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 branches/DEV_r1826_DOC/DOC/TexFiles/Chapters – NEMO

source: branches/DEV_r1826_DOC/DOC/TexFiles/Chapters/Chap_Model_Basics.tex @ 1831

Last change on this file since 1831 was 1831, checked in by gm, 14 years ago

cover, namelist, rigid-lid, e3t, appendices, see ticket: #658

  • Property svn:executable set to *
File size: 66.4 KB
Line 
1% ================================================================
2% Chapter 1 Ñ Model Basics
3% ================================================================
4
5\chapter{Model basics}
6\label{PE}
7\minitoc
8
9
10\newpage
11$\ $\newline    % force a new ligne
12
13% ================================================================
14% Primitive Equations
15% ================================================================
16\section{Primitive Equations}
17\label{PE_PE}
18
19% -------------------------------------------------------------------------------------------------------------
20%        Vector Invariant Formulation
21% -------------------------------------------------------------------------------------------------------------
22
23\subsection{Vector Invariant Formulation}
24\label{PE_Vector}
25
26
27The ocean is a fluid that can be described to a good approximation by the primitive
28equations, $i.e.$ the Navier-Stokes equations along with a nonlinear equation of
29state which couples the two active tracers (temperature and salinity) to the fluid
30velocity, plus the following additional assumptions made from scale considerations:
31
32\textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to
33be spheres so that gravity (local vertical) is parallel to the earth's radius
34
35\textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius
36
37\textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect
38of small scale processes on the large-scale) are expressed in terms of large-scale features
39
40\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their
41contribution to the buoyancy force
42
43\textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a
44balance between the vertical pressure gradient and the buoyancy force (this removes
45convective processes from the initial Navier-Stokes equations and so convective processes
46must be parameterized instead)
47
48\textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity
49vector is assumed to be zero.
50
51Because the gravitational force is so dominant in the equations of large-scale motions,
52it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked
53to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two
54vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define
55the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$ 
56(the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane),
57$T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density.
58The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k})
59vector system provides the following six equations (namely the momentum balance, the
60hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation
61equations and an equation of state):
62\begin{subequations} \label{Eq_PE}
63  \begin{equation}     \label{Eq_PE_dyn}
64\frac{\partial {\rm {\bf U}}_h }{\partial t}=
65-\left[    {\left( {\nabla \times {\rm {\bf U}}} \right)\times {\rm {\bf U}}
66            +\frac{1}{2}\nabla \left( {{\rm {\bf U}}^2} \right)}    \right]_h
67 -f\;{\rm {\bf k}}\times {\rm {\bf U}}_h
68-\frac{1}{\rho _o }\nabla _h p + {\rm {\bf D}}^{\rm {\bf U}} + {\rm {\bf F}}^{\rm {\bf U}}
69  \end{equation}
70  \begin{equation}     \label{Eq_PE_hydrostatic}
71\frac{\partial p }{\partial z} = - \rho \ g
72  \end{equation}
73  \begin{equation}     \label{Eq_PE_continuity}
74\nabla \cdot {\bf U}=  0
75  \end{equation}
76\begin{equation} \label{Eq_PE_tra_T}
77\frac{\partial T}{\partial t} = - \nabla \cdot  \left( T \ \rm{\bf U} \right) + D^T + F^T
78  \end{equation}
79  \begin{equation}     \label{Eq_PE_tra_S}
80\frac{\partial S}{\partial t} = - \nabla \cdot  \left( S \ \rm{\bf U} \right) + D^S + F^S
81  \end{equation}
82  \begin{equation}     \label{Eq_PE_eos}
83\rho = \rho \left( T,S,p \right)
84  \end{equation}
85\end{subequations}
86where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions,
87$t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by
88the equation of state (\ref{Eq_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,
89$f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's
90angular velocity vector), and $g$ is the gravitational acceleration.
91${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterisations of small-scale
92physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ 
93and $F^S$ surface forcing terms. Their nature and formulation are discussed in
94\S\ref{PE_zdf_ldf} and page \S\ref{PE_boundary_condition}.
95
96.
97
98% -------------------------------------------------------------------------------------------------------------
99% Boundary condition
100% -------------------------------------------------------------------------------------------------------------
101\subsection{Boundary Conditions}
102\label{PE_boundary_condition}
103
104An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea
105or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$ 
106and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height
107of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$,
108chosen as a mean sea surface (Fig.~\ref{Fig_ocean_bc}). Through these two boundaries,
109the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth,
110the continental margins, the sea ice and the atmosphere. However, some of these fluxes are
111so weak that even on climatic time scales of thousands of years they can be neglected.
112In the following, we briefly review the fluxes exchanged at the interfaces between the ocean
113and the other components of the earth system.
114
115%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
116\begin{figure}[!ht] \label{Fig_ocean_bc}  \begin{center}
117\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_I_ocean_bc.pdf}
118\caption{The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ 
119is the depth of the sea floor and $\eta$ the height of the sea surface. Both $H$ and $\eta $ 
120are referenced to $z=0$.}
121\end{center}   \end{figure}
122%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
123
124
125\begin{description}
126\item[Land - ocean interface:] the major flux between continental margins and the ocean is
127a mass exchange of fresh water through river runoff. Such an exchange modifies the sea
128surface salinity especially in the vicinity of major river mouths. It can be neglected for short
129range integrations but has to be taken into account for long term integrations as it influences
130the characteristics of water masses formed (especially at high latitudes). It is required in order
131to close the water cycle of the climate system. It is usually specified as a fresh water flux at
132the air-sea interface in the vicinity of river mouths.
133\item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small,
134except in special areas of little extent. They are usually neglected in the model
135\footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling
136($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world
137ocean (see \ref{TRA_bbc}).}.
138The boundary condition is thus set to no flux of heat and salt across solid boundaries.
139For momentum, the situation is different. There is no flow across solid boundaries,
140$i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words,
141the bottom velocity is parallel to solid boundaries). This kinematic boundary condition
142can be expressed as:
143\begin{equation} \label{Eq_PE_w_bbc}
144w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right)
145\end{equation}
146In addition, the ocean exchanges momentum with the earth through frictional processes.
147Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized
148in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification
149depends on the nature of the physical parameterisation used for ${\rm {\bf D}}^{\rm {\bf U}}$ 
150in \eqref{Eq_PE_dyn}. It is discussed in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9.
151\item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux
152of fresh water PE  (the precipitation minus evaporation budget) leads to:
153\begin{equation} \label{Eq_PE_w_sbc}
154w = \frac{\partial \eta }{\partial t} 
155    + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right)
156    + P-E
157\end{equation}
158The dynamic boundary condition, neglecting the surface tension (which removes capillary
159waves from the system) leads to the continuity of pressure across the interface $z=\eta$.
160The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
161\item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water
162and momentum. The sea surface temperature is constrained to be at the freezing point
163at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the
164ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and
165salt fluxes that cannot be neglected.
166\end{description}
167
168
169\newpage
170$\ $\newline    % force a new ligne
171
172% ================================================================
173% The Horizontal Pressure Gradient
174% ================================================================
175\section{The Horizontal Pressure Gradient }
176\label{PE_hor_pg}
177
178% -------------------------------------------------------------------------------------------------------------
179% Pressure Formulation
180% -------------------------------------------------------------------------------------------------------------
181\subsection{Pressure Formulation}
182\label{PE_p_formulation}
183
184The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a
185reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that:
186$p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\ref{Eq_PE_hydrostatic}),
187assuming that pressure in decibars can be approximated by depth in meters in (\ref{Eq_PE_eos}).
188The hydrostatic pressure is then given by:
189\begin{equation} \label{Eq_PE_pressure}
190p_h \left( {i,j,z,t} \right)
191 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,\varsigma} \right)\;d\varsigma } 
192\end{equation}
193 Two strategies can be considered for the surface pressure term: $(a)$ introduce of a
194 new variable $\eta$, the free-surface elevation, for which a prognostic equation can be
195 established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the
196 pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used,
197 one solution of the free-surface elevation consists of the excitation of external gravity waves.
198 The flow is barotropic and the surface moves up and down with gravity as the restoring force.
199 The phase speed of such waves is high (some hundreds of metres per second) so that
200 the time step would have to be very short if they were present in the model. The latter
201 strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$ 
202 the sea surface is the surface $z=0$. This well known approximation increases the surface
203 wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic
204 Rossby or planetary waves). The rigid-lid hypothesis is an obsolescent feature in modern
205 OGCMs. It has been available until the release 3.1 of  \NEMO, and it has been removed
206 in release 3.2 and followings. Only the free surface formulation is now described in the
207 this document (see the next sub-section).
208
209% -------------------------------------------------------------------------------------------------------------
210% Free Surface Formulation
211% -------------------------------------------------------------------------------------------------------------
212\subsection{Free Surface Formulation}
213\label{PE_free_surface}
214
215In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced
216which describes the shape of the air-sea interface. This variable is solution of a
217prognostic equation which is established by forming the vertical average of the kinematic
218surface condition (\ref{Eq_PE_w_bbc}):
219\begin{equation} \label{Eq_PE_ssh}
220\frac{\partial \eta }{\partial t}=-D+P-E
221   \quad \text{where} \
222D=\nabla \cdot \left[ {\left( {H+\eta } \right) \; {\rm{\bf \overline{U}}}_h \,} \right]
223\end{equation}
224and using (\ref{Eq_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
225
226Allowing the air-sea interface to move introduces the external gravity waves (EGWs)
227as a class of solution of the primitive equations. These waves are barotropic because
228of hydrostatic assumption, and their phase speed is quite high. Their time scale is
229short with respect to the other processes described by the primitive equations.
230
231Two choices can be made regarding the implementation of the free surface in the model,
232depending on the physical processes of interest.
233
234$\bullet$ If one is interested in EGWs, in particular the tides and their interaction
235with the baroclinic structure of the ocean (internal waves) possibly in shallow seas,
236then a non linear free surface is the most appropriate. This means that no
237approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the ocean
238volume is fully taken into account. Note that in order to study the fast time scales
239associated with EGWs it is necessary to minimize time filtering effects (use an
240explicit time scheme with very small time step, or a split-explicit scheme with
241reasonably small time step, see \S\ref{DYN_spg_exp} or \S\ref{DYN_spg_ts}.
242
243$\bullet$ If one is not interested in EGW but rather sees them as high frequency
244noise, it is possible to apply an explicit filter to slow down the fastest waves while
245not altering the slow barotropic Rossby waves. If further, an approximative conservation
246of heat and salt contents is sufficient for the problem solved, then it is
247sufficient to solve a linearized version of (\ref{Eq_PE_ssh}), which still allows
248to take into account freshwater fluxes applied at the ocean surface \citep{Roullet_Madec_JGR00}.
249
250The filtering of EGWs in models with a free surface is usually a matter of discretisation
251of the temporal derivatives, using the time splitting method \citep{Killworth_al_JPO91, Zhang_Endoh_JGR92} 
252or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach
253developed by \citet{Roullet_Madec_JGR00}: the damping of EGWs is ensured by introducing an
254additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:
255\begin{equation} \label{Eq_PE_flt}
256\frac{\partial {\rm {\bf U}}_h }{\partial t}= {\rm {\bf M}}
257- g \nabla \left( \tilde{\rho} \ \eta \right)
258- g \ T_c \nabla \left( \widetilde{\rho} \ \partial_t \eta \right)
259\end{equation}
260where $T_c$, is a parameter with dimensions of time which characterizes the force,
261$\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and $\rm {\bf M}$ 
262represents the collected contributions of the Coriolis, hydrostatic pressure gradient,
263non-linear and viscous terms in \eqref{Eq_PE_dyn}.
264
265The new force can be interpreted as a diffusion of vertically integrated volume flux divergence.
266The time evolution of $D$ is thus governed by a balance of two terms, $-g$ \textbf{A} $\eta$ 
267and $g \, T_c \,$ \textbf{A} $D$, associated with a propagative regime and a diffusive regime
268in the temporal spectrum, respectively. In the diffusive regime, the EGWs no longer propagate,
269$i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than
270$T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs
271can be damped by choosing $T_c > \Delta t$. \citet{Roullet_Madec_JGR00} demonstrate that
272(\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which
273has to be computed implicitly. This is not surprising since the use of a large time step has a
274necessarily numerical cost. Two gains arise in comparison with the previous formulations.
275Firstly, the damping of EGWs can be quantified through the magnitude of the additional term.
276Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as
277soon as $T_c > \Delta t$.
278
279When the variations of free surface elevation are small compared to the thickness of the first
280model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized
281by \citet{Roullet_Madec_JGR00} the linearization of (\ref{Eq_PE_ssh}) has consequences on the
282conservation of salt in the model. With the nonlinear free surface equation, the time evolution
283of the total salt content is
284\begin{equation} \label{Eq_PE_salt_content}
285    \frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
286                        =\int\limits_S {S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds}
287\end{equation}
288where $S$ is the salinity, and the total salt is integrated over the whole ocean volume
289$D_\eta$ bounded by the time-dependent free surface. The right hand side (which is an
290integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh})
291is satisfied, so that the salt is perfectly conserved. When the free surface equation is
292linearized, \citet{Roullet_Madec_JGR00} show that the total salt content integrated in the fixed
293volume $D$ (bounded by the surface $z=0$) is no longer conserved:
294\begin{equation} \label{Eq_PE_salt_content_linear}
295         \frac{\partial }{\partial t}\int\limits_D {S\;dv} 
296               = - \int\limits_S {S\;\frac{\partial \eta }{\partial t}ds} 
297\end{equation}
298
299The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions
300\citep{Roullet_Madec_JGR00}. It can be significant when the freshwater forcing is not balanced and
301the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $} 
302results in a decrease of the salinity in the fixed volume $D$. Even in that case though,
303the total salt integrated in the variable volume $D_{\eta}$ varies much less, since
304(\ref{Eq_PE_salt_content_linear}) can be rewritten as
305\begin{equation} \label{Eq_PE_salt_content_corrected}
306\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
307=\frac{\partial}{\partial t} \left[ \;{\int\limits_D {S\;dv} +\int\limits_S {S\eta \;ds} } \right]
308=\int\limits_S {\eta \;\frac{\partial S}{\partial t}ds}
309\end{equation}
310
311Although the total salt content is not exactly conserved with the linearized free surface,
312its variations are driven by correlations of the time variation of surface salinity with the
313sea surface height, which is a negligible term. This situation contrasts with the case of
314the rigid lid approximation in which case freshwater forcing is represented by a virtual
315salt flux, leading to a spurious source of salt at the ocean surface
316\citep{Huang_JPO93, Roullet_Madec_JGR00}.
317
318
319% ================================================================
320% Curvilinear z-coordinate System
321% ================================================================
322\section{Curvilinear \textit{z-}coordinate System}
323\label{PE_zco}
324
325
326% -------------------------------------------------------------------------------------------------------------
327% Tensorial Formalism
328% -------------------------------------------------------------------------------------------------------------
329\subsection{Tensorial Formalism}
330\label{PE_tensorial}
331
332In many ocean circulation problems, the flow field has regions of enhanced dynamics
333($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts).
334The representation of such dynamical processes can be improved by specifically increasing
335the model resolution in these regions. As well, it may be convenient to use a lateral
336boundary-following coordinate system to better represent coastal dynamics. Moreover,
337the common geographical coordinate system has a singular point at the North Pole that
338cannot be easily treated in a global model without filtering. A solution consists of introducing
339an appropriate coordinate transformation that shifts the singular point onto land
340\citep{Madec_Imbard_CD96, Murray_JCP96}. As a consequence, it is important to solve the primitive
341equations in various curvilinear coordinate systems. An efficient way of introducing an
342appropriate coordinate transform can be found when using a tensorial formalism.
343This formalism is suited to any multidimensional curvilinear coordinate system. Ocean
344modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth
345approximation), with preservation of the local vertical. Here we give the simplified equations
346for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey
347of the conservation laws of fluid dynamics.
348
349Let (\textit{i},\textit{j},\textit{k}) be a set of orthogonal curvilinear coordinates on the sphere
350associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k})
351linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are
352two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (Fig.\ref{Fig_referential}).
353Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined
354by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of
355the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea
356level (Fig.\ref{Fig_referential}). The local deformation of the curvilinear coordinate system is
357given by $e_1$, $e_2$ and $e_3$, the three scale factors:
358\begin{equation} \label{Eq_scale_factors}
359\begin{aligned}
360 e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
361}{\partial i}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
362}{\partial i}} \right)^2} \right]^{1/2} \\ 
363 e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
364}{\partial j}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
365}{\partial j}} \right)^2} \right]^{1/2} \\ 
366 e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\ 
367 \end{aligned}
368 \end{equation}
369
370%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
371\begin{figure}[!tb] \label{Fig_referential}  \begin{center}
372\includegraphics[width=0.60\textwidth]{./TexFiles/Figures/Fig_I_earth_referential.pdf}
373\caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear
374coordinate system (\textbf{i},\textbf{j},\textbf{k}). }
375\end{center}   \end{figure}
376%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
377
378Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by
379$a$ in (\ref{Eq_scale_factors}) (thin-shell approximation). The resulting horizontal scale
380factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single
381function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that
382appear in the primitive equations (Eqs. \eqref{Eq_PE_dyn} to \eqref{Eq_PE_eos}) can
383be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate
384system transformation:
385\begin{subequations} \label{Eq_PE_discrete_operators}
386\begin{equation} \label{Eq_PE_grad}
387\nabla q=\frac{1}{e_1 }\frac{\partial q}{\partial i}\;{\rm {\bf 
388i}}+\frac{1}{e_2 }\frac{\partial q}{\partial j}\;{\rm {\bf j}}+\frac{1}{e_3
389}\frac{\partial q}{\partial k}\;{\rm {\bf k}}    \\
390\end{equation}
391\begin{equation} \label{Eq_PE_div}
392\nabla \cdot {\rm {\bf A}} 
393= \frac{1}{e_1 \; e_2} \left[
394  \frac{\partial \left(e_2 \; a_1\right)}{\partial i }
395+\frac{\partial \left(e_1 \; a_2\right)}{\partial j }       \right]
396+ \frac{1}{e_3} \left[ \frac{\partial a_3}{\partial k }   \right]
397\end{equation}
398\begin{equation} \label{Eq_PE_curl}
399   \begin{split}
400\nabla \times \vect{A} =
401    \left[ {\frac{1}{e_2 }\frac{\partial a_3}{\partial j}
402            -\frac{1}{e_3 }\frac{\partial a_2 }{\partial k}} \right] \; \vect{i}
403&+\left[ {\frac{1}{e_3 }\frac{\partial a_1 }{\partial k}
404           -\frac{1}{e_1 }\frac{\partial a_3 }{\partial i}} \right] \; \vect{j}     \\
405&+\frac{1}{e_1 e_2 } \left[ {\frac{\partial \left( {e_2 a_2 } \right)}{\partial i}
406                                       -\frac{\partial \left( {e_1 a_1 } \right)}{\partial j}} \right] \; \vect{k} 
407   \end{split}
408\end{equation}
409\begin{equation} \label{Eq_PE_lap}
410\Delta q = \nabla \cdot \left\nabla q \right)
411\end{equation}
412\begin{equation} \label{Eq_PE_lap_vector}
413\Delta {\rm {\bf A}} =
414  \nabla \left( \nabla \cdot {\rm {\bf A}} \right)
415- \nabla \times \left\nabla \times {\rm {\bf A}} \right)
416\end{equation}
417\end{subequations}
418where $q$ is a scalar quantity and ${\rm {\bf A}}=(a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinate system.
419
420% -------------------------------------------------------------------------------------------------------------
421% Continuous Model Equations
422% -------------------------------------------------------------------------------------------------------------
423\subsection{Continuous Model Equations}
424\label{PE_zco_Eq}
425
426In order to express the Primitive Equations in tensorial formalism, it is necessary to compute
427the horizontal component of the non-linear and viscous terms of the equation using
428\eqref{Eq_PE_grad}) to \eqref{Eq_PE_lap_vector}.
429Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate
430system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity
431field $\chi$, by:
432\begin{equation} \label{Eq_PE_curl_Uh}
433\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v} 
434\right)}{\partial i}-\frac{\partial \left( {e_1 \,u} \right)}{\partial j}} 
435\right]
436\end{equation}
437\begin{equation} \label{Eq_PE_div_Uh}
438\chi =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,u} 
439\right)}{\partial i}+\frac{\partial \left( {e_1 \,v} \right)}{\partial j}} 
440\right]
441\end{equation}
442
443Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ 
444and that $e_3$  is a function of the single variable $k$, the nonlinear term of
445\eqref{Eq_PE_dyn} can be transformed as follows:
446\begin{flalign*}
447&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
448+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h        &
449\end{flalign*}
450\begin{flalign*}
451&\qquad=\left( {{\begin{array}{*{20}c}
452 {\left[    {   \frac{1}{e_3} \frac{\partial u  }{\partial k}
453         -\frac{1}{e_1} \frac{\partial w  }{\partial i} } \right] w - \zeta \; v }     \\
454      {\zeta \; u - \left[ {   \frac{1}{e_2} \frac{\partial w}{\partial j}
455                     -\frac{1}{e_3} \frac{\partial v}{\partial k} } \right] \ w}  \\
456       \end{array} }} \right)       
457+\frac{1}{2}   \left( {{\begin{array}{*{20}c}
458       { \frac{1}{e_1}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial i}}  \hfill    \\
459       { \frac{1}{e_2}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial j}}  \hfill    \\
460       \end{array} }} \right)       &
461\end{flalign*}
462\begin{flalign*}
463& \qquad =\left( {{  \begin{array}{*{20}c}
464 {-\zeta \; v} \hfill \\
465 { \zeta \; u} \hfill \\
466         \end{array} }} \right)
467+\frac{1}{2}\left( {{   \begin{array}{*{20}c}
468 {\frac{1}{e_1 }\frac{\partial \left( {u^2+v^2} \right)}{\partial i}} \hfill  \\
469 {\frac{1}{e_2 }\frac{\partial \left( {u^2+v^2} \right)}{\partial j}} \hfill  \\
470                  \end{array} }} \right)       
471+\frac{1}{e_3 }\left( {{      \begin{array}{*{20}c}
472 { w \; \frac{\partial u}{\partial k}}    \\
473 { w \; \frac{\partial v}{\partial k}}    \\
474                     \end{array} }} \right
475-\left( {{  \begin{array}{*{20}c}
476 {\frac{w}{e_1}\frac{\partial w}{\partial i}
477 -\frac{1}{2e_1}\frac{\partial w^2}{\partial i}} \hfill \\
478 {\frac{w}{e_2}\frac{\partial w}{\partial j}
479  -\frac{1}{2e_2}\frac{\partial w^2}{\partial j}} \hfill \\
480         \end{array} }} \right)        &
481\end{flalign*}
482
483The last term of the right hand side is obviously zero, and thus the nonlinear term of
484\eqref{Eq_PE_dyn} is written in the $(i,j,k)$ coordinate system:
485\begin{equation} \label{Eq_PE_vector_form}
486\left[ {\left( {  \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
487+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h
488=\zeta 
489\;{\rm {\bf k}}\times {\rm {\bf U}}_h +\frac{1}{2}\nabla _h \left( {{\rm 
490{\bf U}}_h^2 } \right)+\frac{1}{e_3 }w\frac{\partial {\rm {\bf U}}_h
491}{\partial k}     
492\end{equation}
493
494This is the so-called \textit{vector invariant form} of the momentum advection term.
495For some purposes, it can be advantageous to write this term in the so-called flux form,
496$i.e.$ to write it as the divergence of fluxes. For example, the first component of
497\eqref{Eq_PE_vector_form} (the $i$-component) is transformed as follows:
498\begin{flalign*}
499&{ \begin{array}{*{20}l}
500\left[ {\left( {\nabla \times \vect{U}} \right)\times \vect{U}
501          +\frac{1}{2}\nabla \left( {\vect{U}}^2 \right)} \right]_i   % \\
502%\\
503     = - \zeta \;v
504     + \frac{1}{2\;e_1 } \frac{\partial \left( {u^2+v^2} \right)}{\partial i}
505     + \frac{1}{e_3}w \ \frac{\partial u}{\partial k}          \\
506\\
507\qquad =\frac{1}{e_1 \; e_2} \left(    -v\frac{\partial \left( {e_2 \,v} \right)}{\partial i}
508                     +v\frac{\partial \left( {e_1 \,u} \right)}{\partial j}    \right)
509+\frac{1}{e_1 e_2 }\left(  +e_2 \; u\frac{\partial u}{\partial i}
510                     +e_2 \; v\frac{\partial v}{\partial i}              \right)
511+\frac{1}{e_3}       \left(   w\;\frac{\partial u}{\partial k}       \right)   \\
512\end{array} }        &
513\end{flalign*}
514\begin{flalign*}
515&{ \begin{array}{*{20}l}
516\qquad =\frac{1}{e_1 \; e_2}  \left\{ 
517 -\left(        v^\frac{\partial e_2                                }{\partial i} 
518      +e_2 \,v    \frac{\partial v                                   }{\partial i}     \right)
519+\left(           \frac{\partial \left( {e_1 \,u\,v}  \right)}{\partial j}
520      -e_1 \,u    \frac{\partial v                                   }{\partial j}  \right\right.
521\\  \left\qquad \qquad \quad
522+\left(           \frac{\partial \left( {e_2 u\,u}     \right)}{\partial i}
523      -u       \frac{\partial \left( {e_2 u}         \right)}{\partial i}  \right)
524+e_2 v            \frac{\partial v                                    }{\partial i}
525                  \right\} 
526+\frac{1}{e_3} \left(
527               \frac{\partial \left( {w\,u} \right)         }{\partial k}
528       -u         \frac{\partial w                    }{\partial k}  \right) \\
529\end{array} }     &
530\end{flalign*}
531\begin{flalign*}
532&{ \begin{array}{*{20}l}
533\qquad =\frac{1}{e_1 \; e_2}  \left(
534               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
535      +        \frac{\partial \left( {e_1 \,u\,v} \right)}{\partial j}  \right)
536+\frac{1}{e_3 }      \frac{\partial \left( {w\,u       } \right)}{\partial k}
537\\  \qquad \qquad \quad
538+\frac{1}{e_1 e_2 }     \left(
539      -u \left(   \frac{\partial \left( {e_1 v   } \right)}{\partial j}
540               -v\,\frac{\partial e_1 }{\partial j}             \right)
541      -u       \frac{\partial \left( {e_2 u   } \right)}{\partial i}
542                  \right)
543 -\frac{1}{e_3 }     \frac{\partial w}{\partial k} u
544 +\frac{1}{e_1 e_2 }\left(    -v^2\frac{\partial e_2   }{\partial i}     \right)
545\end{array} }     &
546\end{flalign*}
547\begin{flalign*}
548&{ \begin{array}{*{20}l}
549\qquad = \nabla \cdot \left( {{\rm {\bf U}}\,u} \right)
550-   \left( \nabla \cdot {\rm {\bf U}} \right) \ u
551+\frac{1}{e_1 e_2 }\left(
552      -v^2     \frac{\partial e_2 }{\partial i}
553      +uv   \,    \frac{\partial e_1 }{\partial j}    \right) \\
554\end{array} }     &
555\end{flalign*}
556as $\nabla \cdot {\rm {\bf U}}\;=0$ (incompressibility) it comes:
557\begin{flalign*}
558&{ \begin{array}{*{20}l}
559\qquad = \nabla \cdot \left{{\rm {\bf U}}\,u}      \right)
560\frac{1}{e_1 e_2 }   \left( v \; \frac{\partial e_2}{\partial i}
561                         -u \; \frac{\partial e_1}{\partial j}    \right\left( -v \right)
562\end{array} }     &
563\end{flalign*}
564
565The flux form of the momentum advection term is therefore given by:
566\begin{multline} \label{Eq_PE_flux_form}
567      \left[
568  \left(    {\nabla \times {\rm {\bf U}}}    \right) \times {\rm {\bf U}}
569+\frac{1}{2}   \nabla \left{{\rm {\bf U}}^2}    \right)
570      \right]_h
571\\
572= \nabla \cdot    \left( {{\begin{array}{*{20}c}   {\rm {\bf U}} \, u   \hfill \\
573                                    {\rm {\bf U}} \, v   \hfill \\
574                  \end{array} }}   
575            \right)
576+\frac{1}{e_1 e_2 }     \left(
577       v\frac{\partial e_2}{\partial i}
578      -u\frac{\partial e_1}{\partial j} 
579                  \right) {\rm {\bf k}} \times {\rm {\bf U}}_h
580\end{multline}
581
582The flux form has two terms, the first one is expressed as the divergence of momentum
583fluxes (hence the flux form name given to this formulation) and the second one is due to
584the curvilinear nature of the coordinate system used. The latter is called the \emph{metric} 
585term and can be viewed as a modification of the Coriolis parameter:
586\begin{equation} \label{Eq_PE_cor+metric}
587f \to f + \frac{1}{e_1\;e_2}  \left(  v \frac{\partial e_2}{\partial i}
588                        -u \frac{\partial e_1}{\partial j}  \right)
589\end{equation}
590
591Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$ 
592and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of
593the Coriolis parameter $f \to f+(u/a) \tan \varphi$.
594
595
596$\ $\newline    % force a new ligne
597
598To sum up, the equations solved by the ocean model can be written in the following tensorial formalism:
599
600\vspace{+10pt}
601$\bullet$ \textbf{Vector invariant form of the momentum equations} :
602
603\begin{subequations} \label{Eq_PE_dyn_vect}
604\begin{equation} \label{Eq_PE_dyn_vect_u} \begin{split}
605\frac{\partial u}{\partial t} 
606= +   \left( {\zeta +f} \right)\,v                                   
607   -   \frac{1}{2\,e_1}           \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
608   -   \frac{1}{e_3    }  w     \frac{\partial u}{\partial k}      &      \\
609   -   \frac{1}{e_1    }            \frac{\partial}{\partial i} \left( \frac{p_s+p_h }{\rho _o}    \right)   
610   &+   D_u^{\vect{U}}  +   F_u^{\vect{U}}      \\
611\\
612\frac{\partial v}{\partial t} =
613       -   \left( {\zeta +f} \right)\,u   
614       -   \frac{1}{2\,e_2 }        \frac{\partial }{\partial j}\left(  u^2+v^\right)   
615       -   \frac{1}{e_3     }   w  \frac{\partial v}{\partial k}     &      \\
616       -   \frac{1}{e_2     }        \frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)   
617    &+  D_v^{\vect{U}}  +   F_v^{\vect{U}}
618\end{split} \end{equation}
619\end{subequations}
620
621
622\vspace{+10pt}
623$\bullet$ \textbf{flux form of the momentum equations} :
624\begin{subequations} \label{Eq_PE_dyn_flux}
625\begin{multline} \label{Eq_PE_dyn_flux_u}
626\frac{\partial u}{\partial t}=
627+   \left( { f + \frac{1}{e_1 \; e_2}
628               \left(    v \frac{\partial e_2}{\partial i}
629                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, v    \\
630- \frac{1}{e_1 \; e_2}  \left(
631               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
632      +        \frac{\partial \left( {e_1 \,v\,u} \right)}{\partial j}  \right)
633                 - \frac{1}{e_3 }\frac{\partial \left( {         w\,u} \right)}{\partial k}    \\
634-   \frac{1}{e_1 }\frac{\partial}{\partial i}\left( \frac{p_s+p_h }{\rho _o}   \right)
635+   D_u^{\vect{U}} +   F_u^{\vect{U}}
636\end{multline}
637\begin{multline} \label{Eq_PE_dyn_flux_v}
638\frac{\partial v}{\partial t}=
639-   \left( { f + \frac{1}{e_1 \; e_2}
640               \left(    v \frac{\partial e_2}{\partial i}
641                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, u   \\
642 \frac{1}{e_1 \; e_2}   \left(
643               \frac{\partial \left( {e_2 \,u\,v} \right)}{\partial i}
644      +        \frac{\partial \left( {e_1 \,v\,v} \right)}{\partial j}  \right)
645                 - \frac{1}{e_3 } \frac{\partial \left( {        w\,v} \right)}{\partial k}    \\
646-   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}    \right)
647+  D_v^{\vect{U}} +  F_v^{\vect{U}} 
648\end{multline}
649\end{subequations}
650where $\zeta$, the relative vorticity, is given by \eqref{Eq_PE_curl_Uh} and $p_s $,
651the surface pressure, is given by:
652\begin{equation} \label{Eq_PE_spg}
653p_s = \left\{ \begin{split} 
654\rho \,g \,\eta &                                 \qquad  \qquad  \;   \qquad \text{ standard free surface} \\ 
655\rho \,g \,\eta &+ \rho_o \,\mu \,\frac{\partial \eta }{\partial t}      \qquad \text{ filtered     free surface}   
656\end{split} 
657\right.
658\end{equation}
659with $\eta$ is solution of \eqref{Eq_PE_ssh}
660
661The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
662\begin{equation} \label{Eq_w_diag}
663\frac{\partial w}{\partial k}=-\chi \;e_3
664\end{equation}
665\begin{equation} \label{Eq_hp_diag}
666\frac{\partial p_h }{\partial k}=-\rho \;g\;e_3
667\end{equation}
668where the divergence of the horizontal velocity, $\chi$ is given by \eqref{Eq_PE_div_Uh}.
669
670\vspace{+10pt}
671$\bullet$ \textit{tracer equations} :
672\begin{equation} \label{Eq_S}
673\frac{\partial T}{\partial t} =
674-\frac{1}{e_1 e_2 }\left[ {      \frac{\partial \left( {e_2 T\,u} \right)}{\partial i}
675                  +\frac{\partial \left( {e_1 T\,v} \right)}{\partial j}} \right]
676-\frac{1}{e_3 }\frac{\partial \left( {T\,w} \right)}{\partial k} + D^T + F^T
677\end{equation}
678\begin{equation} \label{Eq_T}
679\frac{\partial S}{\partial t} =
680-\frac{1}{e_1 e_2 }\left[    {\frac{\partial \left( {e_2 S\,u} \right)}{\partial i}
681                  +\frac{\partial \left( {e_1 S\,v} \right)}{\partial j}} \right]
682-\frac{1}{e_3 }\frac{\partial \left( {S\,w} \right)}{\partial k} + D^S + F^S
683\end{equation}
684\begin{equation} \label{Eq_rho}
685\rho =\rho \left( {T,S,z(k)} \right)
686\end{equation}
687
688The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale
689parameterisation used. It will be defined in \S\ref{PE_zdf}. The nature and formulation of
690${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed
691in Chapter~\ref{SBC}.
692
693
694\newpage 
695% ================================================================
696% Curvilinear s-coordinate System
697% ================================================================
698\section{Curvilinear \textit{s}-coordinate System}
699\label{PE_sco}
700
701
702
703
704
705% -------------------------------------------------------------------------------------------------------------
706% Curvilinear z*-coordinate System
707% -------------------------------------------------------------------------------------------------------------
708\subsection{Curvilinear \textit{z*}-coordinate System}
709\label{PE_zco_star}
710
711%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
712\begin{figure}[!b] \label{Fig_z_zstar}  \begin{center}
713\includegraphics[width=1.0\textwidth]{./TexFiles/Figures/Fig_z_zstar.pdf}
714\caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear
715free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate
716\citep{Adcroft_Campin_OM04} ).}
717\end{center}   \end{figure}
718%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
719
720
721In that case, the free surface equation is nonlinear, and the variations of volume are fully
722taken into account. These coordinates systems is presented in a report \citep{Levier2007} 
723available on the \NEMO web site.
724
725\gmcomment{
726The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation
727which allows one to deal with large amplitude free-surface
728variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In
729the  \textit{z*} formulation, the variation of the column thickness due to sea-surface
730undulations is not concentrated in the surface level, as in the $z$-coordinate formulation,
731but is equally distributed over the full water column. Thus vertical
732levels naturally follow sea-surface variations, with a linear attenuation with
733depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in
734fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent.
735The definition and modified oceanic equations for the rescaled vertical coordinate
736 \textit{z*}, including the treatment of fresh-water flux at the surface, are
737detailed in Adcroft and Campin (2004). The major points are summarized
738here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as:
739
740$H +  \textit{z*} = (H + z) / r$ and  $\delta \textit{z*} = \delta z / r$ with $r = \frac{H+\eta} {H}$
741
742Since the vertical displacement of the free surface is incorporated in the vertical
743coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position, 
744$\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field
745is no longer zero as shown by the continuity equation:
746
747$\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right)
748      \left( r \; w\textit{*} \right) = 0 $
749
750}
751
752
753\newpage 
754% -------------------------------------------------------------------------------------------------------------
755% Terrain following  coordinate System
756% -------------------------------------------------------------------------------------------------------------
757\subsection{Terrain following \textit{s-}coordinate}
758\label{PE_sco}
759
760% -------------------------------------------------------------------------------------------------------------
761% Introduction
762% -------------------------------------------------------------------------------------------------------------
763\subsubsection{Introduction}
764
765Several important aspects of the ocean circulation are influenced by bottom topography.
766Of course, the most important is that bottom topography determines deep ocean sub-basins,
767barriers, sills and channels that strongly constrain the path of water masses, but more subtle
768effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary
769one along continental slopes. Topographic Rossby waves can be excited and can interact
770with the mean current. In the $z-$coordinate system presented in the previous section
771(\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is
772discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom
773and to large localized depth gradients associated with large localized vertical velocities.
774The response to such a velocity field often leads to numerical dispersion effects.
775One solution to strongly reduce this error is to use a partial step representation of bottom
776topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.
777Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)
778
779The $s$-coordinate avoids the discretisation error in the depth field since the layers of
780computation are gradually adjusted with depth to the ocean bottom. Relatively small
781topographic features as well as  gentle, large-scale slopes of the sea floor in the deep
782ocean, which would be ignored in typical $z$-model applications with the largest grid
783spacing at greatest depths, can easily be represented (with relatively low vertical resolution).
784A terrain-following model (hereafter $s-$model) also facilitates the modelling of the
785boundary layer flows over a large depth range, which in the framework of the $z$-model
786would require high vertical resolution over the whole depth range. Moreover, with a
787$s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface
788as the only boundaries of the domain (nomore lateral boundary condition to specify).
789Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a
790homogeneous ocean, it has strong limitations as soon as stratification is introduced.
791The main two problems come from the truncation error in the horizontal pressure
792gradient and a possibly increased diapycnal diffusion. The horizontal pressure force
793in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}),
794
795\begin{equation} \label{Eq_PE_p_sco}
796\left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial 
797p}{\partial s}\left. {\nabla z} \right|_s
798\end{equation}
799
800The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface
801and introduces a truncation error that is not present in a $z$-model. In the special case
802of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),
803\citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude
804of this truncation error. It depends on topographic slope, stratification, horizontal and
805vertical resolution, the equation of state, and the finite difference scheme. This error
806limits the possible topographic slopes that a model can handle at a given horizontal
807and vertical resolution. This is a severe restriction for large-scale applications using
808realistic bottom topography. The large-scale slopes require high horizontal resolution,
809and the computational cost becomes prohibitive. This problem can be at least partially
810overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec_al_JPO96}. However, the definition of the model
811domain vertical coordinate becomes then a non-trivial thing for a realistic bottom
812topography: a envelope topography is defined in $s$-coordinate on which a full or
813partial step bottom topography is then applied in order to adjust the model depth to
814the observed one (see \S\ref{DOM_zgr}.
815
816For numerical reasons a minimum of diffusion is required along the coordinate surfaces
817of any finite difference model. It causes spurious diapycnal mixing when coordinate
818surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as
819well as for a $s$-model. However, density varies more strongly on $s-$surfaces than
820on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal
821diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a
822$z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal
823circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.
824For example, imagine an isolated bump of topography in an ocean at rest with a horizontally
825uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral
826surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,
827the 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
828($i.e.$ the main thermocline) \citep{Madec_al_JPO96}. An alternate solution consists of rotating
829the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}.
830Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,
831strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).
832
833The $s-$coordinates introduced here \citep{Lott_al_OM90,Madec_al_JPO96} differ mainly in two
834aspects from similar models:  it allows  a representation of bottom topography with mixed
835full or partial step-like/terrain following topography ; It also offers a completely general
836transformation, $s=s(i,j,z)$ for the vertical coordinate.
837
838% -------------------------------------------------------------------------------------------------------------
839% The s-coordinate Formulation
840% -------------------------------------------------------------------------------------------------------------
841\subsubsection{The \textit{s-}coordinate Formulation}
842
843Starting from the set of equations established in \S\ref{PE_zco} for the special case $k=z$ 
844and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes
845$z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and
846$s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed
847equations is given in Appendix~\ref{Apdx_A}. Let us define the vertical scale factor by
848$e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the
849(\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by :
850\begin{equation} \label{Eq_PE_sco_slope}
851\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s
852\quad \text{, and } \quad 
853\sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right|_s
854\end{equation}
855We also introduce  $\omega $, a dia-surface velocity component, defined as the velocity
856relative to the moving $s$-surfaces and normal to them:
857\begin{equation} \label{Eq_PE_sco_w}
858\omega  = w - e_3 \, \frac{\partial s}{\partial t} - \sigma _1 \,u - \sigma _2 \,v    \\
859\end{equation}
860
861The equations solved by the ocean model \eqref{Eq_PE} in $s-$coordinate can be written as follows:
862
863 \vspace{0.5cm}
864* momentum equation:
865\begin{multline} \label{Eq_PE_sco_u}
866\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
867   +   \left( {\zeta +f} \right)\,v                                   
868   -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
869   -   \frac{1}{e_3} \omega \frac{\partial u}{\partial k}       \\
870   -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s + p_h}{\rho _o}    \right)   
871   +  g\frac{\rho }{\rho _o}\sigma _1
872   +   D_u^{\vect{U}}  +   F_u^{\vect{U}} \quad
873\end{multline}
874\begin{multline} \label{Eq_PE_sco_v}
875\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
876   -   \left( {\zeta +f} \right)\,u   
877   -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^\right)       
878   -   \frac{1}{e_3 } \omega \frac{\partial v}{\partial k}         \\
879   -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)
880    +  g\frac{\rho }{\rho _o }\sigma _2   
881   +  D_v^{\vect{U}}  +   F_v^{\vect{U}} \quad
882\end{multline}
883where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic
884pressure have the same expressions as in $z$-coordinates although they do not represent
885exactly the same quantities. $\omega$ is provided by the continuity equation
886(see Appendix~\ref{Apdx_A}):
887
888\begin{equation} \label{Eq_PE_sco_continuity}
889\frac{\partial e_3}{\partial t} + e_3 \; \chi + \frac{\partial \omega }{\partial s} = 0   
890\qquad \text{with }\;\; 
891\chi =\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3 \,u} 
892\right)}{\partial i}+\frac{\partial \left( {e_1 e_3 \,v} \right)}{\partial 
893j}} \right]
894\end{equation}
895
896 \vspace{0.5cm}
897* tracer equations:
898\begin{multline} \label{Eq_PE_sco_t}
899\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
900-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,T} \right)}{\partial i}
901                                           +\frac{\partial \left( {e_1 e_3\,v\,T} \right)}{\partial j}} \right]   \\
902-\frac{1}{e_3 }\frac{\partial \left( {T\,\omega } \right)}{\partial k}   + D^T + F^S   \qquad
903\end{multline}
904
905\begin{multline} \label{Eq_PE_sco_s}
906\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
907-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,S} \right)}{\partial i}
908                                           +\frac{\partial \left( {e_1 e_3\,v\,S} \right)}{\partial j}} \right]    \\
909-\frac{1}{e_3 }\frac{\partial \left( {S\,\omega } \right)}{\partial k}     + D^S + F^S   \qquad
910\end{multline}
911
912The equation of state has the same expression as in $z$-coordinate, and similar expressions
913are used for mixing and forcing terms.
914
915\gmcomment{
916\colorbox{yellow}{ to be updated $= = >$}
917Add a few works on z and zps and s and underlies the differences between all of them
918\colorbox{yellow}{ $< = =$ end update}  }
919
920
921
922\newpage 
923% ================================================================
924% Subgrid Scale Physics
925% ================================================================
926\section{Subgrid Scale Physics}
927\label{PE_zdf_ldf}
928
929The primitive equations describe the behaviour of a geophysical fluid at
930space and time scales larger than a few kilometres in the horizontal, a few
931meters in the vertical and a few minutes. They are usually solved at larger
932scales: the specified grid spacing and time step of the numerical model. The
933effects of smaller scale motions (coming from the advective terms in the
934Navier-Stokes equations) must be represented entirely in terms of
935large-scale patterns to close the equations. These effects appear in the
936equations as the divergence of turbulent fluxes ($i.e.$ fluxes associated with
937the mean correlation of small scale perturbations). Assuming a turbulent
938closure hypothesis is equivalent to choose a formulation for these fluxes.
939It is usually called the subgrid scale physics. It must be emphasized that
940this is the weakest part of the primitive equations, but also one of the
941most important for long-term simulations as small scale processes \textit{in fine} 
942balance the surface input of kinetic energy and heat.
943
944The control exerted by gravity on the flow induces a strong anisotropy
945between the lateral and vertical motions. Therefore subgrid-scale physics 
946\textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \eqref{Eq_PE_dyn},
947\eqref{Eq_PE_tra_T} and \eqref{Eq_PE_tra_S} are divided into a lateral part 
948\textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part 
949\textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms
950and their underlying physics are briefly discussed in the next two subsections.
951
952% -------------------------------------------------------------------------------------------------------------
953% Vertical Subgrid Scale Physics
954% -------------------------------------------------------------------------------------------------------------
955\subsection{Vertical Subgrid Scale Physics}
956\label{PE_zdf}
957
958The model resolution is always larger than the scale at which the major
959sources of vertical turbulence occur (shear instability, internal wave
960breaking...). Turbulent motions are thus never explicitly solved, even
961partially, but always parameterized. The vertical turbulent fluxes are
962assumed to depend linearly on the gradients of large-scale quantities (for
963example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$,
964where $A^{vT}$ is an eddy coefficient). This formulation is
965analogous to that of molecular diffusion and dissipation. This is quite
966clearly a necessary compromise: considering only the molecular viscosity
967acting on large scale severely underestimates the role of turbulent
968diffusion and dissipation, while an accurate consideration of the details of
969turbulent motions is simply impractical. The resulting vertical momentum and
970tracer diffusive operators are of second order:
971\begin{equation} \label{Eq_PE_zdf}
972   \begin{split}
973{\vect{D}}^{v \vect{U}} &=\frac{\partial }{\partial z}\left( {A^{vm}\frac{\partial {\vect{U}}_h }{\partial z}} \right) \ , \\         
974D^{vT}                         &= \frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial T}{\partial z}} \right) \ ,
975\quad
976D^{vS}=\frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial S}{\partial z}} \right)
977   \end{split}
978\end{equation}
979where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients,
980respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat
981and salt must be specified (see Chap.~\ref{SBC} and \ref{ZDF} and \S\ref{TRA_bbl}).
982All the vertical physics is embedded in the specification of the eddy coefficients.
983They can be assumed to be either constant, or function of the local fluid properties
984($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a
985turbulent closure model. The choices available in \NEMO are discussed in \S\ref{ZDF}).
986
987% -------------------------------------------------------------------------------------------------------------
988% Lateral Diffusive and Viscous Operators Formulation
989% -------------------------------------------------------------------------------------------------------------
990\subsection{Lateral Diffusive and Viscous Operators Formulation}
991\label{PE_ldf}
992
993Lateral turbulence can be roughly divided into a mesoscale turbulence
994associated with eddies (which can be solved explicitly if the resolution is
995sufficient since their underlying physics are included in the primitive
996equations), and a sub mesoscale turbulence which is never explicitly solved
997even partially, but always parameterized. The formulation of lateral eddy
998fluxes depends on whether the mesoscale is below or above the grid-spacing
999($i.e.$ the model is eddy-resolving or not).
1000
1001In non-eddy-resolving configurations, the closure is similar to that used
1002for the vertical physics. The lateral turbulent fluxes are assumed to depend
1003linearly on the lateral gradients of large-scale quantities. The resulting
1004lateral diffusive and dissipative operators are of second order.
1005Observations show that lateral mixing induced by mesoscale turbulence tends
1006to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987})
1007rather than across them.
1008As the slope of neutral surfaces is small in the ocean, a common
1009approximation is to assume that the `lateral' direction is the horizontal,
1010$i.e.$ the lateral mixing is performed along geopotential surfaces. This leads
1011to a geopotential second order operator for lateral subgrid scale physics.
1012This assumption can be relaxed: the eddy-induced turbulent fluxes can be
1013better approached by assuming that they depend linearly on the gradients of
1014large-scale quantities computed along neutral surfaces. In such a case,
1015the diffusive operator is an isoneutral second order operator and it has
1016components in the three space directions. However, both horizontal and
1017isoneutral operators have no effect on mean ($i.e.$ large scale) potential
1018energy whereas potential energy is a main source of turbulence (through
1019baroclinic instabilities). \citet{Gent1990} have proposed a
1020parameterisation of mesoscale eddy-induced turbulence which associates an
1021eddy-induced velocity to the isoneutral diffusion. Its mean effect is to
1022reduce the mean potential energy of the ocean. This leads to a formulation
1023of lateral subgrid-scale physics made up of an isoneutral second order
1024operator and an eddy induced advective part. In all these lateral diffusive
1025formulations, the specification of the lateral eddy coefficients remains the
1026problematic point as there is no really satisfactory formulation of these
1027coefficients as a function of large-scale features.
1028
1029In eddy-resolving configurations, a second order operator can be used, but
1030usually a more scale selective one (biharmonic operator) is preferred as the
1031grid-spacing is usually not small enough compared to the scale of the
1032eddies. The role devoted to the subgrid-scale physics is to dissipate the
1033energy that cascades toward the grid scale and thus ensures the stability of
1034the model while not interfering with the solved mesoscale activity. Another approach
1035is becoming more and more popular: instead of specifying explicitly a sub-grid scale
1036term in the momentum and tracer time evolution equations, one uses a advective
1037scheme which is diffusive enough to maintain the model stability. It must be emphasised
1038that then, all the sub-grid scale physics is in this case include in the formulation of the
1039advection scheme.
1040
1041All these parameterisations of subgrid scale physics present advantages and
1042drawbacks. There are not all available in \NEMO. In the $z$-coordinate
1043formulation, five options are offered for active tracers (temperature and
1044salinity): second order geopotential operator, second order isoneutral
1045operator, \citet{Gent1990} parameterisation, fourth order
1046geopotential operator, and various slightly diffusive advection schemes.
1047The same options are available for momentum, except
1048\citet{Gent1990} parameterisation which only involves tracers. In the
1049$s$-coordinate formulation, additional options are offered for tracers: second
1050order operator acting along $s-$surfaces, and for momentum: fourth order
1051operator acting along $s-$surfaces (see \S\ref{LDF}).
1052
1053\subsubsection{lateral second order tracer diffusive operator}
1054
1055The lateral second order tracer diffusive operator is defined by (see Appendix~\ref{Apdx_B}):
1056\begin{equation} \label{Eq_PE_iso_tensor}
1057D^{lT}=\nabla {\rm {\bf .}}\left( {A^{lT}\;\Re \;\nabla T} \right) \qquad 
1058\mbox{with}\quad \;\;\Re =\left( {{\begin{array}{*{20}c}
1059 1 \hfill & 0 \hfill & {-r_1 } \hfill \\
1060 0 \hfill & 1 \hfill & {-r_2 } \hfill \\
1061 {-r_1 } \hfill & {-r_2 } \hfill & {r_1 ^2+r_2 ^2} \hfill \\
1062\end{array} }} \right)
1063\end{equation}
1064where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along
1065which the diffusive operator acts and the model level ($e. g.$ $z$- or
1066$s$-surfaces). Note that the formulation \eqref{Eq_PE_iso_tensor} is exact for the
1067rotation between geopotential and $s$-surfaces, while it is only an approximation
1068for the rotation between isoneutral and $z$- or $s$-surfaces. Indeed, in the latter
1069case, two assumptions are made to simplify  \eqref{Eq_PE_iso_tensor} \citep{Cox1987}.
1070First, the horizontal contribution of the dianeutral mixing is neglected since the ratio
1071between iso and dia-neutral diffusive coefficients is known to be several orders of
1072magnitude smaller than unity. Second, the two isoneutral directions of diffusion are
1073assumed to be independent since the slopes are generally less than $10^{-2}$ in the
1074ocean (see Appendix~\ref{Apdx_B}).
1075
1076For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the
1077geopotential and computational surfaces: in $z$-coordinates they are zero
1078($r_1 = r_2 = 0$) while in $s$-coordinate (including $\textit{z*}$ case) they are
1079equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ).
1080
1081For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral
1082and computational surfaces. Therefore, they have a same expression in $z$- and $s$-coordinates:
1083\begin{equation} \label{Eq_PE_iso_slopes}
1084r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
1085                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1} \ , \quad
1086r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
1087                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1}
1088\end{equation}
1089
1090When the \textit{Eddy Induced Velocity} parametrisation (eiv) \citep{Gent1990} is used,
1091an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
1092\begin{equation} \label{Eq_PE_iso+eiv}
1093D^{lT}=\nabla \cdot \left( {A^{lT}\;\Re \;\nabla T} \right)
1094           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right)
1095\end{equation}
1096where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent,
1097eddy-induced transport velocity. This velocity field is defined by:
1098\begin{equation} \label{Eq_PE_eiv}
1099   \begin{split}
1100 u^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_1 } \right] \\ 
1101 v^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_2 } \right] \\ 
1102 w^\ast &=  -\frac{1}{e_1 e_2 }\left[
1103                      \frac{\partial }{\partial i}\left( {A^{eiv}\;e_2\,\tilde{r}_1 } \right)
1104                    +\frac{\partial }{\partial j}\left( {A^{eiv}\;e_1\,\tilde{r}_2 } \right)      \right]
1105   \end{split}
1106\end{equation}
1107where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral
1108thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes
1109between isoneutral and \emph{geopotential} surfaces and thus depends on the coordinate
1110considered:
1111\begin{align} \label{Eq_PE_slopes_eiv}
1112\tilde{r}_n = \begin{cases}
1113   r_n                  &      \text{in $z$-coordinate}    \\
1114   r_n + \sigma_n &      \text{in \textit{z*} and $s$-coordinates} 
1115                   \end{cases}
1116\quad \text{where } n=1,2
1117\end{align}
1118
1119The normal component of the eddy induced velocity is zero at all the boundaries.
1120This can be achieved in a model by tapering either the eddy coefficient or the slopes
1121to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. Chap.~\ref{LDF}).
1122
1123\subsubsection{lateral fourth order tracer diffusive operator}
1124
1125The lateral fourth order tracer diffusive operator is defined by:
1126\begin{equation} \label{Eq_PE_bilapT}
1127D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
1128\qquad \text{where} \  D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
1129 \end{equation}
1130
1131It is the second order operator given by \eqref{Eq_PE_iso_tensor} applied twice with
1132the eddy diffusion coefficient correctly placed.
1133
1134
1135\subsubsection{lateral second order momentum diffusive operator}
1136
1137The second order momentum diffusive operator along $z$- or $s$-surfaces is found by
1138applying \eqref{Eq_PE_lap_vector} to the horizontal velocity vector (see Appendix~\ref{Apdx_B}):
1139\begin{equation} \label{Eq_PE_lapU}
1140\begin{split}
1141{\rm {\bf D}}^{l{\rm {\bf U}}} 
1142&= \quad \  \nabla _h \left( {A^{lm}\chi } \right)
1143   \ - \ \nabla _h \times \left( {A^{lm}\,\zeta \;{\rm {\bf k}}} \right)     \\
1144&=   \left(      \begin{aligned}
1145             \frac{1}{e_1      } \frac{\partial \left( A^{lm} \chi          \right)}{\partial i} 
1146         &-\frac{1}{e_2 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial j}  \\
1147             \frac{1}{e_2      }\frac{\partial \left( {A^{lm} \chi         } \right)}{\partial j}   
1148         &+\frac{1}{e_1 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial i}
1149        \end{aligned}    \right)
1150\end{split}
1151\end{equation}
1152
1153Such a formulation ensures a complete separation between the vorticity and
1154horizontal divergence fields (see Appendix~\ref{Apdx_C}). Unfortunately, it is not
1155available for geopotential diffusion in $s-$coordinates and for isoneutral
1156diffusion in both $z$- and $s$-coordinates ($i.e.$ when a rotation is required).
1157In these two cases, the $u$ and $v-$fields are considered as independent scalar
1158fields, so that the diffusive operator is given by:
1159\begin{equation} \label{Eq_PE_lapU_iso}
1160\begin{split}
1161 D_u^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla u} \right) \\ 
1162 D_v^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla v} \right)
1163 \end{split}
1164 \end{equation}
1165where $\Re$ is given by  \eqref{Eq_PE_iso_tensor}. It is the same expression as
1166those used for diffusive operator on tracers. It must be emphasised that such a
1167formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or
1168$\beta-$plane, not on the sphere. It is also a very good approximation in vicinity
1169of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}.
1170
1171\subsubsection{lateral fourth order momentum diffusive operator}
1172
1173As for tracers, the fourth order momentum diffusive operator along $z$ or $s$-surfaces
1174is a re-entering second order operator \eqref{Eq_PE_lapU} or \eqref{Eq_PE_lapU} 
1175with the eddy viscosity coefficient correctly placed:
1176
1177geopotential diffusion in $z$-coordinate:
1178\begin{equation} \label{Eq_PE_bilapU}
1179\begin{split}
1180{\rm {\bf D}}^{l{\rm {\bf U}}} &=\nabla _h \left\{ {\;\nabla _h {\rm {\bf 
1181.}}\left[ {A^{lm}\,\nabla _h \left( \chi \right)} \right]\;} 
1182\right\}\;   \\
1183&+\nabla _h \times \left\{ {\;{\rm {\bf k}}\cdot \nabla \times 
1184\left[ {A^{lm}\,\nabla _h \times \left( {\zeta \;{\rm {\bf k}}} \right)} 
1185\right]\;} \right\}
1186\end{split}
1187\end{equation}
1188
1189\gmcomment{  change the position of the coefficient, both here and in the code}
1190
1191geopotential diffusion in $s$-coordinate:
1192\begin{equation} \label{Eq_bilapU_iso}
1193   \left\{   \begin{aligned}
1194         D_u^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta u} \right) \\ 
1195         D_v^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta v} \right)
1196   \end{aligned}    \right.
1197   \quad \text{where} \quad 
1198   \Delta \left( \bullet \right) = \nabla \cdot \left( \Re \nabla(\bullet) \right)
1199\end{equation}
1200
Note: See TracBrowser for help on using the repository browser.