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 tags/nemo_v3_2/nemo_v3_2/DOC/TexFiles/Chapters – NEMO

source: tags/nemo_v3_2/nemo_v3_2/DOC/TexFiles/Chapters/Chap_Model_Basics.tex @ 1878

Last change on this file since 1878 was 1878, checked in by flavoni, 14 years ago

initial test for nemogcm

File size: 72.9 KB
Line 
1% ================================================================
2% Chapter 1 Ñ Model Basics
3% ================================================================
4
5\chapter{Model basics}
6\label{PE}
7\minitoc
8
9
10% ================================================================
11% Primitive Equations
12% ================================================================
13\section{Primitive Equations}
14\label{PE_PE}
15
16% -------------------------------------------------------------------------------------------------------------
17%        Vector Invariant Formulation
18% -------------------------------------------------------------------------------------------------------------
19
20\subsection{Vector Invariant Formulation}
21\label{PE_Vector}
22
23
24The ocean is a fluid that can be described to a good approximation by the primitive
25equations, $i.e.$ the Navier-Stokes equations along with a nonlinear equation of
26state which couples the two active tracers (temperature and salinity) to the fluid
27velocity, plus the following additional assumptions made from scale considerations:
28
29\textit{(1) spherical earth approximation: }the geopotential surfaces are assumed to
30be spheres so that gravity (local vertical) is parallel to the earth's radius
31
32\textit{(2) thin-shell approximation: }the ocean depth is neglected compared to the earth's radius
33
34\textit{(3) turbulent closure hypothesis: }the turbulent fluxes (which represent the effect
35of small scale processes on the large-scale) are expressed in terms of large-scale features
36
37\textit{(4) Boussinesq hypothesis:} density variations are neglected except in their
38contribution to the buoyancy force
39
40\textit{(5) Hydrostatic hypothesis: }the vertical momentum equation is reduced to a
41balance between the vertical pressure gradient and the buoyancy force (this removes
42convective processes from the initial Navier-Stokes equations and so convective processes
43must be parameterized instead)
44
45\textit{(6) Incompressibility hypothesis: }the three dimensional divergence of the velocity
46vector is assumed to be zero.
47
48Because the gravitational force is so dominant in the equations of large-scale motions,
49it is useful to choose an orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k}) linked
50to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are two
51vectors orthogonal to \textbf{k}, $i.e.$ tangent to the geopotential surfaces. Let us define
52the following variables: \textbf{U} the vector velocity, $\textbf{U}=\textbf{U}_h + w\, \textbf{k}$ 
53(the subscript $h$ denotes the local horizontal vector, $i.e.$ over the (\textbf{i},\textbf{j}) plane),
54$T$ the potential temperature, $S$ the salinity, \textit{$\rho $} the \textit{in situ} density.
55The vector invariant form of the primitive equations in the (\textbf{i},\textbf{j},\textbf{k})
56vector system provides the following six equations (namely the momentum balance, the
57hydrostatic equilibrium, the incompressibility equation, the heat and salt conservation
58equations and an equation of state):
59\begin{subequations} \label{Eq_PE}
60  \begin{equation}     \label{Eq_PE_dyn}
61\frac{\partial {\rm {\bf U}}_h }{\partial t}=
62-\left[    {\left( {\nabla \times {\rm {\bf U}}} \right)\times {\rm {\bf U}}
63            +\frac{1}{2}\nabla \left( {{\rm {\bf U}}^2} \right)}    \right]_h
64 -f\;{\rm {\bf k}}\times {\rm {\bf U}}_h
65-\frac{1}{\rho _o }\nabla _h p + {\rm {\bf D}}^{\rm {\bf U}} + {\rm {\bf F}}^{\rm {\bf U}}
66  \end{equation}
67  \begin{equation}     \label{Eq_PE_hydrostatic}
68\frac{\partial p }{\partial z} = - \rho \ g
69  \end{equation}
70  \begin{equation}     \label{Eq_PE_continuity}
71\nabla \cdot {\bf U}=  0
72  \end{equation}
73\begin{equation} \label{Eq_PE_tra_T}
74\frac{\partial T}{\partial t} = - \nabla \cdot  \left( T \ \rm{\bf U} \right) + D^T + F^T
75  \end{equation}
76  \begin{equation}     \label{Eq_PE_tra_S}
77\frac{\partial S}{\partial t} = - \nabla \cdot  \left( S \ \rm{\bf U} \right) + D^S + F^S
78  \end{equation}
79  \begin{equation}     \label{Eq_PE_eos}
80\rho = \rho \left( T,S,p \right)
81  \end{equation}
82\end{subequations}
83where $\nabla$ is the generalised derivative vector operator in $(\bf i,\bf j, \bf k)$ directions,
84$t$ is the time, $z$ is the vertical coordinate, $\rho $ is the \textit{in situ} density given by
85the equation of state (\ref{Eq_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,
86$f=2 \bf \Omega \cdot \bf k$ is the Coriolis acceleration (where $\bf \Omega$ is the Earth's
87angular velocity vector), and $g$ is the gravitational acceleration.
88${\rm {\bf D}}^{\rm {\bf U}}$, $D^T$ and $D^S$ are the parameterisations of small-scale
89physics for momentum, temperature and salinity, and ${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ 
90and $F^S$ surface forcing terms. Their nature and formulation are discussed in
91\S\ref{PE_zdf_ldf} and page \S\ref{PE_boundary_condition}.
92
93.
94
95% -------------------------------------------------------------------------------------------------------------
96% Boundary condition
97% -------------------------------------------------------------------------------------------------------------
98\subsection{Boundary Conditions}
99\label{PE_boundary_condition}
100
101An ocean is bounded by complex coastlines, bottom topography at its base and an air-sea
102or ice-sea interface at its top. These boundaries can be defined by two surfaces, $z=-H(i,j)$ 
103and $z=\eta(i,j,k,t)$, where $H$ is the depth of the ocean bottom and $\eta$ is the height
104of the sea surface. Both $H$ and $\eta$ are usually referenced to a given surface, $z=0$,
105chosen as a mean sea surface (Fig.~\ref{Fig_ocean_bc}). Through these two boundaries,
106the ocean can exchange fluxes of heat, fresh water, salt, and momentum with the solid earth,
107the continental margins, the sea ice and the atmosphere. However, some of these fluxes are
108so weak that even on climatic time scales of thousands of years they can be neglected.
109In the following, we briefly review the fluxes exchanged at the interfaces between the ocean
110and the other components of the earth system.
111
112%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
113\begin{figure}[!ht] \label{Fig_ocean_bc}  \begin{center}
114\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_I_ocean_bc.pdf}
115\caption{The ocean is bounded by two surfaces, $z=-H(i,j)$ and $z=\eta(i,j,k,t)$, where $H$ 
116is the depth of the sea floor and $\eta$ the height of the sea surface. Both $H$ and $\eta $ 
117are referenced to $z=0$.}
118\end{center}   \end{figure}
119%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
120
121
122\begin{description}
123\item[Land - ocean interface:] the major flux between continental margins and the ocean is
124a mass exchange of fresh water through river runoff. Such an exchange modifies the sea
125surface salinity especially in the vicinity of major river mouths. It can be neglected for short
126range integrations but has to be taken into account for long term integrations as it influences
127the characteristics of water masses formed (especially at high latitudes). It is required in order
128to close the water cycle of the climate system. It is usually specified as a fresh water flux at
129the air-sea interface in the vicinity of river mouths.
130\item[Solid earth - ocean interface:] heat and salt fluxes through the sea floor are small,
131except in special areas of little extent. They are usually neglected in the model
132\footnote{In fact, it has been shown that the heat flux associated with the solid Earth cooling
133($i.e.$the geothermal heating) is not negligible for the thermohaline circulation of the world
134ocean (see \ref{TRA_bbc}).}.
135The boundary condition is thus set to no flux of heat and salt across solid boundaries.
136For momentum, the situation is different. There is no flow across solid boundaries,
137$i.e.$ the velocity normal to the ocean bottom and coastlines is zero (in other words,
138the bottom velocity is parallel to solid boundaries). This kinematic boundary condition
139can be expressed as:
140\begin{equation} \label{Eq_PE_w_bbc}
141w = -{\rm {\bf U}}_h \cdot  \nabla _h \left( H \right)
142\end{equation}
143In addition, the ocean exchanges momentum with the earth through frictional processes.
144Such momentum transfer occurs at small scales in a boundary layer. It must be parameterized
145in terms of turbulent fluxes using bottom and/or lateral boundary conditions. Its specification
146depends on the nature of the physical parameterisation used for ${\rm {\bf D}}^{\rm {\bf U}}$ 
147in \eqref{Eq_PE_dyn}. It is discussed in \S\ref{PE_zdf}, page~\pageref{PE_zdf}.% and Chap. III.6 to 9.
148\item[Atmosphere - ocean interface:] the kinematic surface condition plus the mass flux
149of fresh water PE  (the precipitation minus evaporation budget) leads to:
150\begin{equation} \label{Eq_PE_w_sbc}
151w = \frac{\partial \eta }{\partial t} 
152    + \left. {{\rm {\bf U}}_h } \right|_{z=\eta } \cdot  \nabla _h \left( \eta \right)
153    + P-E
154\end{equation}
155The dynamic boundary condition, neglecting the surface tension (which removes capillary
156waves from the system) leads to the continuity of pressure across the interface $z=\eta$.
157The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
158\item[Sea ice - ocean interface:] the ocean and sea ice exchange heat, salt, fresh water
159and momentum. The sea surface temperature is constrained to be at the freezing point
160at the interface. Sea ice salinity is very low ($\sim4-6 \,psu$) compared to those of the
161ocean ($\sim34 \,psu$). The cycle of freezing/melting is associated with fresh water and
162salt fluxes that cannot be neglected.
163\end{description}
164
165
166% ================================================================
167% The Horizontal Pressure Gradient
168% ================================================================
169\section{The Horizontal Pressure Gradient }
170\label{PE_hor_pg}
171
172% -------------------------------------------------------------------------------------------------------------
173% Pressure Formulation
174% -------------------------------------------------------------------------------------------------------------
175\subsection{Pressure Formulation}
176\label{PE_p_formulation}
177
178The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at a
179reference geopotential surface ($z=0$) and a hydrostatic pressure $p_h$ such that:
180$p(i,j,k,t)=p_s(i,j,t)+p_h(i,j,k,t)$. The latter is computed by integrating (\ref{Eq_PE_hydrostatic}),
181assuming that pressure in decibars can be approximated by depth in meters in (\ref{Eq_PE_eos}).
182The hydrostatic pressure is then given by:
183\begin{equation} \label{Eq_PE_pressure}
184p_h \left( {i,j,z,t} \right)
185 = \int_{\varsigma =z}^{\varsigma =0} {g\;\rho \left( {T,S,\varsigma} \right)\;d\varsigma } 
186\end{equation}
187 Two strategies can be considered for the surface pressure term: $(a)$ introduce of a
188 new variable $\eta$, the free-surface elevation, for which a prognostic equation can be
189 established and solved; $(b)$ assume that the ocean surface is a rigid lid, on which the
190 pressure (or its horizontal gradient) can be diagnosed. When the former strategy is used,
191 one solution of the free-surface elevation consists of the excitation of external gravity waves.
192 The flow is barotropic and the surface moves up and down with gravity as the restoring force.
193 The phase speed of such waves is high (some hundreds of metres per second) so that
194 the time step would have to be very short if they were present in the model. The latter
195 strategy filters out these waves since the rigid lid approximation implies $\eta=0$, $i.e.$ 
196 the sea surface is the surface $z=0$. This well known approximation increases the surface
197 wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic
198 Rossby or planetary waves). In the present release of \NEMO, both strategies are still available.
199 They are further described in the next two sub-sections.
200
201% -------------------------------------------------------------------------------------------------------------
202% Free Surface Formulation
203% -------------------------------------------------------------------------------------------------------------
204\subsection{Free Surface Formulation}
205\label{PE_free_surface}
206
207In the free surface formulation, a variable $\eta$, the sea-surface height, is introduced
208which describes the shape of the air-sea interface. This variable is solution of a
209prognostic equation which is established by forming the vertical average of the kinematic
210surface condition (\ref{Eq_PE_w_bbc}):
211\begin{equation} \label{Eq_PE_ssh}
212\frac{\partial \eta }{\partial t}=-D+P-E
213   \quad \text{where} \
214D=\nabla \cdot \left[ {\left( {H+\eta } \right) \; {\rm{\bf \overline{U}}}_h \,} \right]
215\end{equation}
216and using (\ref{Eq_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
217
218Allowing the air-sea interface to move introduces the external gravity waves (EGWs)
219as a class of solution of the primitive equations. These waves are barotropic because
220of hydrostatic assumption, and their phase speed is quite high. Their time scale is
221short with respect to the other processes described by the primitive equations.
222
223Three choices can be made regarding the implementation of the free surface in the model,
224depending on the physical processes of interest.
225
226$\bullet$ If one is interested in EGWs, in particular the tides and their interaction
227with the baroclinic structure of the ocean (internal waves) possibly in shallow seas,
228then a non linear free surface is the most appropriate. This means that no
229approximation is made in (\ref{Eq_PE_ssh}) and that the variation of the ocean
230volume is fully taken into account. Note that in order to study the fast time scales
231associated with EGWs it is necessary to minimize time filtering effects (use an
232explicit time scheme with very small time step, or a split-explicit scheme with
233reasonably small time step, see \S\ref{DYN_spg_exp} or \S\ref{DYN_spg_ts}.
234
235$\bullet$ If one is not interested in EGW but rather sees them as high frequency
236noise, it is possible to apply an explicit filter to slow down the fastest waves while
237not altering the slow barotropic Rossby waves. If further, an approximative conservation
238of heat and salt contents is sufficient for the problem solved, then it is
239sufficient to solve a linearized version of (\ref{Eq_PE_ssh}), which still allows
240to take into account freshwater fluxes applied at the ocean surface \citep{Roullet2000}.
241
242$\bullet$ For process studies not involving external waves nor surface freshwater
243fluxes, it is possible to use the rigid lid approximation see (next
244section). The ocean surface is then considered as a fixed surface, so that all
245external waves are removed from the system.
246
247The filtering of EGWs in models with a free surface is usually a matter of discretisation
248of the temporal derivatives, using the time splitting method \citep{Killworth1991, Zhang1992} 
249or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach
250developed by \citet{Roullet2000}: the damping of EGWs is ensured by introducing an
251additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:
252\begin{equation} \label{Eq_PE_flt}
253\frac{\partial {\rm {\bf U}}_h }{\partial t}= {\rm {\bf M}}
254- g \nabla \left( \tilde{\rho} \ \eta \right)
255- g \ T_c \nabla \left( \widetilde{\rho} \ \partial_t \eta \right)
256\end{equation}
257where $T_c$, is a parameter with dimensions of time which characterizes the force,
258$\widetilde{\rho} = \rho / \rho_o$ is the dimensionless density, and $\rm {\bf M}$ 
259represents the collected contributions of the Coriolis, hydrostatic pressure gradient,
260non-linear and viscous terms in \eqref{Eq_PE_dyn}.
261
262The new force can be interpreted as a diffusion of vertically integrated volume flux divergence.
263The time evolution of $D$ is thus governed by a balance of two terms, $-g$ \textbf{A} $\eta$ 
264and $g \, T_c \,$ \textbf{A} $D$, associated with a propagative regime and a diffusive regime
265in the temporal spectrum, respectively. In the diffusive regime, the EGWs no longer propagate,
266$i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than
267$T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs
268can be damped by choosing $T_c > \Delta t$. \citet{Roullet2000} demonstrate that
269(\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which
270has to be computed implicitly. This is not surprising since the use of a large time step has a
271necessarily numerical cost. Two gains arise in comparison with the previous formulations.
272Firstly, the damping of EGWs can be quantified through the magnitude of the additional term.
273Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as
274soon as $T_c > \Delta t$.
275
276When the variations of free surface elevation are small compared to the thickness of the first
277model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized
278by \citet{Roullet2000} the linearization of (\ref{Eq_PE_ssh}) has consequences on the
279conservation of salt in the model. With the nonlinear free surface equation, the time evolution
280of the total salt content is
281\begin{equation} \label{Eq_PE_salt_content}
282    \frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
283                        =\int\limits_S {S\;(-\frac{\partial \eta }{\partial t}-D+P-E)\;ds}
284\end{equation}
285where $S$ is the salinity, and the total salt is integrated over the whole ocean volume
286$D_\eta$ bounded by the time-dependent free surface. The right hand side (which is an
287integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh})
288is satisfied, so that the salt is perfectly conserved. When the free surface equation is
289linearized, \citet{Roullet2000} show that the total salt content integrated in the fixed
290volume $D$ (bounded by the surface $z=0$) is no longer conserved:
291\begin{equation} \label{Eq_PE_salt_content_linear}
292         \frac{\partial }{\partial t}\int\limits_D {S\;dv} 
293               = - \int\limits_S {S\;\frac{\partial \eta }{\partial t}ds} 
294\end{equation}
295
296The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions
297\citep{Roullet2000}. It can be significant when the freshwater forcing is not balanced and
298the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $} 
299results in a decrease of the salinity in the fixed volume $D$. Even in that case though,
300the total salt integrated in the variable volume $D_{\eta}$ varies much less, since
301(\ref{Eq_PE_salt_content_linear}) can be rewritten as
302\begin{equation} \label{Eq_PE_salt_content_corrected}
303\frac{\partial }{\partial t}\int\limits_{D\eta } {S\;dv} 
304=\frac{\partial}{\partial t} \left[ \;{\int\limits_D {S\;dv} +\int\limits_S {S\eta \;ds} } \right]
305=\int\limits_S {\eta \;\frac{\partial S}{\partial t}ds}
306\end{equation}
307
308Although the total salt content is not exactly conserved with the linearized free surface,
309its variations are driven by correlations of the time variation of surface salinity with the
310sea surface height, which is a negligible term. This situation contrasts with the case of
311the rigid lid approximation (following section) in which case freshwater forcing is
312represented by a virtual salt flux, leading to a spurious source of salt at the ocean
313surface \citep{Roullet2000}.
314
315% -------------------------------------------------------------------------------------------------------------
316% Rigid-Lid Formulation
317% -------------------------------------------------------------------------------------------------------------
318\subsection{Rigid-Lid formulation}
319\label{PE_rigid_lid}
320
321With the rigid lid approximation, we assume that the ocean surface ($z=0$) is a rigid lid
322on which a pressure $p_s$ is exerted. This implies that the vertical velocity at the surface
323is equal to zero. From the continuity equation \eqref{Eq_PE_continuity} and the kinematic
324condition at the bottom \eqref{Eq_PE_w_bbc} (no flux across the bottom), it can be shown
325that the vertically integrated flow $H{\rm {\bf \bar {U}}}_h$ is non-divergent (where the
326overbar indicates a vertical average over the whole water column, i.e. from $z=-H$,
327the ocean bottom, to $z=0$ , the rigid-lid). Thus, $\rm {\bf \bar {U}}_h$ can be derived
328from a volume transport streamfunction $\psi$:
329\begin{equation} \label{Eq_PE_u_psi}
330\overline{\vect{U}}_h =\frac{1}{H}\left(   \vect{k} \times \nabla \psi   \right)
331\end{equation}
332
333As $p_s$ does not depend on depth, its horizontal gradient is obtained by forming the
334vertical average of \eqref{Eq_PE_dyn} and using \eqref{Eq_PE_u_psi}:
335
336\begin{equation} \label{Eq_PE_u_barotrope}
337\frac{1}{\rho _o }\nabla _h p_s
338=\overline{\vect{M}} -\frac{\partial \overline{\vect{U}} _h }{\partial t}
339=\overline{\vect{M}} 
340-\frac{1}{H} \left[   \vect{k} \times \nabla \left( \frac{\partial \psi}{\partial t} \right)   \right]
341\end{equation}
342
343Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ represents the collected contributions of the
344Coriolis, hydrostatic pressure gradient, nonlinear and viscous terms in \eqref{Eq_PE_dyn}.
345The time derivative of $\psi $ is the solution of an elliptic equation which is obtained from
346the vertical component of the curl of (\ref{Eq_PE_u_barotrope}):
347\begin{equation} \label{Eq_PE_psi}
348\left[   {\nabla \times \left[ {\frac{1}{H} \vect{\bf k} 
349  \times \nabla \left(   {\frac{\partial \psi }{\partial t}} \right)}   \right]} \; \right]_z
350=\left[   {\nabla \times \overline{\vect{M}} }   \right]_z
351\end{equation}
352
353Using the proper boundary conditions, \eqref{Eq_PE_psi} can be solved to find $\partial_t \psi$ 
354and thus using \eqref{Eq_PE_u_barotrope} the horizontal surface pressure gradient.
355It should be noted that $p_s$ can be computed by taking the divergence of
356\eqref{Eq_PE_u_barotrope} and solving the resulting elliptic equation. Thus the surface
357pressure is a diagnostic quantity that can be recovered for analysis purposes.
358
359A difficulty lies in the determination of the boundary condition on $\partial_t \psi$.
360The boundary condition on velocity is that there is no flow normal to a solid wall,
361$i.e.$ the coastlines are streamlines. Therefore \eqref{Eq_PE_psi} is solved with
362the following Dirichlet boundary condition: $\partial_t \psi$ is constant along each
363coastline of the same continent or of the same island. When all the coastlines are
364connected (there are no islands), the constant value of $\partial_t \psi$ along the
365coast can be arbitrarily chosen to be zero. When islands are present in the domain,
366the value of the barotropic streamfunction will generally be different for each island
367and for the continent, and will vary with respect to time. So the boundary condition is:
368$\psi=0$ along the continent and $\psi=\mu_n$ along island $n$ ($1 \leq n \leq Q$),
369where $Q$ is the number of islands present in the domain and $\mu_n$ is a time
370dependent variable. A time evolution equation of the unknown $\mu_n$ can be found
371by evaluating the circulation of the time derivative of the vertical average (barotropic)
372velocity field along a closed contour around each island. Since the circulation of a
373gradient field along a closed contour is zero, from \eqref{Eq_PE_u_barotrope} we have:
374\begin{equation} \label{Eq_PE_isl_circulation}
375\oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
376{\frac{\partial \psi }{\partial t}} \right)} \right] \cdot {\rm {\bf d}}\ell } 
377= \oint_n {\overline {\rm {\bf M}} \cdot {\rm {\bf d}}\ell } 
378\qquad  1 \leq n \leq Q
379\end{equation}
380
381Since (\ref{Eq_PE_psi}) is linear, its solution \textit{$\psi $} can be decomposed
382as follows:
383\begin{equation} \label{Eq_PE_psi_isl}
384\psi =\psi _o +\sum\limits_{n=1}^{n=Q} {\mu _n \psi _n } 
385\end{equation}
386where $\psi _o$ is the solution of \eqref{Eq_PE_psi} with $\psi _o=0$ long all
387the coastlines, and where $\psi _n$ is the solution of \eqref{Eq_PE_psi} with
388the right-hand side equal to $0$, and with $\psi _n =1$ long the island $n$,
389$\psi _n =0$ along the other boundaries. The function $\psi _n$ is thus
390independent of time. Introducing \eqref{Eq_PE_psi_isl} into
391\eqref{Eq_PE_isl_circulation} yields:
392\begin{multline} \label{Eq_PE_psi_isl_circulation}
393\left[ {\oint_n {\frac{1}{H}  \left[ {{\rm {\bf k}}\times \nabla \psi _m } \right]\cdot {\rm {\bf d}}\ell } } \right]_{1\leq m\leqslant Q \atop 1\leq n\leqslant Q  }
394 \left( {\frac{\partial \mu _n }{\partial t}} 
395\right)_{1\leqslant n\leqslant Q}        \\
396 =\left( {\oint_n {\left[ {\overline {\rm 
397{\bf M}} -\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(
398{\frac{\partial \psi _o }{\partial t}} \right)} \right]} \right]\cdot {\rm 
399{\bf d}}\ell } } \right)_{1\leqslant n\leqslant Q} 
400\end{multline}
401which can be rewritten as:
402\begin{equation} \label{Eq_PE_psi_isl_matrix}
403{\rm {\bf A}}\;\left( {\frac{\partial \mu _n }{\partial t}} 
404\right)_{1\leqslant n\leqslant Q} ={\rm {\bf B}}
405\end{equation}
406where \textbf{A} is a $\times Q$ matrix and \textbf{B} is a time dependent vector.
407As \textbf{A} is independent of time, it can be calculated and inverted once. The time
408derivative of the streamfunction when islands are present is thus given by:
409\begin{equation} \label{Eq_PE_psi_isl_dt}
410\frac{\partial \psi }{\partial t}=\frac{\partial \psi _o }{\partial 
411t}+\sum\limits_{n=1}^{n=Q} {{\rm {\bf A}}^{-1}{\rm {\bf B}}\;\psi _n } 
412\end{equation}
413
414
415
416% ================================================================
417% Curvilinear z-coordinate System
418% ================================================================
419\section{Curvilinear \textit{z-}coordinate System}
420\label{PE_zco}
421
422
423% -------------------------------------------------------------------------------------------------------------
424% Tensorial Formalism
425% -------------------------------------------------------------------------------------------------------------
426\subsection{Tensorial Formalism}
427\label{PE_tensorial}
428
429In many ocean circulation problems, the flow field has regions of enhanced dynamics
430($i.e.$ surface layers, western boundary currents, equatorial currents, or ocean fronts).
431The representation of such dynamical processes can be improved by specifically increasing
432the model resolution in these regions. As well, it may be convenient to use a lateral
433boundary-following coordinate system to better represent coastal dynamics. Moreover,
434the common geographical coordinate system has a singular point at the North Pole that
435cannot be easily treated in a global model without filtering. A solution consists of introducing
436an appropriate coordinate transformation that shifts the singular point onto land
437\citep{MadecImb1996, Murray1996}. As a consequence, it is important to solve the primitive
438equations in various curvilinear coordinate systems. An efficient way of introducing an
439appropriate coordinate transform can be found when using a tensorial formalism.
440This formalism is suited to any multidimensional curvilinear coordinate system. Ocean
441modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth
442approximation), with preservation of the local vertical. Here we give the simplified equations
443for this particular case. The general case is detailed by \citet{Eiseman1980} in their survey
444of the conservation laws of fluid dynamics.
445
446Let (\textbf{i},\textbf{j},\textbf{k}) be a set of orthogonal curvilinear coordinates on the sphere
447associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k})
448linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are
449two vectors orthogonal to \textbf{k}, $i.e.$ along geopotential surfaces (Fig.\ref{Fig_referential}).
450Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined
451by the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and the distance from the centre of
452the earth $a+z(k)$ where $a$ is the earth's radius and $z$ the altitude above a reference sea
453level (Fig.\ref{Fig_referential}). The local deformation of the curvilinear coordinate system is
454given by $e_1$, $e_2$ and $e_3$, the three scale factors:
455\begin{equation} \label{Eq_scale_factors}
456\begin{aligned}
457 e_1 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
458}{\partial i}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
459}{\partial i}} \right)^2} \right]^{1/2} \\ 
460 e_2 &=\left( {a+z} \right)\;\left[ {\left( {\frac{\partial \lambda 
461}{\partial j}\cos \varphi } \right)^2+\left( {\frac{\partial \varphi 
462}{\partial j}} \right)^2} \right]^{1/2} \\ 
463 e_3 &=\left( {\frac{\partial z}{\partial k}} \right) \\ 
464 \end{aligned}
465 \end{equation}
466
467%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
468\begin{figure}[!tb] \label{Fig_referential}  \begin{center}
469\includegraphics[width=0.60\textwidth]{./TexFiles/Figures/Fig_I_earth_referential.pdf}
470\caption{the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear
471coordinate system (\textbf{i},\textbf{j},\textbf{k}). }
472\end{center}   \end{figure}
473%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
474
475Since the ocean depth is far smaller than the earth's radius, $a+z$, can be replaced by
476$a$ in (\ref{Eq_scale_factors}) (thin-shell approximation). The resulting horizontal scale
477factors $e_1$, $e_2$  are independent of $k$ while the vertical scale factor is a single
478function of $k$ as \textbf{k} is parallel to \textbf{z}. The scalar and vector operators that
479appear in the primitive equations (Eqs. \eqref{Eq_PE_dyn} to \eqref{Eq_PE_eos}) can
480be written in the tensorial form, invariant in any orthogonal horizontal curvilinear coordinate
481system transformation:
482\begin{subequations} \label{Eq_PE_discrete_operators}
483\begin{equation} \label{Eq_PE_grad}
484\nabla q=\frac{1}{e_1 }\frac{\partial q}{\partial i}\;{\rm {\bf 
485i}}+\frac{1}{e_2 }\frac{\partial q}{\partial j}\;{\rm {\bf j}}+\frac{1}{e_3
486}\frac{\partial q}{\partial k}\;{\rm {\bf k}}    \\
487\end{equation}
488\begin{equation} \label{Eq_PE_div}
489\nabla \cdot {\rm {\bf A}} 
490= \frac{1}{e_1 \; e_2} \left[
491  \frac{\partial \left(e_2 \; a_1\right)}{\partial i }
492+\frac{\partial \left(e_1 \; a_2\right)}{\partial j }       \right]
493+ \frac{1}{e_3} \left[ \frac{\partial a_3}{\partial k }   \right]
494\end{equation}
495\begin{equation} \label{Eq_PE_curl}
496   \begin{split}
497\nabla \times \vect{A} =
498    \left[ {\frac{1}{e_2 }\frac{\partial a_3}{\partial j}
499            -\frac{1}{e_3 }\frac{\partial a_2 }{\partial k}} \right] \; \vect{i}
500&+\left[ {\frac{1}{e_3 }\frac{\partial a_1 }{\partial k}
501           -\frac{1}{e_1 }\frac{\partial a_3 }{\partial i}} \right] \; \vect{j}     \\
502&+\frac{1}{e_1 e_2 } \left[ {\frac{\partial \left( {e_2 a_2 } \right)}{\partial i}
503                                       -\frac{\partial \left( {e_1 a_1 } \right)}{\partial j}} \right] \; \vect{k} 
504   \end{split}
505\end{equation}
506\begin{equation} \label{Eq_PE_lap}
507\Delta q = \nabla \cdot \left\nabla q \right)
508\end{equation}
509\begin{equation} \label{Eq_PE_lap_vector}
510\Delta {\rm {\bf A}} =
511  \nabla \left( \nabla \cdot {\rm {\bf A}} \right)
512- \nabla \times \left\nabla \times {\rm {\bf A}} \right)
513\end{equation}
514\end{subequations}
515where $q$ is a scalar quantity and ${\rm {\bf A}}=(a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinate system.
516
517% -------------------------------------------------------------------------------------------------------------
518% Continuous Model Equations
519% -------------------------------------------------------------------------------------------------------------
520\subsection{Continuous Model Equations}
521\label{PE_zco_Eq}
522
523In order to express the Primitive Equations in tensorial formalism, it is necessary to compute
524the horizontal component of the non-linear and viscous terms of the equation using
525\eqref{Eq_PE_grad}) to \eqref{Eq_PE_lap_vector}.
526Let us set $\vect U=(u,v,w)={\vect{U}}_h +w\;\vect{k}$, the velocity in the $(i,j,k)$ coordinate
527system and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity
528field $\chi$, by:
529\begin{equation} \label{Eq_PE_curl_Uh}
530\zeta =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,v} 
531\right)}{\partial i}-\frac{\partial \left( {e_1 \,u} \right)}{\partial j}} 
532\right]
533\end{equation}
534\begin{equation} \label{Eq_PE_div_Uh}
535\chi =\frac{1}{e_1 e_2 }\left[ {\frac{\partial \left( {e_2 \,u} 
536\right)}{\partial i}+\frac{\partial \left( {e_1 \,v} \right)}{\partial j}} 
537\right]
538\end{equation}
539
540Using the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ 
541and that $e_3$  is a function of the single variable $k$, the nonlinear term of
542\eqref{Eq_PE_dyn} can be transformed as follows:
543\begin{flalign*}
544&\left[ {\left( { \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
545+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h        &
546\end{flalign*}
547\begin{flalign*}
548&\qquad=\left( {{\begin{array}{*{20}c}
549 {\left[    {   \frac{1}{e_3} \frac{\partial u  }{\partial k}
550         -\frac{1}{e_1} \frac{\partial w  }{\partial i} } \right] w - \zeta \; v }     \\
551      {\zeta \; u - \left[ {   \frac{1}{e_2} \frac{\partial w}{\partial j}
552                     -\frac{1}{e_3} \frac{\partial v}{\partial k} } \right] \ w}  \\
553       \end{array} }} \right)       
554+\frac{1}{2}   \left( {{\begin{array}{*{20}c}
555       { \frac{1}{e_1}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial i}}  \hfill    \\
556       { \frac{1}{e_2}  \frac{\partial \left( u^2+v^2+w^2 \right)}{\partial j}}  \hfill    \\
557       \end{array} }} \right)       &
558\end{flalign*}
559\begin{flalign*}
560& \qquad =\left( {{  \begin{array}{*{20}c}
561 {-\zeta \; v} \hfill \\
562 { \zeta \; u} \hfill \\
563         \end{array} }} \right)
564+\frac{1}{2}\left( {{   \begin{array}{*{20}c}
565 {\frac{1}{e_1 }\frac{\partial \left( {u^2+v^2} \right)}{\partial i}} \hfill  \\
566 {\frac{1}{e_2 }\frac{\partial \left( {u^2+v^2} \right)}{\partial j}} \hfill  \\
567                  \end{array} }} \right)       
568+\frac{1}{e_3 }\left( {{      \begin{array}{*{20}c}
569 { w \; \frac{\partial u}{\partial k}}    \\
570 { w \; \frac{\partial v}{\partial k}}    \\
571                     \end{array} }} \right
572-\left( {{  \begin{array}{*{20}c}
573 {\frac{w}{e_1}\frac{\partial w}{\partial i}
574 -\frac{1}{2e_1}\frac{\partial w^2}{\partial i}} \hfill \\
575 {\frac{w}{e_2}\frac{\partial w}{\partial j}
576  -\frac{1}{2e_2}\frac{\partial w^2}{\partial j}} \hfill \\
577         \end{array} }} \right)        &
578\end{flalign*}
579
580The last term of the right hand side is obviously zero, and thus the nonlinear term of
581\eqref{Eq_PE_dyn} is written in the $(i,j,k)$ coordinate system:
582\begin{equation} \label{Eq_PE_vector_form}
583\left[ {\left( {  \nabla \times {\rm {\bf U}}    } \right) \times {\rm {\bf U}}
584+\frac{1}{2}   \nabla \left( {{\rm {\bf U}}^2} \right)}   \right]_h
585=\zeta 
586\;{\rm {\bf k}}\times {\rm {\bf U}}_h +\frac{1}{2}\nabla _h \left( {{\rm 
587{\bf U}}_h^2 } \right)+\frac{1}{e_3 }w\frac{\partial {\rm {\bf U}}_h
588}{\partial k}     
589\end{equation}
590
591This is the so-called \textit{vector invariant form} of the momentum advection term.
592For some purposes, it can be advantageous to write this term in the so-called flux form,
593$i.e.$ to write it as the divergence of fluxes. For example, the first component of
594\eqref{Eq_PE_vector_form} (the $i$-component) is transformed as follows:
595\begin{flalign*}
596&{ \begin{array}{*{20}l}
597\left[ {\left( {\nabla \times \vect{U}} \right)\times \vect{U}
598          +\frac{1}{2}\nabla \left( {\vect{U}}^2 \right)} \right]_i   % \\
599%\\
600     = - \zeta \;v
601     + \frac{1}{2\;e_1 } \frac{\partial \left( {u^2+v^2} \right)}{\partial i}
602     + \frac{1}{e_3}w \ \frac{\partial u}{\partial k}          \\
603\\
604\qquad =\frac{1}{e_1 \; e_2} \left(    -v\frac{\partial \left( {e_2 \,v} \right)}{\partial i}
605                     +v\frac{\partial \left( {e_1 \,u} \right)}{\partial j}    \right)
606+\frac{1}{e_1 e_2 }\left(  +e_2 \; u\frac{\partial u}{\partial i}
607                     +e_2 \; v\frac{\partial v}{\partial i}              \right)
608+\frac{1}{e_3}       \left(   w\;\frac{\partial u}{\partial k}       \right)   \\
609\end{array} }        &
610\end{flalign*}
611\begin{flalign*}
612&{ \begin{array}{*{20}l}
613\qquad =\frac{1}{e_1 \; e_2}  \left\{ 
614 -\left(        v^\frac{\partial e_2                                }{\partial i} 
615      +e_2 \,v    \frac{\partial v                                   }{\partial i}     \right)
616+\left(           \frac{\partial \left( {e_1 \,u\,v}  \right)}{\partial j}
617      -e_1 \,u    \frac{\partial v                                   }{\partial j}  \right\right.
618\\  \left\qquad \qquad \quad
619+\left(           \frac{\partial \left( {e_2 u\,u}     \right)}{\partial i}
620      -u       \frac{\partial \left( {e_2 u}         \right)}{\partial i}  \right)
621+e_2 v            \frac{\partial v                                    }{\partial i}
622                  \right\} 
623+\frac{1}{e_3} \left(
624               \frac{\partial \left( {w\,u} \right)         }{\partial k}
625       -u         \frac{\partial w                    }{\partial k}  \right) \\
626\end{array} }     &
627\end{flalign*}
628\begin{flalign*}
629&{ \begin{array}{*{20}l}
630\qquad =\frac{1}{e_1 \; e_2}  \left(
631               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
632      +        \frac{\partial \left( {e_1 \,u\,v} \right)}{\partial j}  \right)
633+\frac{1}{e_3 }      \frac{\partial \left( {w\,u       } \right)}{\partial k}
634\\  \qquad \qquad \quad
635+\frac{1}{e_1 e_2 }     \left(
636      -u \left(   \frac{\partial \left( {e_1 v   } \right)}{\partial j}
637               -v\,\frac{\partial e_1 }{\partial j}             \right)
638      -u       \frac{\partial \left( {e_2 u   } \right)}{\partial i}
639                  \right)
640 -\frac{1}{e_3 }     \frac{\partial w}{\partial k} u
641 +\frac{1}{e_1 e_2 }\left(    -v^2\frac{\partial e_2   }{\partial i}     \right)
642\end{array} }     &
643\end{flalign*}
644\begin{flalign*}
645&{ \begin{array}{*{20}l}
646\qquad = \nabla \cdot \left( {{\rm {\bf U}}\,u} \right)
647-   \left( \nabla \cdot {\rm {\bf U}} \right) \ u
648+\frac{1}{e_1 e_2 }\left(
649      -v^2     \frac{\partial e_2 }{\partial i}
650      +uv   \,    \frac{\partial e_1 }{\partial j}    \right) \\
651\end{array} }     &
652\end{flalign*}
653as $\nabla \cdot {\rm {\bf U}}\;=0$ (incompressibility) it comes:
654\begin{flalign*}
655&{ \begin{array}{*{20}l}
656\qquad = \nabla \cdot \left{{\rm {\bf U}}\,u}      \right)
657\frac{1}{e_1 e_2 }   \left( v \; \frac{\partial e_2}{\partial i}
658                         -u \; \frac{\partial e_1}{\partial j}    \right\left( -v \right)
659\end{array} }     &
660\end{flalign*}
661
662The flux form of the momentum advection term is therefore given by:
663\begin{multline} \label{Eq_PE_flux_form}
664      \left[
665  \left(    {\nabla \times {\rm {\bf U}}}    \right) \times {\rm {\bf U}}
666+\frac{1}{2}   \nabla \left{{\rm {\bf U}}^2}    \right)
667      \right]_h
668\\
669= \nabla \cdot    \left( {{\begin{array}{*{20}c}   {\rm {\bf U}} \, u   \hfill \\
670                                    {\rm {\bf U}} \, v   \hfill \\
671                  \end{array} }}   
672            \right)
673+\frac{1}{e_1 e_2 }     \left(
674       v\frac{\partial e_2}{\partial i}
675      -u\frac{\partial e_1}{\partial j} 
676                  \right) {\rm {\bf k}} \times {\rm {\bf U}}_h
677\end{multline}
678
679The flux form has two terms, the first one is expressed as the divergence of momentum
680fluxes (hence the flux form name given to this formulation) and the second one is due to
681the curvilinear nature of the coordinate system used. The latter is called the \emph{metric} 
682term and can be viewed as a modification of the Coriolis parameter:
683\begin{equation} \label{Eq_PE_cor+metric}
684f \to f + \frac{1}{e_1 \; e_2}   \left(    v \frac{\partial e_2}{\partial i}
685                              -u \frac{\partial e_1}{\partial j}  \right)
686\end{equation}
687
688Note that in the case of geographical coordinate, $i.e.$ when $(i,j) \to (\lambda ,\varphi )$ 
689and $(e_1 ,e_2) \to (a \,\cos \varphi ,a)$, we recover the commonly used modification of
690the Coriolis parameter $f \to f+(u/a) \tan \varphi$.
691
692To sum up, the equations solved by the ocean model can be written in the following tensorial formalism:
693
694\vspace{+10pt}
695$\bullet$ \textit{momentum equations} :
696
697vector invariant form :
698\begin{subequations} \label{Eq_PE_dyn_vect}
699\begin{multline} \label{Eq_PE_dyn_vect_u}
700\frac{\partial u}{\partial t}=
701   +   \left( {\zeta +f} \right)\,v                                   
702   -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
703   -   \frac{1}{e_3} w \frac{\partial u}{\partial k}       \\
704   -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s+p_h }{\rho _o}    \right)   
705+   D_u^{\vect{U}}  +   F_u^{\vect{U}}
706\end{multline}
707\begin{multline} \label{Eq_PE_dyn_vect_v}
708\frac{\partial v}{\partial t}=
709   -   \left( {\zeta +f} \right)\,u   
710   -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^\right)        -   \frac{1}{e_3 }w\frac{\partial v}{\partial k}         \\
711   -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)   
712+  D_v^{\vect{U}}  +   F_v^{\vect{U}}
713\end{multline}
714\end{subequations}
715
716flux form:
717\begin{subequations} \label{Eq_PE_dyn_flux}
718\begin{multline} \label{Eq_PE_dyn_flux_u}
719\frac{\partial u}{\partial t}=
720+   \left( { f + \frac{1}{e_1 \; e_2}
721               \left(    v \frac{\partial e_2}{\partial i}
722                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, v    \\
723- \frac{1}{e_1 \; e_2}  \left(
724               \frac{\partial \left( {e_2 \,u\,u} \right)}{\partial i}
725      +        \frac{\partial \left( {e_1 \,v\,u} \right)}{\partial j}  \right)
726                 - \frac{1}{e_3 }\frac{\partial \left( {         w\,u} \right)}{\partial k}    \\
727-   \frac{1}{e_1 }\frac{\partial}{\partial i}\left( \frac{p_s+p_h }{\rho _o}   \right)
728+   D_u^{\vect{U}} +   F_u^{\vect{U}}
729\end{multline}
730\begin{multline} \label{Eq_PE_dyn_flux_v}
731\frac{\partial v}{\partial t}=
732-   \left( { f + \frac{1}{e_1 \; e_2}
733               \left(    v \frac{\partial e_2}{\partial i}
734                  -u \frac{\partial e_1}{\partial j}  \right)}    \right) \, u   \\
735 \frac{1}{e_1 \; e_2}   \left(
736               \frac{\partial \left( {e_2 \,u\,v} \right)}{\partial i}
737      +        \frac{\partial \left( {e_1 \,v\,v} \right)}{\partial j}  \right)
738                 - \frac{1}{e_3 } \frac{\partial \left( {        w\,v} \right)}{\partial k}    \\
739-   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}    \right)
740+  D_v^{\vect{U}} +  F_v^{\vect{U}} 
741\end{multline}
742\end{subequations}
743where $\zeta$ is given by \eqref{Eq_PE_curl_Uh} and the surface pressure gradient formulation
744depends on the one of the free surface:
745
746$*$ free surface formulation
747\begin{equation}\label{Eq_PE_dyn_sco}
748\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
749 {\frac{g}{\;e_1 }\frac{\partial \eta }{\partial i}} \hfill \\
750 {\frac{g}{\;e_2 }\frac{\partial \eta }{\partial j}} \hfill \\
751\end{array} }} \right)
752\qquad \text{where $\eta$ is solution of \eqref{Eq_PE_ssh} }
753\end{equation}
754
755$*$ rigid-lid approximation
756\begin{equation}\label{Eq_PE_dyn_zco}
757\frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c}
758 {\overline M _u +\frac{1}{H\;e_2 }\frac{\partial }{\partial j}\left(
759{\frac{\partial \psi }{\partial t}} \right)}     \\
760 {\overline M _v -\frac{1}{H\;e_1 }\frac{\partial }{\partial i}\left(
761{\frac{\partial \psi }{\partial t}} \right)}        \\
762\end{array} }} \right)
763\end{equation}
764where ${\vect{M}}= \left( M_u,M_v \right)$ represents the collected contributions of nonlinear,
765viscosity and hydrostatic pressure gradient terms in \eqref{Eq_PE_dyn_vect} and the overbar
766indicates a vertical average over the whole water column ($i.e.$ from $z=-H$, the ocean bottom,
767to $z=0$, the rigid-lid), and where the time derivative of $\psi$ is the solution of an elliptic equation:
768\begin{multline} \label{Eq_psi_total}
769  \frac{\partial }{\partial i}\left[ {\frac{e_2 }{H\,e_1}\frac{\partial}{\partial i}
770                         \left( {\frac{\partial \psi }{\partial t}} \right)}   \right]
771+\frac{\partial }{\partial j}\left[ {\frac{e_1 }{H\,e_2}\frac{\partial }{\partial j}
772                         \left( {\frac{\partial \psi }{\partial t}} \right)} \right]
773= \\
774  \frac{\partial }{\partial i}\left( {e_2 \overline M _v } \right)
775- \frac{\partial }{\partial j}\left( {e_1 \overline M _u } \right)
776\end{multline}
777
778The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
779\begin{equation} \label{Eq_w_diag}
780\frac{\partial w}{\partial k}=-\chi \;e_3
781\end{equation}
782\begin{equation} \label{Eq_hp_diag}
783\frac{\partial p_h }{\partial k}=-\rho \;g\;e_3
784\end{equation}
785
786where the divergence of the horizontal velocity, $\chi$ is given by \eqref{Eq_PE_div_Uh}.
787
788\vspace{+10pt}
789$\bullet$ \textit{tracer equations} :
790\begin{equation} \label{Eq_S}
791\frac{\partial T}{\partial t} =
792-\frac{1}{e_1 e_2 }\left[ {      \frac{\partial \left( {e_2 T\,u} \right)}{\partial i}
793                  +\frac{\partial \left( {e_1 T\,v} \right)}{\partial j}} \right]
794-\frac{1}{e_3 }\frac{\partial \left( {T\,w} \right)}{\partial k} + D^T + F^T
795\end{equation}
796\begin{equation} \label{Eq_T}
797\frac{\partial S}{\partial t} =
798-\frac{1}{e_1 e_2 }\left[    {\frac{\partial \left( {e_2 S\,u} \right)}{\partial i}
799                  +\frac{\partial \left( {e_1 S\,v} \right)}{\partial j}} \right]
800-\frac{1}{e_3 }\frac{\partial \left( {S\,w} \right)}{\partial k} + D^S + F^S
801\end{equation}
802\begin{equation} \label{Eq_rho}
803\rho =\rho \left( {T,S,z(k)} \right)
804\end{equation}
805
806The expression of \textbf{D}$^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale
807parameterisation used. It will be defined in \S\ref{PE_zdf}. The nature and formulation of
808${\rm {\bf F}}^{\rm {\bf U}}$, $F^T$ and $F^S$, the surface forcing terms, are discussed
809in Chapter~\ref{SBC}.
810
811\newpage 
812% ================================================================
813% Curvilinear z*-coordinate System
814% ================================================================
815\section{Curvilinear \textit{z*}-coordinate System}
816
817%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
818\begin{figure}[!b] \label{Fig_z_zstar}  \begin{center}
819\includegraphics[width=1.0\textwidth]{./TexFiles/Figures/Fig_z_zstar.pdf}
820\caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear
821free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate
822\citep{Adcroft_Campin_OM04} ).}
823\end{center}   \end{figure}
824%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
825
826
827In that case, the free surface equation is nonlinear, and the variations of volume are fully
828taken into account. These coordinates systems is presented in a report \citep{Levier2007} 
829available on the \NEMO web site.
830
831\gmcomment{
832The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation
833which allows one to deal with large amplitude free-surface
834variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In
835the  \textit{z*} formulation, the variation of the column thickness due to sea-surface
836undulations is not concentrated in the surface level, as in the $z$-coordinate formulation,
837but is equally distributed over the full water column. Thus vertical
838levels naturally follow sea-surface variations, with a linear attenuation with
839depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in
840fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent.
841The definition and modified oceanic equations for the rescaled vertical coordinate
842 \textit{z*}, including the treatment of fresh-water flux at the surface, are
843detailed in Adcroft and Campin (2004). The major points are summarized
844here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as:
845
846$H +  \textit{z*} = (H + z) / r$ and  $\delta \textit{z*} = \delta z / r$ with $r = \frac{H+\eta} {H}$
847
848Since the vertical displacement of the free surface is incorporated in the vertical
849coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position, 
850$\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field
851is no longer zero as shown by the continuity equation:
852
853$\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right)
854      \left( r \; w\textit{*} \right) = 0 $
855
856}
857
858
859\newpage 
860% ================================================================
861% Curvilinear s-coordinate System
862% ================================================================
863\section{Curvilinear \textit{s}-coordinate System}
864\label{PE_sco}
865
866% -------------------------------------------------------------------------------------------------------------
867% Introduction
868% -------------------------------------------------------------------------------------------------------------
869\subsection{Introduction}
870
871Several important aspects of the ocean circulation are influenced by bottom topography.
872Of course, the most important is that bottom topography determines deep ocean sub-basins,
873barriers, sills and channels that strongly constrain the path of water masses, but more subtle
874effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary
875one along continental slopes. Topographic Rossby waves can be excited and can interact
876with the mean current. In the $z-$coordinate system presented in the previous section
877(\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is
878discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom
879and to large localized depth gradients associated with large localized vertical velocities.
880The response to such a velocity field often leads to numerical dispersion effects.
881One solution to strongly reduce this error is to use a partial step representation of bottom
882topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.
883Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)
884
885The $s$-coordinate avoids the discretisation error in the depth field since the layers of
886computation are gradually adjusted with depth to the ocean bottom. Relatively small
887topographic features as well as  gentle, large-scale slopes of the sea floor in the deep
888ocean, which would be ignored in typical $z$-model applications with the largest grid
889spacing at greatest depths, can easily be represented (with relatively low vertical resolution).
890A terrain-following model (hereafter $s-$model) also facilitates the modelling of the
891boundary layer flows over a large depth range, which in the framework of the $z$-model
892would require high vertical resolution over the whole depth range. Moreover, with a
893$s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface
894as the only boundaries of the domain (nomore lateral boundary condition to specify).
895Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a
896homogeneous ocean, it has strong limitations as soon as stratification is introduced.
897The main two problems come from the truncation error in the horizontal pressure
898gradient and a possibly increased diapycnal diffusion. The horizontal pressure force
899in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}),
900
901\begin{equation} \label{Eq_PE_p_sco}
902\left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial 
903p}{\partial s}\left. {\nabla z} \right|_s
904\end{equation}
905
906The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface
907and introduces a truncation error that is not present in a $z$-model. In the special case
908of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),
909\citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude
910of this truncation error. It depends on topographic slope, stratification, horizontal and
911vertical resolution, the equation of state, and the finite difference scheme. This error
912limits the possible topographic slopes that a model can handle at a given horizontal
913and vertical resolution. This is a severe restriction for large-scale applications using
914realistic bottom topography. The large-scale slopes require high horizontal resolution,
915and the computational cost becomes prohibitive. This problem can be at least partially
916overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec1996}. However, the definition of the model
917domain vertical coordinate becomes then a non-trivial thing for a realistic bottom
918topography: a envelope topography is defined in $s$-coordinate on which a full or
919partial step bottom topography is then applied in order to adjust the model depth to
920the observed one (see \S\ref{DOM_zgr}.
921
922For numerical reasons a minimum of diffusion is required along the coordinate surfaces
923of any finite difference model. It causes spurious diapycnal mixing when coordinate
924surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as
925well as for a $s$-model. However, density varies more strongly on $s-$surfaces than
926on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal
927diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a
928$z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal
929circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.
930For example, imagine an isolated bump of topography in an ocean at rest with a horizontally
931uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral
932surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,
933the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column
934($i.e.$ the main thermocline) \citep{Madec1996}. An alternate solution consists of rotating
935the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}.
936Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,
937strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).
938
939The $s-$coordinates introduced here \citep{Lott1990,Madec1996} differ mainly in two
940aspects from similar models:  it allows  a representation of bottom topography with mixed
941full or partial step-like/terrain following topography ; It also offers a completely general
942transformation, $s=s(i,j,z)$ for the vertical coordinate.
943
944% -------------------------------------------------------------------------------------------------------------
945% The s-coordinate Formulation
946% -------------------------------------------------------------------------------------------------------------
947\subsection{The \textit{s-}coordinate Formulation}
948
949Starting from the set of equations established in \S\ref{PE_zco} for the special case $k=z$ 
950and thus $e_3=1$, we introduce an arbitrary vertical coordinate $s=s(i,j,k,t)$, which includes
951$z$-, \textit{z*}- and $\sigma-$coordinates as special cases ($s=z$, $s=\textit{z*}$, and
952$s=\sigma=z/H$ or $=z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed
953equations is given in Appendix~\ref{Apdx_A}. Let us define the vertical scale factor by
954$e_3=\partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ), and the slopes in the
955(\textbf{i},\textbf{j}) directions between $s-$ and $z-$surfaces by :
956\begin{equation} \label{Eq_PE_sco_slope}
957\sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right|_s
958\quad \text{, and } \quad 
959\sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right|_s
960\end{equation}
961We also introduce  $\omega $, a dia-surface velocity component, defined as the velocity
962relative to the moving $s$-surfaces and normal to them:
963\begin{equation} \label{Eq_PE_sco_w}
964\omega  = w - e_3 \, \frac{\partial s}{\partial t} - \sigma _1 \,u - \sigma _2 \,v    \\
965\end{equation}
966
967The equations solved by the ocean model \eqref{Eq_PE} in $s-$coordinate can be written as follows:
968
969 \vspace{0.5cm}
970* momentum equation:
971\begin{multline} \label{Eq_PE_sco_u}
972\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
973   +   \left( {\zeta +f} \right)\,v                                   
974   -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)
975   -   \frac{1}{e_3} \omega \frac{\partial u}{\partial k}       \\
976   -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s + p_h}{\rho _o}    \right)   
977   +  g\frac{\rho }{\rho _o}\sigma _1
978   +   D_u^{\vect{U}}  +   F_u^{\vect{U}} \quad
979\end{multline}
980\begin{multline} \label{Eq_PE_sco_v}
981\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
982   -   \left( {\zeta +f} \right)\,u   
983   -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^\right)       
984   -   \frac{1}{e_3 } \omega \frac{\partial v}{\partial k}         \\
985   -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)
986    +  g\frac{\rho }{\rho _o }\sigma _2   
987   +  D_v^{\vect{U}}  +   F_v^{\vect{U}} \quad
988\end{multline}
989where the relative vorticity, \textit{$\zeta $}, the surface pressure gradient, and the hydrostatic
990pressure have the same expressions as in $z$-coordinates although they do not represent
991exactly the same quantities. $\omega$ is provided by the continuity equation
992(see Appendix~\ref{Apdx_A}):
993
994\begin{equation} \label{Eq_PE_sco_continuity}
995\frac{\partial e_3}{\partial t} + e_3 \; \chi + \frac{\partial \omega }{\partial s} = 0   
996\qquad \text{with }\;\; 
997\chi =\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3 \,u} 
998\right)}{\partial i}+\frac{\partial \left( {e_1 e_3 \,v} \right)}{\partial 
999j}} \right]
1000\end{equation}
1001
1002 \vspace{0.5cm}
1003* tracer equations:
1004\begin{multline} \label{Eq_PE_sco_t}
1005\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
1006-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,T} \right)}{\partial i}
1007                                           +\frac{\partial \left( {e_1 e_3\,v\,T} \right)}{\partial j}} \right]   \\
1008-\frac{1}{e_3 }\frac{\partial \left( {T\,\omega } \right)}{\partial k}   + D^T + F^S   \qquad
1009\end{multline}
1010
1011\begin{multline} \label{Eq_PE_sco_s}
1012\frac{1}{e_3} \frac{\partial \left(  e_3\,\right) }{\partial t}=
1013-\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e_2 e_3\,u\,S} \right)}{\partial i}
1014                                           +\frac{\partial \left( {e_1 e_3\,v\,S} \right)}{\partial j}} \right]    \\
1015-\frac{1}{e_3 }\frac{\partial \left( {S\,\omega } \right)}{\partial k}     + D^S + F^S   \qquad
1016\end{multline}
1017
1018The equation of state has the same expression as in $z$-coordinate, and similar expressions
1019are used for mixing and forcing terms.
1020
1021\gmcomment{
1022\colorbox{yellow}{ to be updated $= = >$}
1023Add a few works on z and zps and s and underlies the differences between all of them
1024\colorbox{yellow}{ $< = =$ end update}  }
1025
1026\newpage 
1027% ================================================================
1028% Subgrid Scale Physics
1029% ================================================================
1030\section{Subgrid Scale Physics}
1031\label{PE_zdf_ldf}
1032
1033The primitive equations describe the behaviour of a geophysical fluid at
1034space and time scales larger than a few kilometres in the horizontal, a few
1035meters in the vertical and a few minutes. They are usually solved at larger
1036scales: the specified grid spacing and time step of the numerical model. The
1037effects of smaller scale motions (coming from the advective terms in the
1038Navier-Stokes equations) must be represented entirely in terms of
1039large-scale patterns to close the equations. These effects appear in the
1040equations as the divergence of turbulent fluxes ($i.e.$ fluxes associated with
1041the mean correlation of small scale perturbations). Assuming a turbulent
1042closure hypothesis is equivalent to choose a formulation for these fluxes.
1043It is usually called the subgrid scale physics. It must be emphasized that
1044this is the weakest part of the primitive equations, but also one of the
1045most important for long-term simulations as small scale processes \textit{in fine} 
1046balance the surface input of kinetic energy and heat.
1047
1048The control exerted by gravity on the flow induces a strong anisotropy
1049between the lateral and vertical motions. Therefore subgrid-scale physics 
1050\textbf{D}$^{\vect{U}}$, $D^{S}$ and $D^{T}$  in \eqref{Eq_PE_dyn},
1051\eqref{Eq_PE_tra_T} and \eqref{Eq_PE_tra_S} are divided into a lateral part 
1052\textbf{D}$^{l \vect{U}}$, $D^{lS}$ and $D^{lT}$ and a vertical part 
1053\textbf{D}$^{vU}$, $D^{vS}$ and $D^{vT}$. The formulation of these terms
1054and their underlying physics are briefly discussed in the next two subsections.
1055
1056% -------------------------------------------------------------------------------------------------------------
1057% Vertical Subgrid Scale Physics
1058% -------------------------------------------------------------------------------------------------------------
1059\subsection{Vertical Subgrid Scale Physics}
1060\label{PE_zdf}
1061
1062The model resolution is always larger than the scale at which the major
1063sources of vertical turbulence occur (shear instability, internal wave
1064breaking...). Turbulent motions are thus never explicitly solved, even
1065partially, but always parameterized. The vertical turbulent fluxes are
1066assumed to depend linearly on the gradients of large-scale quantities (for
1067example, the turbulent heat flux is given by $\overline{T'w'}=-A^{vT} \partial_z \overline T$,
1068where $A^{vT}$ is an eddy coefficient). This formulation is
1069analogous to that of molecular diffusion and dissipation. This is quite
1070clearly a necessary compromise: considering only the molecular viscosity
1071acting on large scale severely underestimates the role of turbulent
1072diffusion and dissipation, while an accurate consideration of the details of
1073turbulent motions is simply impractical. The resulting vertical momentum and
1074tracer diffusive operators are of second order:
1075\begin{equation} \label{Eq_PE_zdf}
1076   \begin{split}
1077{\vect{D}}^{v \vect{U}} &=\frac{\partial }{\partial z}\left( {A^{vm}\frac{\partial {\vect{U}}_h }{\partial z}} \right) \ , \\         
1078D^{vT}                         &= \frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial T}{\partial z}} \right) \ ,
1079\quad
1080D^{vS}=\frac{\partial }{\partial z}\left( {A^{vT}\frac{\partial S}{\partial z}} \right)
1081   \end{split}
1082\end{equation}
1083where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients,
1084respectively. At the sea surface and at the bottom, turbulent fluxes of momentum, heat
1085and salt must be specified (see Chap.~\ref{SBC} and \ref{ZDF} and \S\ref{TRA_bbl}).
1086All the vertical physics is embedded in the specification of the eddy coefficients.
1087They can be assumed to be either constant, or function of the local fluid properties
1088($e.g.$ Richardson number, Brunt-Vais\"{a}l\"{a} frequency...), or computed from a
1089turbulent closure model. The choices available in \NEMO are discussed in \S\ref{ZDF}).
1090
1091% -------------------------------------------------------------------------------------------------------------
1092% Lateral Diffusive and Viscous Operators Formulation
1093% -------------------------------------------------------------------------------------------------------------
1094\subsection{Lateral Diffusive and Viscous Operators Formulation}
1095\label{PE_ldf}
1096
1097Lateral turbulence can be roughly divided into a mesoscale turbulence
1098associated with eddies (which can be solved explicitly if the resolution is
1099sufficient since their underlying physics are included in the primitive
1100equations), and a sub mesoscale turbulence which is never explicitly solved
1101even partially, but always parameterized. The formulation of lateral eddy
1102fluxes depends on whether the mesoscale is below or above the grid-spacing
1103($i.e.$ the model is eddy-resolving or not).
1104
1105In non-eddy-resolving configurations, the closure is similar to that used
1106for the vertical physics. The lateral turbulent fluxes are assumed to depend
1107linearly on the lateral gradients of large-scale quantities. The resulting
1108lateral diffusive and dissipative operators are of second order.
1109Observations show that lateral mixing induced by mesoscale turbulence tends
1110to be along isopycnal surfaces (or more precisely neutral surfaces \cite{McDougall1987})
1111rather than across them.
1112As the slope of neutral surfaces is small in the ocean, a common
1113approximation is to assume that the `lateral' direction is the horizontal,
1114$i.e.$ the lateral mixing is performed along geopotential surfaces. This leads
1115to a geopotential second order operator for lateral subgrid scale physics.
1116This assumption can be relaxed: the eddy-induced turbulent fluxes can be
1117better approached by assuming that they depend linearly on the gradients of
1118large-scale quantities computed along neutral surfaces. In such a case,
1119the diffusive operator is an isoneutral second order operator and it has
1120components in the three space directions. However, both horizontal and
1121isoneutral operators have no effect on mean ($i.e.$ large scale) potential
1122energy whereas potential energy is a main source of turbulence (through
1123baroclinic instabilities). \citet{Gent1990} have proposed a
1124parameterisation of mesoscale eddy-induced turbulence which associates an
1125eddy-induced velocity to the isoneutral diffusion. Its mean effect is to
1126reduce the mean potential energy of the ocean. This leads to a formulation
1127of lateral subgrid-scale physics made up of an isoneutral second order
1128operator and an eddy induced advective part. In all these lateral diffusive
1129formulations, the specification of the lateral eddy coefficients remains the
1130problematic point as there is no really satisfactory formulation of these
1131coefficients as a function of large-scale features.
1132
1133In eddy-resolving configurations, a second order operator can be used, but
1134usually a more scale selective one (biharmonic operator) is preferred as the
1135grid-spacing is usually not small enough compared to the scale of the
1136eddies. The role devoted to the subgrid-scale physics is to dissipate the
1137energy that cascades toward the grid scale and thus ensures the stability of
1138the model while not interfering with the solved mesoscale activity. Another approach
1139is becoming more and more popular: instead of specifying explicitly a sub-grid scale
1140term in the momentum and tracer time evolution equations, one uses a advective
1141scheme which is diffusive enough to maintain the model stability. It must be emphasised
1142that then, all the sub-grid scale physics is in this case include in the formulation of the
1143advection scheme.
1144
1145All these parameterisations of subgrid scale physics present advantages and
1146drawbacks. There are not all available in \NEMO. In the $z$-coordinate
1147formulation, five options are offered for active tracers (temperature and
1148salinity): second order geopotential operator, second order isoneutral
1149operator, \citet{Gent1990} parameterisation, fourth order
1150geopotential operator, and various slightly diffusive advection schemes.
1151The same options are available for momentum, except
1152\citet{Gent1990} parameterisation which only involves tracers. In the
1153$s$-coordinate formulation, additional options are offered for tracers: second
1154order operator acting along $s-$surfaces, and for momentum: fourth order
1155operator acting along $s-$surfaces (see \S\ref{LDF}).
1156
1157\subsubsection{lateral second order tracer diffusive operator}
1158
1159The lateral second order tracer diffusive operator is defined by (see Appendix~\ref{Apdx_B}):
1160\begin{equation} \label{Eq_PE_iso_tensor}
1161D^{lT}=\nabla {\rm {\bf .}}\left( {A^{lT}\;\Re \;\nabla T} \right) \qquad 
1162\mbox{with}\quad \;\;\Re =\left( {{\begin{array}{*{20}c}
1163 1 \hfill & 0 \hfill & {-r_1 } \hfill \\
1164 0 \hfill & 1 \hfill & {-r_2 } \hfill \\
1165 {-r_1 } \hfill & {-r_2 } \hfill & {r_1 ^2+r_2 ^2} \hfill \\
1166\end{array} }} \right)
1167\end{equation}
1168where $r_1 \;\mbox{and}\;r_2 $ are the slopes between the surface along
1169which the diffusive operator acts and the model level ($e. g.$ $z$- or
1170$s$-surfaces). Note that the formulation \eqref{Eq_PE_iso_tensor} is exact for the
1171rotation between geopotential and $s$-surfaces, while it is only an approximation
1172for the rotation between isoneutral and $z$- or $s$-surfaces. Indeed, in the latter
1173case, two assumptions are made to simplify  \eqref{Eq_PE_iso_tensor} \citep{Cox1987}.
1174First, the horizontal contribution of the dianeutral mixing is neglected since the ratio
1175between iso and dia-neutral diffusive coefficients is known to be several orders of
1176magnitude smaller than unity. Second, the two isoneutral directions of diffusion are
1177assumed to be independent since the slopes are generally less than $10^{-2}$ in the
1178ocean (see Appendix~\ref{Apdx_B}).
1179
1180For \textit{geopotential} diffusion, $r_1$ and $r_2 $ are the slopes between the
1181geopotential and computational surfaces: in $z$-coordinates they are zero
1182($r_1 = r_2 = 0$) while in $s$-coordinate (including $\textit{z*}$ case) they are
1183equal to $\sigma _1$ and $\sigma _2$, respectively (see \eqref{Eq_PE_sco_slope} ).
1184
1185For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral
1186and computational surfaces. Therefore, they have a same expression in $z$- and $s$-coordinates:
1187\begin{equation} \label{Eq_PE_iso_slopes}
1188r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
1189                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1} \ , \quad
1190r_1 =\frac{e_3 }{e_1 }  \left( {\frac{\partial \rho }{\partial i}} \right)
1191                  \left( {\frac{\partial \rho }{\partial k}} \right)^{-1}
1192\end{equation}
1193
1194When the \textit{Eddy Induced Velocity} parametrisation (eiv) \citep{Gent1990} is used,
1195an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
1196\begin{equation} \label{Eq_PE_iso+eiv}
1197D^{lT}=\nabla \cdot \left( {A^{lT}\;\Re \;\nabla T} \right)
1198           +\nabla \cdot \left( {{\vect{U}}^\ast \,T} \right)
1199\end{equation}
1200where ${\vect{U}}^\ast =\left( {u^\ast ,v^\ast ,w^\ast } \right)$ is a non-divergent,
1201eddy-induced transport velocity. This velocity field is defined by:
1202\begin{equation} \label{Eq_PE_eiv}
1203   \begin{split}
1204 u^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_1 } \right] \\ 
1205 v^\ast  &= +\frac{1}{e_3       }\frac{\partial }{\partial k}\left[ {A^{eiv}\;\tilde{r}_2 } \right] \\ 
1206 w^\ast &=  -\frac{1}{e_1 e_2 }\left[
1207                      \frac{\partial }{\partial i}\left( {A^{eiv}\;e_2\,\tilde{r}_1 } \right)
1208                    +\frac{\partial }{\partial j}\left( {A^{eiv}\;e_1\,\tilde{r}_2 } \right)      \right]
1209   \end{split}
1210\end{equation}
1211where $A^{eiv}$ is the eddy induced velocity coefficient (or equivalently the isoneutral
1212thickness diffusivity coefficient), and $\tilde{r}_1$ and $\tilde{r}_2$ are the slopes
1213between isoneutral and \emph{geopotential} surfaces and thus depends on the coordinate
1214considered:
1215\begin{align} \label{Eq_PE_slopes_eiv}
1216\tilde{r}_n = \begin{cases}
1217   r_n                  &      \text{in $z$-coordinate}    \\
1218   r_n + \sigma_n &      \text{in \textit{z*} and $s$-coordinates} 
1219                   \end{cases}
1220\quad \text{where } n=1,2
1221\end{align}
1222
1223The normal component of the eddy induced velocity is zero at all the boundaries.
1224This can be achieved in a model by tapering either the eddy coefficient or the slopes
1225to zero in the vicinity of the boundaries. The latter strategy is used in \NEMO (cf. Chap.~\ref{LDF}).
1226
1227\subsubsection{lateral fourth order tracer diffusive operator}
1228
1229The lateral fourth order tracer diffusive operator is defined by:
1230\begin{equation} \label{Eq_PE_bilapT}
1231D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
1232\qquad \text{where} \  D^{lT}=\Delta \left( {A^{lT}\;\Delta T} \right)
1233 \end{equation}
1234
1235It is the second order operator given by \eqref{Eq_PE_iso_tensor} applied twice with
1236the eddy diffusion coefficient correctly placed.
1237
1238
1239\subsubsection{lateral second order momentum diffusive operator}
1240
1241The second order momentum diffusive operator along $z$- or $s$-surfaces is found by
1242applying \eqref{Eq_PE_lap_vector} to the horizontal velocity vector (see Appendix~\ref{Apdx_B}):
1243\begin{equation} \label{Eq_PE_lapU}
1244\begin{split}
1245{\rm {\bf D}}^{l{\rm {\bf U}}} 
1246&= \quad \  \nabla _h \left( {A^{lm}\chi } \right)
1247   \ - \ \nabla _h \times \left( {A^{lm}\,\zeta \;{\rm {\bf k}}} \right)     \\
1248&=   \left(      \begin{aligned}
1249             \frac{1}{e_1      } \frac{\partial \left( A^{lm} \chi          \right)}{\partial i} 
1250         &-\frac{1}{e_2 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial j}  \\
1251             \frac{1}{e_2      }\frac{\partial \left( {A^{lm} \chi         } \right)}{\partial j}   
1252         &+\frac{1}{e_1 e_3}\frac{\partial \left( {A^{lm} \;e_3 \zeta} \right)}{\partial i}
1253        \end{aligned}    \right)
1254\end{split}
1255\end{equation}
1256
1257Such a formulation ensures a complete separation between the vorticity and
1258horizontal divergence fields (see Appendix~\ref{Apdx_C}). Unfortunately, it is not
1259available for geopotential diffusion in $s-$coordinates and for isoneutral
1260diffusion in both $z$- and $s$-coordinates ($i.e.$ when a rotation is required).
1261In these two cases, the $u$ and $v-$fields are considered as independent scalar
1262fields, so that the diffusive operator is given by:
1263\begin{equation} \label{Eq_PE_lapU_iso}
1264\begin{split}
1265 D_u^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla u} \right) \\ 
1266 D_v^{l{\rm {\bf U}}} &= \nabla .\left( {\Re \;\nabla v} \right)
1267 \end{split}
1268 \end{equation}
1269where $\Re$ is given by  \eqref{Eq_PE_iso_tensor}. It is the same expression as
1270those used for diffusive operator on tracers. It must be emphasised that such a
1271formulation is only exact in a Cartesian coordinate system, $i.e.$ on a $f-$ or
1272$\beta-$plane, not on the sphere. It is also a very good approximation in vicinity
1273of the Equator in a geographical coordinate system \citep{Lengaigne_al_JGR03}.
1274
1275\subsubsection{lateral fourth order momentum diffusive operator}
1276
1277As for tracers, the fourth order momentum diffusive operator along $z$ or $s$-surfaces
1278is a re-entering second order operator \eqref{Eq_PE_lapU} or \eqref{Eq_PE_lapU} 
1279with the eddy viscosity coefficient correctly placed:
1280
1281geopotential diffusion in $z$-coordinate:
1282\begin{equation} \label{Eq_PE_bilapU}
1283\begin{split}
1284{\rm {\bf D}}^{l{\rm {\bf U}}} &=\nabla _h \left\{ {\;\nabla _h {\rm {\bf 
1285.}}\left[ {A^{lm}\,\nabla _h \left( \chi \right)} \right]\;} 
1286\right\}\;   \\
1287&+\nabla _h \times \left\{ {\;{\rm {\bf k}}\cdot \nabla \times 
1288\left[ {A^{lm}\,\nabla _h \times \left( {\zeta \;{\rm {\bf k}}} \right)} 
1289\right]\;} \right\}
1290\end{split}
1291\end{equation}
1292
1293\gmcomment{  change the position of the coefficient, both here and in the code}
1294
1295geopotential diffusion in $s$-coordinate:
1296\begin{equation} \label{Eq_bilapU_iso}
1297   \left\{   \begin{aligned}
1298         D_u^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta u} \right) \\ 
1299         D_v^{l{\rm {\bf U}}} =\Delta \left( {A^{lm}\;\Delta v} \right)
1300   \end{aligned}    \right.
1301   \quad \text{where} \quad 
1302   \Delta \left( \bullet \right) = \nabla \cdot \left( \Re \nabla(\bullet) \right)
1303\end{equation}
1304
Note: See TracBrowser for help on using the repository browser.