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/NEMO/DOC/BETA/Chapters – NEMO

source: trunk/NEMO/DOC/BETA/Chapters/Chap_Model_Basics.tex @ 707

Last change on this file since 707 was 707, checked in by smasson, 17 years ago

error during changeset:705... related to ticket:1

  • Property svn:executable set to *
File size: 66.7 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 equations, i.e. the Navier-Stokes equations along with a nonlinear equation of state which couples the two active tracers (temperature and salinity) to the fluid velocity, plus the following additional assumptions made from scale considerations:
25
26\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
27
28\textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius
29
30\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
31
32\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their contribution to the buoyancy force
33
34\textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a balance between the vertical pressure gradient and buoyancy force (this removes convective processes from
35the initial Navier-Stokes equations: they must be parameterized)
36
37\textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity vector is assumed to be zero.
38
39Because the gravitational force is so dominant in the equations of large-scale motions, it is quite 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}) plan), $T$ 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, the heat and salt conservation and an equation of state):
40\begin{subequations} \label{Eq_PE}
41  \begin{equation}     \label{Eq_PE_dyn}
42\frac{\partial {\rm {\bf U}}_h }{\partial t}=
43-\left[    {\left( {\nabla \times {\rm {\bf U}}} \right)\times {\rm {\bf U}}
44            +\frac{1}{2}\nabla \left( {{\rm {\bf U}}^2} \right)}    \right]_h
45 -f\;{\rm {\bf k}}\times {\rm {\bf U}}_h
46-\frac{1}{\rho _o }\nabla _h p + {\rm {\bf D}}^{\rm {\bf U}}
47  \end{equation}
48  \begin{equation}     \label{Eq_PE_hydrostatic}
49\frac{\partial p }{\partial z} = - \rho \ g
50  \end{equation}
51  \begin{equation}     \label{Eq_PE_continuity}
52\nabla \cdot {\bf U}=  0
53  \end{equation}
54\begin{equation} \label{Eq_PE_tra_T}
55\frac{\partial T}{\partial t} = - \nabla \cdot  \left( T \ \rm{\bf U} \right) + D^T
56  \end{equation}
57  \begin{equation}     \label{Eq_PE_tra_S}
58\frac{\partial S}{\partial t} = - \nabla \cdot  \left( S \ \rm{\bf U} \right) + D^S
59  \end{equation}
60  \begin{equation}     \label{Eq_PE_eos}
61\rho = \rho \left( T,S,p \right)
62  \end{equation}
63\end{subequations}
64where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions, $t$ the time, $z$ the vertical coordinate, $\rho $ the \textit{in situ} density given by the equation of state (\ref{Eq_PE_eos}), $\rho_o$ a reference density, $p$ the pressure, $f=2 \bf \Omega \cdot \bf k$ the Coriolis acceleration (where $\bf \Omega$ is the Earth angular velocity vector), and $g$ the gravitational acceleration. ${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterizations of small-scale physics for momentum, temperature and salinity, including surface forcing terms. Their nature and formulation are discussed in \S\ref{PE_zdf_ldf}, page~\pageref{PE_zdf_ldf}.
65
66.
67
68% -------------------------------------------------------------------------------------------------------------
69% Boundary condition
70% -------------------------------------------------------------------------------------------------------------
71\subsection{Boundary Conditions}
72\label{PE_boundary_condition}
73
74An ocean is bounded by complex coastlines and bottom topography at its base and by 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 $\eta$ 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 surfaces, 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.
75
76%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
77\begin{figure}[!ht] \label{Fig_ocean_bc}  \begin{center}
78\includegraphics[width=0.90\textwidth]{./Figures/Fig_I_ocean_bc.pdf}
79\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 $z=0$.}
80\end{center}   \end{figure}
81%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
82
83\begin{description}
84\item[Land - ocean interface:] the major flux between continental surfaces and the ocean is a mass exchange of fresh water through river runoff. Such an exchange modifies locally 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 to close the water cycle of the climatic system. It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths.
85\item[Solid earth - ocean interface:] heat and salt fluxes across the sea floor are negligibly small, except in special areas of little extent. They are always neglected in the model. 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:
86\begin{equation} \label{Eq_PE_w_bbc}
87w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right)
88\end{equation}
89In addition, the ocean exchanges momentum with the earth through friction processes. Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized in terms of turbulent fluxes through 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}. They are discussed in  in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9.
90\item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux of fresh water PE  (the precipitation minus evaporation budget) leads to:
91\begin{equation} \label{Eq_PE_w_sbc}
92w = \frac{\partial \eta }{\partial t} 
93    + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right)
94    + P-E
95\end{equation}
96The 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.
97\item[Sea ice - ocean interface:] the two media 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 ($\sim5 \,psu$) 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.
98\end{description}
99
100
101% ================================================================
102% The Horizontal Pressure Gradient
103% ================================================================
104\section{The Horizontal Pressure Gradient }
105\label{PE_hor_pg}
106
107% -------------------------------------------------------------------------------------------------------------
108% Pressure Formulation
109% -------------------------------------------------------------------------------------------------------------
110\subsection{Pressure Formulation}
111\label{PE_p_formulation}
112
113The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a reference geopotential surface ($z=0$) and an 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:
114\begin{equation} \label{Eq_PE_pressure}
115p_h \left( {i,j,z,t} \right)
116 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,z} \right)\;d\varsigma } 
117\end{equation}
118The surface pressure requires a more specific treatment. Two strategies can be considered: $(a)$ the introduction of a new variable $\eta$, the free-surface elevation, for which a prognostic equation can be established and solved; $(b)$ the assumption 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, a solution of the free-surface elevation consists in 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 these waves as 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 OPA, both strategies are still available. They are further described in the next two sub-sections.
119
120% -------------------------------------------------------------------------------------------------------------
121% Free Surface Formulation
122% -------------------------------------------------------------------------------------------------------------
123\subsection{Free Surface Formulation}
124\label{PE_free_surface}
125
126In 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 kinematics surface condition (\ref{Eq_PE_w_bbc}):
127\begin{equation} \label{Eq_PE_ssh}
128\frac{\partial \eta }{\partial t}=-D+P-E
129   \quad \text{where} \
130D=\nabla \cdot \left[ {\left( {H+\eta } \right) \; {\rm{\bf \overline{U}}}_h \,} \right]
131\end{equation}
132and using (\ref{Eq_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
133
134Allowing 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.
135
136Three choices can be made regarding the implementation of the free surface in the model, depending on the physical processes of interest.
137
138$\bullet$ If one is interested in EGWs, in particular the tides and their interaction
139with the baroclinic structure of the ocean (internal waves) possibly in
140shallow seas, then a non linear free surface is the most adequate: this
141means that no approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the
142ocean volume is fully taken into account. Note that in order to study the
143fast time scales associated with EGWs it is necessary to minimize time
144filtering effects (use an explicit time scheme with very small time step, or
145a split-explicit scheme with reasonably small time step, see \S\ref{DYN_spg_exp} or
146\S\ref{DYN_spg_ts}.
147
148$\bullet$ If one is not interested in EGW but rather sees them as high frequency
149noise, it is possible to apply a filter to slow down the fastest waves while
150not altering the slow barotropic Rossby waves. In that case it is also
151generally sufficient to solve a linearized version of (\ref{Eq_PE_ssh}), which allows
152to take into account freshwater fluxes applied at the ocean surface \citep{Roullet2000}.
153
154$\bullet$ For process studies not involving external waves nor surface freshwater
155fluxes, it is possible to use the rigid lid approximation see (next
156section). The ocean surface is considered as a fixed surface, so that all
157external waves are removed from the system.
158
159The 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 OPA, 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:
160\begin{equation} \label{Eq_PE_flt}
161\frac{\partial {\rm {\bf U}}_h }{\partial t}= {\rm {\bf M}}
162- g \nabla \left( \tilde{\rho} \ \eta \right)
163- g \ T_c \nabla \left( \widetilde{\rho} \ \partial_t \eta \right)
164\end{equation}
165where $T_c$, is a parameter homogeneous to a time which characterizes the force, $\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and   represents the collected contributions of the Coriolis, hydrostatic pressure gradient, non-linear and viscous terms in \eqref{Eq_PE_dyn}.
166
167The 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 propagates, $i.e.$ 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 $T_c > \Delta t$.
168
169When the variations of free surface elevation are small compared to the thickness of the model layers, 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
170\begin{equation} \label{Eq_PE_salt_content}
171\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} =\int\limits_S
172{S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds}
173\end{equation}
174where $S$ is the salinity, and the total salt is integrated in 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:
175\begin{equation} \label{Eq_PE_salt_content_linear}
176\frac{\partial }{\partial t}\int\limits_D {S\;dv} =-\int\limits_S
177{S\;\frac{\partial \eta }{\partial t}ds} 
178\end{equation}
179
180The 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
181\begin{equation} \label{Eq_PE_salt_content_corrected}
182\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
183=\frac{\partial}{\partial t} \left[ \;{\int\limits_D {S\;dv} +\int\limits_S {S\eta \;ds} } \right]
184=\int\limits_S {\eta \;\frac{\partial S}{\partial t}ds}
185\end{equation}
186
187Although 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
188the case of the rigid lid approximation (following section) in which case freshwater forcing is represented by a virtual salt flux, leading to spurious sources or sinks of salt \citep{Roullet2000}.
189
190% -------------------------------------------------------------------------------------------------------------
191% Rigid-Lid Formulation
192% -------------------------------------------------------------------------------------------------------------
193\subsection{Rigid-Lid formulation}
194\label{PE_rigid_lid}
195
196With 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 (\ref{Eq_PE_continuity}) and the kinematic condition at the bottom (\ref{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,  can be derived from a volume transport streamfunction $\psi$:
197\begin{equation} \label{Eq_PE_u_psi}
198\overline{\vect{U}}_h =\frac{1}{H}\left(   \vect{k} \times \nabla \psi   \right)
199\end{equation}
200
201As $p_s$ 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}:
202
203\begin{equation} \label{Eq_PE_u_barotrope}
204\frac{1}{\rho _o }\nabla _h p_s
205=\overline{\vect{M}} -\frac{\partial \overline{\vect{U}} _h }{\partial t}
206=\overline{\vect{M}} 
207-\frac{1}{H} \left[   \vect{k} \times \nabla \left( \frac{\partial \psi}{\partial t} \right)   \right]
208\end{equation}
209
210Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ 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}):
211\begin{equation} \label{Eq_PE_psi}
212\left[   {\nabla \times \left[ {\frac{1}{H} \vect{\bf k} 
213  \times \nabla \left(   {\frac{\partial \psi }{\partial t}} \right)}   \right]} \; \right]_z
214=\left[   {\nabla \times \overline{\vect{M}} }   \right]_z
215\end{equation}
216
217Using the proper boundary conditions, (\ref{Eq_PE_psi}) can be solved to find $\partial_t \psi$ and thus using (\ref{Eq_PE_u_barotrope}) the horizontal surface pressure gradient. It should be noted that $p_s$ can be computed by taking the divergence of (\ref{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.
218
219A 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 (I.2.7 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
220gradient field along a closed contour is zero, from (\ref{Eq_PE_u_barotrope}) we have:
221\begin{equation} \label{Eq_PE_isl_circulation}
222\oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
223{\frac{\partial \psi }{\partial t}} \right)} \right] \cdot {\rm {\bf d}}\ell } 
224= \oint_n {\overline {\rm {\bf M}} \cdot {\rm {\bf d}}\ell } 
225\qquad  1 \leq n \leq Q
226\end{equation}
227
228Since (\ref{Eq_PE_psi}) is linear, its solution \textit{$\psi $} can be decomposed as follows:
229\begin{equation} \label{Eq_PE_psi_isl}
230\psi =\psi _o +\sum\limits_{n=1}^{n=Q} {\mu _n \psi _n } 
231\end{equation}
232where $\psi _o$ is the solution of (\ref{Eq_PE_psi}) with $\psi _o=0$ long all the coastlines, and where $\psi _n$ is the solution of (\ref{Eq_PE_psi}) with the right-hand side equal to $0$, and with $\psi _n =1$ long the island $n$, $\psi _n =0$ along the other boundaries. The function $\psi _n$ is thus independent of time. Introducing (\ref{Eq_PE_psi_isl}) into (\ref{Eq_PE_isl_circulation}) yields:
233\begin{multline} \label{Eq_PE_psi_isl_circulation}
234\left[ {\oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \psi _m } 
235\right]\cdot {\rm {\bf d}}\ell } } \right]_{1\leq m\leqslant Q \atop 1\leq n\leqslant Q  }
236 \left( {\frac{\partial \mu _n }{\partial t}} 
237\right)_{1\leqslant n\leqslant Q}        \\
238 =\left( {\oint_n {\left[ {\overline {\rm 
239{\bf M}} -\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
240{\frac{\partial \psi _o }{\partial t}} \right)} \right]} \right]\cdot {\rm 
241{\bf d}}\ell } } \right)_{1\leqslant n\leqslant Q} 
242\end{multline}
243which can be rewritten as:
244\begin{equation} \label{Eq_PE_psi_isl_matrix}
245{\rm {\bf A}}\;\left( {\frac{\partial \mu _n }{\partial t}} 
246\right)_{1\leqslant n\leqslant Q} ={\rm {\bf B}}
247\end{equation}
248where \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:
249\begin{equation} \label{Eq_PE_psi_isl_dt}
250\frac{\partial \psi }{\partial t}=\frac{\partial \psi _o }{\partial 
251t}+\sum\limits_{n=1}^{n=Q} {{\rm {\bf A}}^{-1}{\rm {\bf B}}\;\psi _n } 
252\end{equation}
253
254
255
256% ================================================================
257% Curvilinear z-coordinate System
258% ================================================================
259\section{Curvilinear \textit{z-}coordinate System}
260\label{PE_zco}
261
262
263
264% -------------------------------------------------------------------------------------------------------------
265% Tensorial Formalism
266% -------------------------------------------------------------------------------------------------------------
267\subsection{Tensorial Formalism}
268\label{PE_tensorial}
269
270In 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 in introducing an appropriate coordinate transformation that shifts the singular point on land \citep{MadecImb1996, Murray1996}. As a conclusion, 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, with conservation 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.
271
272Let (\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 (\ref{Fig_referential}). Let $(\lambda,\varphi,z)$ be the geographical coordinates 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
273earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea level (\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:
274\begin{equation} \label{Eq_scale_factors}
275\begin{aligned}
276 e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
277}{\partial i}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
278}{\partial i}} \right)^2} \right]^{1/2} \\ 
279 e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
280}{\partial j}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
281}{\partial j}} \right)^2} \right]^{1/2} \\ 
282 e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\ 
283 \end{aligned}
284 \end{equation}
285
286%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
287\begin{figure}[!tb] \label{Fig_referential}  \begin{center}
288\includegraphics[width=0.60\textwidth]{./Figures/Fig_I_earth_referential.pdf}
289\caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear coordinate system (\textbf{i},\textbf{j},\textbf{k}). }
290\end{center}   \end{figure}
291%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
292
293Since 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:
294\begin{subequations} \label{Eq_PE_discrete_operators}
295\begin{equation} \label{Eq_PE_grad}
296\nabla q=\frac{1}{e_1 }\frac{\partial q}{\partial i}\;{\rm {\bf 
297i}}+\frac{1}{e_2 }\frac{\partial q}{\partial j}\;{\rm {\bf j}}+\frac{1}{e_3
298}\frac{\partial q}{\partial k}\;{\rm {\bf k}}    \\
299\end{equation}
300\begin{equation} \label{Eq_PE_div}
301\nabla \cdot {\rm {\bf A}} 
302= \frac{1}{e_1 \; e_2} \left[
303  \frac{\partial \left(e_2 \; a_1\right)}{\partial i }
304+\frac{\partial \left(e_1 \; a_2\right)}{\partial j }       \right]
305+ \frac{1}{e_3} \left[ \frac{\partial a_3}{\partial k }   \right]
306\end{equation}
307\begin{equation} \label{Eq_PE_curl}
308   \begin{split}
309\nabla \times \vect{A} =
310    \left[ {\frac{1}{e_2 }\frac{\partial a_3}{\partial j}
311            -\frac{1}{e_3 }\frac{\partial a_2 }{\partial k}} \right] \; \vect{i}
312&+\left[ {\frac{1}{e_3 }\frac{\partial a_1 }{\partial k}
313           -\frac{1}{e_1 }\frac{\partial a_3 }{\partial i}} \right] \; \vect{j}     \\
314&+\frac{1}{e_1 e_2 } \left[ {\frac{\partial \left( {e_2 a_2 } \right)}{\partial i}
315                                       -\frac{\partial \left( {e_1 a_1 } \right)}{\partial j}} \right] \; \vect{k} 
316   \end{split}
317\end{equation}
318\begin{equation} \label{Eq_PE_lap}
319\Delta q = \nabla \cdot \left\nabla q \right)
320\end{equation}
321\begin{equation} \label{Eq_PE_lap_vector}
322\Delta {\rm {\bf A}} =
323  \nabla \left( \nabla \cdot {\rm {\bf A}} \right)
324- \nabla \times \left\nabla \times {\rm {\bf A}} \right)
325\end{equation}
326\end{subequations}
327where $q$ is a scalar quantity and ${\rm {\bf A}}=(a_1,a_2,a_3)$ a vector in
328the $(i,j,k)$ coordinate system.
329
330% -------------------------------------------------------------------------------------------------------------
331% Continuous Model Equations
332% -------------------------------------------------------------------------------------------------------------
333\subsection{Continuous Model Equations}
334\label{PE_zco_Eq}
335
336In 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:
337\begin{equation} \label{Eq_PE_curl_Uh}
338\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v} 
339\right)}{\partial i}-\frac{\partial \left( {e_1 \,u} \right)}{\partial j}} 
340\right]
341\end{equation}
342\begin{equation} \label{Eq_PE_div_Uh}
343\chi =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,u} 
344\right)}{\partial i}+\frac{\partial \left( {e_1 \,v} \right)}{\partial j}} 
345\right]
346\end{equation}
347
348Using the fact that 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:
349\begin{flalign*}
350&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
351+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h        &
352\end{flalign*}
353\begin{flalign*}
354&=\left( {{\begin{array}{*{20}c}
355 {\left[    {   \frac{1}{e_3} \frac{\partial u  }{\partial k}
356         -\frac{1}{e_1} \frac{\partial w  }{\partial i} } \right] w - \zeta \; v }     \\
357      {\zeta \; u - \left[ {   \frac{1}{e_2} \frac{\partial w}{\partial j}
358                     -\frac{1}{e_3} \frac{\partial v}{\partial k} } \right] \ w}  \\
359       \end{array} }} \right)       
360+\frac{1}{2}   \left( {{\begin{array}{*{20}c}
361       { \frac{1}{e_1}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial i}}  \hfill    \\
362       { \frac{1}{e_2}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial j}}  \hfill    \\
363       \end{array} }} \right)       &
364\end{flalign*}
365\begin{flalign*}
366&=\left( {{ \begin{array}{*{20}c}
367 {-\zeta \; v} \hfill \\
368 { \zeta \; u} \hfill \\
369         \end{array} }} \right)
370+\frac{1}{2}\left( {{   \begin{array}{*{20}c}
371 {\frac{1}{e_1 }\frac{\partial \left( {u^2+v^2} \right)}{\partial i}} \hfill  \\
372 {\frac{1}{e_2 }\frac{\partial \left( {u^2+v^2} \right)}{\partial j}} \hfill  \\
373                  \end{array} }} \right)       
374+\frac{1}{e_3 }\left( {{      \begin{array}{*{20}c}
375 { w \; \frac{\partial u}{\partial k}}    \\
376 { w \; \frac{\partial v}{\partial k}}    \\
377                     \end{array} }} \right
378-\left( {{  \begin{array}{*{20}c}
379 {\frac{w}{e_1}\frac{\partial w}{\partial i}
380 -\frac{1}{2e_1}\frac{\partial w^2}{\partial i}} \hfill \\
381 {\frac{w}{e_2}\frac{\partial w}{\partial j}
382  -\frac{1}{2e_2}\frac{\partial w^2}{\partial j}} \hfill \\
383         \end{array} }} \right)        &
384\end{flalign*}
385
386The 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:
387\begin{equation} \label{Eq_PE_vector_form}
388\left[ {\left( {  \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
389+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h
390=\zeta 
391\;{\rm {\bf k}}\times {\rm {\bf U}}_h +\frac{1}{2}\nabla _h \left( {{\rm 
392{\bf U}}_h^2 } \right)+\frac{1}{e_3 }w\frac{\partial {\rm {\bf U}}_h
393}{\partial k}     
394\end{equation}
395
396This is the so-called \textit{vector invariant form} of the momentum advection. 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 (\ref{Eq_PE_vector_form}) (the $i$-component) is transformed as follows:
397\begin{flalign*}
398&{ \begin{array}{*{20}l}
399\left[ {\left( {\nabla \times \vect{U}} \right)\times \vect{U}
400          +\frac{1}{2}\nabla \left( {\vect{U}}^2 \right)} \right]_i    \\
401\\
402     = - \zeta \;v
403     + \frac{1}{2\;e_1 } \frac{\partial \left( {u^2+v^2} \right)}{\partial i}
404     + \frac{1}{e_3}w \ \frac{\partial u}{\partial k}          \\
405\\
406=\frac{1}{e_1 \; e_2} \left(  -v\frac{\partial \left( {e_2 \,v} \right)}{\partial i}
407                     +v\frac{\partial \left( {e_1 \,u} \right)}{\partial j}    \right)
408+\frac{1}{e_1 e_2 }\left(  +e_2 \; u\frac{\partial u}{\partial i}
409                     +e_2 \; v\frac{\partial v}{\partial i}              \right)
410+\frac{1}{e_3}       \left(   w\;\frac{\partial u}{\partial k}       \right)   \\
411\end{array} }        &
412\end{flalign*}
413\begin{flalign*}
414&{ \begin{array}{*{20}l}
415=\frac{1}{e_1 \; e_2}   \left\{ 
416 -\left(        v^\frac{\partial e_2                                }{\partial i} 
417      +e_2 \,v    \frac{\partial v                                   }{\partial i}     \right)
418+\left(           \frac{\partial \left( {e_1 \,u\,v}  \right)}{\partial j}
419      -e_1 \,u    \frac{\partial v                                   }{\partial j}  \right\right.
420\\  \left\qquad \qquad \quad
421+\left(           \frac{\partial \left( {e_2 u\,u}     \right)}{\partial i}
422      -u       \frac{\partial \left( {e_2 u}         \right)}{\partial i}  \right)
423+e_2 v            \frac{\partial v                                    }{\partial i}
424                  \right\} 
425+\frac{1}{e_3} \left(
426               \frac{\partial \left( {w\,v} \right)         }{\partial k}
427       -u         \frac{\partial w                    }{\partial k}  \right) \\
428\end{array} }     &
429\end{flalign*}
430\begin{flalign*}
431&{ \begin{array}{*{20}l}
432=\frac{1}{e_1 \; e_2}   \left(
433               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
434      +        \frac{\partial \left( {e_1 \,u\,v} \right)}{\partial j}  \right)
435+\frac{1}{e_3 }      \frac{\partial \left( {w\,v       } \right)}{\partial k}
436\\  \qquad \qquad \quad
437+\frac{1}{e_1 e_2 }     \left(
438      -u \left(   \frac{\partial \left( {e_1 v   } \right)}{\partial j}
439               -v\,\frac{\partial e_1 }{\partial j}             \right)
440      -u       \frac{\partial \left( {e_2 u   } \right)}{\partial i}
441                  \right)
442 -\frac{1}{e_3 }     \frac{\partial w}{\partial k} u
443 +\frac{1}{e_1 e_2 }\left(    -v^2\frac{\partial e_2   }{\partial i}     \right)
444\end{array} }     &
445\end{flalign*}
446\begin{flalign*}
447&{ \begin{array}{*{20}l}
448= \nabla \cdot \left( {{\rm {\bf U}}\,u}  \right)
449\nabla \cdot {\rm {\bf U}} \ u
450+\frac{1}{e_1 e_2 }\left(
451      -v^2     \frac{\partial e_2 }{\partial i}
452      +uv   \,    \frac{\partial e_1 }{\partial j}    \right) \\
453\end{array} }     &
454\end{flalign*}
455as $\nabla \cdot {\rm {\bf U}}\;=0$ (incompressibility) it comes:
456\begin{flalign*}
457&{ \begin{array}{*{20}l}
458= \nabla \cdot \left(   {{\rm {\bf U}}\,u}      \right)
459\frac{1}{e_1 e_2 }   \left( v \; \frac{\partial e_2}{\partial i}
460                         -u \; \frac{\partial e_1}{\partial j}    \right\left( -v \right)
461\end{array} }     &
462\end{flalign*}
463
464The flux form of the momentum advection is therefore given by:
465\begin{multline} \label{Eq_PE_flux_form}
466      \left[
467  \left(    {\nabla \times {\rm {\bf U}}}    \right) \times {\rm {\bf U}}
468+\frac{1}{2}   \nabla \left{{\rm {\bf U}}^2}    \right)
469      \right]_h
470\nabla \cdot   \left( {{\begin{array}{*{20}c}   {\rm {\bf U}} \, u   \hfill \\
471                                    {\rm {\bf U}} \, v   \hfill \\
472                  \end{array} }}   
473            \right)
474\\
475+\frac{1}{e_1 e_2 }     \left(
476       v\frac{\partial e_2}{\partial i}
477      -u\frac{\partial e_1}{\partial j} 
478                  \right) {\rm {\bf k}} \times {\rm {\bf U}}_h
479\end{multline}
480
481The flux form has two terms, the first one is expressed as the divergence of momentum fluxes (so 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:
482\begin{equation} \label{Eq_PE_cor+metric}
483f \to f + \frac{1}{e_1 \; e_2}   \left(    v \frac{\partial e_2}{\partial i}
484                              -u \frac{\partial e_1}{\partial j}  \right)
485\end{equation}
486
487Note 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 $f \to f+(u/a) \tan \varphi$.
488
489The equations solved by the ocean model can be written in the following tensorial formalism:
490
491$\bullet$ vector invariant form of the momentum equations :
492\begin{subequations} \label{Eq_PE_dyn_vect}
493\begin{equation} \label{Eq_PE_dyn_vect_u}
494\frac{\partial u}{\partial t}=
495+   \left( {\zeta +f} \right)\,v   -   \frac{1}{e_3 }w\frac{\partial u}{\partial k}
496-   \frac{1}{e_1 }\frac{\partial}{\partial i}\left(
497      {\frac{1}{2}\left( {u^2+v^2} \right)+\frac{p_h+p_s }{\rho _o }}    \right)
498+   D_u^{\vect{U}} 
499\end{equation}
500\begin{equation} \label{Eq_PE_dyn_vect_v}
501\frac{\partial v}{\partial t}=
502-   \left( {\zeta +f} \right)\,u   -   \frac{1}{e_3 }w\frac{\partial v}{\partial k}
503-   \frac{1}{e_2 }\frac{\partial }{\partial j}\left(
504      {\frac{1}{2}\left( {u^2+v^2} \right)+\frac{p_h+p_s }{\rho _o }}    \right)
505+  D_v^{\vect{U}} 
506\end{equation}
507\end{subequations}
508
509where $\zeta$ is given by (\ref{Eq_PE_curl_Uh}) and the surface pressure gradient is given by:
510
511$*$ free surface formulation
512\begin{equation}\label{Eq_PE_dyn_sco}
513\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
514 {\frac{g}{\;e_1 }\frac{\partial \eta }{\partial i}} \hfill \\
515 {\frac{g}{\;e_2 }\frac{\partial \eta }{\partial j}} \hfill \\
516\end{array} }} \right)
517\qquad \text{where $\eta$ is solution of \eqref{Eq_PE_ssh} }
518\end{equation}
519
520$*$ rigid-lid approximation
521\begin{equation}\label{Eq_PE_dyn_zco}
522\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
523 {\overline M _u +\frac{1}{H\;e_2 }\frac{\partial }{\partial j}\left(
524{\frac{\partial \psi }{\partial t}} \right)}     \\
525 {\overline M _v -\frac{1}{H\;e_1 }\frac{\partial }{\partial i}\left(
526{\frac{\partial \psi }{\partial t}} \right)}        \\
527\end{array} }} \right)
528\end{equation}
529where ${\vect{M}}= \left( M_u,M_v \right)$ represents the collected contributions of nonlinear,
530viscous 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:
531\begin{multline} \label{Eq_psi_total}
532  \frac{\partial }{\partial i}\left[ {\frac{e_2 }{H\,e_1}\frac{\partial}{\partial i}
533                         \left( {\frac{\partial \psi }{\partial t}} \right)}   \right]
534+\frac{\partial }{\partial j}\left[ {\frac{e_1 }{H\,e_2}\frac{\partial }{\partial j}
535                         \left( {\frac{\partial \psi }{\partial t}} \right)} \right]
536= \\
537  \frac{\partial }{\partial i}\left( {e_2 \overline M _v } \right)
538- \frac{\partial }{\partial j}\left( {e_1 \overline M _u } \right)
539\end{multline}
540
541The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
542\begin{equation} \label{Eq_w_diag}
543\frac{\partial w}{\partial k}=-\chi \;e_3
544\end{equation}
545\begin{equation} \label{Eq_hp_diag}
546\frac{\partial p_h }{\partial k}=-\rho \;g\;e_3
547\end{equation}
548
549where the divergence of the horizontal velocity, $\chi$ is given by (I.3.8).
550
551$\bullet$ tracer equations:
552\begin{equation} \label{Eq_S}
553\frac{\partial T}{\partial t}=-\frac{1}{e_1 e_2 }\left[ {\frac{\partial 
554\left( {e_2 T\,u} \right)}{\partial i}+\frac{\partial \left( {e_1 T\,v} 
555\right)}{\partial j}} \right]-\frac{1}{e_3 }\frac{\partial \left( {T\,w} 
556\right)}{\partial k}+D^T
557\end{equation}
558\begin{equation} \label{Eq_T}
559\frac{\partial S}{\partial t}=-\frac{1}{e_1 e_2 }\left[ {\frac{\partial 
560\left( {e_2 S\,u} \right)}{\partial i}+\frac{\partial \left( {e_1 S\,v} 
561\right)}{\partial j}} \right]-\frac{1}{e_3 }\frac{\partial \left( {S\,w} 
562\right)}{\partial k}+D^S
563\end{equation}
564\begin{equation} \label{Eq_rho}
565\rho =\rho \left( {T,S,z(k)} \right)
566\end{equation}
567
568The expression of \textbf{D}$^{U}$, $D^{S}$ and$ D^{T}$ depends on the subgrid
569scale parameterization used. It will be defined in \S\ref{PE_zdf}.
570
571\newpage 
572% ================================================================
573% Curvilinear s-coordinate System
574% ================================================================
575\section{Curvilinear \textit{s}-coordinate System}
576\label{PE_sco}
577
578% -------------------------------------------------------------------------------------------------------------
579% Introduction
580% -------------------------------------------------------------------------------------------------------------
581\subsection{Introduction}
582
583Several 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
584example, 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.
585
586A terrain-following coordinate system (hereafter $s-$coordinates) avoids the discretisation error in the depth field since the layers of computation are gradually adjusted with depth to the ocean bottom. Relatively shallow topographic features 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) as can gentle, large-scale slopes of the sea floor. 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 $s-$coordinates it is possible, at least in principle, to have the bottom and the sea surface as the only boundaries of the domain. Nevertheless, $s-$coordinates also have 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 $s-$coordinates consists of two terms (see \colorbox{yellow}{Appendix A}),
587
588\begin{equation} \label{Eq_PE_p_sco}
589\left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial 
590p}{\partial s}\left. {\nabla z} \right|_s
591\end{equation}
592
593The second term in (\ref{Eq_PE_p_sco}) depends on the tilt of the coordinate surface and
594introduces a truncation error that is not present in a $z-$model. In the special case of $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$), \cite{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude of this truncation error. It depends on topographic slope, stratification, horizontal and vertical resolution, 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 $s-$coordinates and step-like representation of bottom topography \citep{Madec1996}. However, another problem is then raised in the definition of the model domain.
595
596\colorbox{yellow}{Aike Beckmann's solution}
597
598A minimum of diffusion along the coordinate surfaces of any finite difference model is always required for numerical reasons. 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.
599However, 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
600water column (i.e. the main thermocline) \citep{Madec1996}. An alternate solution consists in rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_zdf} and \colorbox{yellow}{Appendix B}.
601
602%AMT
603\amtcomment{ 
604The $s-$coordinates introduced here \citep{Lott1990,Madec1996} 
605\colorbox{yellow}{ to be update $==>$}
606differ mainly in two aspects from
607similar models. It combines the properties which make OPA suitable for
608climate applications with a good representation of bottom topography
609allowing mixed step-like/terrain following topography. It also offers a
610completely general transformation, $s=s(i,j,z)$, for the vertical coordinate which goes beyond those of previous hybrid models except the GFDL version developed by \citep{Gerdes1993a,Gerdes1993b} 
611which has similar properties as the OPA release presented here.
612\colorbox{yellow}{ <== end update}
613}
614% -------------------------------------------------------------------------------------------------------------
615% The s-coordinate Formulation
616% -------------------------------------------------------------------------------------------------------------
617\subsection{The \textit{s-}coordinate Formulation}
618
619Starting 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 $s=s(i,j,k)$, which includes $z-$ and $\sigma-$coordinates as special cases ($s=z$ and $s=\sigma=z/H$, resp.). A formal derivation of the transformed equations is given in \colorbox{yellow}{Appendix A}. Let us define the vertical scale factor by $e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k)$ ), and the slopes in the (\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by :
620\begin{equation} \label{Eq_PE_sco_slope}
621\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s
622\quad \text{, and } \quad 
623\sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right|_s
624\end{equation}
625
626We also introduce a "vertical" velocity $\omega$ defined as the velocity normal to $s-$surfaces:
627\begin{equation} \label{Eq_PE_sco_w}
628\omega = w - \sigma_1 \; u - \sigma_2 \; v
629\end{equation}
630The equations solved by the ocean model \eqref{Eq_PE} in $s-$coordinates can be written as follows:
631
632* momentum equation:
633\begin{multline} \label{Eq_PE_sco_u}
634\frac{\partial u}{\partial t}=+\left( {\zeta +f} \right)\,v-\frac{1}{e_3
635}\omega \frac{\partial u}{\partial k}-\frac{1}{e_1 }\frac{\partial 
636}{\partial i}\left( {\frac{1}{2}\left( {u^2+v^2} \right)+\frac{p_h }{\rho _o}} \right)   \\
637+g\frac{\rho }{\rho _o }\sigma _1 -\frac{1}{\rho _o e_1
638}\frac{\partial p_s }{\partial i}+D_u^{\rm {\bf U}} 
639\end{multline}
640\begin{multline} \label{Eq_PE_sco_v}
641\frac{\partial v}{\partial t}=-\left( {\zeta +f} \right)\,u-\frac{1}{e_3
642}\omega \frac{\partial v}{\partial k}-\frac{1}{e_2 }\frac{\partial 
643}{\partial j}\left( {\frac{1}{2}\left( {u^2+v^2} \right)+\frac{p_h }{\rho _o}} \right)   \\
644+g\frac{\rho }{\rho _o }\sigma _2 -\frac{1}{\rho _o e_2
645}\frac{\partial p_s }{\partial j}+D_v^{\rm {\bf U}} 
646\end{multline}
647where 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 same equation as w, i.e. (\ref{Eq_w_diag}), with $\chi$, the divergence of the horizontal velocity field given by:
648
649\begin{equation} \label{Eq_PE_sco_div}
650\chi =\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3 \,u} 
651\right)}{\partial i}+\frac{\partial \left( {e_1 e_3 \,v} \right)}{\partial 
652j}} \right]
653\end{equation}
654
655* tracer equations:
656
657\begin{equation} \label{Eq_PE_sco_t}
658\frac{\partial T}{\partial t}=-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial 
659\left( {e_2 e_3 T\,u} \right)}{\partial i}+\frac{\partial \left( {e_1 e_3
660T\,v} \right)}{\partial j}} \right]-\frac{1}{e_3 }\frac{\partial \left(
661{T\,\omega } \right)}{\partial k}+D^T
662\end{equation}
663
664\begin{equation} \label{Eq_PE_sco_s}
665\frac{\partial S}{\partial t}=-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial 
666\left( {e_2 e_3 S\,u} \right)}{\partial i}+\frac{\partial \left( {e_1 e_3
667S\,v} \right)}{\partial j}} \right]-\frac{1}{e_3 }\frac{\partial \left(
668{S\,\omega } \right)}{\partial k}+D^S
669\end{equation}
670
671The equation of state has the same expression as in $z-$coordinates. The expression of \textbf{D}$^{U}$, $D^{S}$ and$ D^{T}$ depends on the subgrid scale parameterization used. It will be defined in \S\ref{PE_zdf_ldf}.
672
673\colorbox{yellow}{ to be updated $= = >$}
674The whole set of the continuous equations solved by the model in the $s-$coordinate system is summarised in Table~I.2.
675
676Add a few works on z and zps and s and underlies the differences between all of them
677\colorbox{yellow}{ $< = =$ end update}
678
679\newpage 
680% ================================================================
681% Curvilinear z*- s*-coordinate System
682% ================================================================
683\section{Curvilinear \textit{z*}- or \textit{s*} coordinate System}
684
685% -------------------------------------------------------------------------------------------------------------
686% ????
687% -------------------------------------------------------------------------------------------------------------
688
689\colorbox{yellow}{ to be updated $= = >$}
690
691In that case, the free surface equation is nonlinear, and the variations of
692volume are fully taken into account. These coordinates systems is presented in a report
693\citep{Levier2007} available on the \NEMO web site.
694
695\colorbox{yellow}{ $< = =$ end update}
696
697\newpage 
698% ================================================================
699% Subgrid Scale Physics
700% ================================================================
701\section{Subgrid Scale Physics}
702\label{PE_zdf_ldf}
703
704The primitive equations describe the behaviour of a geophysical fluid at
705space and time scales larger than a few kilometres in the horizontal, a few
706meters in the vertical and a few minutes. They are usually solved at larger
707scales, the specified grid spacing and time step of the numerical model. The
708effects of smaller scale motions (coming from the advective terms in the
709Navier-Stokes equations) must be represented entirely in terms of
710large-scale patterns to close the equations. These effects appear in the
711equations as the divergence of turbulent fluxes (i.e. fluxes associated with
712the mean correlation of small scale perturbations). Assuming a turbulent
713closure hypothesis is equivalent to choose a formulation for these fluxes.
714It is usually called the subgrid scale physics. It must be emphasized that
715this is the weakest part of the primitive equations, but also one of the
716most important for long-term simulations as small scale processes \textit{in fine} balance
717the surface input of kinetic energy and heat.
718
719The control exerted by gravity on the flow induces a strong anisotropy
720between 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.
721
722% -------------------------------------------------------------------------------------------------------------
723% Vertical Subgrid Scale Physics
724% -------------------------------------------------------------------------------------------------------------
725\subsection{Vertical Subgrid Scale Physics}
726\label{PE_zdf}
727
728The model resolution is always larger than the scale at which the major
729sources of vertical turbulence occurs (shear instability, internal wave
730breaking...). Turbulent motions are thus never explicitly solved, even
731partially, but always parameterized. The vertical turbulent fluxes are
732assumed to depend linearly on the gradients of large-scale quantities (for
733example, 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
734analogous to that of molecular diffusion and dissipation. This is quite
735clearly a necessary compromise: considering only the molecular viscosity
736acting on large scale severely underestimates the role of turbulent
737diffusion and dissipation, while an accurate consideration of the details of
738turbulent motions is simply impractical. The resulting vertical momentum and
739tracer diffusive operators are of second order:
740\begin{equation} \label{Eq_PE_zdf}
741   \begin{split}
742{\vect{D}}^{v \vect{U}}
743&=\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) \ ,
744\quad
745D^{vS}=\frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial S}{\partial z}} \right)
746   \end{split}
747\end{equation}
748where $A^{vm}$ 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
749be specified (see Chap.~\ref{SBC}). 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 (as Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a turbulent closure model. The choices available in OPA are discussed in \S\ref{ZDF}).
750
751% -------------------------------------------------------------------------------------------------------------
752% Lateral Diffusive and Viscous Operators Formulation
753% -------------------------------------------------------------------------------------------------------------
754\subsection{Lateral Diffusive and Viscous Operators Formulation}
755\label{PE_ldf}
756
757Lateral turbulence can be roughly divided into a mesoscale turbulence
758associated to eddies which can be solved explicitly if the resolution is
759sufficient as their underlying physics are included in the primitive
760equations, and a sub mesoscale turbulence which is never explicitly solved
761even partially, but always parameterized. The formulation of lateral eddy
762fluxes depends on whether the mesoscale is below or above the grid-spacing
763(i.e. the model is eddy-resolving or not).
764
765In non-eddy- resolving configurations, the closure is similar to that used
766for the vertical physics. The lateral turbulent fluxes are assumed to depend
767linearly on the lateral gradients of large-scale quantities. The resulting
768lateral diffusive and dissipative operators are of second order.
769Observations show that lateral mixing induced by mesoscale turbulence tends
770to be along isoneutral surfaces (or more precisely neutral surfaces, i.e.
771isoneutral surfaces referenced at the local depth) rather than across them.
772As the slope of isoneutral surfaces is small in the ocean, a common
773approximation is to assume that the `lateral' direction is the horizontal,
774i.e. the lateral mixing is performed along geopotential surfaces. This leads
775to a geopotential second order operator for lateral subgrid scale physics.
776This assumption can be relaxed: the eddy-induced turbulent fluxes can be
777better approached by assuming that they depend linearly on the gradients of
778large-scale quantities computed along isoneutral surfaces. In such a case,
779the diffusive operator is an isoneutral second order operator and it has
780components in the three space directions. However, both horizontal and
781isoneutral operators have no effect on mean (i.e. large scale) potential
782energy whereas potential energy is a main source of turbulence (through
783baroclinic instabilities). \citet{Gent1990} have proposed a
784parameterization of mesoscale eddy-induced turbulence which associates an
785eddy-induced velocity to the isoneutral diffusion. Its mean effect is to
786reduce the mean potential energy of the ocean. This leads to a formulation
787of lateral subgrid-scale physics made up of an isoneutral second order
788operator and an eddy induced advective part. In all these lateral diffusive
789formulations, the specification of the lateral eddy coefficients remains the
790problematic point as there is no satisfactory formulation of these
791coefficients as a function of large-scale features.
792
793In eddy-resolving configurations, a second order operator can be used, but
794usually a more scale selective one (biharmonic operator) is preferred as the
795grid-spacing is usually not small enough compared to the scale of the
796eddies. The role devoted to the subgrid-scale physics is to dissipate the
797energy that cascades toward the grid scale and thus ensures the stability of
798the model while not interfering with the solved mesoscale activity.
799
800All these parameterizations of subgrid scale physics present advantages and
801disadvantages. There are not all available in OPA. In the $z-$coordinate
802formulation, four options are offered for active tracers (temperature and
803salinity): second order geopotential operator, second order isoneutral
804operator, \citet{Gent1990} parameterization and fourth order
805geopotential operator. The same options are available for momentum, except
806\citet{Gent1990} parameterization which only involves tracers. In
807$s-$coordinate formulation, an additional option is offered for tracers: second
808order operator acting along $s-$surfaces, and for momentum: fourth order
809operator acting along $s-$surfaces (see \S\ref{LDF}).
810
811\subsubsection{lateral second order tracer diffusive operator}
812
813The lateral second order tracer diffusive operator is defined by (see \colorbox{yellow}{Appendix B}):
814\begin{equation} \label{Eq_PE_iso_tensor}
815D^{lT}=\nabla {\rm {\bf .}}\left( {A^{lT}\;\Re \;\nabla T} \right) \qquad 
816\mbox{with}\quad \;\;\Re =\left( {{\begin{array}{*{20}c}
817 1 \hfill & 0 \hfill & {-r_1 } \hfill \\
818 0 \hfill & 1 \hfill & {-r_2 } \hfill \\
819 {-r_1 } \hfill & {-r_2 } \hfill & {r_1 ^2+r_2 ^2} \hfill \\
820\end{array} }} \right)
821\end{equation}
822where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along
823which the diffusive operator acts and the surface of computation ($z-$ or
824$s-$surfaces), and $r_1$ and $r_2$ is the differential operator defined
825in \S\ref{PE_zco} or \S\ref{PE_sco} depending on the vertical coordinate used. Note that the formulation of $r_1$ and $r_2 $ is exact for the
826slopes between geopotential and $s-$surfaces, while it is only an approximation
827for the slopes between isoneutral and $z$ or $s-$surfaces. Indeed, in the latter
828case, two assumptions are made to simplify $r_1$ and $r_2$ \citep{Cox1987}: the ratio between lateral and vertical diffusive coefficients is
829known to be several orders of magnitude smaller than unity, and the slopes
830are, generally less than $10^{2}$ in the ocean (see \colorbox{yellow}{Appendix B}). This leads to
831the linear tensor (\ref{Eq_PE_iso_tensor}) where the two isoneutral directions of diffusion
832are independent and where the diapycnal diffusivity contribution is solely along the vertical.
833
834For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the
835geopotential and computational surfaces: in $z$-coordinates they are zero ($r_1$ and $r_2)$ while in $s-$coordinate they are equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ).
836
837For \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 $s-$coor-dinates:
838\begin{equation} \label{Eq_PE_iso_slopes}
839r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
840                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1} \ , \quad
841r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
842                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1}
843\end{equation}
844
845When 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:
846\begin{equation} \label{Eq_PE_iso+eiv}
847D^{lT}=\nabla \cdot \left( {A^{lT}\;\Re \;\nabla T} \right)
848           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right)
849\end{equation}
850where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent, eddy-induced transport velocity. This velocity field is defined by:
851\begin{equation} \label{Eq_PE_eiv}
852   \begin{split}
853 u^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_1 } \right] \\ 
854 v^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_2 } \right] \\ 
855 w^\ast &=  -\frac{1}{e_1 e_2 }\left[
856                      \frac{\partial }{\partial i}\left( {A^{eiv}\;e_2\,\tilde{r}_1 } \right)
857                    +\frac{\partial }{\partial j}\left( {A^{eiv}\;e_1\,\tilde{r}_2 } \right)      \right]
858   \end{split}
859\end{equation}
860where $A^{eiv}$ 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:
861\begin{align} \label{Eq_PE_slopes_eiv}
862\tilde{r}_n = \begin{cases}
863   r_n                  &      \text{in $z-$coordinate}    \\
864   r_n + \sigma_n &      \text{in $s-$coordinate} 
865                   \end{cases}
866\quad \text{where } n=1,2
867\end{align}
868
869The normal component of the eddy induced velocity is zero at all the boundaries. this can be achieved by tapering either the eddy coefficient or the slopes to zero in the vicinity of the boundaries.
870
871\subsubsection{lateral fourth order tracer diffusive operator}
872
873The lateral fourth order tracer diffusive operator is defined by:
874\begin{equation} \label{Eq_PE_bilapT}
875D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
876\qquad \text{where} \  D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
877 \end{equation}
878
879It is the second order operator given by (\ref{Eq_PE_iso_tensor}) applied twice with the eddy
880diffusion coefficient correctly placed.
881
882\subsubsection{lateral second order momentum diffusive operator}
883
884The second order momentum diffusive operator along $z-$ or $s-$surfaces is found by
885applying \eqref{Eq_PE_curl} to the horizontal velocity vector (see Appendix B):
886\begin{equation} \label{Eq_PE_lapU}
887\begin{split}
888{\rm {\bf D}}^{l{\rm {\bf U}}} 
889&= \quad \  \nabla _h \left( {A^{lm}\chi } \right)
890   \ - \ \nabla _h \times \left( {A^{lm}\,\zeta \;{\rm {\bf k}}} \right)     \\
891&=   \left(      \begin{aligned}
892             \frac{1}{e_1      } \frac{\partial \left( A^{lm} \chi          \right)}{\partial i} 
893         &-\frac{1}{e_2 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial j}  \\
894             \frac{1}{e_2      }\frac{\partial \left( {A^{lm} \chi         } \right)}{\partial j}   
895         &+\frac{1}{e_1 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial i}
896        \end{aligned}    \right)
897\end{split}
898\end{equation}
899
900Such a formulation ensures a complete separation between the vorticity and
901horizontal divergence fields ({\S}~II.4c). Unfortunately, it is not
902available for geopotential diffusion in $s-$coordinates and for isoneutral
903diffusion. In these two cases, the $u$ and $v-$fields are considered as independent
904scalar fields, so that the diffusive operator is given by:
905\begin{equation} \label{Eq_PE_lapU_iso}
906\begin{split}
907 D_u^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla u} \right) \\ 
908 D_v^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla v} \right)
909 \end{split}
910 \end{equation}
911where $\Re$ is given by (I.5.2). It is the same expression as those used
912for diffusive operator on tracers.
913
914\subsubsection{lateral fourth order momentum diffusive operator}
915
916As 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:
917
918geopotential diffusion in $z-$coordinates:
919\begin{equation} \label{Eq_PE_bilapU}
920\begin{split}
921{\rm {\bf D}}^{l{\rm {\bf U}}} &=\nabla _h \left\{ {\;\nabla _h {\rm {\bf 
922.}}\left[ {A^{lm}\,\nabla _h \left( \chi \right)} \right]\;} 
923\right\}\;   \\
924&+\nabla _h \times \left\{ {\;{\rm {\bf k}}\cdot \nabla \times 
925\left[ {A^{lm}\,\nabla _h \times \left( {\zeta \;{\rm {\bf k}}} \right)} 
926\right]\;} \right\}
927\end{split}
928\end{equation}
929
930geopotential diffusion in $s-$coordinates:
931\begin{equation} \label{Eq_bilapU_iso}
932   \left\{   \begin{aligned}
933         D_u^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta u} \right) \\ 
934         D_v^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta v} \right)
935   \end{aligned}    \right.
936   \quad \text{where} \quad 
937   \Delta \left( \bullet \right) = \nabla \cdot \left( \Re \nabla(\bullet) \right)
938\end{equation}
939
Note: See TracBrowser for help on using the repository browser.