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 @ 11596

Last change on this file since 11596 was 11596, checked in by nicolasmartin, 5 years ago

Application of some coding rules

  • Replace comments before sectioning cmds by a single line of 100 characters long to display when every line should break
  • Replace multi blank lines by one single blank line
  • For list environment, put \item, label and content on the same line
  • Remove \newpage and comments line around figure envs
File size: 64.6 KB
Line 
1
2\documentclass[../main/NEMO_manual]{subfiles}
3
4\begin{document}
5
6\chapter{Model Basics}
7\label{chap:MB}
8
9\chaptertoc
10
11%% =================================================================================================
12\section{Primitive equations}
13\label{sec:MB_PE}
14
15%% =================================================================================================
16\subsection{Vector invariant formulation}
17\label{subsec:MB_PE_vector}
18
19The ocean is a fluid that can be described to a good approximation by the primitive equations,
20\ie\ the Navier-Stokes equations along with a nonlinear equation of state which
21couples the two active tracers (temperature and salinity) to the fluid velocity,
22plus the following additional assumptions made from scale considerations:
23
24\begin{enumerate}
25\item \textit{spherical Earth approximation}: the geopotential surfaces are assumed to be oblate spheriods
26  that follow the Earth's bulge; these spheroids are approximated by spheres with
27  gravity locally vertical (parallel to the Earth's radius) and independent of latitude
28  \citep[][section 2]{white.hoskins.ea_QJRMS05}.
29\item \textit{thin-shell approximation}: the ocean depth is neglected compared to the earth's radius
30\item \textit{turbulent closure hypothesis}: the turbulent fluxes
31  (which represent the effect of small scale processes on the large-scale)
32  are expressed in terms of large-scale features
33\item \textit{Boussinesq hypothesis}: density variations are neglected except in their contribution to
34  the buoyancy force
35  \begin{equation}
36    \label{eq:MB_PE_eos}
37    \rho = \rho \ (T,S,p)
38  \end{equation}
39\item \textit{Hydrostatic hypothesis}: the vertical momentum equation is reduced to a balance between
40  the vertical pressure gradient and the buoyancy force
41  (this removes convective processes from the initial Navier-Stokes equations and so
42  convective processes must be parameterized instead)
43  \begin{equation}
44    \label{eq:MB_PE_hydrostatic}
45    \pd[p]{z} = - \rho \ g
46  \end{equation}
47\item \textit{Incompressibility hypothesis}: the three dimensional divergence of the velocity vector $\vect U$
48  is assumed to be zero.
49  \begin{equation}
50    \label{eq:MB_PE_continuity}
51    \nabla \cdot \vect U = 0
52  \end{equation}
53\item \textit{Neglect of additional Coriolis terms}: the Coriolis terms that vary with the cosine of latitude are neglected.
54  These terms may be non-negligible where the Brunt-Vaisala frequency $N$ is small, either in the deep ocean or
55  in the sub-mesoscale motions of the mixed layer, or near the equator \citep[][section 1]{white.hoskins.ea_QJRMS05}.
56  They can be consistently included as part of the ocean dynamics \citep[][section 3(d)]{white.hoskins.ea_QJRMS05} and are
57  retained in the MIT ocean model.
58\end{enumerate}
59
60Because the gravitational force is so dominant in the equations of large-scale motions,
61it is useful to choose an orthogonal set of unit vectors $(i,j,k)$ linked to the Earth such that
62$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
63\ie\ tangent to the geopotential surfaces.
64Let us define the following variables: $\vect U$ the vector velocity, $\vect U = \vect U_h + w \, \vect k$
65(the subscript $h$ denotes the local horizontal vector, \ie\ over the $(i,j)$ plane),
66$T$ the potential temperature, $S$ the salinity, $\rho$ the \textit{in situ} density.
67The vector invariant form of the primitive equations in the $(i,j,k)$ vector system provides
68the following equations:
69\begin{subequations}
70  \label{eq:MB_PE}
71  \begin{gather}
72    \intertext{$-$ the momentum balance}
73    \label{eq:MB_PE_dyn}
74    \pd[\vect U_h]{t} = - \lt[ (\nabla \times \vect U) \times \vect U + \frac{1}{2} \nabla \lt( \vect U^2 \rt) \rt]_h
75                        - f \; k \times \vect U_h - \frac{1}{\rho_o} \nabla_h p
76                        + \vect D^{\vect U} + \vect F^{\vect U} \\
77    \intertext{$-$ the heat and salt conservation equations}
78    \label{eq:MB_PE_tra_T}
79    \pd[T]{t} = - \nabla \cdot (T \ \vect U) + D^T + F^T \\
80    \label{eq:MB_PE_tra_S}
81    \pd[S]{t} = - \nabla \cdot (S \ \vect U) + D^S + F^S
82  \end{gather}
83\end{subequations}
84where $\nabla$ is the generalised derivative vector operator in $(i,j,k)$ directions, $t$ is the time,
85$z$ is the vertical coordinate, $\rho$ is the \textit{in situ} density given by the equation of state
86(\autoref{eq:MB_PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,
87$f = 2 \vect \Omega \cdot k$ is the Coriolis acceleration
88(where $\vect \Omega$ is the Earth's angular velocity vector), and $g$ is the gravitational acceleration.
89$\vect D^{\vect U}$, $D^T$ and $D^S$ are the parameterisations of small-scale physics for momentum,
90temperature and salinity, and $\vect F^{\vect U}$, $F^T$ and $F^S$ surface forcing terms.
91Their nature and formulation are discussed in \autoref{sec:MB_zdf_ldf} and \autoref{subsec:MB_boundary_condition}.
92
93%% =================================================================================================
94\subsection{Boundary conditions}
95\label{subsec:MB_boundary_condition}
96
97An ocean is bounded by complex coastlines, bottom topography at its base and
98an air-sea or ice-sea interface at its top.
99These boundaries can be defined by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,k,t)$,
100where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface
101(discretisation can introduce additional artificial ``side-wall'' boundaries).
102Both $H$ and $\eta$ are referenced to a surface of constant geopotential (\ie\ a mean sea surface height) on which $z = 0$.
103(\autoref{fig:MB_ocean_bc}).
104Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt, and momentum with
105the solid earth, the continental margins, the sea ice and the atmosphere.
106However, some of these fluxes are so weak that even on climatic time scales of thousands of years
107they can be neglected.
108In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and
109the other components of the earth system.
110
111\begin{figure}[!ht]
112  \centering
113  \includegraphics[width=0.66\textwidth]{Fig_I_ocean_bc}
114  \caption[Ocean boundary conditions]{
115    The ocean is bounded by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,t)$,
116    where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface.
117    Both $H$ and $\eta$ are referenced to $z = 0$.}
118  \label{fig:MB_ocean_bc}
119\end{figure}
120
121\begin{description}
122\item [Land - ocean interface:]  the major flux between continental margins and the ocean is a mass exchange of fresh water through river runoff.
123  Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths.
124  It can be neglected for short range integrations but has to be taken into account for long term integrations as
125  it influences the characteristics of water masses formed (especially at high latitudes).
126  It is required in order to close the water cycle of the climate system.
127  It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths.
128\item [Solid earth - ocean interface:]  heat and salt fluxes through the sea floor are small, except in special areas of little extent.
129  They are usually neglected in the model
130  \footnote{
131    In fact, it has been shown that the heat flux associated with the solid Earth cooling
132    (\ie\ the geothermal heating) is not negligible for the thermohaline circulation of the world ocean
133    (see \autoref{subsec:TRA_bbc}).
134  }.
135  The boundary condition is thus set to no flux of heat and salt across solid boundaries.
136  For momentum, the situation is different. There is no flow across solid boundaries,
137  \ie\ the velocity normal to the ocean bottom and coastlines is zero (in other words,
138  the bottom velocity is parallel to solid boundaries). This kinematic boundary condition
139  can be expressed as:
140  \begin{equation}
141    \label{eq:MB_w_bbc}
142    w = - \vect U_h \cdot \nabla_h (H)
143  \end{equation}
144  In addition, the ocean exchanges momentum with the earth through frictional processes.
145  Such momentum transfer occurs at small scales in a boundary layer.
146  It must be parameterized in terms of turbulent fluxes using bottom and/or lateral boundary conditions.
147  Its specification depends on the nature of the physical parameterisation used for
148  $\vect D^{\vect U}$ in \autoref{eq:MB_PE_dyn}.
149  It is discussed in \autoref{eq:MB_zdf}.% and Chap. III.6 to 9.
150\item [Atmosphere - ocean interface:]  the kinematic surface condition plus the mass flux of fresh water PE (the precipitation minus evaporation budget)
151  leads to:
152  \[
153    % \label{eq:MB_w_sbc}
154    w = \pd[\eta]{t} + \lt. \vect U_h \rt|_{z = \eta} \cdot \nabla_h (\eta) + P - E
155  \]
156  The dynamic boundary condition, neglecting the surface tension (which removes capillary waves from the system)
157  leads to the continuity of pressure across the interface $z = \eta$.
158  The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
159\item [Sea ice - ocean interface:]  the ocean and sea ice exchange heat, salt, fresh water and momentum.
160  The sea surface temperature is constrained to be at the freezing point at the interface.
161  Sea ice salinity is very low ($\sim4-6 \, psu$) compared to those of the ocean ($\sim34 \, psu$).
162  The cycle of freezing/melting is associated with fresh water and salt fluxes that cannot be neglected.
163\end{description}
164
165%% =================================================================================================
166\section{Horizontal pressure gradient}
167\label{sec:MB_hor_pg}
168
169%% =================================================================================================
170\subsection{Pressure formulation}
171\label{subsec:MB_p_formulation}
172
173The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at
174a reference geopotential surface ($z = 0$) and a hydrostatic pressure $p_h$ such that:
175$p(i,j,k,t) = p_s(i,j,t) + p_h(i,j,k,t)$.
176The latter is computed by integrating (\autoref{eq:MB_PE_hydrostatic}),
177assuming that pressure in decibars can be approximated by depth in meters in (\autoref{eq:MB_PE_eos}).
178The hydrostatic pressure is then given by:
179\[
180  % \label{eq:MB_pressure}
181  p_h (i,j,z,t) = \int_{\varsigma = z}^{\varsigma = 0} g \; \rho (T,S,\varsigma) \; d \varsigma
182\]
183Two strategies can be considered for the surface pressure term:
184$(a)$ introduce of a  new variable $\eta$, the free-surface elevation,
185for which a prognostic equation can be established and solved;
186$(b)$ assume that the ocean surface is a rigid lid,
187on which the pressure (or its horizontal gradient) can be diagnosed.
188When the former strategy is used, one solution of the free-surface elevation consists of
189the excitation of external gravity waves.
190The flow is barotropic and the surface moves up and down with gravity as the restoring force.
191The phase speed of such waves is high (some hundreds of metres per second) so that
192the time step has to be very short when they are present in the model.
193The latter strategy filters out these waves since the rigid lid approximation implies $\eta = 0$,
194\ie\ the sea surface is the surface $z = 0$.
195This well known approximation increases the surface wave speed to infinity and
196modifies certain other longwave dynamics (\eg\ barotropic Rossby or planetary waves).
197The rigid-lid hypothesis is an obsolescent feature in modern OGCMs.
198It has been available until the release 3.1 of \NEMO, and it has been removed in release 3.2 and followings.
199Only the free surface formulation is now described in this document (see the next sub-section).
200
201%% =================================================================================================
202\subsection{Free surface formulation}
203\label{subsec:MB_free_surface}
204
205In the free surface formulation, a variable $\eta$, the sea-surface height,
206is introduced which describes the shape of the air-sea interface.
207This variable is solution of a prognostic equation which is established by forming the vertical average of
208the kinematic surface condition (\autoref{eq:MB_w_bbc}):
209\begin{equation}
210  \label{eq:MB_ssh}
211  \pd[\eta]{t} = - D + P - E \quad \text{where} \quad D = \nabla \cdot \lt[ (H + \eta) \; \overline{U}_h \, \rt]
212\end{equation}
213and using (\autoref{eq:MB_PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
214
215Allowing the air-sea interface to move introduces the external gravity waves (EGWs) as
216a class of solution of the primitive equations.
217These waves are barotropic (\ie\ nearly independent of depth) and their phase speed is quite high.
218Their time scale is short with respect to the other processes described by the primitive equations.
219
220Two choices can be made regarding the implementation of the free surface in the model,
221depending on the physical processes of interest.
222
223$\bullet$ If one is interested in EGWs, in particular the tides and their interaction with
224the baroclinic structure of the ocean (internal waves) possibly in shallow seas,
225then a non linear free surface is the most appropriate.
226This means that no approximation is made in \autoref{eq:MB_ssh} and that
227the variation of the ocean volume is fully taken into account.
228Note that in order to study the fast time scales associated with EGWs it is necessary to
229minimize time filtering effects
230(use an explicit time scheme with very small time step, or a split-explicit scheme with reasonably small time step,
231see \autoref{subsec:DYN_spg_exp} or \autoref{subsec:DYN_spg_ts}).
232
233$\bullet$ If one is not interested in EGW but rather sees them as high frequency noise,
234it is possible to apply an explicit filter to slow down the fastest waves while
235not altering the slow barotropic Rossby waves.
236If further, an approximative conservation of heat and salt contents is sufficient for the problem solved,
237then it is sufficient to solve a linearized version of \autoref{eq:MB_ssh},
238which still allows to take into account freshwater fluxes applied at the ocean surface \citep{roullet.madec_JGR00}.
239Nevertheless, with the linearization, an exact conservation of heat and salt contents is lost.
240
241The filtering of EGWs in models with a free surface is usually a matter of discretisation of
242the temporal derivatives,
243using a split-explicit method \citep{killworth.webb.ea_JPO91, zhang.endoh_JGR92} or
244the implicit scheme \citep{dukowicz.smith_JGR94} or
245the addition of a filtering force in the momentum equation \citep{roullet.madec_JGR00}.
246With the present release, \NEMO\  offers the choice between
247an explicit free surface (see \autoref{subsec:DYN_spg_exp}) or
248a split-explicit scheme strongly inspired the one proposed by \citet{shchepetkin.mcwilliams_OM05}
249(see \autoref{subsec:DYN_spg_ts}).
250
251%% =================================================================================================
252\section{Curvilinear \textit{z-}coordinate system}
253\label{sec:MB_zco}
254
255%% =================================================================================================
256\subsection{Tensorial formalism}
257\label{subsec:MB_tensorial}
258
259In many ocean circulation problems, the flow field has regions of enhanced dynamics
260(\ie\ surface layers, western boundary currents, equatorial currents, or ocean fronts).
261The representation of such dynamical processes can be improved by
262specifically increasing the model resolution in these regions.
263As well, it may be convenient to use a lateral boundary-following coordinate system to
264better represent coastal dynamics.
265Moreover, the common geographical coordinate system has a singular point at the North Pole that
266cannot be easily treated in a global model without filtering.
267A solution consists of introducing an appropriate coordinate transformation that
268shifts the singular point onto land \citep{madec.imbard_CD96, murray_JCP96}.
269As a consequence, it is important to solve the primitive equations in various curvilinear coordinate systems.
270An efficient way of introducing an appropriate coordinate transform can be found when using a tensorial formalism.
271This formalism is suited to any multidimensional curvilinear coordinate system.
272Ocean modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth approximation),
273with preservation of the local vertical. Here we give the simplified equations for this particular case.
274The general case is detailed by \citet{eiseman.stone_SR80} in their survey of the conservation laws of fluid dynamics.
275
276Let $(i,j,k)$ be a set of orthogonal curvilinear coordinates on
277the sphere associated with the positively oriented orthogonal set of unit vectors
278$(i,j,k)$ linked to the earth such that
279$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
280\ie\ along geopotential surfaces (\autoref{fig:MB_referential}).
281Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by
282the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and
283the distance from the centre of the earth $a + z(k)$ where $a$ is the earth's radius and
284$z$ the altitude above a reference sea level (\autoref{fig:MB_referential}).
285The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$,
286the three scale factors:
287\begin{equation}
288  \label{eq:MB_scale_factors}
289  \begin{aligned}
290    e_1 &= (a + z) \lt[ \lt( \pd[\lambda]{i} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{i} \rt)^2 \rt]^{1/2} \\
291    e_2 &= (a + z) \lt[ \lt( \pd[\lambda]{j} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{j} \rt)^2 \rt]^{1/2} \\
292    e_3 &= \lt( \pd[z]{k} \rt)
293  \end{aligned}
294\end{equation}
295
296% >>>>>>>>>>>>>>>>>>>>>>>>>>>>
297\begin{figure}[!tb]
298  \centering
299  \includegraphics[width=0.66\textwidth]{Fig_I_earth_referential}
300  \caption[Geographical and curvilinear coordinate systems]{
301    the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear
302    coordinate system $(i,j,k)$.}
303  \label{fig:MB_referential}
304\end{figure}
305
306Since the ocean depth is far smaller than the earth's radius, $a + z$, can be replaced by $a$ in
307(\autoref{eq:MB_scale_factors}) (thin-shell approximation).
308The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while
309the vertical scale factor is a single function of $k$ as $k$ is parallel to $z$.
310The scalar and vector operators that appear in the primitive equations
311(\autoref{eq:MB_PE_dyn} to \autoref{eq:MB_PE_eos}) can then be written in the tensorial form,
312invariant in any orthogonal horizontal curvilinear coordinate system transformation:
313\begin{subequations}
314  % \label{eq:MB_discrete_operators}
315  \begin{gather}
316    \label{eq:MB_grad}
317    \nabla q =   \frac{1}{e_1} \pd[q]{i} \; \vect i
318               + \frac{1}{e_2} \pd[q]{j} \; \vect j
319               + \frac{1}{e_3} \pd[q]{k} \; \vect k \\
320    \label{eq:MB_div}
321    \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]
322                           + \frac{1}{e_3} \lt[ \pd[a_3]{k} \rt]
323  \end{gather}
324  \begin{multline}
325    \label{eq:MB_curl}
326      \nabla \times \vect{A} =   \lt[ \frac{1}{e_2} \pd[a_3]{j} - \frac{1}{e_3} \pd[a_2]{k}   \rt] \vect i \\
327                               + \lt[ \frac{1}{e_3} \pd[a_1]{k} - \frac{1}{e_1} \pd[a_3]{i}   \rt] \vect j \\
328                               + \frac{1}{e_1 e_2} \lt[ \pd[(e_2 a_2)]{i} - \pd[(e_1 a_1)]{j} \rt] \vect k
329  \end{multline}
330  \begin{gather}
331    \label{eq:MB_lap}
332    \Delta q = \nabla \cdot (\nabla q) \\
333    \label{eq:MB_lap_vector}
334    \Delta \vect A = \nabla (\nabla \cdot \vect A) - \nabla \times (\nabla \times \vect A)
335  \end{gather}
336\end{subequations}
337where $q$ is a scalar quantity and $\vect A = (a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinates system.
338
339%% =================================================================================================
340\subsection{Continuous model equations}
341\label{subsec:MB_zco_Eq}
342
343In order to express the Primitive Equations in tensorial formalism,
344it is necessary to compute the horizontal component of the non-linear and viscous terms of the equation using
345\autoref{eq:MB_grad}) to \autoref{eq:MB_lap_vector}.
346Let us set $\vect U = (u,v,w) = \vect U_h + w \; \vect k $, the velocity in the $(i,j,k)$ coordinates system and
347define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$, by:
348\begin{gather}
349  \label{eq:MB_curl_Uh}
350  \zeta = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, v)]{i} - \pd[(e_1 \, u)]{j} \rt] \\
351  \label{eq:MB_div_Uh}
352  \chi  = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, u)]{i} + \pd[(e_1 \, v)]{j} \rt]
353\end{gather}
354
355Using again the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that
356$e_3$  is a function of the single variable $k$,
357$NLT$ the nonlinear term of \autoref{eq:MB_PE_dyn} can be transformed as follows:
358\begin{alignat*}{2}
359  &NLT &=   &\lt[ (\nabla \times {\vect U}) \times {\vect U} + \frac{1}{2} \nabla \lt( {\vect U}^2 \rt) \rt]_h \\
360  &    &=   &\lt(
361    \begin{array}{*{20}c}
362                \lt[ \frac{1}{e_3} \pd[u]{k} - \frac{1}{e_1} \pd[w]{i} \rt] w - \zeta \; v   \\
363                \zeta \; u - \lt[ \frac{1}{e_2} \pd[w]{j} - \frac{1}{e_3} \pd[v]{k} \rt] \ w
364    \end{array}
365                                                                                             \rt)
366          + \frac{1}{2} \lt(
367    \begin{array}{*{20}c}
368                             \frac{1}{e_1} \pd[(u^2 + v^2 + w^2)]{i} \\
369                             \frac{1}{e_2} \pd[(u^2 + v^2 + w^2)]{j}
370    \end{array}
371                                                                     \rt) \\
372  &    &=   &\lt(
373    \begin{array}{*{20}c}
374                  -\zeta \; v \\
375                   \zeta \; u
376    \end{array}
377                              \rt)
378          + \frac{1}{2} \lt(
379    \begin{array}{*{20}c}
380                             \frac{1}{e_1} \pd[(u^2 + v^2)]{i} \\
381                             \frac{1}{e_2} \pd[(u^2 + v^2)]{j}
382    \end{array}
383                                                               \rt) \\
384  &    &  &+ \frac{1}{e_3} \lt(
385    \begin{array}{*{20}c}
386                                w \; \pd[u]{k} \\
387                                w \; \pd[v]{k}
388    \end{array}
389                                               \rt)
390           - \lt(
391    \begin{array}{*{20}c}
392                  \frac{w}{e_1} \pd[w]{i} - \frac{1}{2 e_1} \pd[w^2]{i} \\
393                  \frac{w}{e_2} \pd[w]{j} - \frac{1}{2 e_2} \pd[w^2]{j}
394    \end{array}
395                                                                        \rt)
396\end{alignat*}
397The last term of the right hand side is obviously zero, and thus the nonlinear term of
398\autoref{eq:MB_PE_dyn} is written in the $(i,j,k)$ coordinate system:
399\begin{equation}
400  \label{eq:MB_vector_form}
401  NLT =   \zeta \; \vect k \times \vect U_h + \frac{1}{2} \nabla_h \lt( \vect U_h^2 \rt)
402        + \frac{1}{e_3} w \pd[\vect U_h]{k}
403\end{equation}
404
405This is the so-called \textit{vector invariant form} of the momentum advection term.
406For some purposes, it can be advantageous to write this term in the so-called flux form,
407\ie\ to write it as the divergence of fluxes.
408For example, the first component of \autoref{eq:MB_vector_form} (the $i$-component) is transformed as follows:
409\begin{alignat*}{2}
410  &NLT_i &= &- \zeta \; v + \frac{1}{2 \; e_1} \pd[ (u^2 + v^2) ]{i} + \frac{1}{e_3} w \ \pd[u]{k} \\
411  &      &&\frac{1}{e_1 \; e_2} \lt( -v \pd[(e_2 \, v)]{i} + v \pd[(e_1 \, u)]{j} \rt)
412            + \frac{1}{e_1 e_2} \lt( e_2 \; u \pd[u]{i} + e_2 \; v \pd[v]{i} \rt) \\
413  &      & &+ \frac{1}{e_3} \lt( w \; \pd[u]{k} \rt) \\
414  &      &&\frac{1}{e_1 \; e_2} \lt[ - \lt( v^2 \pd[e_2]{i} + e_2 \, v \pd[v]{i} \rt)
415                                     + \lt( \pd[ \lt( e_1 \, u \, v \rt)]{j} -         e_1 \, u \pd[v]{j} \rt) \rt. \\
416  &      &                       &\lt. + \lt( \pd[ \lt( e_2 \, u \, u \rt)]{i} - u \pd[ \lt( e_2 u \rt)]{i} \rt)
417                                     + e_2 v \pd[v]{i}                                                         \rt] \\
418  &      & &+ \frac{1}{e_3} \lt( \pd[(w \, u)]{k} - u \pd[w]{k} \rt) \\
419  &      &&\frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, u \, v)]{j} \rt)
420            + \frac{1}{e_3} \pd[(w \, u)]{k} \\
421  &      & &+ \frac{1}{e_1 e_2} \lt[ - u \lt( \pd[(e_1 v)]{j} - v \, \pd[e_1]{j} \rt)
422                                  - u \pd[(e_2 u)]{i}                              \rt]
423            - \frac{1}{e_3} \pd[w]{k} u \\
424  &      & &+ \frac{1}{e_1 e_2} \lt( - v^2 \pd[e_2]{i} \rt) \\
425  &      &= &\nabla \cdot (\vect U \, u) - (\nabla \cdot \vect U) \ u
426            + \frac{1}{e_1 e_2} \lt( -v^2 \pd[e_2]{i} + u v \, \pd[e_1]{j} \rt) \\
427  \intertext{as $\nabla \cdot {\vect U} \; = 0$ (incompressibility) it becomes:}
428  &      &= &\, \nabla \cdot (\vect U \, u) + \frac{1}{e_1 e_2} \lt( v \; \pd[e_2]{i} - u \; \pd[e_1]{j} \rt) (-v)
429\end{alignat*}
430
431The flux form of the momentum advection term is therefore given by:
432\begin{equation}
433  \label{eq:MB_flux_form}
434  NLT =   \nabla \cdot \lt(
435    \begin{array}{*{20}c}
436                            \vect U \, u \\
437                            \vect U \, v
438    \end{array}
439                                         \rt)
440        + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \vect k \times \vect U_h
441\end{equation}
442
443The flux form has two terms,
444the first one is expressed as the divergence of momentum fluxes (hence the flux form name given to this formulation)
445and the second one is due to the curvilinear nature of the coordinate system used.
446The latter is called the \textit{metric} term and can be viewed as a modification of the Coriolis parameter:
447\[
448  % \label{eq:MB_cor+metric}
449  f \to f + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt)
450\]
451
452Note that in the case of geographical coordinate,
453\ie\ when $(i,j) \to (\lambda,\varphi)$ and $(e_1,e_2) \to (a \, \cos \varphi,a)$,
454we recover the commonly used modification of the Coriolis parameter $f \to f + (u / a) \tan \varphi$.
455
456To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be written in
457the following tensorial formalism:
458
459\begin{itemize}
460\item \textbf{Vector invariant form of the momentum equations}:
461  \begin{equation}
462    \label{eq:MB_dyn_vect}
463    \begin{split}
464    % \label{eq:MB_dyn_vect_u}
465      \pd[u]{t} = &+ (\zeta + f) \, v - \frac{1}{2 e_1} \pd[]{i} (u^2 + v^2)
466                   - \frac{1}{e_3} w \pd[u]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) \\
467                  &+ D_u^{\vect U} + F_u^{\vect U} \\
468      \pd[v]{t} = &- (\zeta + f) \, u - \frac{1}{2 e_2} \pd[]{j} (u^2 + v^2)
469                   - \frac{1}{e_3} w \pd[v]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) \\
470                  &+ D_v^{\vect U} + F_v^{\vect U}
471    \end{split}
472  \end{equation}
473\item \textbf{flux form of the momentum equations}:
474  % \label{eq:MB_dyn_flux}
475  \begin{multline*}
476    % \label{eq:MB_dyn_flux_u}
477    \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 \\
478                - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, v \, u)]{j} \rt) \\
479                - \frac{1}{e_3} \pd[(w \, u)]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt)
480                + D_u^{\vect U} + F_u^{\vect U}
481  \end{multline*}
482  \begin{multline*}
483    % \label{eq:MB_dyn_flux_v}
484    \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 \\
485                - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, v)]{i} + \pd[(e_1 \, v \, v)]{j} \rt) \\
486                - \frac{1}{e_3} \pd[(w \, v)]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt)
487                + D_v^{\vect U} + F_v^{\vect U}
488  \end{multline*}
489  where $\zeta$, the relative vorticity, is given by \autoref{eq:MB_curl_Uh} and $p_s$, the surface pressure,
490  is given by:
491  \[
492  % \label{eq:MB_spg}
493    p_s = \rho \,g \, \eta
494  \]
495  and $\eta$ is the solution of \autoref{eq:MB_ssh}.
496
497  The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
498  \[
499  % \label{eq:MB_w_diag}
500    \pd[w]{k} = - \chi \; e_3 \qquad
501  % \label{eq:MB_hp_diag}
502    \pd[p_h]{k} = - \rho \; g \; e_3
503  \]
504  where the divergence of the horizontal velocity, $\chi$ is given by \autoref{eq:MB_div_Uh}.
505
506\item \textbf{tracer equations}:
507  \begin{equation}
508  \begin{split}
509    \pd[T]{t} = & - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 T \, u)]{i} + \pd[(e_1 T \, v)]{j} \rt]
510                - \frac{1}{e_3} \pd[(T \, w)]{k} + D^T + F^T \\
511    \pd[S]{t} = & - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 S \, u)]{i} + \pd[(e_1 S \, v)]{j} \rt]
512                - \frac{1}{e_3} \pd[(S \, w)]{k} + D^S + F^S \\
513    \rho = & \rho \big( T,S,z(k) \big)
514  \end{split}
515  \end{equation}
516\end{itemize}
517
518The expression of $\vect D^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale parameterisation used.
519It will be defined in \autoref{eq:MB_zdf}.
520The nature and formulation of $\vect F^{\vect U}$, $F^T$ and $F^S$, the surface forcing terms,
521are discussed in \autoref{chap:SBC}.
522
523%% =================================================================================================
524\section{Curvilinear generalised vertical coordinate system}
525\label{sec:MB_gco}
526
527The ocean domain presents a huge diversity of situation in the vertical.
528First the ocean surface is a time dependent surface (moving surface).
529Second the ocean floor depends on the geographical position,
530varying from more than 6,000 meters in abyssal trenches to zero at the coast.
531Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.
532Therefore, in order to represent the ocean with respect to
533the first point a space and time dependent vertical coordinate that follows the variation of the sea surface height
534\eg\ an \zstar-coordinate;
535for the second point, a space variation to fit the change of bottom topography
536\eg\ a terrain-following or $\sigma$-coordinate;
537and for the third point, one will be tempted to use a space and time dependent coordinate that
538follows the isopycnal surfaces, \eg\ an isopycnic coordinate.
539
540In order to satisfy two or more constraints one can even be tempted to mixed these coordinate systems, as in
541HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and $\sigma$ at
542the ocean bottom) \citep{chassignet.smith.ea_JPO03} or
543OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and $\sigma$-coordinate elsewhere)
544\citep{madec.delecluse.ea_JPO96} among others.
545
546In fact one is totally free to choose any space and time vertical coordinate by
547introducing an arbitrary vertical coordinate :
548\begin{equation}
549  \label{eq:MB_s}
550  s = s(i,j,k,t)
551\end{equation}
552with the restriction that the above equation gives a single-valued monotonic relationship between $s$ and $k$,
553when $i$, $j$ and $t$ are held fixed.
554\autoref{eq:MB_s} is a transformation from the $(i,j,k,t)$ coordinate system with independent variables into
555the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through
556\autoref{eq:MB_s}.
557This so-called \textit{generalised vertical coordinate} \citep{kasahara_MWR74} is in fact
558an Arbitrary Lagrangian--Eulerian (ALE) coordinate.
559Indeed, one has a great deal of freedom in the choice of expression for $s$. The choice determines
560which part of the vertical velocity (defined from a fixed referential) will cross the levels (Eulerian part) and
561which part will be used to move them (Lagrangian part).
562The coordinate is also sometime referenced as an adaptive coordinate \citep{hofmeister.burchard.ea_OM10},
563since the coordinate system is adapted in the course of the simulation.
564Its most often used implementation is via an ALE algorithm,
565in which a pure lagrangian step is followed by regridding and remapping steps,
566the latter step implicitly embedding the vertical advection
567\citep{hirt.amsden.ea_JCP74, chassignet.smith.ea_JPO03, white.adcroft.ea_JCP09}.
568Here we follow the \citep{kasahara_MWR74} strategy:
569a regridding step (an update of the vertical coordinate) followed by an Eulerian step with
570an explicit computation of vertical advection relative to the moving s-surfaces.
571
572%\gmcomment{
573%A key point here is that the $s$-coordinate depends on $(i,j)$ ==> horizontal pressure gradient...
574The generalized vertical coordinates used in ocean modelling are not orthogonal,
575which contrasts with many other applications in mathematical physics.
576Hence, it is useful to keep in mind the following properties that may seem odd on initial encounter.
577
578The horizontal velocity in ocean models measures motions in the horizontal plane,
579perpendicular to the local gravitational field.
580That is, horizontal velocity is mathematically the same regardless of the vertical coordinate, be it geopotential,
581isopycnal, pressure, or terrain following.
582The key motivation for maintaining the same horizontal velocity component is that
583the hydrostatic and geostrophic balances are dominant in the large-scale ocean.
584Use of an alternative quasi -horizontal velocity, for example one oriented parallel to the generalized surface,
585would lead to unacceptable numerical errors.
586Correspondingly, the vertical direction is anti -parallel to the gravitational force in
587all of the coordinate systems.
588We do not choose the alternative of a quasi -vertical direction oriented normal to
589the surface of a constant generalized vertical coordinate.
590
591It is the method used to measure transport across the generalized vertical coordinate surfaces which differs between
592the vertical coordinate choices.
593That is, computation of the dia-surface velocity component represents the fundamental distinction between
594the various coordinates.
595In some models, such as geopotential, pressure, and terrain following, this transport is typically diagnosed from
596volume or mass conservation.
597In other models, such as isopycnal layered models, this transport is prescribed based on assumptions about
598the physical processes producing a flux across the layer interfaces.
599
600In this section we first establish the PE in the generalised vertical $s$-coordinate,
601then we discuss the particular cases available in \NEMO, namely $z$, \zstar, $s$, and \ztilde.
602%}
603
604%% =================================================================================================
605\subsection{\textit{S}-coordinate formulation}
606
607Starting from the set of equations established in \autoref{sec:MB_zco} for the special case $k = z$ and
608thus $e_3 = 1$, we introduce an arbitrary vertical coordinate $s = s(i,j,k,t)$,
609which includes $z$-, \zstar- and $\sigma$-coordinates as special cases
610($s = z$, $s = \zstar$, and $s = \sigma = z / H$ or $ = z / \lt( H + \eta \rt)$, resp.).
611A formal derivation of the transformed equations is given in \autoref{apdx:SCOORD}.
612Let us define the vertical scale factor by $e_3 = \partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ),
613and the slopes in the $(i,j)$ directions between $s$- and $z$-surfaces by:
614\begin{equation}
615  \label{eq:MB_sco_slope}
616  \sigma_1 = \frac{1}{e_1} \; \lt. \pd[z]{i} \rt|_s \quad \text{and} \quad
617  \sigma_2 = \frac{1}{e_2} \; \lt. \pd[z]{j} \rt|_s
618\end{equation}
619We also introduce $\omega$, a dia-surface velocity component, defined as the velocity
620relative to the moving $s$-surfaces and normal to them:
621\[
622  % \label{eq:MB_sco_w}
623  \omega = w -  \, \lt. \pd[z]{t} \rt|_s - \sigma_1 \, u - \sigma_2 \, v
624\]
625
626The equations solved by the ocean model \autoref{eq:MB_PE} in $s$-coordinate can be written as follows
627(see \autoref{sec:SCOORD_momentum}):
628
629\begin{itemize}
630\item \textbf{Vector invariant form of the momentum equation}:
631  \begin{multline*}
632  % \label{eq:MB_sco_u_vector}
633    \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} \\
634                - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_1
635                + D_u^{\vect U} + F_u^{\vect U}
636  \end{multline*}
637  \begin{multline*}
638  % \label{eq:MB_sco_v_vector}
639    \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} \\
640                - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_2
641                + D_v^{\vect U} + F_v^{\vect U}
642  \end{multline*}
643\item \textbf{Flux form of the momentum equation}:
644  \begin{multline*}
645  % \label{eq:MB_sco_u_flux}
646    \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 \\
647                                       - \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) \\
648                                       - \frac{1}{e_3} \pd[(\omega \, u)]{k}
649                                       - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt)
650                                       - g \frac{\rho}{\rho_o} \sigma_1 + D_u^{\vect U} + F_u^{\vect U}
651  \end{multline*}
652  \begin{multline*}
653  % \label{eq:MB_sco_v_flux}
654    \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 \\
655                                       - \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) \\
656                                       - \frac{1}{e_3} \pd[(\omega \, v)]{k}
657                                       - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt)
658                                       - g \frac{\rho}{\rho_o}\sigma_2 + D_v^{\vect U} + F_v^{\vect U}
659  \end{multline*}
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 \textit{tracer equations}:
670  \begin{multline*}
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}
673                                                                    + \pd[(e_1 e_3 \, v \, T)]{j} \rt) \\
674                                       - \frac{1}{e_3} \pd[(T \, \omega)]{k} + D^T + F^S
675  \end{multline*}
676  \begin{multline}
677  % \label{eq:MB_sco_s}
678    \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}
679                                                                    + \pd[(e_1 e_3 \, v \, S)]{j} \rt) \\
680                                       - \frac{1}{e_3} \pd[(S \, \omega)]{k} + D^S + F^S
681  \end{multline}
682\end{itemize}
683The equation of state has the same expression as in $z$-coordinate,
684and similar expressions are used for mixing and forcing terms.
685
686\gmcomment{
687  \colorbox{yellow}{ to be updated $= = >$}
688  Add a few works on z and zps and s and underlies the differences between all of them
689  \colorbox{yellow}{$< = =$ end update}
690}
691
692%% =================================================================================================
693\subsection{Curvilinear \zstar-coordinate system}
694\label{subsec:MB_zco_star}
695
696\begin{figure}[!b]
697  \centering
698  \includegraphics[width=0.66\textwidth]{Fig_z_zstar}
699  \caption[Curvilinear z-coordinate systems (\{non-\}linear free-surface cases and re-scaled \zstar)]{
700    (a) $z$-coordinate in linear free-surface case ;
701    (b) $z$-coordinate in non-linear free surface case ;
702    (c) re-scaled height coordinate
703    (become popular as the \zstar-coordinate \citep{adcroft.campin_OM04}).}
704  \label{fig:MB_z_zstar}
705\end{figure}
706
707In this case, the free surface equation is nonlinear, and the variations of volume are fully taken into account.
708These coordinates systems is presented in a report \citep{levier.treguier.ea_rpt07} available on the \NEMO\ web site.
709
710The \zstar coordinate approach is an unapproximated, non-linear free surface implementation which allows one to
711deal with large amplitude free-surface variations relative to the vertical resolution \citep{adcroft.campin_OM04}.
712In the \zstar formulation,
713the variation of the column thickness due to sea-surface undulations is not concentrated in the surface level,
714as in the $z$-coordinate formulation, but 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}, the bottom-following $z$ coordinate and \zstar are equivalent.
718The definition and modified oceanic equations for the rescaled vertical coordinate \zstar,
719including the treatment of fresh-water flux at the surface, are detailed in Adcroft and Campin (2004).
720The major points are summarized here.
721The position (\zstar) and vertical discretization (\zstar) are expressed as:
722\[
723  % \label{eq:MB_z-star}
724  H + \zstar = (H + z)  / r \quad \text{and}  \quad \delta \zstar
725              = \delta z / r \quad \text{with} \quad r
726              = \frac{H + \eta}{H} .
727\]
728Simple re-organisation of the above expressions gives
729\[
730  % \label{eq:MB_zstar_2}
731  \zstar = H \lt( \frac{z - \eta}{H + \eta} \rt) .
732\]
733Since the vertical displacement of the free surface is incorporated in the vertical coordinate \zstar,
734the upper and lower boundaries are at fixed  \zstar position,
735$\zstar = 0$ and $\zstar = -H$ respectively.
736Also the divergence of the flow field is no longer zero as shown by the continuity equation:
737\[
738  \pd[r]{t} = \nabla_{\zstar} \cdot \lt( r \; \vect U_h \rt) + \pd[r \; w^*]{\zstar} = 0 .
739\]
740This \zstar coordinate is closely related to the "eta" coordinate used in many atmospheric models
741(see Black (1994) for a review of eta coordinate atmospheric models).
742It was originally used in ocean models by Stacey et al. (1995) for studies of tides next to shelves,
743and it has been recently promoted by Adcroft and Campin (2004) for global climate modelling.
744
745The surfaces of constant \zstar are quasi -horizontal.
746Indeed, the \zstar coordinate reduces to $z$ when $\eta$ is zero.
747In general, when noting the large differences between
748undulations of the bottom topography versus undulations in the surface height,
749it is clear that surfaces constant \zstar are very similar to the depth surfaces.
750These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to
751terrain following sigma models discussed in \autoref{subsec:MB_sco}.
752Additionally, since $\zstar = z$ when $\eta = 0$,
753no flow is spontaneously generated in an unforced ocean starting from rest, regardless the bottom topography.
754This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in the presence of
755nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state,
756depending on the sophistication of the pressure gradient solver.
757The quasi -horizontal nature of the coordinate surfaces also facilitates the implementation of
758neutral physics parameterizations in \zstar  models using the same techniques as in $z$-models
759(see Chapters 13-16 of \cite{griffies_bk04}) for a discussion of neutral physics in $z$-models,
760as well as \autoref{sec:LDF_slp} in this document for treatment in \NEMO).
761
762The range over which \zstar  varies is time independent $-H \leq \zstar \leq 0$.
763Hence, all cells remain nonvanishing, so long as the surface height maintains $\eta > -H$.
764This is a minor constraint relative to that encountered on the surface height when using $s = z$ or $s = z - \eta$.
765
766Because \zstar  has a time independent range, all grid cells have static increments ds,
767and the sum of the vertical increments yields the time independent ocean depth. %k ds = H.
768The \zstar coordinate is therefore invisible to undulations of the free surface,
769since it moves along with the free surface.
770This property means that no spurious vertical transport is induced across surfaces of constant \zstar  by
771the motion of external gravity waves.
772Such spurious transport can be a problem in z-models, especially those with tidal forcing.
773Quite generally, the time independent range for the \zstar  coordinate is a very convenient property that
774allows for a nearly arbitrary vertical resolution even in the presence of large amplitude fluctuations of
775the surface height, again so long as $\eta > -H$.
776%end MOM doc %%%
777
778%% =================================================================================================
779\subsection{Curvilinear terrain-following \textit{s}--coordinate}
780\label{subsec:MB_sco}
781
782%% =================================================================================================
783\subsubsection{Introduction}
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, sills and
787channels that strongly constrain the path of water masses, but more subtle effects exist.
788For example, the topographic $\beta$-effect is usually larger than the planetary one along continental slopes.
789Topographic Rossby waves can be excited and can interact with the mean current.
790In the $z$-coordinate system presented in the previous section (\autoref{sec:MB_zco}),
791$z$-surfaces are geopotential surfaces.
792The bottom topography is discretised by steps.
793This often leads to a misrepresentation of a gradually sloping bottom and to
794large localized depth gradients associated with large localized vertical velocities.
795The response to such a velocity field often leads to numerical dispersion effects.
796One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of
797a full step one \cite{pacanowski.gnanadesikan_MWR98}.
798Another solution is to introduce a terrain-following coordinate system (hereafter $s$-coordinate).
799
800The $s$-coordinate avoids the discretisation error in the depth field since the layers of
801computation are gradually adjusted with depth to the ocean bottom.
802Relatively small topographic features as well as  gentle, large-scale slopes of the sea floor in the deep ocean,
803which would be ignored in typical $z$-model applications with the largest grid spacing at greatest depths,
804can easily be represented (with relatively low vertical resolution).
805A terrain-following model (hereafter $s$-model) also facilitates the modelling of the boundary layer flows over
806a large depth range, which in the framework of the $z$-model would require high vertical resolution over
807the whole depth range.
808Moreover, with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as
809the only boundaries of the domain (no more lateral boundary condition to specify).
810Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean,
811it has strong limitations as soon as stratification is introduced.
812The main two problems come from the truncation error in the horizontal pressure gradient and
813a possibly increased diapycnal diffusion.
814The horizontal pressure force in $s$-coordinate consists of two terms (see \autoref{apdx:SCOORD}),
815
816\begin{equation}
817  \label{eq:MB_p_sco}
818  \nabla p |_z = \nabla p |_s - \frac{1}{e_3} \pd[p]{s} \nabla z |_s
819\end{equation}
820
821The second term in \autoref{eq:MB_p_sco} depends on the tilt of the coordinate surface and
822leads to a truncation error that is not present in a $z$-model.
823In the special case of a $\sigma$-coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),
824\citet{haney_JPO91} and \citet{beckmann.haidvogel_JPO93} have given estimates of the magnitude of this truncation error.
825It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state,
826and the finite difference scheme.
827This error limits the possible topographic slopes that a model can handle at
828a given horizontal and vertical resolution.
829This is a severe restriction for large-scale applications using realistic bottom topography.
830The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive.
831This problem can be at least partially overcome by mixing $s$-coordinate and
832step-like representation of bottom topography \citep{gerdes_JGR93*a,gerdes_JGR93*b,madec.delecluse.ea_JPO96}.
833However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for
834a realistic bottom topography:
835an envelope topography is defined in $s$-coordinate on which a full or
836partial step bottom topography is then applied in order to adjust the model depth to the observed one
837(see \autoref{subsec:DOM_zgr}.
838
839For numerical reasons a minimum of diffusion is required along the coordinate surfaces of
840any finite difference model.
841It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces.
842This is the case for a $z$-model as well as for a $s$-model.
843However, density varies more strongly on $s$-surfaces than on horizontal surfaces in regions of
844large topographic slopes, implying larger diapycnal diffusion in a $s$-model than in a $z$-model.
845Whereas such a diapycnal diffusion in a $z$-model tends to weaken horizontal density (pressure) gradients and thus
846the horizontal circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.
847For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification.
848Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography,
849and thus will generate there a baroclinic eddy.
850In contrast, the ocean will stay at rest in a $z$-model.
851As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below
852the strongly stratified portion of the water column (\ie\ the main thermocline) \citep{madec.delecluse.ea_JPO96}.
853An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces
854(see \autoref{subsec:MB_ldf}).
855Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,
856strongly exceeding the stability limit of such a operator when it is discretized (see \autoref{chap:LDF}).
857
858The $s$-coordinates introduced here \citep{lott.madec.ea_OM90,madec.delecluse.ea_JPO96} differ mainly in two aspects from
859similar models:
860it allows a representation of bottom topography with mixed full or partial step-like/terrain following topography;
861It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate.
862
863%% =================================================================================================
864\subsection{\texorpdfstring{Curvilinear \ztilde-coordinate}{}}
865\label{subsec:MB_zco_tilde}
866
867The \ztilde -coordinate has been developed by \citet{leclair.madec_OM11}.
868It is available in \NEMO\ since the version 3.4 and is more robust in version 4.0 than previously.
869Nevertheless, it is currently not robust enough to be used in all possible configurations.
870Its use is therefore not recommended.
871
872%% =================================================================================================
873\section{Subgrid scale physics}
874\label{sec:MB_zdf_ldf}
875
876The hydrostatic primitive equations describe the behaviour of a geophysical fluid at space and time scales larger than
877a few kilometres in the horizontal, a few meters in the vertical and a few minutes.
878They are usually solved at larger scales: the specified grid spacing and time step of the numerical model.
879The effects of smaller scale motions (coming from the advective terms in the Navier-Stokes equations)
880must be represented entirely in terms of large-scale patterns to close the equations.
881These effects appear in the equations as the divergence of turbulent fluxes
882(\ie\ fluxes associated with the mean correlation of small scale perturbations).
883Assuming a turbulent closure hypothesis is equivalent to choose a formulation for these fluxes.
884It is usually called the subgrid scale physics.
885It must be emphasized that this is the weakest part of the primitive equations,
886but also one of the most important for long-term simulations as
887small scale processes \textit{in fine} balance the surface input of kinetic energy and heat.
888
889The control exerted by gravity on the flow induces a strong anisotropy between the lateral and vertical motions.
890Therefore subgrid-scale physics \textbf{D}$^{\vect U}$, $D^{S}$ and $D^{T}$  in
891\autoref{eq:MB_PE_dyn}, \autoref{eq:MB_PE_tra_T} and \autoref{eq:MB_PE_tra_S} are divided into
892a lateral part \textbf{D}$^{l \vect U}$, $D^{l S}$ and $D^{l T}$ and
893a vertical part \textbf{D}$^{v \vect U}$, $D^{v S}$ and $D^{v T}$.
894The formulation of these terms and their underlying physics are briefly discussed in the next two subsections.
895
896%% =================================================================================================
897\subsection{Vertical subgrid scale physics}
898\label{subsec:MB_zdf}
899
900The model resolution is always larger than the scale at which the major sources of vertical turbulence occur
901(shear instability, internal wave breaking...).
902Turbulent motions are thus never explicitly solved, even partially, but always parameterized.
903The vertical turbulent fluxes are assumed to depend linearly on the gradients of large-scale quantities
904(for example, the turbulent heat flux is given by $\overline{T' w'} = -A^{v T} \partial_z \overline T$,
905where $A^{v T}$ is an eddy coefficient).
906This formulation is analogous to that of molecular diffusion and dissipation.
907This is quite clearly a necessary compromise: considering only the molecular viscosity acting on
908large scale severely underestimates the role of turbulent diffusion and dissipation,
909while an accurate consideration of the details of turbulent motions is simply impractical.
910The resulting vertical momentum and tracer diffusive operators are of second order:
911\begin{equation}
912  \label{eq:MB_zdf}
913  \begin{gathered}
914    \vect D^{v \vect U} = \pd[]{z} \lt( A^{vm} \pd[\vect U_h]{z} \rt) \ , \\
915          D^{vT}       = \pd[]{z} \lt( A^{vT} \pd[T]{z}         \rt) \quad \text{and} \quad
916          D^{vS}       = \pd[]{z} \lt( A^{vT} \pd[S]{z}         \rt)
917  \end{gathered}
918\end{equation}
919where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients, respectively.
920At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified
921(see \autoref{chap:SBC} and \autoref{chap:ZDF} and \autoref{sec:TRA_bbl}).
922All the vertical physics is embedded in the specification of the eddy coefficients.
923They can be assumed to be either constant, or function of the local fluid properties
924(\eg\ Richardson number, Brunt-Vais\"{a}l\"{a} frequency, distance from the boundary ...),
925or computed from a turbulent closure model.
926The choices available in \NEMO\ are discussed in \autoref{chap:ZDF}).
927
928%% =================================================================================================
929\subsection{Formulation of the lateral diffusive and viscous operators}
930\label{subsec:MB_ldf}
931
932Lateral turbulence can be roughly divided into a mesoscale turbulence associated with eddies
933(which can be solved explicitly if the resolution is sufficient since
934their underlying physics are included in the primitive equations),
935and a sub mesoscale turbulence which is never explicitly solved even partially, but always parameterized.
936The formulation of lateral eddy fluxes depends on whether the mesoscale is below or above the grid-spacing
937(\ie\ the model is eddy-resolving or not).
938
939In non-eddy-resolving configurations, the closure is similar to that used for the vertical physics.
940The lateral turbulent fluxes are assumed to depend linearly on the lateral gradients of large-scale quantities.
941The resulting lateral diffusive and dissipative operators are of second order.
942Observations show that lateral mixing induced by mesoscale turbulence tends to be along isopycnal surfaces
943(or more precisely neutral surfaces \cite{mcdougall_JPO87}) rather than across them.
944As the slope of neutral surfaces is small in the ocean, a common approximation is to assume that
945the `lateral' direction is the horizontal, \ie\ the lateral mixing is performed along geopotential surfaces.
946This leads to a geopotential second order operator for lateral subgrid scale physics.
947This assumption can be relaxed: the eddy-induced turbulent fluxes can be better approached by assuming that
948they depend linearly on the gradients of large-scale quantities computed along neutral surfaces.
949In such a case, the diffusive operator is an isoneutral second order operator and
950it has components in the three space directions.
951However,
952both horizontal and isoneutral operators have no effect on mean (\ie\ large scale) potential energy whereas
953potential energy is a main source of turbulence (through baroclinic instabilities).
954\citet{gent.mcwilliams_JPO90} proposed a parameterisation of mesoscale eddy-induced turbulence which
955associates an eddy-induced velocity to the isoneutral diffusion.
956Its mean effect is to reduce the mean potential energy of the ocean.
957This leads to a formulation of lateral subgrid-scale physics made up of an isoneutral second order operator and
958an eddy induced advective part.
959In all these lateral diffusive formulations,
960the specification of the lateral eddy coefficients remains the problematic point as
961there is no really satisfactory formulation of these coefficients as a function of large-scale features.
962
963In eddy-resolving configurations, a second order operator can be used,
964but usually the more scale selective biharmonic operator is preferred as
965the grid-spacing is usually not small enough compared to the scale of the eddies.
966The role devoted to the subgrid-scale physics is to dissipate the energy that
967cascades toward the grid scale and thus to ensure the stability of the model while
968not interfering with the resolved mesoscale activity.
969Another approach is becoming more and more popular:
970instead of specifying explicitly a sub-grid scale term in the momentum and tracer time evolution equations,
971one uses an advective scheme which is diffusive enough to maintain the model stability.
972It must be emphasised that then, all the sub-grid scale physics is included in the formulation of
973the advection scheme.
974
975All these parameterisations of subgrid scale physics have advantages and drawbacks.
976They are not all available in \NEMO. For active tracers (temperature and salinity) the main ones are:
977Laplacian and bilaplacian operators acting along geopotential or iso-neutral surfaces,
978\citet{gent.mcwilliams_JPO90} parameterisation, and various slightly diffusive advection schemes.
979For momentum, the main ones are: Laplacian and bilaplacian operators acting along geopotential surfaces,
980and UBS advection schemes when flux form is chosen for the momentum advection.
981
982%% =================================================================================================
983\subsubsection{Lateral laplacian tracer diffusive operator}
984
985The lateral Laplacian tracer diffusive operator is defined by (see \autoref{apdx:DIFFOPERS}):
986\begin{equation}
987  \label{eq:MB_iso_tensor}
988  D^{lT} = \nabla \vect . \lt( A^{lT} \; \Re \; \nabla T \rt) \quad \text{with} \quad
989  \Re =
990    \begin{pmatrix}
991      1    & 0    & -r_1          \\
992      0    & 1    & -r_2          \\
993      -r_1 & -r_2 & r_1^2 + r_2^2 \\
994    \end{pmatrix}
995\end{equation}
996where $r_1$ and $r_2$ are the slopes between the surface along which the diffusive operator acts and
997the model level (\eg\ $z$- or $s$-surfaces).
998Note that the formulation \autoref{eq:MB_iso_tensor} is exact for
999the rotation between geopotential and $s$-surfaces,
1000while it is only an approximation for the rotation between isoneutral and $z$- or $s$-surfaces.
1001Indeed, in the latter case, two assumptions are made to simplify \autoref{eq:MB_iso_tensor} \citep{cox_OM87}.
1002First, the horizontal contribution of the dianeutral mixing is neglected since the ratio between iso and
1003dia-neutral diffusive coefficients is known to be several orders of magnitude smaller than unity.
1004Second, the two isoneutral directions of diffusion are assumed to be independent since
1005the slopes are generally less than $10^{-2}$ in the ocean (see \autoref{apdx:DIFFOPERS}).
1006
1007For \textit{iso-level} diffusion, $r_1$ and $r_2 $ are zero.
1008$\Re$ reduces to the identity in the horizontal direction, no rotation is applied.
1009
1010For \textit{geopotential} diffusion,
1011$r_1$ and $r_2 $ are the slopes between the geopotential and computational surfaces:
1012they are equal to $\sigma_1$ and $\sigma_2$, respectively (see \autoref{eq:MB_sco_slope}).
1013
1014For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral and computational surfaces.
1015Therefore, they are different quantities, but have similar expressions in $z$- and $s$-coordinates.
1016In $z$-coordinates:
1017\begin{equation}
1018  \label{eq:MB_iso_slopes}
1019  r_1 = \frac{e_3}{e_1} \lt( \pd[\rho]{i} \rt) \lt( \pd[\rho]{k} \rt)^{-1} \quad
1020  r_2 = \frac{e_3}{e_2} \lt( \pd[\rho]{j} \rt) \lt( \pd[\rho]{k} \rt)^{-1}
1021\end{equation}
1022while in $s$-coordinates $\pd[]{k}$ is replaced by $\pd[]{s}$.
1023
1024%% =================================================================================================
1025\subsubsection{Eddy induced velocity}
1026
1027When the \textit{eddy induced velocity} parametrisation (eiv) \citep{gent.mcwilliams_JPO90} is used,
1028an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
1029\[
1030  % \label{eq:MB_iso+eiv}
1031  D^{lT} = \nabla \cdot \lt( A^{lT} \; \Re \; \nabla T \rt) + \nabla \cdot \lt( \vect U^\ast \, T \rt)
1032\]
1033where $ \vect U^\ast = \lt( u^\ast,v^\ast,w^\ast \rt)$ is a non-divergent,
1034eddy-induced transport velocity. This velocity field is defined by:
1035\begin{gather}
1036  % \label{eq:MB_eiv}
1037  u^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_1 \rt) \\
1038  v^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_2 \rt) \\
1039  w^\ast = - \frac{1}{e_1 e_2} \lt[   \pd[]{i} \lt( A^{eiv} \; e_2 \, \tilde{r}_1 \rt)
1040                                     + \pd[]{j} \lt( A^{eiv} \; e_1 \, \tilde{r}_2 \rt) \rt]
1041\end{gather}
1042where $A^{eiv}$ is the eddy induced velocity coefficient
1043(or equivalently the isoneutral thickness diffusivity coefficient),
1044and $\tilde r_1$ and $\tilde r_2$ are the slopes between isoneutral and \textit{geopotential} surfaces.
1045Their values are thus independent of the vertical coordinate, but their expression depends on the coordinate:
1046\begin{align}
1047  \label{eq:MB_slopes_eiv}
1048  \tilde{r}_n =
1049    \begin{cases}
1050      r_n            & \text{in $z$-coordinate}                \\
1051      r_n + \sigma_n & \text{in \zstar- and $s$-coordinates}
1052    \end{cases}
1053  \quad \text{where~} n = 1, 2
1054\end{align}
1055
1056The normal component of the eddy induced velocity is zero at all the boundaries.
1057This can be achieved in a model by tapering either the eddy coefficient or the slopes to zero in the vicinity of
1058the boundaries.
1059The latter strategy is used in \NEMO\ (cf. \autoref{chap:LDF}).
1060
1061%% =================================================================================================
1062\subsubsection{Lateral bilaplacian tracer diffusive operator}
1063
1064The lateral bilaplacian tracer diffusive operator is defined by:
1065\[
1066  % \label{eq:MB_bilapT}
1067  D^{lT}= - \Delta \; (\Delta T) \quad \text{where} \quad
1068  \Delta \bullet = \nabla \lt( \sqrt{B^{lT}} \; \Re \; \nabla \bullet \rt)
1069\]
1070It is the Laplacian operator given by \autoref{eq:MB_iso_tensor} applied twice with
1071the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1072
1073%% =================================================================================================
1074\subsubsection{Lateral Laplacian momentum diffusive operator}
1075
1076The Laplacian momentum diffusive operator along $z$- or $s$-surfaces is found by
1077applying \autoref{eq:MB_lap_vector} to the horizontal velocity vector (see \autoref{apdx:DIFFOPERS}):
1078\begin{align*}
1079  % \label{eq:MB_lapU}
1080  \vect D^{l \vect U} &=   \nabla_h        \big( A^{lm}    \chi             \big)
1081                         - \nabla_h \times \big( A^{lm} \, \zeta \; \vect k \big) \\
1082                      &= \lt(   \frac{1}{e_1}     \pd[ \lt( A^{lm}    \chi      \rt) ]{i} \rt.
1083                              - \frac{1}{e_2 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{j} ,
1084                                \frac{1}{e_2}     \pd[ \lt( A^{lm}    \chi      \rt) ]{j}
1085                         \lt. + \frac{1}{e_1 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{i} \rt)
1086\end{align*}
1087
1088Such a formulation ensures a complete separation between the vorticity and horizontal divergence fields
1089(see \autoref{apdx:INVARIANTS}).
1090Unfortunately, it is only available in \textit{iso-level} direction.
1091When a rotation is required
1092(\ie\ geopotential diffusion in $s$-coordinates or isoneutral diffusion in both $z$- and $s$-coordinates),
1093the $u$ and $v$-fields are considered as independent scalar fields, so that the diffusive operator is given by:
1094\begin{gather*}
1095  % \label{eq:MB_lapU_iso}
1096    D_u^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla u \rt) \\
1097    D_v^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla v \rt)
1098\end{gather*}
1099where $\Re$ is given by \autoref{eq:MB_iso_tensor}.
1100It is the same expression as those used for diffusive operator on tracers.
1101It must be emphasised that such a formulation is only exact in a Cartesian coordinate system,
1102\ie\ on a $f$- or $\beta$-plane, not on the sphere.
1103It is also a very good approximation in vicinity of the Equator in
1104a geographical coordinate system \citep{lengaigne.madec.ea_JGR03}.
1105
1106%% =================================================================================================
1107\subsubsection{Lateral bilaplacian momentum diffusive operator}
1108
1109As for tracers, the bilaplacian order momentum diffusive operator is a re-entering Laplacian operator with
1110the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1111Nevertheless it is currently not available in the iso-neutral case.
1112
1113\onlyinsubfile{\input{../../global/epilogue}}
1114
1115\end{document}
Note: See TracBrowser for help on using the repository browser.