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 NEMO/trunk/doc/latex/NEMO/subfiles – NEMO

source: NEMO/trunk/doc/latex/NEMO/subfiles/chap_model_basics.tex

Last change on this file was 14257, checked in by nicolasmartin, 3 years ago

Overall review of LaTeX sources (not tested completely as of now):

  • Reworking global files: main document.tex, add glossary.tex, cosmetic changes...
  • Ignore missing namelists (namsbc_isf, namsbc_iscpl and namptr)
  • Removal of references for unused indices (\hfile, \ifile and \jp)
  • Update of .svnignore and svn:ignore properties accordingly
  • Split of manual abstract into a common NEMO abs for all and a specific one for each engine
  • Shrinking variables names used in the frontmatter
File size: 62.9 KB
Line 
1\documentclass[../main/NEMO_manual]{subfiles}
2
3\begin{document}
4
5\chapter{Model Basics}
6\label{chap:MB}
7
8\chaptertoc
9
10\paragraph{Changes record} ~\\
11
12{\footnotesize
13  \begin{tabular}{l||l|l}
14    Release          & Author(s)                                   & Modifications       \\
15    \hline
16    {\em        4.0} & {\em Mike Bell                            } & {\em Review       } \\
17    {\em        3.6} & {\em Tim Graham and Gurvan Madec          } & {\em Updates      } \\
18    {\em $\leq$ 3.4} & {\em Gurvan Madec and S\'{e}bastien Masson} & {\em First version} \\
19  \end{tabular}
20}
21
22\clearpage
23
24%% =================================================================================================
25\section{Primitive equations}
26\label{sec:MB_PE}
27
28%% =================================================================================================
29\subsection{Vector invariant formulation}
30\label{subsec:MB_PE_vector}
31
32The ocean is a fluid that can be described to a good approximation by the primitive equations,
33\ie\ the Navier-Stokes equations along with a nonlinear equation of state which
34couples the two active tracers (temperature and salinity) to the fluid velocity,
35plus the following additional assumptions made from scale considerations:
36
37\begin{labeling}{Neglect of additional Coriolis terms}
38\item [\textit{Spherical Earth approximation}] The geopotential surfaces are assumed to
39  be oblate spheroids that follow the Earth's bulge;
40  these spheroids are approximated by spheres with gravity locally vertical
41  (parallel to the Earth's radius) and independent of latitude
42  \citep[][section 2]{white.hoskins.ea_QJRMS05}.
43\item [\textit{Thin-shell approximation}] The ocean depth is neglected compared to the earth's radius
44\item [\textit{Turbulent closure hypothesis}] The turbulent fluxes
45  (which represent the effect of small scale processes on the large-scale)
46  are expressed in terms of large-scale features
47\item [\textit{Boussinesq hypothesis}] Density variations are neglected except in
48  their contribution to the buoyancy force
49  \begin{equation}
50    \label{eq:MB_PE_eos}
51    \rho = \rho \ (T,S,p)
52  \end{equation}
53\item [\textit{Hydrostatic hypothesis}] The vertical momentum equation is reduced to
54  a balance between the vertical pressure gradient and the buoyancy force
55  (this removes convective processes from the initial Navier-Stokes equations and so
56  convective processes must be parameterized instead)
57  \begin{equation}
58    \label{eq:MB_PE_hydrostatic}
59    \pd[p]{z} = - \rho \ g
60  \end{equation}
61\item [\textit{Incompressibility hypothesis}] The three dimensional divergence of
62  the velocity vector $\vect U$ is assumed to be zero.
63  \begin{equation}
64    \label{eq:MB_PE_continuity}
65    \nabla \cdot \vect U = 0
66  \end{equation}
67\item [\textit{Neglect of additional Coriolis terms}] The Coriolis terms that vary with
68  the cosine of latitude are neglected.
69  These terms may be non-negligible where the Brunt-V\"{a}is\"{a}l\"{a} frequency $N$ is small,
70  either in the deep ocean or in the sub-mesoscale motions of the mixed layer,
71  or near the equator \citep[][section 1]{white.hoskins.ea_QJRMS05}.
72  They can be consistently included as part of the ocean dynamics
73  \citep[][section 3(d)]{white.hoskins.ea_QJRMS05} and are retained in the MIT ocean model.
74\end{labeling}
75
76Because the gravitational force is so dominant in the equations of large-scale motions,
77it is useful to choose an orthogonal set of unit vectors $(i,j,k)$ linked to the Earth such that
78$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
79\ie\ tangent to the geopotential surfaces.
80Let us define the following variables:
81$\vect U$ the vector velocity, $\vect U = \vect U_h + w \, \vect k$
82(the subscript $h$ denotes the local horizontal vector, \ie\ over the $(i,j)$ plane),
83$T$ the potential temperature, $S$ the salinity, $\rho$ the \textit{in situ} density.
84The vector invariant form of the primitive equations in the $(i,j,k)$ vector system provides
85the following equations:
86\begin{subequations}
87  \label{eq:MB_PE}
88  \begin{gather}
89    \shortintertext{$-$ the momentum balance}
90    \label{eq:MB_PE_dyn}
91    \pd[\vect U_h]{t} = - \lt[ (\nabla \times \vect U) \times \vect U + \frac{1}{2} \nabla \lt( \vect U^2 \rt) \rt]_h - f \; k \times \vect U_h - \frac{1}{\rho_o} \nabla_h p + \vect D^{\vect U} + \vect F^{\vect U}
92    \shortintertext{$-$ the heat and salt conservation equations}
93    \label{eq:MB_PE_tra_T}
94    \pd[T]{t} = - \nabla \cdot (T \ \vect U) + D^T + F^T \\
95    \label{eq:MB_PE_tra_S}
96    \pd[S]{t} = - \nabla \cdot (S \ \vect U) + D^S + F^S
97  \end{gather}
98\end{subequations}
99where $\nabla$ is the generalised derivative vector operator in $(i,j,k)$ directions, $t$ is the time,
100$z$ is the vertical coordinate, $\rho$ is the \textit{in situ} density given by the equation of state
101(\autoref{eq:MB_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,
102$f = 2 \vect \Omega \cdot k$ is the Coriolis acceleration
103(where $\vect \Omega$ is the Earth's angular velocity vector),
104and $g$ is the gravitational acceleration.
105$\vect D^{\vect U}$, $D^T$ and $D^S$ are the parameterisations of small-scale physics for momentum,
106temperature and salinity, and $\vect F^{\vect U}$, $F^T$ and $F^S$ surface forcing terms.
107Their nature and formulation are discussed in \autoref{sec:MB_zdf_ldf} and
108\autoref{subsec:MB_boundary_condition}.
109
110%% =================================================================================================
111\subsection{Boundary conditions}
112\label{subsec:MB_boundary_condition}
113
114An ocean is bounded by complex coastlines, bottom topography at its base and
115an air-sea or ice-sea interface at its top.
116These boundaries can be defined by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,k,t)$,
117where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface
118(discretisation can introduce additional artificial ``side-wall'' boundaries).
119Both $H$ and $\eta$ are referenced to a surface of constant geopotential
120(\ie\ a mean sea surface height) on which $z = 0$ (\autoref{fig:MB_ocean_bc}).
121Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt,
122and momentum with the solid earth, the continental margins, the sea ice and the atmosphere.
123However, some of these fluxes are so weak that
124even on climatic time scales of thousands of years they can be neglected.
125In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and
126the other components of the earth system.
127
128\begin{figure}
129  \centering
130  \includegraphics[width=0.66\textwidth]{MB_ocean_bc}
131  \caption[Ocean boundary conditions]{
132    The ocean is bounded by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,t)$,
133    where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface.
134    Both $H$ and $\eta$ are referenced to $z = 0$.}
135  \label{fig:MB_ocean_bc}
136\end{figure}
137
138\begin{description}
139\item [Land - ocean] The major flux between continental margins and the ocean is a mass exchange of
140  fresh water through river runoff.
141  Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths.
142  It can be neglected for short range integrations but
143  has to be taken into account for long term integrations as
144  it influences the characteristics of water masses formed (especially at high latitudes).
145  It is required in order to close the water cycle of the climate system.
146  It is usually specified as a fresh water flux at the air-sea interface in
147  the vicinity of river mouths.
148\item [Solid earth - ocean] Heat and salt fluxes through the sea floor are small,
149  except in special areas of little extent.
150  They are usually neglected in the model \footnote{
151    In fact, it has been shown that the heat flux associated with the solid Earth cooling
152    (\ie\ the geothermal heating) is not negligible for
153    the thermohaline circulation of the world ocean (see \autoref{subsec:TRA_bbc}).
154  }.
155  The boundary condition is thus set to no flux of heat and salt across solid boundaries.
156  For momentum, the situation is different.
157  There is no flow across solid boundaries,
158  \ie\ the velocity normal to the ocean bottom and coastlines is zero
159  (in other words, the bottom velocity is parallel to solid boundaries).
160  This kinematic boundary condition can be expressed as:
161  \begin{equation}
162    \label{eq:MB_w_bbc}
163    w = - \vect U_h \cdot \nabla_h (H)
164  \end{equation}
165  In addition, the ocean exchanges momentum with the earth through frictional processes.
166  Such momentum transfer occurs at small scales in a boundary layer.
167  It must be parameterized in terms of turbulent fluxes using
168  bottom and/or lateral boundary conditions.
169  Its specification depends on the nature of the physical parameterisation used for
170  $\vect D^{\vect U}$ in \autoref{eq:MB_PE_dyn}.
171  It is discussed in \autoref{eq:MB_zdf}. % and Chap. III.6 to 9.
172\item [Atmosphere - ocean] The kinematic surface condition plus the mass flux of fresh water PE
173  (the precipitation minus evaporation budget) leads to:
174  \[
175    % \label{eq:MB_w_sbc}
176    w = \pd[\eta]{t} + \lt. \vect U_h \rt|_{z = \eta} \cdot \nabla_h (\eta) + P - E
177  \]
178  The dynamic boundary condition, neglecting the surface tension
179  (which removes capillary waves from the system) leads to
180  the continuity of pressure across the interface $z = \eta$.
181  The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
182\item [Sea ice - ocean] The ocean and sea ice exchange heat, salt, fresh water and momentum.
183  The sea surface temperature is constrained to be at the freezing point at the interface.
184  Sea ice salinity is very low ($\sim4-6 \, psu$) compared to those of the ocean ($\sim34 \, psu$).
185  The cycle of freezing/melting is associated with fresh water and salt fluxes that
186  cannot be neglected.
187\end{description}
188
189%% =================================================================================================
190\section{Horizontal pressure gradient}
191\label{sec:MB_hor_pg}
192
193%% =================================================================================================
194\subsection{Pressure formulation}
195\label{subsec:MB_p_formulation}
196
197The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at
198a reference geopotential surface ($z = 0$) and a hydrostatic pressure $p_h$ such that:
199$p(i,j,k,t) = p_s(i,j,t) + p_h(i,j,k,t)$.
200The latter is computed by integrating (\autoref{eq:MB_PE_hydrostatic}),
201assuming that pressure in decibars can be approximated by depth in meters in (\autoref{eq:MB_PE_eos}).
202The hydrostatic pressure is then given by:
203\[
204  % \label{eq:MB_pressure}
205  p_h (i,j,z,t) = \int_{\varsigma = z}^{\varsigma = 0} g \; \rho (T,S,\varsigma) \; d \varsigma
206\]
207Two strategies can be considered for the surface pressure term:
208\begin{enumerate*}[label=(\textit{\alph*})]
209\item introduce of a new variable $\eta$, the free-surface elevation,
210for which a prognostic equation can be established and solved;
211\item assume that the ocean surface is a rigid lid,
212on which the pressure (or its horizontal gradient) can be diagnosed.
213\end{enumerate*}
214When the former strategy is used, one solution of the free-surface elevation consists of
215the excitation of external gravity waves.
216The flow is barotropic and the surface moves up and down with gravity as the restoring force.
217The phase speed of such waves is high (some hundreds of metres per second) so that
218the time step has to be very short when they are present in the model.
219The latter strategy filters out these waves since the rigid lid approximation implies $\eta = 0$,
220\ie\ the sea surface is the surface $z = 0$.
221This well known approximation increases the surface wave speed to infinity and
222modifies certain other longwave dynamics (\eg\ barotropic Rossby or planetary waves).
223The rigid-lid hypothesis is an obsolescent feature in modern OGCMs.
224It has been available until the release 3.1 of \NEMO,
225and it has been removed in release 3.2 and followings.
226Only the free surface formulation is now described in this document (see the next sub-section).
227
228%% =================================================================================================
229\subsection{Free surface formulation}
230\label{subsec:MB_free_surface}
231
232In the free surface formulation, a variable $\eta$, the sea-surface height,
233is introduced which describes the shape of the air-sea interface.
234This variable is solution of a prognostic equation which is established by
235forming the vertical average of the kinematic surface condition (\autoref{eq:MB_w_bbc}):
236\begin{equation}
237  \label{eq:MB_ssh}
238  \pd[\eta]{t} = - D + P - E \quad \text{where} \quad D = \nabla \cdot \lt[ (H + \eta) \; \overline{U}_h \, \rt]
239\end{equation}
240and using (\autoref{eq:MB_PE_hydrostatic}) the surface pressure is given by:
241$p_s = \rho \, g \, \eta$.
242
243Allowing the air-sea interface to move introduces
244the \textbf{E}xternal \textbf{G}ravity \textbf{W}aves (EGWs) as
245a class of solution of the primitive equations.
246These waves are barotropic (\ie\ nearly independent of depth) and their phase speed is quite high.
247Their time scale is short with respect to the other processes described by the primitive equations.
248
249Two choices can be made regarding the implementation of the free surface in the model,
250depending on the physical processes of interest.
251\begin{itemize}
252\item If one is interested in EGWs, in particular the tides and their interaction with
253  the baroclinic structure of the ocean (internal waves) possibly in shallow seas,
254  then a non linear free surface is the most appropriate.
255  This means that no approximation is made in \autoref{eq:MB_ssh} and that
256  the variation of the ocean volume is fully taken into account.
257  Note that in order to study the fast time scales associated with EGWs it is necessary to
258  minimize time filtering effects
259  (use an explicit time scheme with very small time step,
260  or a split-explicit scheme with reasonably small time step,
261  see \autoref{subsec:DYN_spg_exp} or \autoref{subsec:DYN_spg_ts}).
262\item If one is not interested in EGWs but rather sees them as high frequency noise,
263  it is possible to apply an explicit filter to slow down the fastest waves while
264  not altering the slow barotropic Rossby waves.
265  If further, an approximative conservation of heat and salt contents is sufficient for
266  the problem solved, then it is sufficient to solve a linearized version of \autoref{eq:MB_ssh},
267  which still allows to take into account freshwater fluxes applied at the ocean surface
268  \citep{roullet.madec_JGR00}.
269  Nevertheless, with the linearization, an exact conservation of heat and salt contents is lost.
270\end{itemize}
271The filtering of EGWs in models with a free surface is usually
272a matter of discretisation of the temporal derivatives,
273using a split-explicit method \citep{killworth.webb.ea_JPO91, zhang.endoh_JGR92} or
274the implicit scheme \citep{dukowicz.smith_JGR94} or the addition of a filtering force in
275the momentum equation \citep{roullet.madec_JGR00}.
276With the present release, \NEMO\ offers the choice between an explicit free surface
277(see \autoref{subsec:DYN_spg_exp}) or a split-explicit scheme strongly inspired the one proposed by
278\citet{shchepetkin.mcwilliams_OM05} (see \autoref{subsec:DYN_spg_ts}).
279
280%% =================================================================================================
281\section{Curvilinear \textit{z}-coordinate system}
282\label{sec:MB_zco}
283
284%% =================================================================================================
285\subsection{Tensorial formalism}
286\label{subsec:MB_tensorial}
287
288In many ocean circulation problems, the flow field has regions of enhanced dynamics
289(\ie\ surface layers, western boundary currents, equatorial currents, or ocean fronts).
290The representation of such dynamical processes can be improved by
291specifically increasing the model resolution in these regions.
292As well, it may be convenient to use a lateral boundary-following coordinate system to
293better represent coastal dynamics.
294Moreover, the common geographical coordinate system has a singular point at the North Pole that
295cannot be easily treated in a global model without filtering.
296A solution consists of introducing an appropriate coordinate transformation that
297shifts the singular point onto land \citep{madec.imbard_CD96, murray_JCP96}.
298As a consequence,
299it is important to solve the primitive equations in various curvilinear coordinate systems.
300An efficient way of introducing an appropriate coordinate transform can be found when
301using a tensorial formalism.
302This formalism is suited to any multidimensional curvilinear coordinate system.
303Ocean modellers mainly use three-dimensional orthogonal grids on the sphere
304(spherical earth approximation), with preservation of the local vertical.
305Here we give the simplified equations for this particular case.
306The general case is detailed by \citet{eiseman.stone_SR80} in
307their survey of the conservation laws of fluid dynamics.
308
309Let $(i,j,k)$ be a set of orthogonal curvilinear coordinates on
310the sphere associated with the positively oriented orthogonal set of unit vectors
311$(i,j,k)$ linked to the earth such that
312$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
313\ie\ along geopotential surfaces (\autoref{fig:MB_referential}).
314Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by
315the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and
316the distance from the centre of the earth $a + z(k)$ where $a$ is the earth's radius and
317$z$ the altitude above a reference sea level (\autoref{fig:MB_referential}).
318The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$,
319the three scale factors:
320\begin{equation}
321  \label{eq:MB_scale_factors}
322    e_1 = (a + z) \lt[ \lt( \pd[\lambda]{i} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{i} \rt)^2 \rt]^{1/2} \quad e_2 = (a + z) \lt[ \lt( \pd[\lambda]{j} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{j} \rt)^2 \rt]^{1/2} \quad e_3 = \lt( \pd[z]{k} \rt)
323\end{equation}
324
325\begin{figure}
326  \centering
327  \includegraphics[width=0.33\textwidth]{MB_earth_referential}
328  \caption[Geographical and curvilinear coordinate systems]{
329    the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear
330    coordinate system $(i,j,k)$.}
331  \label{fig:MB_referential}
332\end{figure}
333
334Since the ocean depth is far smaller than the earth's radius, $a + z$, can be replaced by $a$ in
335(\autoref{eq:MB_scale_factors}) (thin-shell approximation).
336The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while
337the vertical scale factor is a single function of $k$ as $k$ is parallel to $z$.
338The scalar and vector operators that appear in the primitive equations
339(\autoref{eq:MB_PE_dyn} to \autoref{eq:MB_PE_eos}) can then be written in the tensorial form,
340invariant in any orthogonal horizontal curvilinear coordinate system transformation:
341\begin{subequations}
342  % \label{eq:MB_discrete_operators}
343  \begin{align}
344    \label{eq:MB_grad}
345    \nabla q &=   \frac{1}{e_1} \pd[q]{i} \; \vect i + \frac{1}{e_2} \pd[q]{j} \; \vect j + \frac{1}{e_3} \pd[q]{k} \; \vect k \\
346    \label{eq:MB_div}
347    \nabla \cdot \vect A &=   \frac{1}{e_1 \; e_2} \lt[ \pd[(e_2 \; a_1)]{\partial i} + \pd[(e_1 \; a_2)]{j} \rt] + \frac{1}{e_3} \lt[ \pd[a_3]{k} \rt] \\
348    \label{eq:MB_curl}
349      \nabla \times \vect{A} &=   \lt[ \frac{1}{e_2} \pd[a_3]{j} - \frac{1}{e_3} \pd[a_2]{k}   \rt] \vect i + \lt[ \frac{1}{e_3} \pd[a_1]{k} - \frac{1}{e_1} \pd[a_3]{i}   \rt] \vect j + \frac{1}{e_1 e_2} \lt[ \pd[(e_2 a_2)]{i} - \pd[(e_1 a_1)]{j} \rt] \vect k \\
350    \label{eq:MB_lap}
351    \Delta q &= \nabla \cdot (\nabla q) \\
352    \label{eq:MB_lap_vector}
353    \Delta \vect A &= \nabla (\nabla \cdot \vect A) - \nabla \times (\nabla \times \vect A)
354  \end{align}
355\end{subequations}
356where $q$ is a scalar quantity and
357$\vect A = (a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinates system.
358
359%% =================================================================================================
360\subsection{Continuous model equations}
361\label{subsec:MB_zco_Eq}
362
363In order to express the Primitive Equations in tensorial formalism,
364it is necessary to compute the horizontal component of the non-linear and viscous terms of
365the equation using \autoref{eq:MB_grad}) to \autoref{eq:MB_lap_vector}.
366Let us set $\vect U = (u,v,w) = \vect U_h + w \; \vect k $,
367the velocity in the $(i,j,k)$ coordinates system,
368and define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$,
369by:
370\begin{gather}
371  \label{eq:MB_curl_Uh}
372  \zeta = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, v)]{i} - \pd[(e_1 \, u)]{j} \rt] \\
373  \label{eq:MB_div_Uh}
374  \chi  = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, u)]{i} + \pd[(e_1 \, v)]{j} \rt]
375\end{gather}
376
377Using again the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that
378$e_3$  is a function of the single variable $k$,
379$NLT$ the nonlinear term of \autoref{eq:MB_PE_dyn} can be transformed as follows:
380\begin{align*}
381  NLT &= \lt[ (\nabla \times {\vect U}) \times {\vect U} + \frac{1}{2} \nabla \lt( {\vect U}^2 \rt) \rt]_h \\
382      &= \lt(
383        \begin{array}{*{20}c}
384          \lt[ \frac{1}{e_3} \pd[u]{k} - \frac{1}{e_1} \pd[w]{i} \rt] w - \zeta \; v   \\
385          \zeta \; u - \lt[ \frac{1}{e_2} \pd[w]{j} - \frac{1}{e_3} \pd[v]{k} \rt] \ w
386        \end{array}
387  \rt)
388  + \frac{1}{2} \lt(
389  \begin{array}{*{20}c}
390    \frac{1}{e_1} \pd[(u^2 + v^2 + w^2)]{i} \\
391    \frac{1}{e_2} \pd[(u^2 + v^2 + w^2)]{j}
392  \end{array}
393  \rt) \\
394      &= \lt(
395        \begin{array}{*{20}c}
396          -\zeta \; v \\
397          \zeta \; u
398        \end{array}
399  \rt)
400  + \frac{1}{2} \lt(
401  \begin{array}{*{20}c}
402    \frac{1}{e_1} \pd[(u^2 + v^2)]{i} \\
403    \frac{1}{e_2} \pd[(u^2 + v^2)]{j}
404  \end{array}
405  \rt)
406  + \frac{1}{e_3} \lt(
407  \begin{array}{*{20}c}
408    w \; \pd[u]{k} \\
409    w \; \pd[v]{k}
410  \end{array}
411  \rt)
412  - \lt(
413  \begin{array}{*{20}c}
414    \frac{w}{e_1} \pd[w]{i} - \frac{1}{2 e_1} \pd[w^2]{i} \\
415    \frac{w}{e_2} \pd[w]{j} - \frac{1}{2 e_2} \pd[w^2]{j}
416  \end{array}
417  \rt)
418\end{align*}
419The last term of the right hand side is obviously zero,
420and thus the \textbf{N}on\textbf{L}inear \textbf{T}erm ($NLT$) of \autoref{eq:MB_PE_dyn} is written in
421the $(i,j,k)$ coordinate system:
422\begin{equation}
423  \label{eq:MB_vector_form}
424  NLT = \zeta \; \vect k \times \vect U_h + \frac{1}{2} \nabla_h \lt( \vect U_h^2 \rt) + \frac{1}{e_3} w \pd[\vect U_h]{k}
425\end{equation}
426
427This is the so-called \textit{vector invariant form} of the momentum advection term.
428For some purposes, it can be advantageous to write this term in the so-called flux form,
429\ie\ to write it as the divergence of fluxes.
430For example,
431the first component of \autoref{eq:MB_vector_form} (the $i$-component) is transformed as follows:
432\begin{alignat*}{3}
433  &NLT_i &= &- \zeta \; v + \frac{1}{2 \; e_1} \pd[ (u^2 + v^2) ]{i} + \frac{1}{e_3} w \ \pd[u]{k} \\
434  &      &= &\frac{1}{e_1 \; e_2} \lt( -v \pd[(e_2 \, v)]{i} + v \pd[(e_1 \, u)]{j} \rt) + \frac{1}{e_1 e_2} \lt( e_2 \; u \pd[u]{i} + e_2 \; v \pd[v]{i} \rt) + \frac{1}{e_3} \lt( w \; \pd[u]{k} \rt) \\
435  &      &= &\frac{1}{e_1 \; e_2} \lt[ - \lt( v^2 \pd[e_2]{i} + e_2 \, v \pd[v]{i} \rt) + \lt( \pd[ \lt( e_1 \, u \, v \rt)]{j} - e_1 \, u \pd[v]{j} \rt) \rt. \lt. + \lt( \pd[ \lt( e_2 \, u \, u \rt)]{i} - u \pd[ \lt( e_2 u \rt)]{i} \rt) + e_2 v \pd[v]{i} \rt] \\
436  &      & &+ \frac{1}{e_3} \lt( \pd[(w \, u)]{k} - u \pd[w]{k} \rt) \\
437  &      &= &\frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, u \, v)]{j} \rt) + \frac{1}{e_3} \pd[(w \, u)]{k} + \frac{1}{e_1 e_2} \lt[ - u \lt( \pd[(e_1 v)]{j} - v \, \pd[e_1]{j} \rt) - u \pd[(e_2 u)]{i} \rt] - \frac{1}{e_3} \pd[w]{k} u \\
438  &      & &+ \frac{1}{e_1 e_2} \lt( - v^2 \pd[e_2]{i} \rt) \\
439  &      &= &\nabla \cdot (\vect U \, u) - (\nabla \cdot \vect U) \ u + \frac{1}{e_1 e_2} \lt( -v^2 \pd[e_2]{i} + u v \, \pd[e_1]{j} \rt) \\
440  \shortintertext{as $\nabla \cdot {\vect U} \; = 0$ (incompressibility) it becomes:}
441  &      &= &\, \nabla \cdot (\vect U \, u) + \frac{1}{e_1 e_2} \lt( v \; \pd[e_2]{i} - u \; \pd[e_1]{j} \rt) (-v)
442\end{alignat*}
443
444The flux form of the momentum advection term is therefore given by:
445\begin{equation}
446  \label{eq:MB_flux_form}
447  NLT = \nabla \cdot \lt(
448  \begin{array}{*{20}c}
449    \vect U \, u \\
450    \vect U \, v
451  \end{array}
452  \rt)
453  + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \vect k \times \vect U_h
454\end{equation}
455
456The flux form has two terms,
457the first one is expressed as the divergence of momentum fluxes
458(hence the flux form name given to this formulation) and
459the second one is due to the curvilinear nature of the coordinate system used.
460The latter is called the \textit{metric} term and can be viewed as
461a modification of the Coriolis parameter:
462\[
463  % \label{eq:MB_cor+metric}
464  f \to f + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt)
465\]
466
467Note that in the case of geographical coordinate,
468\ie\ when $(i,j) \to (\lambda,\varphi)$ and $(e_1,e_2) \to (a \, \cos \varphi,a)$,
469we recover the commonly used modification of the Coriolis parameter $f \to f + (u / a) \tan \varphi$.
470
471To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be written in
472the following tensorial formalism:
473
474\begin{description}
475\item [Vector invariant form of the momentum equations]
476  \begin{equation}
477    \label{eq:MB_dyn_vect}
478    \begin{gathered}
479    % \label{eq:MB_dyn_vect_u}
480      \pd[u]{t} = + (\zeta + f) \, v - \frac{1}{2 e_1} \pd[]{i} (u^2 + v^2) - \frac{1}{e_3} w \pd[u]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) + D_u^{\vect U} + F_u^{\vect U} \\
481      \pd[v]{t} = - (\zeta + f) \, u - \frac{1}{2 e_2} \pd[]{j} (u^2 + v^2) - \frac{1}{e_3} w \pd[v]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) + D_v^{\vect U} + F_v^{\vect U}
482    \end{gathered}
483  \end{equation}
484\item [Flux form of the momentum equations]
485  % \label{eq:MB_dyn_flux}
486  \begin{alignat*}{2}
487    % \label{eq:MB_dyn_flux_u}
488    \pd[u]{t} = &+ \lt[ f + \frac{1}{e_1 \; e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \rt] \, v - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, v \, u)]{j} \rt) - \frac{1}{e_3} \pd[(w \, u)]{k} \\
489    &- \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) + D_u^{\vect U} + F_u^{\vect U}
490  \end{alignat*}
491  \begin{alignat*}{2}
492    % \label{eq:MB_dyn_flux_v}
493    \pd[v]{t} = &- \lt[ f + \frac{1}{e_1 \; e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \rt] \, u - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, v)]{i} + \pd[(e_1 \, v \, v)]{j} \rt) - \frac{1}{e_3} \pd[(w \, v)]{k} \\
494    &- \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) + D_v^{\vect U} + F_v^{\vect U}
495  \end{alignat*}
496  where $\zeta$, the relative vorticity, is given by \autoref{eq:MB_curl_Uh} and
497  $p_s$, the surface pressure, is given by:
498  \[
499  % \label{eq:MB_spg}
500    p_s = \rho \,g \, \eta
501  \]
502  and $\eta$ is the solution of \autoref{eq:MB_ssh}.
503
504  The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
505  \[
506  % \label{eq:MB_w_diag}
507    \pd[w]{k} = - \chi \; e_3 \qquad
508  % \label{eq:MB_hp_diag}
509    \pd[p_h]{k} = - \rho \; g \; e_3
510  \]
511  where the divergence of the horizontal velocity, $\chi$ is given by \autoref{eq:MB_div_Uh}.
512\item [Tracer equations]
513  \begin{gather*}
514    \pd[T]{t} = - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 T \, u)]{i} + \pd[(e_1 T \, v)]{j} \rt] - \frac{1}{e_3} \pd[(T \, w)]{k} + D^T + F^T \\
515    \pd[S]{t} = - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 S \, u)]{i} + \pd[(e_1 S \, v)]{j} \rt] - \frac{1}{e_3} \pd[(S \, w)]{k} + D^S + F^S \\
516    \rho = \rho \big( T,S,z(k) \big)
517  \end{gather*}
518\end{description}
519
520The expression of $\vect D^{U}$, $D^{S}$ and $D^{T}$ depends on
521the subgrid scale parameterisation used.
522It will be defined in \autoref{eq:MB_zdf}.
523The nature and formulation of $\vect F^{\vect U}$, $F^T$ and $F^S$, the surface forcing terms,
524are discussed in \autoref{chap:SBC}.
525
526%% =================================================================================================
527\section{Curvilinear generalised vertical coordinate system}
528\label{sec:MB_gco}
529
530The ocean domain presents a huge diversity of situation in the vertical.
531First the ocean surface is a time dependent surface (moving surface).
532Second the ocean floor depends on the geographical position,
533varying from more than 6,000 meters in abyssal trenches to zero at the coast.
534Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.
535Therefore, in order to represent the ocean with respect to the first point
536a space and time dependent vertical coordinate that follows the variation of the sea surface height
537\eg\ an \zstar-coordinate;
538for the second point,
539a space variation to fit the change of bottom topography
540\eg\ a terrain-following or $\sigma$-coordinate;
541and for the third point,
542one will be tempted to use a space and time dependent coordinate that follows the isopycnal surfaces,
543\eg\ an isopycnic coordinate.
544
545In order to satisfy two or more constraints one can even be tempted to mixed these coordinate systems,
546as in HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and
547$\sigma$ at the ocean bottom) \citep{chassignet.smith.ea_JPO03} or
548OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and
549$\sigma$-coordinate elsewhere) \citep{madec.delecluse.ea_JPO96} among others.
550
551In fact one is totally free to choose any space and time vertical coordinate by
552introducing an arbitrary vertical coordinate :
553\begin{equation}
554  \label{eq:MB_s}
555  s = s(i,j,k,t)
556\end{equation}
557with the restriction that the above equation gives a single-valued monotonic relationship between
558$s$ and $k$, when $i$, $j$ and $t$ are held fixed.
559\autoref{eq:MB_s} is a transformation from
560the $(i,j,k,t)$ coordinate system with independent variables into
561the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through
562\autoref{eq:MB_s}.
563This so-called \textit{generalised vertical coordinate} \citep{kasahara_MWR74} is in fact
564an \textbf{A}rbitrary \textbf{L}agrangian--\textbf{E}ulerian (ALE) coordinate.
565Indeed, one has a great deal of freedom in the choice of expression for $s$.
566The choice determines which part of the vertical velocity (defined from a fixed referential)
567will cross the levels (Eulerian part) and which part will be used to move them (Lagrangian part).
568The coordinate is also sometimes referenced as an adaptive coordinate
569\citep{hofmeister.burchard.ea_OM10}, since the coordinate system is adapted in
570the course of the simulation.
571Its most often used implementation is via an ALE algorithm,
572in which a pure lagrangian step is followed by regridding and remapping steps,
573the latter step implicitly embedding the vertical advection
574\citep{hirt.amsden.ea_JCP74, chassignet.smith.ea_JPO03, white.adcroft.ea_JCP09}.
575Here we follow the \citep{kasahara_MWR74} strategy:
576a regridding step (an update of the vertical coordinate) followed by an Eulerian step with
577an explicit computation of vertical advection relative to the moving s-surfaces.
578
579\cmtgm{A key point here is that the $s$-coordinate depends on $(i,j)$
580  ==> horizontal pressure gradient...}
581The generalized vertical coordinates used in ocean modelling are not orthogonal,
582which contrasts with many other applications in mathematical physics.
583Hence, it is useful to keep in mind the following properties that may seem odd on initial encounter.
584
585The horizontal velocity in ocean models measures motions in the horizontal plane,
586perpendicular to the local gravitational field.
587That is, horizontal velocity is mathematically the same regardless of the vertical coordinate,
588be it geopotential, isopycnal, pressure, or terrain following.
589The key motivation for maintaining the same horizontal velocity component is that
590the hydrostatic and geostrophic balances are dominant in the large-scale ocean.
591Use of an alternative quasi -horizontal velocity,
592for example one oriented parallel to the generalized surface,
593would lead to unacceptable numerical errors.
594Correspondingly, the vertical direction is anti -parallel to the gravitational force in
595all of the coordinate systems.
596We do not choose the alternative of a quasi -vertical direction oriented normal to
597the surface of a constant generalized vertical coordinate.
598
599It is the method used to measure transport across the generalized vertical coordinate surfaces which
600differs between the vertical coordinate choices.
601That is,
602computation of the dia-surface velocity component represents the fundamental distinction between
603the various coordinates.
604In some models, such as geopotential, pressure, and terrain following,
605this transport is typically diagnosed from volume or mass conservation.
606In other models, such as isopycnal layered models,
607this transport is prescribed based on assumptions about the physical processes producing
608a flux across the layer interfaces.
609
610In this section we first establish the PE in the generalised vertical $s$-coordinate,
611then we discuss the particular cases available in \NEMO, namely $z$, \zstar, $s$, and \ztilde.
612%}
613
614%% =================================================================================================
615\subsection{\textit{S}-coordinate formulation}
616
617Starting from the set of equations established in \autoref{sec:MB_zco} for
618the special case $k = z$ and thus $e_3 = 1$,
619we introduce an arbitrary vertical coordinate $s = s(i,j,k,t)$,
620which includes $z$-, \zstar- and $\sigma$-coordinates as special cases
621($s = z$, $s = \zstar$, and $s = \sigma = z / H$ or $ = z / \lt( H + \eta \rt)$, resp.).
622A formal derivation of the transformed equations is given in \autoref{apdx:SCOORD}.
623Let us define the vertical scale factor by $e_3 = \partial_s z$
624($e_3$ is now a function of $(i,j,k,t)$ ),
625and the slopes in the $(i,j)$ directions between $s$- and $z$-surfaces by:
626\begin{equation}
627  \label{eq:MB_sco_slope}
628  \sigma_1 = \frac{1}{e_1} \; \lt. \pd[z]{i} \rt|_s \quad \text{and} \quad
629  \sigma_2 = \frac{1}{e_2} \; \lt. \pd[z]{j} \rt|_s
630\end{equation}
631We also introduce $\omega$, a dia-surface velocity component,
632defined as the velocity relative to the moving $s$-surfaces and normal to them:
633\[
634  % \label{eq:MB_sco_w}
635  \omega = w -  \, \lt. \pd[z]{t} \rt|_s - \sigma_1 \, u - \sigma_2 \, v
636\]
637
638The equations solved by the ocean model \autoref{eq:MB_PE} in $s$-coordinate can be written as follows
639(see \autoref{sec:SCOORD_momentum}):
640
641\begin{description}
642\item [Vector invariant form of the momentum equation]
643  \begin{gather*}
644    % \label{eq:MB_sco_u_vector}
645    \pd[u]{t} = + (\zeta + f) \, v - \frac{1}{2 \, e_1} \pd[]{i} (u^2 + v^2) - \frac{1}{e_3} \omega \pd[u]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_1 + D_u^{\vect U} + F_u^{\vect U} \\
646    % \label{eq:MB_sco_v_vector}
647    \pd[v]{t} = - (\zeta + f) \, u - \frac{1}{2 \, e_2} \pd[]{j} (u^2 + v^2) - \frac{1}{e_3} \omega \pd[v]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_2 + D_v^{\vect U} + F_v^{\vect U}
648  \end{gather*}
649\item [Flux form of the momentum equation]
650  \begin{alignat*}{2}
651    % \label{eq:MB_sco_u_flux}
652    \frac{1}{e_3} \pd[(e_3 \, u)]{t} = &+ \lt[ f + \frac{1}{e_1 \; e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \rt] \, v - \frac{1}{e_1 \; e_2 \; e_3} \lt( \pd[(e_2 \, e_3 \, u \, u)]{i} + \pd[(e_1 \, e_3 \, v \, u)]{j} \rt) - \frac{1}{e_3} \pd[(\omega \, u)]{k} \\
653    &- \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_1 + D_u^{\vect U} + F_u^{\vect U}
654  \end{alignat*}
655  \begin{alignat*}{2}
656  % \label{eq:MB_sco_v_flux}
657    \frac{1}{e_3} \pd[(e_3 \, v)]{t} = &- \lt[ f + \frac{1}{e_1 \; e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \rt] \, u - \frac{1}{e_1 \; e_2 \; e_3} \lt( \pd[( e_2 \; e_3 \, u \, v)]{i} + \pd[(e_1 \; e_3 \, v \, v)]{j} \rt) - \frac{1}{e_3} \pd[(\omega \, v)]{k} \\
658    &- \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o}\sigma_2 + D_v^{\vect U} + F_v^{\vect U}
659  \end{alignat*}
660  where the relative vorticity, $\zeta$, the surface pressure gradient,
661  and the hydrostatic pressure have the same expressions as in $z$-coordinates although
662  they do not represent exactly the same quantities.
663  $\omega$ is provided by the continuity equation (see \autoref{apdx:SCOORD}):
664  \[
665  % \label{eq:MB_sco_continuity}
666    \pd[e_3]{t} + e_3 \; \chi + \pd[\omega]{s} = 0 \quad \text{with} \quad
667    \chi = \frac{1}{e_1 e_2 e_3} \lt( \pd[(e_2 e_3 \, u)]{i} + \pd[(e_1 e_3 \, v)]{j} \rt)
668  \]
669\item [Tracer equations]
670  \begin{gather*}
671    % \label{eq:MB_sco_t}
672    \frac{1}{e_3} \pd[(e_3 \, T)]{t} = - \frac{1}{e_1 e_2 e_3} \lt(   \pd[(e_2 e_3 \, u \, T)]{i} + \pd[(e_1 e_3 \, v \, T)]{j} \rt) - \frac{1}{e_3} \pd[(T \, \omega)]{k} + D^T + F^S \\
673    % \label{eq:MB_sco_s}
674    \frac{1}{e_3} \pd[(e_3 \, S)]{t} = - \frac{1}{e_1 e_2 e_3} \lt(   \pd[(e_2 e_3 \, u \, S)]{i} + \pd[(e_1 e_3 \, v \, S)]{j} \rt) - \frac{1}{e_3} \pd[(S \, \omega)]{k} + D^S + F^S
675  \end{gather*}
676\end{description}
677The equation of state has the same expression as in $z$-coordinate,
678and similar expressions are used for mixing and forcing terms.
679
680\cmtgm{
681  \colorbox{yellow}{ to be updated $= = >$}
682  Add a few works on z and zps and s and underlies the differences between all of them
683  \colorbox{yellow}{$< = =$ end update}
684}
685
686%% =================================================================================================
687\subsection{Curvilinear \zstar-coordinate system}
688\label{subsec:MB_zco_star}
689
690\begin{figure}
691  \centering
692  \includegraphics[width=0.66\textwidth]{MB_z_zstar}
693  \caption[Curvilinear z-coordinate systems (\{non-\}linear free-surface cases and re-scaled \zstar)]{
694    \begin{enumerate*}[label=(\textit{\alph*})]
695    \item $z$-coordinate in linear free-surface case;
696    \item $z$-coordinate in non-linear free surface case;
697    \item re-scaled height coordinate
698      (become popular as the \zstar-coordinate \citep{adcroft.campin_OM04}).
699    \end{enumerate*}
700  }
701  \label{fig:MB_z_zstar}
702\end{figure}
703
704In this case, the free surface equation is nonlinear,
705and the variations of volume are fully taken into account.
706These coordinates systems is presented in a report \citep{levier.treguier.ea_trpt07} available on
707the \NEMO\ web site.
708
709The \zstar\ coordinate approach is an unapproximated, non-linear free surface implementation which
710allows one to deal with large amplitude free-surface variations relative to the vertical resolution
711\citep{adcroft.campin_OM04}.
712In the \zstar\ formulation, the variation of the column thickness due to sea-surface undulations is
713not concentrated in the surface level, as in the $z$-coordinate formulation,
714but is equally distributed over the full water column.
715Thus vertical levels naturally follow sea-surface variations, with a linear attenuation with depth,
716as illustrated by \autoref{fig:MB_z_zstar}.
717Note that with a flat bottom, such as in \autoref{fig:MB_z_zstar},
718the bottom-following $z$ coordinate and \zstar\ are equivalent.
719The definition and modified oceanic equations for the rescaled vertical coordinate \zstar,
720including the treatment of fresh-water flux at the surface, are detailed in Adcroft and Campin (2004).
721The major points are summarized here.
722The position (\zstar) and vertical discretization ($\delta \zstar$) are expressed as:
723\[
724  % \label{eq:MB_z-star}
725  H + \zstar = (H + z)  / r \quad \text{and}  \quad \delta \zstar = \delta z / r \quad \text{with} \quad r = \frac{H + \eta}{H}
726\]
727Simple re-organisation of the above expressions gives
728\[
729  % \label{eq:MB_zstar_2}
730  \zstar = H \lt( \frac{z - \eta}{H + \eta} \rt)
731\]
732Since the vertical displacement of the free surface is incorporated in the vertical coordinate \zstar,
733the upper and lower boundaries are at fixed \zstar\ position,
734$\zstar = 0$ and $\zstar = -H$ respectively.
735Also the divergence of the flow field is no longer zero as shown by the continuity equation:
736\[
737  \pd[r]{t} = \nabla_{\zstar} \cdot \lt( r \; \vect U_h \rt) + \pd[r \; w^*]{\zstar} = 0
738\]
739This \zstar\ coordinate is closely related to the $\eta$ coordinate used in many atmospheric models
740(see Black (1994) for a review of $\eta$ coordinate atmospheric models).
741It was originally used in ocean models by Stacey et al. (1995) for studies of tides next to shelves,
742and it has been recently promoted by Adcroft and Campin (2004) for global climate modelling.
743
744The surfaces of constant \zstar\ are quasi-horizontal.
745Indeed, the \zstar\ coordinate reduces to $z$ when $\eta$ is zero.
746In general, when noting the large differences between
747undulations of the bottom topography versus undulations in the surface height,
748it is clear that surfaces constant \zstar\ are very similar to the depth surfaces.
749These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to
750terrain following sigma models discussed in \autoref{subsec:MB_sco}.
751Additionally, since $\zstar = z$ when $\eta = 0$,
752no flow is spontaneously generated in an unforced ocean starting from rest,
753regardless the bottom topography.
754This behaviour is in contrast to the case with ``s''-models,
755where pressure gradient errors in the presence of nontrivial topographic variations can generate
756nontrivial spontaneous flow from a resting state,
757depending on the sophistication of the pressure gradient solver.
758The quasi-horizontal nature of the coordinate surfaces also facilitates the implementation of
759neutral physics parameterizations in \zstar\ models using the same techniques as in $z$-models
760(see Chapters 13-16 of \cite{griffies_bk04}) for a discussion of neutral physics in $z$-models,
761as well as \autoref{sec:LDF_slp} in this document for treatment in \NEMO).
762
763The range over which \zstar\ varies is time independent $-H \leq \zstar \leq 0$.
764Hence, all cells remain nonvanishing, so long as the surface height maintains $\eta > -H$.
765This is a minor constraint relative to that encountered on the surface height when
766using $s = z$ or $s = z - \eta$.
767
768Because \zstar\ has a time independent range, all grid cells have static increments ds,
769and the sum of the vertical increments yields the time independent ocean depth. %k ds = H.
770The \zstar\ coordinate is therefore invisible to undulations of the free surface,
771since it moves along with the free surface.
772This property means that no spurious vertical transport is induced across
773surfaces of constant \zstar\  by the motion of external gravity waves.
774Such spurious transport can be a problem in z-models, especially those with tidal forcing.
775Quite generally,
776the time independent range for the \zstar\ coordinate is a very convenient property that
777allows for a nearly arbitrary vertical resolution even in
778the presence of large amplitude fluctuations of the surface height, again so long as $\eta > -H$.
779%end MOM doc %%%
780
781%% =================================================================================================
782\subsection{Curvilinear terrain-following \textit{s}--coordinate}
783\label{subsec:MB_sco}
784
785Several important aspects of the ocean circulation are influenced by bottom topography.
786Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers,
787sills and channels that strongly constrain the path of water masses, but more subtle effects exist.
788For example,
789the topographic $\beta$-effect is usually larger than the planetary one along continental slopes.
790Topographic Rossby waves can be excited and can interact with the mean current.
791In the $z$-coordinate system presented in the previous section (\autoref{sec:MB_zco}),
792$z$-surfaces are geopotential surfaces.
793The bottom topography is discretised by steps.
794This often leads to a misrepresentation of a gradually sloping bottom and to
795large localized depth gradients associated with large localized vertical velocities.
796The response to such a velocity field often leads to numerical dispersion effects.
797One solution to strongly reduce this error is to use a partial step representation of
798bottom topography instead of a full step one \cite{pacanowski.gnanadesikan_MWR98}.
799Another solution is to introduce a terrain-following coordinate system (hereafter $s$-coordinate).
800
801The $s$-coordinate avoids the discretisation error in the depth field since
802the layers of computation are gradually adjusted with depth to the ocean bottom.
803Relatively small topographic features as well as
804gentle, large-scale slopes of the sea floor in the deep ocean,
805which would be ignored in typical $z$-model applications with
806the largest grid spacing at greatest depths,
807can easily be represented (with relatively low vertical resolution).
808A terrain-following model (hereafter $s$-model) also facilitates
809the modelling of the boundary layer flows over a large depth range,
810which in the framework of the $z$-model would require high vertical resolution over
811the whole depth range.
812Moreover,
813with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as
814the only boundaries of the domain (no more lateral boundary condition to specify).
815Nevertheless, a $s$-coordinate also has its drawbacks.
816Perfectly adapted to a homogeneous ocean,
817it has strong limitations as soon as stratification is introduced.
818The main two problems come from the truncation error in the horizontal pressure gradient and
819a possibly increased diapycnal diffusion.
820The horizontal pressure force in $s$-coordinate consists of two terms (see \autoref{apdx:SCOORD}),
821\begin{equation}
822  \label{eq:MB_p_sco}
823  \nabla p |_z = \nabla p |_s - \frac{1}{e_3} \pd[p]{s} \nabla z |_s
824\end{equation}
825
826The second term in \autoref{eq:MB_p_sco} depends on the tilt of the coordinate surface and
827leads to a truncation error that is not present in a $z$-model.
828In the special case of a $\sigma$-coordinate
829(i.e. a depth-normalised coordinate system $\sigma = z/H$),
830\citet{haney_JPO91} and \citet{beckmann.haidvogel_JPO93} have given estimates of
831the magnitude of this truncation error.
832It depends on topographic slope, stratification, horizontal and vertical resolution,
833the equation of state, and the finite difference scheme.
834This error limits the possible topographic slopes that a model can handle at
835a given horizontal and vertical resolution.
836This is a severe restriction for large-scale applications using realistic bottom topography.
837The large-scale slopes require high horizontal resolution,
838and the computational cost becomes prohibitive.
839This problem can be at least partially overcome by mixing $s$-coordinate and
840step-like representation of bottom topography
841\citep{gerdes_JGR93,gerdes_JGR93*a,madec.delecluse.ea_JPO96}.
842However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for
843a realistic bottom topography:
844an envelope topography is defined in $s$-coordinate on which
845a full or partial step bottom topography is then applied in order to
846adjust the model depth to the observed one (see \autoref{subsec:DOM_zgr}).
847
848For numerical reasons a minimum of diffusion is required along
849the coordinate surfaces of any finite difference model.
850It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces.
851This is the case for a $z$-model as well as for a $s$-model.
852However, density varies more strongly on $s$-surfaces than on horizontal surfaces in
853regions of large topographic slopes,
854implying larger diapycnal diffusion in a $s$-model than in a $z$-model.
855Whereas such a diapycnal diffusion in a $z$-model tends to
856weaken horizontal density (pressure) gradients and thus the horizontal circulation,
857it usually reinforces these gradients in a $s$-model, creating spurious circulation.
858For example, imagine an isolated bump of topography in
859an ocean at rest with a horizontally uniform stratification.
860Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography,
861and thus will generate there a baroclinic eddy.
862In contrast, the ocean will stay at rest in a $z$-model.
863As for the truncation error, the problem can be reduced by
864introducing the terrain-following coordinate below the strongly stratified portion of the water column
865(\ie\ the main thermocline) \citep{madec.delecluse.ea_JPO96}.
866An alternate solution consists of rotating the lateral diffusive tensor to
867geopotential or to isoneutral surfaces (see \autoref{subsec:MB_ldf}).
868Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,
869strongly exceeding the stability limit of such a operator when it is discretized
870(see \autoref{chap:LDF}).
871
872The $s$-coordinates introduced here \citep{lott.madec.ea_OM90,madec.delecluse.ea_JPO96}
873differ mainly in two aspects from similar models:
874it allows a representation of bottom topography with
875mixed full or partial step-like/terrain following topography;
876It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate.
877
878%% =================================================================================================
879\subsection{Curvilinear \ztilde-coordinate}
880\label{subsec:MB_zco_tilde}
881
882The \ztilde-coordinate has been developed by \citet{leclair.madec_OM11}.
883It is available in \NEMO\ since the version 3.4 and is more robust in version 4.0 than previously.
884Nevertheless, it is currently not robust enough to be used in all possible configurations.
885Its use is therefore not recommended.
886
887%% =================================================================================================
888\section{Subgrid scale physics}
889\label{sec:MB_zdf_ldf}
890
891The hydrostatic primitive equations describe the behaviour of a geophysical fluid at
892space and time scales larger than a few kilometres in the horizontal,
893a few meters in the vertical and a few minutes.
894They are usually solved at larger scales: the specified grid spacing and time step of
895the numerical model.
896The effects of smaller scale motions (coming from the advective terms in the Navier-Stokes equations)
897must be represented entirely in terms of large-scale patterns to close the equations.
898These effects appear in the equations as the divergence of turbulent fluxes
899(\ie\ fluxes associated with the mean correlation of small scale perturbations).
900Assuming a turbulent closure hypothesis is equivalent to choose a formulation for these fluxes.
901It is usually called the subgrid scale physics.
902It must be emphasized that this is the weakest part of the primitive equations,
903but also one of the most important for long-term simulations as
904small scale processes \textit{in fine} balance the surface input of kinetic energy and heat.
905
906The control exerted by gravity on the flow induces a strong anisotropy between
907the lateral and vertical motions.
908Therefore subgrid-scale physics \textbf{D}$^{\vect U}$, $D^{S}$ and $D^{T}$  in
909\autoref{eq:MB_PE_dyn}, \autoref{eq:MB_PE_tra_T} and \autoref{eq:MB_PE_tra_S} are divided into
910a  lateral part \textbf{D}$^{l \vect U}$, $D^{l S}$ and $D^{l T}$ and
911a vertical part \textbf{D}$^{v \vect U}$, $D^{v S}$ and $D^{v T}$.
912The formulation of these terms and their underlying physics are briefly discussed in
913the next two subsections.
914
915%% =================================================================================================
916\subsection{Vertical subgrid scale physics}
917\label{subsec:MB_zdf}
918
919The model resolution is always larger than the scale at which
920the major sources of vertical turbulence occur (shear instability, internal wave breaking...).
921Turbulent motions are thus never explicitly solved, even partially, but always parameterized.
922The vertical turbulent fluxes are assumed to depend linearly on
923the gradients of large-scale quantities (for example,
924the turbulent heat flux is given by $\overline{T' w'} = -A^{v T} \partial_z \overline T$,
925where $A^{v T}$ is an eddy coefficient).
926This formulation is analogous to that of molecular diffusion and dissipation.
927This is quite clearly a necessary compromise: considering only the molecular viscosity acting on
928large scale severely underestimates the role of turbulent diffusion and dissipation,
929while an accurate consideration of the details of turbulent motions is simply impractical.
930The resulting vertical momentum and tracer diffusive operators are of second order:
931\begin{equation}
932  \label{eq:MB_zdf}
933  \begin{gathered}
934    \vect D^{v \vect U} = \pd[]{z} \lt( A^{vm} \pd[\vect U_h]{z} \rt) \text{,} \
935          D^{vT}       = \pd[]{z} \lt( A^{vT} \pd[T]{z}         \rt) \ \text{and} \
936          D^{vS}       = \pd[]{z} \lt( A^{vT} \pd[S]{z}         \rt)
937  \end{gathered}
938\end{equation}
939where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients,
940respectively.
941At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified
942(see \autoref{chap:SBC} and \autoref{chap:ZDF} and \autoref{sec:TRA_bbl}).
943All the vertical physics is embedded in the specification of the eddy coefficients.
944They can be assumed to be either constant, or function of the local fluid properties
945(\eg\ Richardson number, Brunt-V\"{a}is\"{a}l\"{a} frequency, distance from the boundary ...),
946or computed from a turbulent closure model.
947The choices available in \NEMO\ are discussed in \autoref{chap:ZDF}).
948
949%% =================================================================================================
950\subsection{Formulation of the lateral diffusive and viscous operators}
951\label{subsec:MB_ldf}
952
953Lateral turbulence can be roughly divided into a mesoscale turbulence associated with eddies
954(which can be solved explicitly if the resolution is sufficient since
955their underlying physics are included in the primitive equations),
956and a sub mesoscale turbulence which is never explicitly solved even partially,
957but always parameterized.
958The formulation of lateral eddy fluxes depends on whether
959the mesoscale is below or above the grid-spacing (\ie\ the model is eddy-resolving or not).
960
961In non-eddy-resolving configurations, the closure is similar to that used for the vertical physics.
962The lateral turbulent fluxes are assumed to depend linearly on
963the lateral gradients of large-scale quantities.
964The resulting lateral diffusive and dissipative operators are of second order.
965Observations show that
966lateral mixing induced by mesoscale turbulence tends to be along isopycnal surfaces
967(or more precisely neutral surfaces \cite{mcdougall_JPO87}) rather than across them.
968As the slope of neutral surfaces is small in the ocean,
969a common approximation is to assume that the ``lateral'' direction is the horizontal,
970\ie\ the lateral mixing is performed along geopotential surfaces.
971This leads to a geopotential second order operator for lateral subgrid scale physics.
972This assumption can be relaxed:
973the eddy-induced turbulent fluxes can be better approached by assuming that
974they depend linearly on the gradients of large-scale quantities computed along neutral surfaces.
975In such a case, the diffusive operator is an isoneutral second order operator and
976it has components in the three space directions.
977However, both horizontal and isoneutral operators have no effect on
978mean (\ie\ large scale) potential energy whereas
979potential energy is a main source of turbulence (through baroclinic instabilities).
980\citet{gent.mcwilliams_JPO90} proposed a parameterisation of mesoscale eddy-induced turbulence which
981associates an eddy-induced velocity to the isoneutral diffusion.
982Its mean effect is to reduce the mean potential energy of the ocean.
983This leads to a formulation of lateral subgrid-scale physics made up of
984an isoneutral second order operator and an eddy induced advective part.
985In all these lateral diffusive formulations,
986the specification of the lateral eddy coefficients remains the problematic point as
987there is no really satisfactory formulation of these coefficients as
988a function of large-scale features.
989
990In eddy-resolving configurations, a second order operator can be used,
991but usually the more scale selective biharmonic operator is preferred as
992the grid-spacing is usually not small enough compared to the scale of the eddies.
993The role devoted to the subgrid-scale physics is to dissipate the energy that
994cascades toward the grid scale and thus to ensure the stability of the model while
995not interfering with the resolved mesoscale activity.
996Another approach is becoming more and more popular:
997instead of specifying explicitly a sub-grid scale term in
998the momentum and tracer time evolution equations,
999one uses an advective scheme which is diffusive enough to maintain the model stability.
1000It must be emphasised that then,
1001all the sub-grid scale physics is included in the formulation of the advection scheme.
1002
1003All these parameterisations of subgrid scale physics have advantages and drawbacks.
1004They are not all available in \NEMO.
1005For active tracers (temperature and salinity) the main ones are:
1006Laplacian and bilaplacian operators acting along geopotential or iso-neutral surfaces,
1007\citet{gent.mcwilliams_JPO90} parameterisation, and various slightly diffusive advection schemes.
1008For momentum, the main ones are:
1009Laplacian and bilaplacian operators acting along geopotential surfaces,
1010and UBS advection schemes when flux form is chosen for the momentum advection.
1011
1012%% =================================================================================================
1013\subsubsection{Lateral laplacian tracer diffusive operator}
1014
1015The lateral Laplacian tracer diffusive operator is defined by (see \autoref{apdx:DIFFOPERS}):
1016\begin{equation}
1017  \label{eq:MB_iso_tensor}
1018  D^{lT} = \nabla \vect . \lt( A^{lT} \; \Re \; \nabla T \rt) \quad \text{with} \quad \Re =
1019  \begin{pmatrix}
1020    1    & 0    & -r_1          \\
1021    0    & 1    & -r_2          \\
1022    -r_1 & -r_2 & r_1^2 + r_2^2 \\
1023  \end{pmatrix}
1024\end{equation}
1025where $r_1$ and $r_2$ are the slopes between the surface along which the diffusive operator acts and
1026the model level (\eg\ $z$- or $s$-surfaces).
1027Note that the formulation \autoref{eq:MB_iso_tensor} is exact for
1028the rotation between geopotential and $s$-surfaces,
1029while it is only an approximation for the rotation between isoneutral and $z$- or $s$-surfaces.
1030Indeed, in the latter case, two assumptions are made to simplify \autoref{eq:MB_iso_tensor}
1031\citep{cox_OM87}.
1032First, the horizontal contribution of the dianeutral mixing is neglected since
1033the ratio between iso and dia-neutral diffusive coefficients is known to be
1034several orders of magnitude smaller than unity.
1035Second, the two isoneutral directions of diffusion are assumed to be independent since
1036the slopes are generally less than $10^{-2}$ in the ocean (see \autoref{apdx:DIFFOPERS}).
1037
1038For \textit{iso-level} diffusion, $r_1$ and $r_2 $ are zero.
1039$\Re$ reduces to the identity in the horizontal direction, no rotation is applied.
1040
1041For \textit{geopotential} diffusion,
1042$r_1$ and $r_2 $ are the slopes between the geopotential and computational surfaces:
1043they are equal to $\sigma_1$ and $\sigma_2$, respectively (see \autoref{eq:MB_sco_slope}).
1044
1045For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between
1046the isoneutral and computational surfaces.
1047Therefore, they are different quantities, but have similar expressions in $z$- and $s$-coordinates.
1048In $z$-coordinates:
1049\begin{equation}
1050  \label{eq:MB_iso_slopes}
1051  r_1 = \frac{e_3}{e_1} \lt( \pd[\rho]{i} \rt) \lt( \pd[\rho]{k} \rt)^{-1} \quad
1052  r_2 = \frac{e_3}{e_2} \lt( \pd[\rho]{j} \rt) \lt( \pd[\rho]{k} \rt)^{-1}
1053\end{equation}
1054while in $s$-coordinates $\pd[]{k}$ is replaced by $\pd[]{s}$.
1055
1056%% =================================================================================================
1057\subsubsection{Eddy induced velocity}
1058
1059When the \textit{eddy induced velocity} parametrisation (eiv) \citep{gent.mcwilliams_JPO90} is used,
1060an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
1061\[
1062  % \label{eq:MB_iso+eiv}
1063  D^{lT} = \nabla \cdot \lt( A^{lT} \; \Re \; \nabla T \rt) + \nabla \cdot \lt( \vect U^\ast \, T \rt)
1064\]
1065where $ \vect U^\ast = \lt( u^\ast,v^\ast,w^\ast \rt)$ is a non-divergent,
1066eddy-induced transport velocity. This velocity field is defined by:
1067\[
1068  % \label{eq:MB_eiv}
1069  u^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_1 \rt) \quad
1070  v^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_2 \rt) \quad
1071  w^\ast = - \frac{1}{e_1 e_2} \lt[   \pd[]{i} \lt( A^{eiv} \; e_2 \, \tilde{r}_1 \rt)
1072                                     + \pd[]{j} \lt( A^{eiv} \; e_1 \, \tilde{r}_2 \rt) \rt]
1073\]
1074where $A^{eiv}$ is the eddy induced velocity coefficient
1075(or equivalently the isoneutral thickness diffusivity coefficient),
1076and $\tilde r_1$ and $\tilde r_2$ are the slopes between
1077isoneutral and \textit{geopotential} surfaces.
1078Their values are thus independent of the vertical coordinate,
1079but their expression depends on the coordinate:
1080\begin{equation}
1081  \label{eq:MB_slopes_eiv}
1082  \tilde{r}_n =
1083    \begin{cases}
1084      r_n            & \text{in $z$-coordinate}                \\
1085      r_n + \sigma_n & \text{in \zstar- and $s$-coordinates}
1086    \end{cases}
1087  \quad \text{where~} n = 1, 2
1088\end{equation}
1089
1090The normal component of the eddy induced velocity is zero at all the boundaries.
1091This can be achieved in a model by tapering either the eddy coefficient or the slopes to zero in
1092the vicinity of the boundaries.
1093The latter strategy is used in \NEMO\ (cf. \autoref{chap:LDF}).
1094
1095%% =================================================================================================
1096\subsubsection{Lateral bilaplacian tracer diffusive operator}
1097
1098The lateral bilaplacian tracer diffusive operator is defined by:
1099\[
1100  % \label{eq:MB_bilapT}
1101  D^{lT}= - \Delta \; (\Delta T) \quad \text{where} \quad
1102  \Delta \bullet = \nabla \lt( \sqrt{B^{lT}} \; \Re \; \nabla \bullet \rt)
1103\]
1104It is the Laplacian operator given by \autoref{eq:MB_iso_tensor} applied twice with
1105the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1106
1107%% =================================================================================================
1108\subsubsection{Lateral Laplacian momentum diffusive operator}
1109
1110The Laplacian momentum diffusive operator along $z$- or $s$-surfaces is found by
1111applying \autoref{eq:MB_lap_vector} to the horizontal velocity vector (see \autoref{apdx:DIFFOPERS}):
1112\begin{align*}
1113  % \label{eq:MB_lapU}
1114  \vect D^{l \vect U} &=   \nabla_h        \big( A^{lm}    \chi             \big)
1115                         - \nabla_h \times \big( A^{lm} \, \zeta \; \vect k \big) \\
1116                      &= \lt(   \frac{1}{e_1}     \pd[ \lt( A^{lm}    \chi      \rt) ]{i} \rt.
1117                              - \frac{1}{e_2 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{j} ,
1118                                \frac{1}{e_2}     \pd[ \lt( A^{lm}    \chi      \rt) ]{j}
1119                         \lt. + \frac{1}{e_1 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{i} \rt)
1120\end{align*}
1121
1122Such a formulation ensures a complete separation between
1123the vorticity and horizontal divergence fields (see \autoref{apdx:INVARIANTS}).
1124Unfortunately, it is only available in \textit{iso-level} direction.
1125When a rotation is required
1126(\ie\ geopotential diffusion in $s$-coordinates or
1127isoneutral diffusion in both $z$- and $s$-coordinates),
1128the $u$ and $v$-fields are considered as independent scalar fields,
1129so that the diffusive operator is given by:
1130\[
1131  % \label{eq:MB_lapU_iso}
1132    D_u^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla u \rt) \quad
1133    D_v^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla v \rt)
1134\]
1135where $\Re$ is given by \autoref{eq:MB_iso_tensor}.
1136It is the same expression as those used for diffusive operator on tracers.
1137It must be emphasised that such a formulation is only exact in a Cartesian coordinate system,
1138\ie\ on a $f$- or $\beta$-plane, not on the sphere.
1139It is also a very good approximation in vicinity of the Equator in
1140a geographical coordinate system \citep{lengaigne.madec.ea_JGR03}.
1141
1142%% =================================================================================================
1143\subsubsection{Lateral bilaplacian momentum diffusive operator}
1144
1145As for tracers,
1146the bilaplacian order momentum diffusive operator is a re-entering Laplacian operator with
1147the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1148Nevertheless it is currently not available in the iso-neutral case.
1149
1150\subinc{\input{../../global/epilogue}}
1151
1152\end{document}
Note: See TracBrowser for help on using the repository browser.