horizontal Pressure Gradient
108% ================================================================
109\section{The Horizontal Pressure Gradient }
110\label{PE_hor_pg}
111
112% -------------------------------------------------------------------------------------------------------------
113% Pressure Formulation
114% -------------------------------------------------------------------------------------------------------------
115\subsection{Pressure Formulation}
116\label{PE_p_formulation}
117
118The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that: $p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\ref{Eq_PE_hydrostatic}), assuming that pressure in decibars can be approximated by depth in meters in (\ref{Eq_PE_eos}). The hydrostatic pressure is then given by:
119\begin{equation} \label{Eq_PE_pressure}
120p_h \left( {i,j,z,t} \right)
121 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,z} \right)\;d\varsigma } 
122\end{equation}
123 Two strategies can be considered for the surface pressure term: $(a)$ introduce of a new variable $\eta$, the free-surface elevation, for which a prognostic equation can be established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used, one solution of the free-surface elevation consists of the excitation of external gravity waves. The flow is barotropic and the surface moves up and down with gravity as the restoring force. The phase speed of such waves is high (some hundreds of metres per second) so that the time step would have to be very short if they were present in the model. The latter strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$ the sea surface is the surface $z=0$. This well known approximation increases the surface wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic Rossby or planetary waves). In the present release of \NEMO, both strategies are still available. They are further described in the next two sub-sections.
124
125% -------------------------------------------------------------------------------------------------------------
126% Free Surface Formulation
127% -------------------------------------------------------------------------------------------------------------
128\subsection{Free Surface Formulation}
129\label{PE_free_surface}
130
131In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced which describes the shape of the air-sea interface. This variable is solution of a prognostic equation which is established by forming the vertical average of the kinematic surface condition (\ref{Eq_PE_w_bbc}):
132\begin{equation} \label{Eq_PE_ssh}
133\frac{\partial \eta }{\partial t}=-D+P-E
134   \quad \text{where} \
135D=\nabla \cdot \left[ {\left( {H+\eta } \right) \; {\rm{\bf \overline{U}}}_h \,} \right]
136\end{equation}
137and using (\ref{Eq_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
138
139Allowing the air-sea interface to move introduces the external gravity waves (EGWs) as a class of solution of the primitive equations. These waves are barotropic because of hydrostatic assumption, and their phase speed is quite high. Their time scale is short with respect to the other processes described by the primitive equations.
140
141Three choices can be made regarding the implementation of the free surface in the model, depending on the physical processes of interest.
142
143$\bullet$ If one is interested in EGWs, in particular the tides and their interaction
144with the baroclinic structure of the ocean (internal waves) possibly in
145shallow seas, then a non linear free surface is the most appropriate. This
146means that no approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the
147ocean volume is fully taken into account. Note that in order to study the
148fast time scales associated with EGWs it is necessary to minimize time
149filtering effects (use an explicit time scheme with very small time step, or
150a split-explicit scheme with reasonably small time step, see \S\ref{DYN_spg_exp} or
151\S\ref{DYN_spg_ts}.
152
153$\bullet$ If one is not interested in EGW but rather sees them as high frequency
154noise, it is possible to apply an explicit filter to slow down the fastest waves while
155not altering the slow barotropic Rossby waves. If further, an approximative conservation
156of heat and salt contents is sufficient for the problem solved, then it is
157sufficient to solve a linearized version of (\ref{Eq_PE_ssh}), which still allows
158to take into account freshwater fluxes applied at the ocean surface \citep{Roullet2000}.
159
160$\bullet$ For process studies not involving external waves nor surface freshwater
161fluxes, it is possible to use the rigid lid approximation see (next
162section). The ocean surface is then considered as a fixed surface, so that all
163external waves are removed from the system.
164
165The filtering of EGWs in models with a free surface is usually a matter of discretisation of the temporal derivatives, using the time splitting method \citep{Killworth1991, Zhang1992} or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach developed by \citet{Roullet2000}: the damping of EGWs is ensured by introducing an additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:
166\begin{equation} \label{Eq_PE_flt}
167\frac{\partial {\rm {\bf U}}_h }{\partial t}= {\rm {\bf M}}
168- g \nabla \left( \tilde{\rho} \ \eta \right)
169- g \ T_c \nabla \left( \widetilde{\rho} \ \partial_t \eta \right)
170\end{equation}
171where $T_c$, is a parameter with dimensions of time which characterizes the force, $\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and $\rm {\bf M}$ represents the collected contributions of the Coriolis, hydrostatic pressure gradient, non-linear and viscous terms in \eqref{Eq_PE_dyn}.
172
173The new force can be interpreted as a diffusion of vertically integrated volume flux divergence. The time evolution of $D$ is thus governed by a balance of two terms, $-g$ \textbf{A} $\eta$ and $g \, T_c \,$ \textbf{A} $D$, associated with a propagative regime and a diffusive regime in the temporal spectrum, respectively. In the diffusive regime, the EGWs no longer propagate, $i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than $T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs can be damped by choosing $T_c > \Delta t$. \citet{Roullet2000} demonstrate that (\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which has to be computed implicitly. This is not surprising since the use of a large time step has a necessarily numerical cost. Two gains arise in comparison with the previous formulations. Firstly, the damping of EGWs can be quantified through the magnitude of the additional term. Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as soon as $T_c > \Delta t$.
174
175When the variations of free surface elevation are small compared to the thickness of the first model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized by \citet{Roullet2000} the linearization of (\ref{Eq_PE_ssh}) has consequences on the conservation of salt in the model. With the nonlinear free surface equation, the time evolution of the total salt content is
176\begin{equation} \label{Eq_PE_salt_content}
177\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} =\int\limits_S
178{S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds}
179\end{equation}
180where $S$ is the salinity, and the total salt is integrated over the whole ocean volume $D_\eta$ bounded by the time-dependent free surface. The right hand side (which is an integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh}) is satisfied, so that the salt is perfectly conserved. When the free surface equation is linearized, \citet{Roullet2000} show that the total salt content integrated in the fixed volume $D$ (bounded by the surface $z=0$) is no longer conserved:
181\begin{equation} \label{Eq_PE_salt_content_linear}
182\frac{\partial }{\partial t}\int\limits_D {S\;dv} =-\int\limits_S
183{S\;\frac{\partial \eta }{\partial t}ds} 
184\end{equation}
185
186The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions \citep{Roullet2000}. It can be significant when the freshwater forcing is not balanced and the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $} results in a decrease of the salinity in the fixed volume $D$. Even in that case though, the total salt integrated in the variable volume $D_{\eta}$ varies much less, since (\ref{Eq_PE_salt_content_linear}) can be rewritten as
187\begin{equation} \label{Eq_PE_salt_content_corrected}
188\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
189=\frac{\partial}{\partial t} \left[ \;{\int\limits_D {S\;dv} +\int\limits_S {S\eta \;ds} } \right]
190=\int\limits_S {\eta \;\frac{\partial S}{\partial t}ds}
191\end{equation}
192
193Although the total salt content is not exactly conserved with the linearized free surface, its variations are driven by correlations of the time variation of surface salinity with the sea surface height, which is a negligible term. This situation contrasts with
194the case of the rigid lid approximation (following section) in which case freshwater forcing is represented by a virtual salt flux, leading to a spurious source of salt at the ocean surface \citep{Roullet2000}.
195
196% -------------------------------------------------------------------------------------------------------------
197% Rigid-Lid Formulation
198% -------------------------------------------------------------------------------------------------------------
199\subsection{Rigid-Lid formulation}
200\label{PE_rigid_lid}
201
202With the rigid lid approximation, we assume that the ocean surface ($z=0$) is a rigid lid on which a pressure $p_s$ is exerted. This implies that the vertical velocity at the surface is equal to zero. From the continuity equation \eqref{Eq_PE_continuity} and the kinematic condition at the bottom \eqref{Eq_PE_w_bbc} (no flux across the bottom), it can be shown that the vertically integrated flow $H{\rm {\bf \bar {U}}}_h$ is non-divergent (where the overbar indicates a vertical average over the whole water column, i.e. from $z=-H$, the ocean bottom, to $z=0$ , the rigid-lid). Thus, $\rm {\bf \bar {U}}_h$ can be derived from a volume transport streamfunction $\psi$:
203\begin{equation} \label{Eq_PE_u_psi}
204\overline{\vect{U}}_h =\frac{1}{H}\left(   \vect{k} \times \nabla \psi   \right)
205\end{equation}
206
207As $p_s$ does not depend on depth, its horizontal gradient is obtained by forming the vertical average of \eqref{Eq_PE_dyn} and using \eqref{Eq_PE_u_psi}:
208
209\begin{equation} \label{Eq_PE_u_barotrope}
210\frac{1}{\rho _o }\nabla _h p_s
211=\overline{\vect{M}} -\frac{\partial \overline{\vect{U}} _h }{\partial t}
212=\overline{\vect{M}} 
213-\frac{1}{H} \left[   \vect{k} \times \nabla \left( \frac{\partial \psi}{\partial t} \right)   \right]
214\end{equation}
215
216Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ represents the collected contributions of the Coriolis, hydrostatic pressure gradient, nonlinear and viscous terms in \eqref{Eq_PE_dyn}. The time derivative of $\psi $ is the solution of an elliptic equation which is obtained from the vertical component of the curl of (\ref{Eq_PE_u_barotrope}):
217\begin{equation} \label{Eq_PE_psi}
218\left[   {\nabla \times \left[ {\frac{1}{H} \vect{\bf k} 
219  \times \nabla \left(   {\frac{\partial \psi }{\partial t}} \right)}   \right]} \; \right]_z
220=\left[   {\nabla \times \overline{\vect{M}} }   \right]_z
221\end{equation}
222
223Using the proper boundary conditions, \eqref{Eq_PE_psi} can be solved to find $\partial_t \psi$ and thus using \eqref{Eq_PE_u_barotrope} the horizontal surface pressure gradient. It should be noted that $p_s$ can be computed by taking the divergence of \eqref{Eq_PE_u_barotrope} and solving the resulting elliptic equation. Thus the surface pressure is a diagnostic quantity that can be recovered for analysis purposes.
224
225A difficulty lies in the determination of the boundary condition on $\partial_t \psi$. The boundary condition on velocity is that there is no flow normal to a solid wall, $i.e.$ the coastlines are streamlines. Therefore \eqref{Eq_PE_psi} is solved with the following Dirichlet boundary condition: $\partial_t \psi$ is constant along each coastline of the same continent or of the same island. When all the coastlines are connected (there are no islands), the constant value of $\partial_t \psi$ along the coast can be arbitrarily chosen to be zero. When islands are present in the domain, the value of the barotropic streamfunction will generally be different for each island and for the continent, and will vary with respect to time. So the boundary condition is: $\psi=0$ along the continent and $\psi=\mu_n$ along island $n$ ($1 \leq n \leq Q$), where $Q$ is the number of islands present in the domain and $\mu_n$ is a time dependent variable. A time evolution equation of the unknown $\mu_n$ can be found by evaluating the circulation of the time derivative of the vertical average (barotropic) velocity field along a closed contour around each island. Since the circulation of a gradient field along a closed contour is zero, from \eqref{Eq_PE_u_barotrope} we have:
226\begin{equation} \label{Eq_PE_isl_circulation}
227\oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
228{\frac{\partial \psi }{\partial t}} \right)} \right] \cdot {\rm {\bf d}}\ell } 
229= \oint_n {\overline {\rm {\bf M}} \cdot {\rm {\bf d}}\ell } 
230\qquad  1 \leq n \leq Q
231\end{equation}
232
233Since (\ref{Eq_PE_psi}) is linear, its solution \textit{$\psi $} can be decomposed
234as follows:
235\begin{equation} \label{Eq_PE_psi_isl}
236\psi =\psi _o +\sum\limits_{n=1}^{n=Q} {\mu _n \psi _n } 
237\end{equation}
238where $\psi _o$ is the solution of \eqref{Eq_PE_psi} with $\psi _o=0$ long all
239the coastlines, and where $\psi _n$ is the solution of \eqref{Eq_PE_psi} with
240the right-hand side equal to $0$, and with $\psi _n =1$ long the island $n$,
241$\psi _n =0$ along the other boundaries. The function $\psi _n$ is thus
242independent of time. Introducing \eqref{Eq_PE_psi_isl} into
243\eqref{Eq_PE_isl_circulation} yields:
244\begin{multline} \label{Eq_PE_psi_isl_circulation}
245\left[ {\oint_n {\frac{1}{H}  \left[ {{\rm {\bf k}}\times \nabla \psi _m } \right]\cdot {\rm {\bf d}}\ell } } \right]_{1\leq m\leqslant Q \atop 1\leq n\leqslant Q  }
246 \left( {\frac{\partial \mu _n }{\partial t}} 
247\right)_{1\leqslant n\leqslant Q}        \\
248 =\left( {\oint_n {\left[ {\overline {\rm 
249{\bf M}} -\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
250{\frac{\partial \psi _o }{\partial t}} \right)} \right]} \right]\cdot {\rm 
251{\bf d}}\ell } } \right)_{1\leqslant n\leqslant Q} 
252\end{multline}
253which can be rewritten as:
254\begin{equation} \label{Eq_PE_psi_isl_matrix}
255{\rm {\bf A}}\;\left( {\frac{\partial \mu _n }{\partial t}} 
256\right)_{1\leqslant n\leqslant Q} ={\rm {\bf B}}
257\end{equation}
258where \textbf{A} is a $\times Q$ matrix and \textbf{B} is a time dependent vector. As \textbf{A} is independent of time, it can be calculated and inverted once. The time derivative of the streamfunction when islands are present is thus given by:
259\begin{equation} \label{Eq_PE_psi_isl_dt}
260\frac{\partial \psi }{\partial t}=\frac{\partial \psi _o }{\partial 
261t}+\sum\limits_{n=1}^{n=Q} {{\rm {\bf A}}^{-1}{\rm {\bf B}}\;\psi _n } 
262\end{equation}
263
264
265
266% ================================================================
267% Curvilinear z-coordinate System
268% ================================================================
269\section{Curvilinear \textit{z-}coordinate System}
270\label{PE_zco}
271
272
273% -------------------------------------------------------------------------------------------------------------
274% Tensorial Formalism
275% -------------------------------------------------------------------------------------------------------------
276\subsection{Tensorial Formalism}
277\label{PE_tensorial}
278
279In many ocean circulation problems, the flow field has regions of enhanced dynamics ($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts). The representation of such dynamical processes can be improved by specifically increasing the model resolution in these regions. As well, it may be convenient to use a lateral boundary-following coordinate system to better represent coastal dynamics. Moreover, the common geographical coordinate system has a singular point at the North Pole that cannot be easily treated in a global model without filtering. A solution consists of introducing an appropriate coordinate transformation that shifts the singular point onto land \citep{MadecImb1996, Murray1996}. As a consequence, it is important to solve the primitive equations in various curvilinear coordinate systems. An efficient way of introducing an appropriate coordinate transform can be found when using a tensorial formalism. This formalism is suited to any multidimensional curvilinear coordinate system. Ocean modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth approximation), with preservation of the local vertical. Here we give the simplified equations for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey of the conservation laws of fluid dynamics.
280
281Let (\textbf{i},\textbf{j},\textbf{k}) be a set of orthogonal curvilinear coordinates on the sphere associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (Fig.\ref{Fig_referential}). Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea level (Fig.\ref{Fig_referential}). The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$, the three scale factors:
282\begin{equation} \label{Eq_scale_factors}
283\begin{aligned}
284 e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
285}{\partial i}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
286}{\partial i}} \right)^2} \right]^{1/2} \\ 
287 e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
288}{\partial j}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
289}{\partial j}} \right)^2} \right]^{1/2} \\ 
290 e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\ 
291 \end{aligned}
292 \end{equation}
293
294%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
295\begin{figure}[!tb] \label{Fig_referential}  \begin{center}
296\includegraphics[width=0.60\textwidth]{./Figures/Fig_I_earth_referential.pdf}
297\caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear coordinate system (\textbf{i},\textbf{j},\textbf{k}). }
298\end{center}   \end{figure}
299%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
300
301Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by $a$ in (\ref{Eq_scale_factors}) (thin-shell approximation). The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that appear in the primitive equations (Eqs. \eqref{Eq_PE_dyn} to \eqref{Eq_PE_eos}) can be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate system transformation:
302\begin{subequations} \label{Eq_PE_discrete_operators}
303\begin{equation} \label{Eq_PE_grad}
304\nabla q=\frac{1}{e_1 }\frac{\partial q}{\partial i}\;{\rm {\bf 
305i}}+\frac{1}{e_2 }\frac{\partial q}{\partial j}\;{\rm {\bf j}}+\frac{1}{e_3
306}\frac{\partial q}{\partial k}\;{\rm {\bf k}}    \\
307\end{equation}
308\begin{equation} \label{Eq_PE_div}
309\nabla \cdot {\rm {\bf A}} 
310= \frac{1}{e_1 \; e_2} \left[
311  \frac{\partial \left(e_2 \; a_1\right)}{\partial i }
312+\frac{\partial \left(e_1 \; a_2\right)}{\partial j }       \right]
313+ \frac{1}{e_3} \left[ \frac{\partial a_3}{\partial k }   \right]
314\end{equation}
315\begin{equation} \label{Eq_PE_curl}
316   \begin{split}
317\nabla \times \vect{A} =
318    \left[ {\frac{1}{e_2 }\frac{\partial a_3}{\partial j}
319            -\frac{1}{e_3 }\frac{\partial a_2 }{\partial k}} \right] \; \vect{i}
320&+\left[ {\frac{1}{e_3 }\frac{\partial a_1 }{\partial k}
321           -\frac{1}{e_1 }\frac{\partial a_3 }{\partial i}} \right] \; \vect{j}     \\
322&+\frac{1}{e_1 e_2 } \left[ {\frac{\partial \left( {e_2 a_2 } \right)}{\partial i}
323                                       -\frac{\partial \left( {e_1 a_1 } \right)}{\partial j}} \right] \; \vect{k} 
324   \end{split}
325\end{equation}
326\begin{equation} \label{Eq_PE_lap}
327\Delta q = \nabla \cdot \left\nabla q \right)
328\end{equation}
329\begin{equation} \label{Eq_PE_lap_vector}
330\Delta {\rm {\bf A}} =
331  \nabla \left( \nabla \cdot {\rm {\bf A}} \right)
332- \nabla \times \left\nabla \times {\rm {\bf A}} \right)
333\end{equation}
334\end{subequations}
335where $q$ is a scalar quantity and ${\rm {\bf A}}=(a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinate system.
336
337% -------------------------------------------------------------------------------------------------------------
338% Continuous Model Equations
339% -------------------------------------------------------------------------------------------------------------
340\subsection{Continuous Model Equations}
341\label{PE_zco_Eq}
342
343In order to express the Primitive Equations in tensorial formalism, it is necessary to compute the horizontal component of the non-linear and viscous terms of the equation using \eqref{Eq_PE_grad}) to \eqref{Eq_PE_lap_vector}. Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$, by:
344\begin{equation} \label{Eq_PE_curl_Uh}
345\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v} 
346\right)}{\partial i}-\frac{\partial \left( {e_1 \,u} \right)}{\partial j}} 
347\right]
348\end{equation}
349\begin{equation} \label{Eq_PE_div_Uh}
350\chi =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,u} 
351\right)}{\partial i}+\frac{\partial \left( {e_1 \,v} \right)}{\partial j}} 
352\right]
353\end{equation}
354
355Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that $e_3$  is a function of the single variable $k$, the nonlinear term of \eqref{Eq_PE_dyn} can be transformed as follows:
356\begin{flalign*}
357&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
358+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h        &
359\end{flalign*}
360\begin{flalign*}
361&\qquad=\left( {{\begin{array}{*{20}c}
362 {\left[    {   \frac{1}{e_3} \frac{\partial u  }{\partial k}
363         -\frac{1}{e_1} \frac{\partial w  }{\partial i} } \right] w - \zeta \; v }     \\
364      {\zeta \; u - \left[ {   \frac{1}{e_2} \frac{\partial w}{\partial j}
365                     -\frac{1}{e_3} \frac{\partial v}{\partial k} } \right] \ w}  \\
366       \end{array} }} \right)       
367+\frac{1}{2}   \left( {{\begin{array}{*{20}c}
368       { \frac{1}{e_1}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial i}}  \hfill    \\
369       { \frac{1}{e_2}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial j}}  \hfill    \\
370       \end{array} }} \right)       &
371\end{flalign*}
372\begin{flalign*}
373& \qquad =\left( {{  \begin{array}{*{20}c}
374 {-\zeta \; v} \hfill \\
375 { \zeta \; u} \hfill \\
376         \end{array} }} \right)
377+\frac{1}{2}\left( {{   \begin{array}{*{20}c}
378 {\frac{1}{e_1 }\frac{\partial \left( {u^2+v^2} \right)}{\partial i}} \hfill  \\
379 {\frac{1}{e_2 }\frac{\partial \left( {u^2+v^2} \right)}{\partial j}} \hfill  \\
380                  \end{array} }} \right)       
381+\frac{1}{e_3 }\left( {{      \begin{array}{*{20}c}
382 { w \; \frac{\partial u}{\partial k}}    \\
383 { w \; \frac{\partial v}{\partial k}}    \\
384                     \end{array} }} \right
385-\left( {{  \begin{array}{*{20}c}
386 {\frac{w}{e_1}\frac{\partial w}{\partial i}
387 -\frac{1}{2e_1}\frac{\partial w^2}{\partial i}} \hfill \\
388 {\frac{w}{e_2}\frac{\partial w}{\partial j}
389  -\frac{1}{2e_2}\frac{\partial w^2}{\partial j}} \hfill \\
390         \end{array} }} \right)        &
391\end{flalign*}
392
393The last term of the right hand side is obviously zero, and thus the nonlinear term of \eqref{Eq_PE_dyn} is written in the $(i,j,k)$ coordinate system:
394\begin{equation} \label{Eq_PE_vector_form}
395\left[ {\left( {  \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
396+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h
397=\zeta 
398\;{\rm {\bf k}}\times {\rm {\bf U}}_h +\frac{1}{2}\nabla _h \left( {{\rm 
399{\bf U}}_h^2 } \right)+\frac{1}{e_3 }w\frac{\partial {\rm {\bf U}}_h
400}{\partial k}     
401\end{equation}
402
403This is the so-called \textit{vector invariant form} of the momentum advection term. For some purposes, it can be advantageous to write this term in the so-called flux form, $i.e.$ to write it as the divergence of fluxes. For example, the first component of \eqref{Eq_PE_vector_form} (the $i$-component) is transformed as follows:
404\begin{flalign*}
405&{ \begin{array}{*{20}l}
406\left[ {\left( {\nabla \times \vect{U}} \right)\times \vect{U}
407          +\frac{1}{2}\nabla \left( {\vect{U}}^2 \right)} \right]_i   % \\
408%\\
409     = - \zeta \;v
410     + \frac{1}{2\;e_1 } \frac{\partial \left( {u^2+v^2} \right)}{\partial i}
411     + \frac{1}{e_3}w \ \frac{\partial u}{\partial k}          \\
412\\
413\qquad =\frac{1}{e_1 \; e_2} \left(    -v\frac{\partial \left( {e_2 \,v} \right)}{\partial i}
414                     +v\frac{\partial \left( {e_1 \,u} \right)}{\partial j}    \right)
415+\frac{1}{e_1 e_2 }\left(  +e_2 \; u\frac{\partial u}{\partial i}
416                     +e_2 \; v\frac{\partial v}{\partial i}              \right)
417+\frac{1}{e_3}       \left(   w\;\frac{\partial u}{\partial k}       \right)   \\
418\end{array} }        &
419\end{flalign*}
420\begin{flalign*}
421&{ \begin{array}{*{20}l}
422\qquad =\frac{1}{e_1 \; e_2}  \left\{ 
423 -\left(        v^\frac{\partial e_2                                }{\partial i} 
424      +e_2 \,v    \frac{\partial v                                   }{\partial i}     \right)
425+\left(           \frac{\partial \left( {e_1 \,u\,v}  \right)}{\partial j}
426      -e_1 \,u    \frac{\partial v                                   }{\partial j}  \right\right.
427\\  \left\qquad \qquad \quad
428+\left(           \frac{\partial \left( {e_2 u\,u}     \right)}{\partial i}
429      -u       \frac{\partial \left( {e_2 u}         \right)}{\partial i}  \right)
430+e_2 v            \frac{\partial v                                    }{\partial i}
431                  \right\} 
432+\frac{1}{e_3} \left(
433               \frac{\partial \left( {w\,u} \right)         }{\partial k}
434       -u         \frac{\partial w                    }{\partial k}  \right) \\
435\end{array} }     &
436\end{flalign*}
437\begin{flalign*}
438&{ \begin{array}{*{20}l}
439\qquad =\frac{1}{e_1 \; e_2}  \left(
440               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
441      +        \frac{\partial \left( {e_1 \,u\,v} \right)}{\partial j}  \right)
442+\frac{1}{e_3 }      \frac{\partial \left( {w\,u       } \right)}{\partial k}
443\\  \qquad \qquad \quad
444+\frac{1}{e_1 e_2 }     \left(
445      -u \left(   \frac{\partial \left( {e_1 v   } \right)}{\partial j}
446               -v\,\frac{\partial e_1 }{\partial j}             \right)
447      -u       \frac{\partial \left( {e_2 u   } \right)}{\partial i}
448                  \right)
449 -\frac{1}{e_3 }     \frac{\partial w}{\partial k} u
450 +\frac{1}{e_1 e_2 }\left(    -v^2\frac{\partial e_2   }{\partial i}     \right)
451\end{array} }     &
452\end{flalign*}
453\begin{flalign*}
454&{ \begin{array}{*{20}l}
455\qquad = \nabla \cdot \left( {{\rm {\bf U}}\,u} \right)
456-   \left( \nabla \cdot {\rm {\bf U}} \right) \ u
457+\frac{1}{e_1 e_2 }\left(
458      -v^2     \frac{\partial e_2 }{\partial i}
459      +uv   \,    \frac{\partial e_1 }{\partial j}    \right) \\
460\end{array} }     &
461\end{flalign*}
462as $\nabla \cdot {\rm {\bf U}}\;=0$ (incompressibility) it comes:
463\begin{flalign*}
464&{ \begin{array}{*{20}l}
465\qquad = \nabla \cdot \left{{\rm {\bf U}}\,u}      \right)
466\frac{1}{e_1 e_2 }   \left( v \; \frac{\partial e_2}{\partial i}
467                         -u \; \frac{\partial e_1}{\partial j}    \right\left( -v \right)
468\end{array} }     &
469\end{flalign*}
470
471The flux form of the momentum advection term is therefore given by:
472\begin{multline} \label{Eq_PE_flux_form}
473      \left[
474  \left(    {\nabla \times {\rm {\bf U}}}    \right) \times {\rm {\bf U}}
475+\frac{1}{2}   \nabla \left{{\rm {\bf U}}^2}    \right)
476      \right]_h
477\\
478= \nabla \cdot    \left( {{\begin{array}{*{20}c}   {\rm {\bf U}} \, u   \hfill \\
479                                    {\rm {\bf U}} \, v   \hfill \\
480                  \end{array} }}   
481            \right)
482+\frac{1}{e_1 e_2 }     \left(
483       v\frac{\partial e_2}{\partial i}
484      -u\frac{\partial e_1}{\partial j} 
485                  \right) {\rm {\bf k}} \times {\rm {\bf U}}_h
486\end{multline}
487
488The flux form has two terms, the first one is expressed as the divergence of momentum fluxes (hence the flux form name given to this formulation) and the second one is due to the curvilinear nature of the coordinate system used. The latter is called the \emph{metric} term and can be viewed as a modification of the Coriolis parameter:
489\begin{equation} \label{Eq_PE_cor+metric}
490f \to f + \frac{1}{e_1 \; e_2}   \left(    v \frac{\partial e_2}{\partial i}
491                              -u \frac{\partial e_1}{\partial j}  \right)
492\end{equation}
493
494Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$ and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of the Coriolis parameter $f \to f+(u/a) \tan \varphi$.
495
496To sum up, the equations solved by the ocean model can be written in the following tensorial formalism:
497
498\vspace{+10pt}
499$\bullet$ \textit{momentum equations} :
500
501vector invariant form :
502\begin{subequations} \label{Eq_PE_dyn_vect}
503\begin{multline} \label{Eq_PE_dyn_vect_u}
504\frac{\partial u}{\partial t}=
505   +   \left( {\zeta +f} \right)\,v                                   
506   -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
507   -   \frac{1}{e_3} w \frac{\partial u}{\partial k}       \\
508   -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s+p_h }{\rho _o}    \right)   
509+   D_u^{\vect{U}}  +   F_u^{\vect{U}}
510\end{multline}
511\begin{multline} \label{Eq_PE_dyn_vect_v}
512\frac{\partial v}{\partial t}=
513   -   \left( {\zeta +f} \right)\,u   
514   -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^\right)        -   \frac{1}{e_3 }w\frac{\partial v}{\partial k}         \\
515   -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)   
516+  D_v^{\vect{U}}  +   F_v^{\vect{U}}
517\end{multline}
518\end{subequations}
519
520flux form:
521\begin{subequations} \label{Eq_PE_dyn_flux}
522\begin{multline} \label{Eq_PE_dyn_flux_u}
523\frac{\partial u}{\partial t}=
524+   \left( { f + \frac{1}{e_1 \; e_2}
525               \left(    v \frac{\partial e_2}{\partial i}
526                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, v    \\
527- \frac{1}{e_1 \; e_2}  \left(
528               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
529      +        \frac{\partial \left( {e_1 \,v\,u} \right)}{\partial j}  \right)
530                 - \frac{1}{e_3 }\frac{\partial \left( {         w\,u} \right)}{\partial k}    \\
531-   \frac{1}{e_1 }\frac{\partial}{\partial i}\left( \frac{p_s+p_h }{\rho _o}   \right)
532+   D_u^{\vect{U}} +   F_u^{\vect{U}}
533\end{multline}
534\begin{multline} \label{Eq_PE_dyn_flux_v}
535\frac{\partial v}{\partial t}=
536-   \left( { f + \frac{1}{e_1 \; e_2}
537               \left(    v \frac{\partial e_2}{\partial i}
538                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, u   \\
539 \frac{1}{e_1 \; e_2}   \left(
540               \frac{\partial \left( {e_2 \,u\,v} \right)}{\partial i}
541      +        \frac{\partial \left( {e_1 \,v\,v} \right)}{\partial j}  \right)
542                 - \frac{1}{e_3 } \frac{\partial \left( {        w\,v} \right)}{\partial k}    \\
543-   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}    \right)
544+  D_v^{\vect{U}} +  F_v^{\vect{U}} 
545\end{multline}
546\end{subequations}
547where $\zeta$ is given by \eqref{Eq_PE_curl_Uh} and the surface pressure gradient formulation depends on the one of the free surface:
548
549$*$ free surface formulation
550\begin{equation}\label{Eq_PE_dyn_sco}
551\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
552 {\frac{g}{\;e_1 }\frac{\partial \eta }{\partial i}} \hfill \\
553 {\frac{g}{\;e_2 }\frac{\partial \eta }{\partial j}} \hfill \\
554\end{array} }} \right)
555\qquad \text{where $\eta$ is solution of \eqref{Eq_PE_ssh} }
556\end{equation}
557
558$*$ rigid-lid approximation
559\begin{equation}\label{Eq_PE_dyn_zco}
560\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
561 {\overline M _u +\frac{1}{H\;e_2 }\frac{\partial }{\partial j}\left(
562{\frac{\partial \psi }{\partial t}} \right)}     \\
563 {\overline M _v -\frac{1}{H\;e_1 }\frac{\partial }{\partial i}\left(
564{\frac{\partial \psi }{\partial t}} \right)}        \\
565\end{array} }} \right)
566\end{equation}
567where ${\vect{M}}= \left( M_u,M_v \right)$ represents the collected contributions of nonlinear,
568viscosity and hydrostatic pressure gradient terms in \eqref{Eq_PE_dyn_vect} and the overbar indicates a vertical average over the whole water column ($i.e.$ from $z=-H$, the ocean bottom, to $z=0$, the rigid-lid), and where the time derivative of $\psi$ is the solution of an elliptic equation:
569\begin{multline} \label{Eq_psi_total}
570  \frac{\partial }{\partial i}\left[ {\frac{e_2 }{H\,e_1}\frac{\partial}{\partial i}
571                         \left( {\frac{\partial \psi }{\partial t}} \right)}   \right]
572+\frac{\partial }{\partial j}\left[ {\frac{e_1 }{H\,e_2}\frac{\partial }{\partial j}
573                         \left( {\frac{\partial \psi }{\partial t}} \right)} \right]
574= \\
575  \frac{\partial }{\partial i}\left( {e_2 \overline M _v } \right)
576- \frac{\partial }{\partial j}\left( {e_1 \overline M _u } \right)
577\end{multline}
578
579The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
580\begin{equation} \label{Eq_w_diag}
581\frac{\partial w}{\partial k}=-\chi \;e_3
582\end{equation}
583\begin{equation} \label{Eq_hp_diag}
584\frac{\partial p_h }{\partial k}=-\rho \;g\;e_3
585\end{equation}
586
587where the divergence of the horizontal velocity, $\chi$ is given by \eqref{Eq_PE_div_Uh}.
588
589\vspace{+10pt}
590$\bullet$ \textit{tracer equations} :
591\begin{equation} \label{Eq_S}
592\frac{\partial T}{\partial t} =
593-\frac{1}{e_1 e_2 }\left[ {      \frac{\partial \left( {e_2 T\,u} \right)}{\partial i}
594                  +\frac{\partial \left( {e_1 T\,v} \right)}{\partial j}} \right]
595-\frac{1}{e_3 }\frac{\partial \left( {T\,w} \right)}{\partial k} + D^T + F^T
596\end{equation}
597\begin{equation} \label{Eq_T}
598\frac{\partial S}{\partial t} =
599-\frac{1}{e_1 e_2 }\left[    {\frac{\partial \left( {e_2 S\,u} \right)}{\partial i}
600                  +\frac{\partial \left( {e_1 S\,v} \right)}{\partial j}} \right]
601-\frac{1}{e_3 }\frac{\partial \left( {S\,w} \right)}{\partial k} + D^S + F^S
602\end{equation}
603\begin{equation} \label{Eq_rho}
604\rho =\rho \left( {T,S,z(k)} \right)
605\end{equation}
606
607The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid
608scale parameterization used. It will be defined in \S\ref{PE_zdf}. The nature and formulation of ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed in Chapter~\ref{SBC}.
609
610\newpage 
611% ================================================================
612% Curvilinear z*-coordinate System
613% ================================================================
614\section{Curvilinear \textit{z*}-coordinate System}
615
616%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
617\begin{figure}[!b] \label{Fig_z_zstar}  \begin{center}
618\includegraphics[width=1.0\textwidth]{./Figures/Fig_z_zstar.pdf}
619\caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate \citep{Adcroft_Campin_OM04} ).}
620\end{center}   \end{figure}
621%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
622
623
624In that case, the free surface equation is nonlinear, and the variations of volume are fully taken into account. These coordinates systems is presented in a report
625\citep{Levier2007} available on the \NEMO web site.
626
627\gmcomment{
628The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation which allows one to deal with large amplitude free-surface
629variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In
630the  \textit{z*} formulation, the variation of the column thickness due to sea-surface
631undulations is not concentrated in the surface level, as in the $z$-coordinate formulation,
632but is equally distributed over the full water column. Thus vertical
633levels naturally follow sea-surface variations, with a linear attenuation with
634depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in
635fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent.
636The definition and modified oceanic equations for the rescaled vertical coordinate
637 \textit{z*}, including the treatment of fresh-water flux at the surface, are
638detailed in Adcroft and Campin (2004). The major points are summarized
639here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as:
640
641$H +  \textit{z*} = (H + z) / r$ and  $\delta \textit{z*} = \delta z / r$ with $r = \frac{H+\eta} {H}$
642
643Since the vertical displacement of the free surface is incorporated in the vertical
644coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,  $\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field is no longer zero as shown by the continuity equation:
645
646$\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right)
647      \left( r \; w\textit{*} \right) = 0 $
648
649}
650
651
652\newpage 
653% ================================================================
654% Curvilinear s-coordinate System
655% ================================================================
656\section{Curvilinear \textit{s}-coordinate System}
657\label{PE_sco}
658
659% -------------------------------------------------------------------------------------------------------------
660% Introduction
661% -------------------------------------------------------------------------------------------------------------
662\subsection{Introduction}
663
664Several important aspects of the ocean circulation are influenced by bottom topography. Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers, sills and channels that strongly constrain the path of water masses, but more subtle effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary one along continental slopes. Topographic Rossby waves can be excited and can interact with the mean current. In the $z-$coordinate system presented in the previous section (\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom and to large localized depth gradients associated with large localized vertical velocities. The response to such a velocity field often leads to numerical dispersion effects. One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}. Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)
665
666The $s$-coordinate avoids the discretisation error in the depth field since the layers of computation are gradually adjusted with depth to the ocean bottom. Relatively small topographic features as well as  gentle, large-scale slopes of the sea floor in the deep ocean, which would be ignored in typical $z$-model applications with the largest grid spacing at greatest depths, can easily be represented (with relatively low vertical resolution). A terrain-following model (hereafter $s-$model) also facilitates the modelling of the boundary layer flows over a large depth range, which in the framework of the $z$-model would require high vertical resolution over the whole depth range. Moreover, with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as the only boundaries of the domain (nomore lateral boundary condition to specify). Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean, it has strong limitations as soon as stratification is introduced. The main two problems come from the truncation error in the horizontal pressure gradient and a possibly increased diapycnal diffusion. The horizontal pressure force in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}),
667
668\begin{equation} \label{Eq_PE_p_sco}
669\left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial 
670p}{\partial s}\left. {\nabla z} \right|_s
671\end{equation}
672
673The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface and introduces a truncation error that is not present in a $z$-model. In the special case of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$), \citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude of this truncation error. It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state, and the finite difference scheme. This error limits the possible topographic slopes that a model can handle at a given horizontal and vertical resolution. This is a severe restriction for large-scale applications using realistic bottom topography. The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive. This problem can be at least partially overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec1996}. However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for a realistic bottom topography: a envelope topography is defined in $s$-coordinate on which a full or partial step bottom topography is then applied in order to adjust the model depth to the observed one (see \S\ref{DOM_zgr}.
674
675For numerical reasons a minimum of diffusion is required along the coordinate surfaces of any finite difference model. It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as well as for a $s$-model.
676However, density varies more strongly on $s-$surfaces than on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a $z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation. For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast, the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column (i.e. the main thermocline) \citep{Madec1996}. An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}. Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large, strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).
677
678The $s-$coordinates introduced here \citep{Lott1990,Madec1996} differ mainly in two aspects from similar models:  it allows  a representation of bottom topography with mixed full or partial step-like/terrain following topography ; It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate.
679
680% -------------------------------------------------------------------------------------------------------------
681% The s-coordinate Formulation
682% -------------------------------------------------------------------------------------------------------------
683\subsection{The \textit{s-}coordinate Formulation}
684
685Starting from the set of equations established in \S\ref{PE_zco} for the special case $k=z$ and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes $z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and $s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed equations is given in Appendix~\ref{Apdx_A}. Let us define the vertical scale factor by $e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the (\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by :
686\begin{equation} \label{Eq_PE_sco_slope}
687\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s
688\quad \text{, and } \quad 
689\sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right|_s
690\end{equation}
691We also introduce  $\omega $, a dia-surface velocity component, defined as the velocity relative to the moving $s$-surfaces and normal to them:
692\begin{equation} \label{Eq_PE_sco_w}
693\omega  = w - e_3 \, \frac{\partial s}{\partial t} - \sigma _1 \,u - \sigma _2 \,v    \\
694\end{equation}
695
696The equations solved by the ocean model \eqref{Eq_PE} in $s-$coordinate can be written as follows:
697
698 \vspace{0.5cm}
699* momentum equation:
700\begin{multline} \label{Eq_PE_sco_u}
701\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
702   +   \left( {\zeta +f} \right)\,v                                   
703   -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
704   -   \frac{1}{e_3} \omega \frac{\partial u}{\partial k}       \\
705   -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s + p_h}{\rho _o}    \right)   
706   +  g\frac{\rho }{\rho _o}\sigma _1
707   +   D_u^{\vect{U}}  +   F_u^{\vect{U}} \quad
708\end{multline}
709\begin{multline} \label{Eq_PE_sco_v}
710\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
711   -   \left( {\zeta +f} \right)\,u   
712   -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^\right)       
713   -   \frac{1}{e_3 } \omega \frac{\partial v}{\partial k}         \\
714   -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)
715    +  g\frac{\rho }{\rho _o }\sigma _2   
716   +  D_v^{\vect{U}}  +   F_v^{\vect{U}} \quad
717\end{multline}
718where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic pressure have the same expressions as in $z$-coordinates although they do not represent exactly the same quantities. $\omega$ is provided by the continuity equation (see Appendix~\ref{Apdx_A}):
719
720\begin{equation} \label{Eq_PE_sco_continuity}
721\frac{\partial e_3}{\partial t} + e_3 \; \chi + \frac{\partial \omega }{\partial s} = 0   
722\qquad \text{with }\;\; 
723\chi =\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3 \,u} 
724\right)}{\partial i}+\frac{\partial \left( {e_1 e_3 \,v} \right)}{\partial 
725j}} \right]
726\end{equation}
727
728 \vspace{0.5cm}
729* tracer equations:
730\begin{multline} \label{Eq_PE_sco_t}
731\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
732-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,T} \right)}{\partial i}
733                                           +\frac{\partial \left( {e_1 e_3\,v\,T} \right)}{\partial j}} \right]   \\
734-\frac{1}{e_3 }\frac{\partial \left( {T\,\omega } \right)}{\partial k}   + D^T + F^S   \qquad
735\end{multline}
736
737\begin{multline} \label{Eq_PE_sco_s}
738\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
739-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,S} \right)}{\partial i}
740                                           +\frac{\partial \left( {e_1 e_3\,v\,S} \right)}{\partial j}} \right]    \\
741-\frac{1}{e_3 }\frac{\partial \left( {S\,\omega } \right)}{\partial k}     + D^S + F^S   \qquad
742\end{multline}
743
744The equation of state has the same expression as in $z$-coordinate, and similar expressions are used for mixing and forcing terms.
745
746\gmcomment{
747\colorbox{yellow}{ to be updated $= = >$}
748Add a few works on z and zps and s and underlies the differences between all of them
749\colorbox{yellow}{ $< = =$ end update}  }
750
751\newpage 
752% ================================================================
753% Subgrid Scale Physics
754% ================================================================
755\section{Subgrid Scale Physics}
756\label{PE_zdf_ldf}
757
758The primitive equations describe the behaviour of a geophysical fluid at
759space and time scales larger than a few kilometres in the horizontal, a few
760meters in the vertical and a few minutes. They are usually solved at larger
761scales: the specified grid spacing and time step of the numerical model. The
762effects of smaller scale motions (coming from the advective terms in the
763Navier-Stokes equations) must be represented entirely in terms of
764large-scale patterns to close the equations. These effects appear in the
765equations as the divergence of turbulent fluxes ($i.e.$ fluxes associated with
766the mean correlation of small scale perturbations). Assuming a turbulent
767closure hypothesis is equivalent to choose a formulation for these fluxes.
768It is usually called the subgrid scale physics. It must be emphasized that
769this is the weakest part of the primitive equations, but also one of the
770most important for long-term simulations as small scale processes \textit{in fine} balance
771the surface input of kinetic energy and heat.
772
773The control exerted by gravity on the flow induces a strong anisotropy
774between the lateral and vertical motions. Therefore subgrid-scale physics  \textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \eqref{Eq_PE_dyn}, \eqref{Eq_PE_tra_T} and \eqref{Eq_PE_tra_S} are divided into a lateral part  \textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part  \textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms and their underlying physics are briefly discussed in the next two subsections.
775
776% -------------------------------------------------------------------------------------------------------------
777% Vertical Subgrid Scale Physics
778% -------------------------------------------------------------------------------------------------------------
779\subsection{Vertical Subgrid Scale Physics}
780\label{PE_zdf}
781
782The model resolution is always larger than the scale at which the major
783sources of vertical turbulence occur (shear instability, internal wave
784breaking...). Turbulent motions are thus never explicitly solved, even
785partially, but always parameterized. The vertical turbulent fluxes are
786assumed to depend linearly on the gradients of large-scale quantities (for
787example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$, where $A^{vT}$ is an eddy coefficient). This formulation is
788analogous to that of molecular diffusion and dissipation. This is quite
789clearly a necessary compromise: considering only the molecular viscosity
790acting on large scale severely underestimates the role of turbulent
791diffusion and dissipation, while an accurate consideration of the details of
792turbulent motions is simply impractical. The resulting vertical momentum and
793tracer diffusive operators are of second order:
794\begin{equation} \label{Eq_PE_zdf}
795   \begin{split}
796{\vect{D}}^{v \vect{U}}
797&=\frac{\partial }{\partial z}\left( {A^{vm}\frac{\partial {\vect{U}}_h }{\partial z}} \right) \ , \\         D^{vT} &= \frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial T}{\partial z}} \right) \ ,
798\quad
799D^{vS}=\frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial S}{\partial z}} \right)
800   \end{split}
801\end{equation}
802where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients, respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified (see Chap.~\ref{SBC} and \ref{ZDF} and \S\ref{TRA_bbl}). All the vertical physics is embedded in the specification of the eddy coefficients. They can be assumed to be either constant, or function of the local fluid properties ($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a turbulent closure model. The choices available in \NEMO are discussed in \S\ref{ZDF}).
803
804% -------------------------------------------------------------------------------------------------------------
805% Lateral Diffusive and Viscous Operators Formulation
806% -------------------------------------------------------------------------------------------------------------
807\subsection{Lateral Diffusive and Viscous Operators Formulation}
808\label{PE_ldf}
809
810Lateral turbulence can be roughly divided into a mesoscale turbulence
811associated with eddies (which can be solved explicitly if the resolution is
812sufficient since their underlying physics are included in the primitive
813equations), and a sub mesoscale turbulence which is never explicitly solved
814even partially, but always parameterized. The formulation of lateral eddy
815fluxes depends on whether the mesoscale is below or above the grid-spacing
816($i.e.$ the model is eddy-resolving or not).
817
818In non-eddy-resolving configurations, the closure is similar to that used
819for the vertical physics. The lateral turbulent fluxes are assumed to depend
820linearly on the lateral gradients of large-scale quantities. The resulting
821lateral diffusive and dissipative operators are of second order.
822Observations show that lateral mixing induced by mesoscale turbulence tends
823to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987}) rather than across them.
824As the slope of neutral surfaces is small in the ocean, a common
825approximation is to assume that the `lateral' direction is the horizontal,
826$i.e.$ the lateral mixing is performed along geopotential surfaces. This leads
827to a geopotential second order operator for lateral subgrid scale physics.
828This assumption can be relaxed: the eddy-induced turbulent fluxes can be
829better approached by assuming that they depend linearly on the gradients of
830large-scale quantities computed along neutral surfaces. In such a case,
831the diffusive operator is an isoneutral second order operator and it has
832components in the three space directions. However, both horizontal and
833isoneutral operators have no effect on mean ($i.e.$ large scale) potential
834energy whereas potential energy is a main source of turbulence (through
835baroclinic instabilities). \citet{Gent1990} have proposed a
836parameterization of mesoscale eddy-induced turbulence which associates an
837eddy-induced velocity to the isoneutral diffusion. Its mean effect is to
838reduce the mean potential energy of the ocean. This leads to a formulation
839of lateral subgrid-scale physics made up of an isoneutral second order
840operator and an eddy induced advective part. In all these lateral diffusive
841formulations, the specification of the lateral eddy coefficients remains the
842problematic point as there is no really satisfactory formulation of these
843coefficients as a function of large-scale features.
844
845In eddy-resolving configurations, a second order operator can be used, but
846usually a more scale selective one (biharmonic operator) is preferred as the
847grid-spacing is usually not small enough compared to the scale of the
848eddies. The role devoted to the subgrid-scale physics is to dissipate the
849energy that cascades toward the grid scale and thus ensures the stability of
850the model while not interfering with the solved mesoscale activity. Another approach
851is becoming more and more popular: instead of specifying explicitly a sub-grid scale
852term in the momentum and tracer time evolution equations, one uses a advective scheme which is diffusive enough to maintain the model stability. It must be emphasised
853that then, all the sub-grid scale physics is in this case include in the formulation of the
854advection scheme.
855
856All these parameterizations of subgrid scale physics present advantages and
857drawbacks. There are not all available in \NEMO. In the $z$-coordinate
858formulation, five options are offered for active tracers (temperature and
859salinity): second order geopotential operator, second order isoneutral
860operator, \citet{Gent1990} parameterization, fourth order
861geopotential operator, and various slightly diffusive advection schemes. The same options are available for momentum, except
862\citet{Gent1990} parameterization which only involves tracers. In the
863$s$-coordinate formulation, additional options are offered for tracers: second
864order operator acting along $s-$surfaces, and for momentum: fourth order
865operator acting along $s-$surfaces (see \S\ref{LDF}).
866
867\subsubsection{lateral second order tracer diffusive operator}
868
869The lateral second order tracer diffusive operator is defined by (see Appendix~\ref{Apdx_B}):
870\begin{equation} \label{Eq_PE_iso_tensor}
871D^{lT}=\nabla {\rm {\bf .}}\left( {A^{lT}\;\Re \;\nabla T} \right) \qquad 
872\mbox{with}\quad \;\;\Re =\left( {{\begin{array}{*{20}c}
873 1 \hfill & 0 \hfill & {-r_1 } \hfill \\
874 0 \hfill & 1 \hfill & {-r_2 } \hfill \\
875 {-r_1 } \hfill & {-r_2 } \hfill & {r_1 ^2+r_2 ^2} \hfill \\
876\end{array} }} \right)
877\end{equation}
878where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along
879which the diffusive operator acts and the model level ($e. g.$ $z$- or
880$s$-surfaces). Note that the formulation \eqref{Eq_PE_iso_tensor} is exact for the
881rotation between geopotential and $s$-surfaces, while it is only an approximation
882for the rotation between isoneutral and $z$- or $s$-surfaces. Indeed, in the latter
883case, two assumptions are made to simplify  \eqref{Eq_PE_iso_tensor} \citep{Cox1987}. First, the horizontal contribution of the dianeutral mixing is neglected since the ratio between iso and dia-neutral diffusive coefficients is known to be several orders of magnitude smaller than unity. Second, the two isoneutral directions of diffusion are assumed to be independent since the slopes are generally less than $10^{-2}$ in the ocean (see Appendix~\ref{Apdx_B}).
884
885For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the
886geopotential and computational surfaces: in $z$-coordinates they are zero ($r_1 = r_2 = 0$) while in $s$-coordinate (including $\textit{z*}$ case) they are equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ).
887
888For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral and computational surfaces. Therefore, they have a same expression in $z$- and $s$-coordinates:
889\begin{equation} \label{Eq_PE_iso_slopes}
890r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
891                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1} \ , \quad
892r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
893                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1}
894\end{equation}
895
896When the \textit{Eddy Induced Velocity} parametrisation (eiv) \citep{Gent1990} is used, an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
897\begin{equation} \label{Eq_PE_iso+eiv}
898D^{lT}=\nabla \cdot \left( {A^{lT}\;\Re \;\nabla T} \right)
899           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right)
900\end{equation}
901where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent, eddy-induced transport velocity. This velocity field is defined by:
902\begin{equation} \label{Eq_PE_eiv}
903   \begin{split}
904 u^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_1 } \right] \\ 
905 v^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_2 } \right] \\ 
906 w^\ast &=  -\frac{1}{e_1 e_2 }\left[
907                      \frac{\partial }{\partial i}\left( {A^{eiv}\;e_2\,\tilde{r}_1 } \right)
908                    +\frac{\partial }{\partial j}\left( {A^{eiv}\;e_1\,\tilde{r}_2 } \right)      \right]
909   \end{split}
910\end{equation}
911where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes between isoneutral and \emph{geopotential} surfaces and thus depends on the coordinate considered:
912\begin{align} \label{Eq_PE_slopes_eiv}
913\tilde{r}_n = \begin{cases}
914   r_n                  &      \text{in $z$-coordinate}    \\
915   r_n + \sigma_n &      \text{in \textit{z*} and $s$-coordinates} 
916                   \end{cases}
917\quad \text{where } n=1,2
918\end{align}
919
920The normal component of the eddy induced velocity is zero at all the boundaries. this can be achieved in a model by tapering either the eddy coefficient or the slopes to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. Chap.~\ref{LDF}).
921
922\subsubsection{lateral fourth order tracer diffusive operator}
923
924The lateral fourth order tracer diffusive operator is defined by:
925\begin{equation} \label{Eq_PE_bilapT}
926D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
927\qquad \text{where} \  D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
928 \end{equation}
929
930It is the second order operator given by \eqref{Eq_PE_iso_tensor} applied twice with the eddy diffusion coefficient correctly placed.
931
932
933\subsubsection{lateral second order momentum diffusive operator}
934
935The second order momentum diffusive operator along $z$- or $s$-surfaces is found by
936applying \eqref{Eq_PE_lap_vector} to the horizontal velocity vector (see Appendix~\ref{Apdx_B}):
937\begin{equation} \label{Eq_PE_lapU}
938\begin{split}
939{\rm {\bf D}}^{l{\rm {\bf U}}} 
940&= \quad \  \nabla _h \left( {A^{lm}\chi } \right)
941   \ - \ \nabla _h \times \left( {A^{lm}\,\zeta \;{\rm {\bf k}}} \right)     \\
942&=   \left(      \begin{aligned}
943             \frac{1}{e_1      } \frac{\partial \left( A^{lm} \chi          \right)}{\partial i} 
944         &-\frac{1}{e_2 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial j}  \\
945             \frac{1}{e_2      }\frac{\partial \left( {A^{lm} \chi         } \right)}{\partial j}   
946         &+\frac{1}{e_1 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial i}
947        \end{aligned}    \right)
948\end{split}
949\end{equation}
950
951Such a formulation ensures a complete separation between the vorticity and
952horizontal divergence fields (see Appendix~\ref{Apdx_C}). Unfortunately, it is not
953available for geopotential diffusion in $s-$coordinates and for isoneutral
954diffusion in both $z$- and $s$-coordinates ($i.e.$ when a rotation is required). In these two cases, the $u$ and $v-$fields are considered as independent scalar fields, so that the diffusive operator is given by:
955\begin{equation} \label{Eq_PE_lapU_iso}
956\begin{split}
957 D_u^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla u} \right) \\ 
958 D_v^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla v} \right)
959 \end{split}
960 \end{equation}
961where $\Re$ is given by  \eqref{Eq_PE_iso_tensor}. It is the same expression as those used for diffusive operator on tracers. It must be emphasised that such a formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or $\beta-$plane, not on the sphere. It is also a very good approximation in vicinity of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}.
962
963\subsubsection{lateral fourth order momentum diffusive operator}
964
965As for tracers, the fourth order momentum diffusive operator along $z$ or $s$-surfaces is a re-entering second order operator \eqref{Eq_PE_lapU} or \eqref{Eq_PE_lapU} with the eddy viscosity coefficient correctly placed:
966
967geopotential diffusion in $z$-coordinate:
968\begin{equation} \label{Eq_PE_bilapU}
969\begin{split}
970{\rm {\bf D}}^{l{\rm {\bf U}}} &=\nabla _h \left\{ {\;\nabla _h {\rm {\bf 
971.}}\left[ {A^{lm}\,\nabla _h \left( \chi \right)} \right]\;} 
972\right\}\;   \\
973&+\nabla _h \times \left\{ {\;{\rm {\bf k}}\cdot \nabla \times 
974\left[ {A^{lm}\,\nabla _h \times \left( {\zeta \;{\rm {\bf k}}} \right)} 
975\right]\;} \right\}
976\end{split}
977\end{equation}
978
979\gmcomment{  change the position of the coefficient, both here and in the code}
980
981geopotential diffusion in $s$-coordinate:
982\begin{equation} \label{Eq_bilapU_iso}
983   \left\{   \begin{aligned}
984         D_u^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta u} \right) \\ 
985         D_v^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta v} \right)
986   \end{aligned}    \right.
987   \quad \text{where} \quad 
988   \Delta \left( \bullet \right) = \nabla \cdot \left( \Re \nabla(\bullet) \right)
989\end{equation}
990
Note: See TracBrowser for help on using the repository browser.