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/branches/2019/dev_r10984_HPC-13_IRRMANN_BDY_optimization/doc/latex/NEMO/subfiles – NEMO

source: NEMO/branches/2019/dev_r10984_HPC-13_IRRMANN_BDY_optimization/doc/latex/NEMO/subfiles/chap_model_basics.tex @ 11512

Last change on this file since 11512 was 11512, checked in by smasson, 5 years ago

dev_r10984_HPC-13 : merge with trunk@11511, see #2285

File size: 66.6 KB
Line 
1
2\documentclass[../main/NEMO_manual]{subfiles}
3
4\begin{document}
5
6% ================================================================
7% Chapter 1  Model Basics
8% ================================================================
9\chapter{Model Basics}
10\label{chap:PE}
11
12\chaptertoc
13
14\newpage
15
16% ================================================================
17% Primitive Equations
18% ================================================================
19\section{Primitive equations}
20\label{sec:PE_PE}
21
22% -------------------------------------------------------------------------------------------------------------
23%        Vector Invariant Formulation
24% -------------------------------------------------------------------------------------------------------------
25
26\subsection{Vector invariant formulation}
27\label{subsec:PE_Vector}
28
29The ocean is a fluid that can be described to a good approximation by the primitive equations,
30\ie\ the Navier-Stokes equations along with a nonlinear equation of state which
31couples the two active tracers (temperature and salinity) to the fluid velocity,
32plus the following additional assumptions made from scale considerations:
33
34\begin{enumerate}
35\item
36  \textit{spherical Earth approximation}: the geopotential surfaces are assumed to be oblate spheriods
37  that follow the Earth's bulge; these spheroids are approximated by spheres with
38  gravity locally vertical (parallel to the Earth's radius) and independent of latitude
39  \citep[][section 2]{white.hoskins.ea_QJRMS05}.
40\item
41  \textit{thin-shell approximation}: the ocean depth is neglected compared to the earth's radius
42\item
43  \textit{turbulent closure hypothesis}: the turbulent fluxes
44  (which represent the effect of small scale processes on the large-scale)
45  are expressed in terms of large-scale features
46\item
47  \textit{Boussinesq hypothesis}: density variations are neglected except in their contribution to
48  the buoyancy force
49  \begin{equation}
50    \label{eq:PE_eos}
51    \rho = \rho \ (T,S,p)
52  \end{equation}
53\item
54  \textit{Hydrostatic hypothesis}: the vertical momentum equation is reduced to a balance between
55  the vertical pressure gradient and the buoyancy force
56  (this removes convective processes from the initial Navier-Stokes equations and so
57  convective processes must be parameterized instead)
58  \begin{equation}
59    \label{eq:PE_hydrostatic}
60    \pd[p]{z} = - \rho \ g
61  \end{equation}
62\item
63  \textit{Incompressibility hypothesis}: the three dimensional divergence of the velocity vector $\vect U$
64  is assumed to be zero.
65  \begin{equation}
66    \label{eq:PE_continuity}
67    \nabla \cdot \vect U = 0
68  \end{equation}
69 \item
70  \textit{Neglect of additional Coriolis terms}: the Coriolis terms that vary with the cosine of latitude are neglected.
71  These terms may be non-negligible where the Brunt-Vaisala frequency $N$ is small, either in the deep ocean or
72  in the sub-mesoscale motions of the mixed layer, or near the equator \citep[][section 1]{white.hoskins.ea_QJRMS05}.
73  They can be consistently included as part of the ocean dynamics \citep[][section 3(d)]{white.hoskins.ea_QJRMS05} and are
74  retained in the MIT ocean model.
75\end{enumerate}
76
77Because the gravitational force is so dominant in the equations of large-scale motions,
78it is useful to choose an orthogonal set of unit vectors $(i,j,k)$ linked to the Earth such that
79$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
80\ie\ tangent to the geopotential surfaces.
81Let us define the following variables: $\vect U$ 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:PE}
88  \begin{gather}
89    \intertext{$-$ the momentum balance}
90    \label{eq: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
92                        - f \; k \times \vect U_h - \frac{1}{\rho_o} \nabla_h p
93                        + \vect D^{\vect U} + \vect F^{\vect U} \\
94    \intertext{$-$ the heat and salt conservation equations}
95    \label{eq:PE_tra_T}
96    \pd[T]{t} = - \nabla \cdot (T \ \vect U) + D^T + F^T \\
97    \label{eq:PE_tra_S}
98    \pd[S]{t} = - \nabla \cdot (S \ \vect U) + D^S + F^S
99  \end{gather}
100\end{subequations}
101where $\nabla$ is the generalised derivative vector operator in $(i,j,k)$ directions, $t$ is the time,
102$z$ is the vertical coordinate, $\rho$ is the \textit{in situ} density given by the equation of state
103(\autoref{eq:PE_eos}), $\rho_o$ is a reference density, $p$ the pressure,
104$f = 2 \vect \Omega \cdot k$ is the Coriolis acceleration
105(where $\vect \Omega$ is the Earth's angular velocity vector), and $g$ is the gravitational acceleration.
106$\vect D^{\vect U}$, $D^T$ and $D^S$ are the parameterisations of small-scale physics for momentum,
107temperature and salinity, and $\vect F^{\vect U}$, $F^T$ and $F^S$ surface forcing terms.
108Their nature and formulation are discussed in \autoref{sec:PE_zdf_ldf} and \autoref{subsec:PE_boundary_condition}.
109
110% -------------------------------------------------------------------------------------------------------------
111% Boundary condition
112% -------------------------------------------------------------------------------------------------------------
113\subsection{Boundary conditions}
114\label{subsec:PE_boundary_condition}
115
116An ocean is bounded by complex coastlines, bottom topography at its base and
117an air-sea or ice-sea interface at its top.
118These boundaries can be defined by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,k,t)$,
119where $H$ is the depth of the ocean bottom and $\eta$ is the height of the sea surface
120(discretisation can introduce additional artificial ``side-wall'' boundaries).
121Both $H$ and $\eta$ are referenced to a surface of constant geopotential (\ie\ a mean sea surface height) on which $z = 0$.
122(\autoref{fig:ocean_bc}).
123Through these two boundaries, the ocean can exchange fluxes of heat, fresh water, salt, and momentum with
124the solid earth, the continental margins, the sea ice and the atmosphere.
125However, some of these fluxes are so weak that even on climatic time scales of thousands of years
126they can be neglected.
127In the following, we briefly review the fluxes exchanged at the interfaces between the ocean and
128the other components of the earth system.
129
130%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
131\begin{figure}[!ht]
132  \begin{center}
133    \includegraphics[width=\textwidth]{Fig_I_ocean_bc}
134    \caption{
135      \protect\label{fig:ocean_bc}
136      The ocean is bounded by two surfaces, $z = - H(i,j)$ and $z = \eta(i,j,t)$,
137      where $H$ is the depth of the sea floor and $\eta$ the height of the sea surface.
138      Both $H$ and $\eta$ are referenced to $z = 0$.
139    }
140  \end{center}
141\end{figure}
142%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
143
144\begin{description}
145\item[Land - ocean interface:]
146  the major flux between continental margins and the ocean is a mass exchange of fresh water through river runoff.
147  Such an exchange modifies the sea surface salinity especially in the vicinity of major river mouths.
148  It can be neglected for short range integrations but has to be taken into account for long term integrations as
149  it influences the characteristics of water masses formed (especially at high latitudes).
150  It is required in order to close the water cycle of the climate system.
151  It is usually specified as a fresh water flux at the air-sea interface in the vicinity of river mouths.
152\item[Solid earth - ocean interface:]
153  heat and salt fluxes through the sea floor are small, except in special areas of little extent.
154  They are usually neglected in the model
155  \footnote{
156    In fact, it has been shown that the heat flux associated with the solid Earth cooling
157    (\ie\ the geothermal heating) is not negligible for the thermohaline circulation of the world ocean
158    (see \autoref{subsec:TRA_bbc}).
159  }.
160  The boundary condition is thus set to no flux of heat and salt across solid boundaries.
161  For momentum, the situation is different. There is no flow across solid boundaries,
162  \ie\ the velocity normal to the ocean bottom and coastlines is zero (in other words,
163  the bottom velocity is parallel to solid boundaries). This kinematic boundary condition
164  can be expressed as:
165  \begin{equation}
166    \label{eq:PE_w_bbc}
167    w = - \vect U_h \cdot \nabla_h (H)
168  \end{equation}
169  In addition, the ocean exchanges momentum with the earth through frictional processes.
170  Such momentum transfer occurs at small scales in a boundary layer.
171  It must be parameterized in terms of turbulent fluxes using bottom and/or lateral boundary conditions.
172  Its specification depends on the nature of the physical parameterisation used for
173  $\vect D^{\vect U}$ in \autoref{eq:PE_dyn}.
174  It is discussed in \autoref{eq:PE_zdf}.% and Chap. III.6 to 9.
175\item[Atmosphere - ocean interface:]
176  the kinematic surface condition plus the mass flux of fresh water PE (the precipitation minus evaporation budget)
177  leads to:
178  \[
179    % \label{eq:PE_w_sbc}
180    w = \pd[\eta]{t} + \lt. \vect U_h \rt|_{z = \eta} \cdot \nabla_h (\eta) + P - E
181  \]
182  The dynamic boundary condition, neglecting the surface tension (which removes capillary waves from the system)
183  leads to the continuity of pressure across the interface $z = \eta$.
184  The atmosphere and ocean also exchange horizontal momentum (wind stress), and heat.
185\item[Sea ice - ocean interface:]
186  the ocean and sea ice exchange heat, salt, fresh water and momentum.
187  The sea surface temperature is constrained to be at the freezing point at the interface.
188  Sea ice salinity is very low ($\sim4-6 \, psu$) compared to those of the ocean ($\sim34 \, psu$).
189  The cycle of freezing/melting is associated with fresh water and salt fluxes that cannot be neglected.
190\end{description}
191
192% ================================================================
193% The Horizontal Pressure Gradient
194% ================================================================
195\section{Horizontal pressure gradient}
196\label{sec:PE_hor_pg}
197
198% -------------------------------------------------------------------------------------------------------------
199% Pressure Formulation
200% -------------------------------------------------------------------------------------------------------------
201\subsection{Pressure formulation}
202\label{subsec:PE_p_formulation}
203
204The total pressure at a given depth $z$ is composed of a surface pressure $p_s$ at
205a reference geopotential surface ($z = 0$) and a hydrostatic pressure $p_h$ such that:
206$p(i,j,k,t) = p_s(i,j,t) + p_h(i,j,k,t)$.
207The latter is computed by integrating (\autoref{eq:PE_hydrostatic}),
208assuming that pressure in decibars can be approximated by depth in meters in (\autoref{eq:PE_eos}).
209The hydrostatic pressure is then given by:
210\[
211  % \label{eq:PE_pressure}
212  p_h (i,j,z,t) = \int_{\varsigma = z}^{\varsigma = 0} g \; \rho (T,S,\varsigma) \; d \varsigma
213\]
214Two strategies can be considered for the surface pressure term:
215$(a)$ introduce of a  new variable $\eta$, the free-surface elevation,
216for which a prognostic equation can be established and solved;
217$(b)$ assume that the ocean surface is a rigid lid,
218on which the pressure (or its horizontal gradient) can be diagnosed.
219When the former strategy is used, one solution of the free-surface elevation consists of
220the excitation of external gravity waves.
221The flow is barotropic and the surface moves up and down with gravity as the restoring force.
222The phase speed of such waves is high (some hundreds of metres per second) so that
223the time step has to be very short when they are present in the model.
224The latter strategy filters out these waves since the rigid lid approximation implies $\eta = 0$,
225\ie\ the sea surface is the surface $z = 0$.
226This well known approximation increases the surface wave speed to infinity and
227modifies certain other longwave dynamics (\eg\ barotropic Rossby or planetary waves).
228The rigid-lid hypothesis is an obsolescent feature in modern OGCMs.
229It has been available until the release 3.1 of \NEMO, and it has been removed in release 3.2 and followings.
230Only the free surface formulation is now described in this document (see the next sub-section).
231
232% -------------------------------------------------------------------------------------------------------------
233% Free Surface Formulation
234% -------------------------------------------------------------------------------------------------------------
235\subsection{Free surface formulation}
236\label{subsec:PE_free_surface}
237
238In the free surface formulation, a variable $\eta$, the sea-surface height,
239is introduced which describes the shape of the air-sea interface.
240This variable is solution of a prognostic equation which is established by forming the vertical average of
241the kinematic surface condition (\autoref{eq:PE_w_bbc}):
242\begin{equation}
243  \label{eq:PE_ssh}
244  \pd[\eta]{t} = - D + P - E \quad \text{where} \quad D = \nabla \cdot \lt[ (H + \eta) \; \overline{U}_h \, \rt]
245\end{equation}
246and using (\autoref{eq:PE_hydrostatic}) the surface pressure is given by: $p_s = \rho \, g \, \eta$.
247
248Allowing the air-sea interface to move introduces the external gravity waves (EGWs) as
249a class of solution of the primitive equations.
250These waves are barotropic (\ie\ nearly independent of depth) and their phase speed is quite high.
251Their time scale is short with respect to the other processes described by the primitive equations.
252
253Two choices can be made regarding the implementation of the free surface in the model,
254depending on the physical processes of interest.
255
256$\bullet$ If one is interested in EGWs, in particular the tides and their interaction with
257the baroclinic structure of the ocean (internal waves) possibly in shallow seas,
258then a non linear free surface is the most appropriate.
259This means that no approximation is made in \autoref{eq:PE_ssh} and that
260the variation of the ocean volume is fully taken into account.
261Note that in order to study the fast time scales associated with EGWs it is necessary to
262minimize time filtering effects
263(use an explicit time scheme with very small time step, or a split-explicit scheme with reasonably small time step,
264see \autoref{subsec:DYN_spg_exp} or \autoref{subsec:DYN_spg_ts}).
265
266$\bullet$ If one is not interested in EGW but rather sees them as high frequency noise,
267it is possible to apply an explicit filter to slow down the fastest waves while
268not altering the slow barotropic Rossby waves.
269If further, an approximative conservation of heat and salt contents is sufficient for the problem solved,
270then it is sufficient to solve a linearized version of \autoref{eq:PE_ssh},
271which still allows to take into account freshwater fluxes applied at the ocean surface \citep{roullet.madec_JGR00}.
272Nevertheless, with the linearization, an exact conservation of heat and salt contents is lost.
273
274The filtering of EGWs in models with a free surface is usually a matter of discretisation of
275the temporal derivatives,
276using a split-explicit method \citep{killworth.webb.ea_JPO91, zhang.endoh_JGR92} or
277the implicit scheme \citep{dukowicz.smith_JGR94} or
278the addition of a filtering force in the momentum equation \citep{roullet.madec_JGR00}.
279With the present release, \NEMO\  offers the choice between
280an explicit free surface (see \autoref{subsec:DYN_spg_exp}) or
281a split-explicit scheme strongly inspired the one proposed by \citet{shchepetkin.mcwilliams_OM05}
282(see \autoref{subsec:DYN_spg_ts}).
283
284% ================================================================
285% Curvilinear z-coordinate System
286% ================================================================
287\section{Curvilinear \textit{z-}coordinate system}
288\label{sec:PE_zco}
289
290% -------------------------------------------------------------------------------------------------------------
291% Tensorial Formalism
292% -------------------------------------------------------------------------------------------------------------
293\subsection{Tensorial formalism}
294\label{subsec:PE_tensorial}
295
296In many ocean circulation problems, the flow field has regions of enhanced dynamics
297(\ie\ surface layers, western boundary currents, equatorial currents, or ocean fronts).
298The representation of such dynamical processes can be improved by
299specifically increasing the model resolution in these regions.
300As well, it may be convenient to use a lateral boundary-following coordinate system to
301better represent coastal dynamics.
302Moreover, the common geographical coordinate system has a singular point at the North Pole that
303cannot be easily treated in a global model without filtering.
304A solution consists of introducing an appropriate coordinate transformation that
305shifts the singular point onto land \citep{madec.imbard_CD96, murray_JCP96}.
306As a consequence, it is important to solve the primitive equations in various curvilinear coordinate systems.
307An efficient way of introducing an appropriate coordinate transform can be found when using a tensorial formalism.
308This formalism is suited to any multidimensional curvilinear coordinate system.
309Ocean modellers mainly use three-dimensional orthogonal grids on the sphere (spherical earth approximation),
310with preservation of the local vertical. Here we give the simplified equations for this particular case.
311The general case is detailed by \citet{eiseman.stone_SR80} in their survey of the conservation laws of fluid dynamics.
312
313Let $(i,j,k)$ be a set of orthogonal curvilinear coordinates on
314the sphere associated with the positively oriented orthogonal set of unit vectors
315$(i,j,k)$ linked to the earth such that
316$k$ is the local upward vector and $(i,j)$ are two vectors orthogonal to $k$,
317\ie\ along geopotential surfaces (\autoref{fig:referential}).
318Let $(\lambda,\varphi,z)$ be the geographical coordinate system in which a position is defined by
319the latitude $\varphi(i,j)$, the longitude $\lambda(i,j)$ and
320the distance from the centre of the earth $a + z(k)$ where $a$ is the earth's radius and
321$z$ the altitude above a reference sea level (\autoref{fig:referential}).
322The local deformation of the curvilinear coordinate system is given by $e_1$, $e_2$ and $e_3$,
323the three scale factors:
324\begin{equation}
325  \label{eq:scale_factors}
326  \begin{aligned}
327    e_1 &= (a + z) \lt[ \lt( \pd[\lambda]{i} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{i} \rt)^2 \rt]^{1/2} \\
328    e_2 &= (a + z) \lt[ \lt( \pd[\lambda]{j} \cos \varphi \rt)^2 + \lt( \pd[\varphi]{j} \rt)^2 \rt]^{1/2} \\
329    e_3 &= \lt( \pd[z]{k} \rt)
330  \end{aligned}
331\end{equation}
332
333% >>>>>>>>>>>>>>>>>>>>>>>>>>>>
334\begin{figure}[!tb]
335  \begin{center}
336    \includegraphics[width=\textwidth]{Fig_I_earth_referential}
337    \caption{
338      \protect\label{fig:referential}
339      the geographical coordinate system $(\lambda,\varphi,z)$ and the curvilinear
340      coordinate system $(i,j,k)$.
341    }
342  \end{center}
343\end{figure}
344%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
345
346Since the ocean depth is far smaller than the earth's radius, $a + z$, can be replaced by $a$ in
347(\autoref{eq:scale_factors}) (thin-shell approximation).
348The resulting horizontal scale factors $e_1$, $e_2$  are independent of $k$ while
349the vertical scale factor is a single function of $k$ as $k$ is parallel to $z$.
350The scalar and vector operators that appear in the primitive equations
351(\autoref{eq:PE_dyn} to \autoref{eq:PE_eos}) can then be written in the tensorial form,
352invariant in any orthogonal horizontal curvilinear coordinate system transformation:
353\begin{subequations}
354  % \label{eq:PE_discrete_operators}
355  \begin{gather}
356    \label{eq:PE_grad}
357    \nabla q =   \frac{1}{e_1} \pd[q]{i} \; \vect i
358               + \frac{1}{e_2} \pd[q]{j} \; \vect j
359               + \frac{1}{e_3} \pd[q]{k} \; \vect k \\
360    \label{eq:PE_div}
361    \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]
362                           + \frac{1}{e_3} \lt[ \pd[a_3]{k} \rt]
363  \end{gather}
364  \begin{multline}
365    \label{eq:PE_curl}
366      \nabla \times \vect{A} =   \lt[ \frac{1}{e_2} \pd[a_3]{j} - \frac{1}{e_3} \pd[a_2]{k}   \rt] \vect i \\
367                               + \lt[ \frac{1}{e_3} \pd[a_1]{k} - \frac{1}{e_1} \pd[a_3]{i}   \rt] \vect j \\
368                               + \frac{1}{e_1 e_2} \lt[ \pd[(e_2 a_2)]{i} - \pd[(e_1 a_1)]{j} \rt] \vect k
369  \end{multline}
370  \begin{gather}
371    \label{eq:PE_lap}
372    \Delta q = \nabla \cdot (\nabla q) \\
373    \label{eq:PE_lap_vector}
374    \Delta \vect A = \nabla (\nabla \cdot \vect A) - \nabla \times (\nabla \times \vect A)
375  \end{gather}
376\end{subequations}
377where $q$ is a scalar quantity and $\vect A = (a_1,a_2,a_3)$ a vector in the $(i,j,k)$ coordinates system.
378
379% -------------------------------------------------------------------------------------------------------------
380% Continuous Model Equations
381% -------------------------------------------------------------------------------------------------------------
382\subsection{Continuous model equations}
383\label{subsec:PE_zco_Eq}
384
385In order to express the Primitive Equations in tensorial formalism,
386it is necessary to compute the horizontal component of the non-linear and viscous terms of the equation using
387\autoref{eq:PE_grad}) to \autoref{eq:PE_lap_vector}.
388Let us set $\vect U = (u,v,w) = \vect U_h + w \; \vect k $, the velocity in the $(i,j,k)$ coordinates system and
389define the relative vorticity $\zeta$ and the divergence of the horizontal velocity field $\chi$, by:
390\begin{gather}
391  \label{eq:PE_curl_Uh}
392  \zeta = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, v)]{i} - \pd[(e_1 \, u)]{j} \rt] \\
393  \label{eq:PE_div_Uh}
394  \chi  = \frac{1}{e_1 e_2} \lt[ \pd[(e_2 \, u)]{i} + \pd[(e_1 \, v)]{j} \rt]
395\end{gather}
396
397Using again the fact that the horizontal scale factors $e_1$ and $e_2$ are independent of $k$ and that
398$e_3$  is a function of the single variable $k$,
399$NLT$ the nonlinear term of \autoref{eq:PE_dyn} can be transformed as follows:
400\begin{alignat*}{2}
401  &NLT &=   &\lt[ (\nabla \times {\vect U}) \times {\vect U} + \frac{1}{2} \nabla \lt( {\vect U}^2 \rt) \rt]_h \\
402  &    &=   &\lt(
403    \begin{array}{*{20}c}
404                \lt[ \frac{1}{e_3} \pd[u]{k} - \frac{1}{e_1} \pd[w]{i} \rt] w - \zeta \; v   \\
405                \zeta \; u - \lt[ \frac{1}{e_2} \pd[w]{j} - \frac{1}{e_3} \pd[v]{k} \rt] \ w
406    \end{array}
407                                                                                             \rt)
408          + \frac{1}{2} \lt(
409    \begin{array}{*{20}c}
410                             \frac{1}{e_1} \pd[(u^2 + v^2 + w^2)]{i} \\
411                             \frac{1}{e_2} \pd[(u^2 + v^2 + w^2)]{j}
412    \end{array}
413                                                                     \rt) \\
414  &    &=   &\lt(
415    \begin{array}{*{20}c}
416                  -\zeta \; v \\
417                   \zeta \; u
418    \end{array}
419                              \rt)
420          + \frac{1}{2} \lt(
421    \begin{array}{*{20}c}
422                             \frac{1}{e_1} \pd[(u^2 + v^2)]{i} \\
423                             \frac{1}{e_2} \pd[(u^2 + v^2)]{j}
424    \end{array}
425                                                               \rt) \\
426  &    &  &+ \frac{1}{e_3} \lt(
427    \begin{array}{*{20}c}
428                                w \; \pd[u]{k} \\
429                                w \; \pd[v]{k}
430    \end{array}
431                                               \rt)
432           - \lt(
433    \begin{array}{*{20}c}
434                  \frac{w}{e_1} \pd[w]{i} - \frac{1}{2 e_1} \pd[w^2]{i} \\
435                  \frac{w}{e_2} \pd[w]{j} - \frac{1}{2 e_2} \pd[w^2]{j}
436    \end{array}
437                                                                        \rt)
438\end{alignat*}
439The last term of the right hand side is obviously zero, and thus the nonlinear term of
440\autoref{eq:PE_dyn} is written in the $(i,j,k)$ coordinate system:
441\begin{equation}
442  \label{eq:PE_vector_form}
443  NLT =   \zeta \; \vect k \times \vect U_h + \frac{1}{2} \nabla_h \lt( \vect U_h^2 \rt)
444        + \frac{1}{e_3} w \pd[\vect U_h]{k}
445\end{equation}
446
447This is the so-called \textit{vector invariant form} of the momentum advection term.
448For some purposes, it can be advantageous to write this term in the so-called flux form,
449\ie\ to write it as the divergence of fluxes.
450For example, the first component of \autoref{eq:PE_vector_form} (the $i$-component) is transformed as follows:
451\begin{alignat*}{2}
452  &NLT_i &= &- \zeta \; v + \frac{1}{2 \; e_1} \pd[ (u^2 + v^2) ]{i} + \frac{1}{e_3} w \ \pd[u]{k} \\
453  &      &&\frac{1}{e_1 \; e_2} \lt( -v \pd[(e_2 \, v)]{i} + v \pd[(e_1 \, u)]{j} \rt)
454            + \frac{1}{e_1 e_2} \lt( e_2 \; u \pd[u]{i} + e_2 \; v \pd[v]{i} \rt) \\
455  &      & &+ \frac{1}{e_3} \lt( w \; \pd[u]{k} \rt) \\
456  &      &&\frac{1}{e_1 \; e_2} \lt[ - \lt( v^2 \pd[e_2]{i} + e_2 \, v \pd[v]{i} \rt)
457                                     + \lt( \pd[ \lt( e_1 \, u \, v \rt)]{j} -         e_1 \, u \pd[v]{j} \rt) \rt. \\
458  &      &                       &\lt. + \lt( \pd[ \lt( e_2 \, u \, u \rt)]{i} - u \pd[ \lt( e_2 u \rt)]{i} \rt)
459                                     + e_2 v \pd[v]{i}                                                         \rt] \\
460  &      & &+ \frac{1}{e_3} \lt( \pd[(w \, u)]{k} - u \pd[w]{k} \rt) \\
461  &      &&\frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, u \, v)]{j} \rt)
462            + \frac{1}{e_3} \pd[(w \, u)]{k} \\
463  &      & &+ \frac{1}{e_1 e_2} \lt[ - u \lt( \pd[(e_1 v)]{j} - v \, \pd[e_1]{j} \rt)
464                                  - u \pd[(e_2 u)]{i}                              \rt]
465            - \frac{1}{e_3} \pd[w]{k} u \\
466  &      & &+ \frac{1}{e_1 e_2} \lt( - v^2 \pd[e_2]{i} \rt) \\
467  &      &= &\nabla \cdot (\vect U \, u) - (\nabla \cdot \vect U) \ u
468            + \frac{1}{e_1 e_2} \lt( -v^2 \pd[e_2]{i} + u v \, \pd[e_1]{j} \rt) \\
469  \intertext{as $\nabla \cdot {\vect U} \; = 0$ (incompressibility) it becomes:}
470  &      &= &\, \nabla \cdot (\vect U \, u) + \frac{1}{e_1 e_2} \lt( v \; \pd[e_2]{i} - u \; \pd[e_1]{j} \rt) (-v)
471\end{alignat*}
472
473The flux form of the momentum advection term is therefore given by:
474\begin{equation}
475  \label{eq:PE_flux_form}
476  NLT =   \nabla \cdot \lt(
477    \begin{array}{*{20}c}
478                            \vect U \, u \\
479                            \vect U \, v
480    \end{array}
481                                         \rt)
482        + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt) \vect k \times \vect U_h
483\end{equation}
484
485The flux form has two terms,
486the first one is expressed as the divergence of momentum fluxes (hence the flux form name given to this formulation)
487and the second one is due to the curvilinear nature of the coordinate system used.
488The latter is called the \textit{metric} term and can be viewed as a modification of the Coriolis parameter:
489\[
490  % \label{eq:PE_cor+metric}
491  f \to f + \frac{1}{e_1 e_2} \lt( v \pd[e_2]{i} - u \pd[e_1]{j} \rt)
492\]
493
494Note that in the case of geographical coordinate,
495\ie\ when $(i,j) \to (\lambda,\varphi)$ and $(e_1,e_2) \to (a \, \cos \varphi,a)$,
496we recover the commonly used modification of the Coriolis parameter $f \to f + (u / a) \tan \varphi$.
497
498To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be written in
499the following tensorial formalism:
500
501\begin{itemize}
502\item
503  \textbf{Vector invariant form of the momentum equations}:
504  \begin{equation}
505    \label{eq:PE_dyn_vect}
506    \begin{split}
507    % \label{eq:PE_dyn_vect_u}
508      \pd[u]{t} = &+ (\zeta + f) \, v - \frac{1}{2 e_1} \pd[]{i} (u^2 + v^2)
509                   - \frac{1}{e_3} w \pd[u]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) \\
510                  &+ D_u^{\vect U} + F_u^{\vect U} \\
511      \pd[v]{t} = &- (\zeta + f) \, u - \frac{1}{2 e_2} \pd[]{j} (u^2 + v^2)
512                   - \frac{1}{e_3} w \pd[v]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) \\
513                  &+ D_v^{\vect U} + F_v^{\vect U}
514    \end{split}
515  \end{equation}
516\item
517  \textbf{flux form of the momentum equations}:
518  % \label{eq:PE_dyn_flux}
519  \begin{multline*}
520    % \label{eq:PE_dyn_flux_u}
521    \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 \\
522                - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, u)]{i} + \pd[(e_1 \, v \, u)]{j} \rt) \\
523                - \frac{1}{e_3} \pd[(w \, u)]{k} - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt)
524                + D_u^{\vect U} + F_u^{\vect U}
525  \end{multline*}
526  \begin{multline*}
527    % \label{eq:PE_dyn_flux_v}
528    \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 \\
529                - \frac{1}{e_1 \; e_2} \lt( \pd[(e_2 \, u \, v)]{i} + \pd[(e_1 \, v \, v)]{j} \rt) \\
530                - \frac{1}{e_3} \pd[(w \, v)]{k} - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt)
531                + D_v^{\vect U} + F_v^{\vect U}
532  \end{multline*}
533  where $\zeta$, the relative vorticity, is given by \autoref{eq:PE_curl_Uh} and $p_s$, the surface pressure,
534  is given by:
535  \[
536  % \label{eq:PE_spg}
537    p_s = \rho \,g \, \eta
538  \]
539  and $\eta$ is the solution of \autoref{eq:PE_ssh}.
540
541  The vertical velocity and the hydrostatic pressure are diagnosed from the following equations:
542  \[
543  % \label{eq:w_diag}
544    \pd[w]{k} = - \chi \; e_3 \qquad
545  % \label{eq:hp_diag}
546    \pd[p_h]{k} = - \rho \; g \; e_3
547  \]
548  where the divergence of the horizontal velocity, $\chi$ is given by \autoref{eq:PE_div_Uh}.
549
550\item
551  \textbf{tracer equations}:
552  \begin{equation}
553  \begin{split}
554    \pd[T]{t} = & - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 T \, u)]{i} + \pd[(e_1 T \, v)]{j} \rt]
555                - \frac{1}{e_3} \pd[(T \, w)]{k} + D^T + F^T \\
556    \pd[S]{t} = & - \frac{1}{e_1 e_2} \lt[ \pd[(e_2 S \, u)]{i} + \pd[(e_1 S \, v)]{j} \rt]
557                - \frac{1}{e_3} \pd[(S \, w)]{k} + D^S + F^S \\
558    \rho = & \rho \big( T,S,z(k) \big)
559  \end{split}
560  \end{equation}
561\end{itemize}
562
563The expression of $\vect D^{U}$, $D^{S}$ and $D^{T}$ depends on the subgrid scale parameterisation used.
564It will be defined in \autoref{eq:PE_zdf}.
565The nature and formulation of $\vect F^{\vect U}$, $F^T$ and $F^S$, the surface forcing terms,
566are discussed in \autoref{chap:SBC}.
567
568\newpage
569
570% ================================================================
571% Curvilinear generalised vertical coordinate System
572% ================================================================
573\section{Curvilinear generalised vertical coordinate system}
574\label{sec:PE_gco}
575
576The ocean domain presents a huge diversity of situation in the vertical.
577First the ocean surface is a time dependent surface (moving surface).
578Second the ocean floor depends on the geographical position,
579varying from more than 6,000 meters in abyssal trenches to zero at the coast.
580Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.
581Therefore, in order to represent the ocean with respect to
582the first point a space and time dependent vertical coordinate that follows the variation of the sea surface height
583\eg\ an \zstar-coordinate;
584for the second point, a space variation to fit the change of bottom topography
585\eg\ a terrain-following or $\sigma$-coordinate;
586and for the third point, one will be tempted to use a space and time dependent coordinate that
587follows the isopycnal surfaces, \eg\ an isopycnic coordinate.
588
589In order to satisfy two or more constraints one can even be tempted to mixed these coordinate systems, as in
590HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and $\sigma$ at
591the ocean bottom) \citep{chassignet.smith.ea_JPO03} or
592OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and $\sigma$-coordinate elsewhere)
593\citep{madec.delecluse.ea_JPO96} among others.
594
595In fact one is totally free to choose any space and time vertical coordinate by
596introducing an arbitrary vertical coordinate :
597\begin{equation}
598  \label{eq:PE_s}
599  s = s(i,j,k,t)
600\end{equation}
601with the restriction that the above equation gives a single-valued monotonic relationship between $s$ and $k$,
602when $i$, $j$ and $t$ are held fixed.
603\autoref{eq:PE_s} is a transformation from the $(i,j,k,t)$ coordinate system with independent variables into
604the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through
605\autoref{eq:PE_s}.
606This so-called \textit{generalised vertical coordinate} \citep{kasahara_MWR74} is in fact
607an Arbitrary Lagrangian--Eulerian (ALE) coordinate.
608Indeed, one has a great deal of freedom in the choice of expression for $s$. The choice determines
609which part of the vertical velocity (defined from a fixed referential) will cross the levels (Eulerian part) and
610which part will be used to move them (Lagrangian part).
611The coordinate is also sometime referenced as an adaptive coordinate \citep{hofmeister.burchard.ea_OM10},
612since the coordinate system is adapted in the course of the simulation.
613Its most often used implementation is via an ALE algorithm,
614in which a pure lagrangian step is followed by regridding and remapping steps,
615the latter step implicitly embedding the vertical advection
616\citep{hirt.amsden.ea_JCP74, chassignet.smith.ea_JPO03, white.adcroft.ea_JCP09}.
617Here we follow the \citep{kasahara_MWR74} strategy:
618a regridding step (an update of the vertical coordinate) followed by an Eulerian step with
619an explicit computation of vertical advection relative to the moving s-surfaces.
620
621%\gmcomment{
622%A key point here is that the $s$-coordinate depends on $(i,j)$ ==> horizontal pressure gradient...
623The generalized vertical coordinates used in ocean modelling are not orthogonal,
624which contrasts with many other applications in mathematical physics.
625Hence, it is useful to keep in mind the following properties that may seem odd on initial encounter.
626
627The horizontal velocity in ocean models measures motions in the horizontal plane,
628perpendicular to the local gravitational field.
629That is, horizontal velocity is mathematically the same regardless of the vertical coordinate, be it geopotential,
630isopycnal, pressure, or terrain following.
631The key motivation for maintaining the same horizontal velocity component is that
632the hydrostatic and geostrophic balances are dominant in the large-scale ocean.
633Use of an alternative quasi -horizontal velocity, for example one oriented parallel to the generalized surface,
634would lead to unacceptable numerical errors.
635Correspondingly, the vertical direction is anti -parallel to the gravitational force in
636all of the coordinate systems.
637We do not choose the alternative of a quasi -vertical direction oriented normal to
638the surface of a constant generalized vertical coordinate.
639
640It is the method used to measure transport across the generalized vertical coordinate surfaces which differs between
641the vertical coordinate choices.
642That is, computation of the dia-surface velocity component represents the fundamental distinction between
643the various coordinates.
644In some models, such as geopotential, pressure, and terrain following, this transport is typically diagnosed from
645volume or mass conservation.
646In other models, such as isopycnal layered models, this transport is prescribed based on assumptions about
647the physical processes producing a flux across the layer interfaces.
648
649In this section we first establish the PE in the generalised vertical $s$-coordinate,
650then we discuss the particular cases available in \NEMO, namely $z$, \zstar, $s$, and \ztilde.
651%}
652
653% -------------------------------------------------------------------------------------------------------------
654% The s-coordinate Formulation
655% -------------------------------------------------------------------------------------------------------------
656\subsection{\textit{S}-coordinate formulation}
657
658Starting from the set of equations established in \autoref{sec:PE_zco} for the special case $k = z$ and
659thus $e_3 = 1$, we introduce an arbitrary vertical coordinate $s = s(i,j,k,t)$,
660which includes $z$-, \zstar- and $\sigma$-coordinates as special cases
661($s = z$, $s = \zstar$, and $s = \sigma = z / H$ or $ = z / \lt( H + \eta \rt)$, resp.).
662A formal derivation of the transformed equations is given in \autoref{apdx:A}.
663Let us define the vertical scale factor by $e_3 = \partial_s z$  ($e_3$ is now a function of $(i,j,k,t)$ ),
664and the slopes in the $(i,j)$ directions between $s$- and $z$-surfaces by:
665\begin{equation}
666  \label{eq:PE_sco_slope}
667  \sigma_1 = \frac{1}{e_1} \; \lt. \pd[z]{i} \rt|_s \quad \text{and} \quad
668  \sigma_2 = \frac{1}{e_2} \; \lt. \pd[z]{j} \rt|_s
669\end{equation}
670We also introduce $\omega$, a dia-surface velocity component, defined as the velocity
671relative to the moving $s$-surfaces and normal to them:
672\[
673  % \label{eq:PE_sco_w}
674  \omega = w -  \, \lt. \pd[z]{t} \rt|_s - \sigma_1 \, u - \sigma_2 \, v
675\]
676
677The equations solved by the ocean model \autoref{eq:PE} in $s$-coordinate can be written as follows
678(see \autoref{sec:A_momentum}):
679
680\begin{itemize}
681\item \textbf{Vector invariant form of the momentum equation}:
682  \begin{multline*}
683  % \label{eq:PE_sco_u_vector}
684    \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} \\
685                - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_1
686                + D_u^{\vect U} + F_u^{\vect U}
687  \end{multline*}
688  \begin{multline*}
689  % \label{eq:PE_sco_v_vector}
690    \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} \\
691                - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt) - g \frac{\rho}{\rho_o} \sigma_2
692                + D_v^{\vect U} + F_v^{\vect U}
693  \end{multline*}
694\item \textbf{Flux form of the momentum equation}:
695  \begin{multline*}
696  % \label{eq:PE_sco_u_flux}
697    \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 \\
698                                       - \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) \\
699                                       - \frac{1}{e_3} \pd[(\omega \, u)]{k}
700                                       - \frac{1}{e_1} \pd[]{i} \lt( \frac{p_s + p_h}{\rho_o} \rt)
701                                       - g \frac{\rho}{\rho_o} \sigma_1 + D_u^{\vect U} + F_u^{\vect U}
702  \end{multline*}
703  \begin{multline*}
704  % \label{eq:PE_sco_v_flux}
705    \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 \\
706                                       - \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) \\
707                                       - \frac{1}{e_3} \pd[(\omega \, v)]{k}
708                                       - \frac{1}{e_2} \pd[]{j} \lt( \frac{p_s + p_h}{\rho_o} \rt)
709                                       - g \frac{\rho}{\rho_o}\sigma_2 + D_v^{\vect U} + F_v^{\vect U}
710  \end{multline*}
711  where the relative vorticity, $\zeta$, the surface pressure gradient,
712  and the hydrostatic pressure have the same expressions as in $z$-coordinates although
713  they do not represent exactly the same quantities.
714  $\omega$ is provided by the continuity equation (see \autoref{apdx:A}):
715  \[
716  % \label{eq:PE_sco_continuity}
717    \pd[e_3]{t} + e_3 \; \chi + \pd[\omega]{s} = 0 \quad \text{with} \quad
718    \chi = \frac{1}{e_1 e_2 e_3} \lt( \pd[(e_2 e_3 \, u)]{i} + \pd[(e_1 e_3 \, v)]{j} \rt)
719  \]
720\item \textit{tracer equations}:
721  \begin{multline*}
722  % \label{eq:PE_sco_t}
723    \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}
724                                                                    + \pd[(e_1 e_3 \, v \, T)]{j} \rt) \\
725                                       - \frac{1}{e_3} \pd[(T \, \omega)]{k} + D^T + F^S
726  \end{multline*}
727  \begin{multline}
728  % \label{eq:PE_sco_s}
729    \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}
730                                                                    + \pd[(e_1 e_3 \, v \, S)]{j} \rt) \\
731                                       - \frac{1}{e_3} \pd[(S \, \omega)]{k} + D^S + F^S
732  \end{multline}
733\end{itemize}
734The equation of state has the same expression as in $z$-coordinate,
735and similar expressions are used for mixing and forcing terms.
736
737\gmcomment{
738  \colorbox{yellow}{ to be updated $= = >$}
739  Add a few works on z and zps and s and underlies the differences between all of them
740  \colorbox{yellow}{$< = =$ end update}
741}
742
743% -------------------------------------------------------------------------------------------------------------
744% Curvilinear \zstar-coordinate System
745% -------------------------------------------------------------------------------------------------------------
746\subsection{Curvilinear \zstar-coordinate system}
747\label{subsec:PE_zco_star}
748
749%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
750\begin{figure}[!b]
751  \begin{center}
752    \includegraphics[width=\textwidth]{Fig_z_zstar}
753    \caption{
754      \protect\label{fig:z_zstar}
755      (a) $z$-coordinate in linear free-surface case ;
756      (b) $z$-coordinate in non-linear free surface case ;
757      (c) re-scaled height coordinate
758      (become popular as the \zstar-coordinate \citep{adcroft.campin_OM04}).
759    }
760  \end{center}
761\end{figure}
762%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
763
764In this case, the free surface equation is nonlinear, and the variations of volume are fully taken into account.
765These coordinates systems is presented in a report \citep{levier.treguier.ea_rpt07} available on the \NEMO\ web site.
766
767The \zstar coordinate approach is an unapproximated, non-linear free surface implementation which allows one to
768deal with large amplitude free-surface variations relative to the vertical resolution \citep{adcroft.campin_OM04}.
769In the \zstar formulation,
770the variation of the column thickness due to sea-surface undulations is not concentrated in the surface level,
771as in the $z$-coordinate formulation, but is equally distributed over the full water column.
772Thus vertical levels naturally follow sea-surface variations, with a linear attenuation with depth,
773as illustrated by \autoref{fig:z_zstar}.
774Note that with a flat bottom, such as in \autoref{fig:z_zstar}, the bottom-following $z$ coordinate and \zstar are equivalent.
775The definition and modified oceanic equations for the rescaled vertical coordinate \zstar,
776including the treatment of fresh-water flux at the surface, are detailed in Adcroft and Campin (2004).
777The major points are summarized here.
778The position (\zstar) and vertical discretization (\zstar) are expressed as:
779\[
780  % \label{eq:PE_z-star}
781  H + \zstar = (H + z)  / r \quad \text{and}  \quad \delta \zstar
782              = \delta z / r \quad \text{with} \quad r
783              = \frac{H + \eta}{H} .
784\]
785Simple re-organisation of the above expressions gives
786\[
787  % \label{eq:PE_zstar_2}
788  \zstar = H \lt( \frac{z - \eta}{H + \eta} \rt) .
789\]
790Since the vertical displacement of the free surface is incorporated in the vertical coordinate \zstar,
791the upper and lower boundaries are at fixed  \zstar position,
792$\zstar = 0$ and $\zstar = -H$ respectively.
793Also the divergence of the flow field is no longer zero as shown by the continuity equation:
794\[
795  \pd[r]{t} = \nabla_{\zstar} \cdot \lt( r \; \vect U_h \rt) + \pd[r \; w^*]{\zstar} = 0 .
796\]
797This \zstar coordinate is closely related to the "eta" coordinate used in many atmospheric models
798(see Black (1994) for a review of eta coordinate atmospheric models).
799It was originally used in ocean models by Stacey et al. (1995) for studies of tides next to shelves,
800and it has been recently promoted by Adcroft and Campin (2004) for global climate modelling.
801
802The surfaces of constant \zstar are quasi -horizontal.
803Indeed, the \zstar coordinate reduces to $z$ when $\eta$ is zero.
804In general, when noting the large differences between
805undulations of the bottom topography versus undulations in the surface height,
806it is clear that surfaces constant \zstar are very similar to the depth surfaces.
807These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to
808terrain following sigma models discussed in \autoref{subsec:PE_sco}.
809Additionally, since $\zstar = z$ when $\eta = 0$,
810no flow is spontaneously generated in an unforced ocean starting from rest, regardless the bottom topography.
811This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in the presence of
812nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state,
813depending on the sophistication of the pressure gradient solver.
814The quasi -horizontal nature of the coordinate surfaces also facilitates the implementation of
815neutral physics parameterizations in \zstar  models using the same techniques as in $z$-models
816(see Chapters 13-16 of \cite{griffies_bk04}) for a discussion of neutral physics in $z$-models,
817as well as \autoref{sec:LDF_slp} in this document for treatment in \NEMO).
818
819The range over which \zstar  varies is time independent $-H \leq \zstar \leq 0$.
820Hence, all cells remain nonvanishing, so long as the surface height maintains $\eta > -H$.
821This is a minor constraint relative to that encountered on the surface height when using $s = z$ or $s = z - \eta$.
822
823Because \zstar  has a time independent range, all grid cells have static increments ds,
824and the sum of the vertical increments yields the time independent ocean depth. %k ds = H.
825The \zstar coordinate is therefore invisible to undulations of the free surface,
826since it moves along with the free surface.
827This property means that no spurious vertical transport is induced across surfaces of constant \zstar  by
828the motion of external gravity waves.
829Such spurious transport can be a problem in z-models, especially those with tidal forcing.
830Quite generally, the time independent range for the \zstar  coordinate is a very convenient property that
831allows for a nearly arbitrary vertical resolution even in the presence of large amplitude fluctuations of
832the surface height, again so long as $\eta > -H$.
833%end MOM doc %%%
834
835\newpage
836
837% -------------------------------------------------------------------------------------------------------------
838% Terrain following  coordinate System
839% -------------------------------------------------------------------------------------------------------------
840\subsection{Curvilinear terrain-following \textit{s}--coordinate}
841\label{subsec:PE_sco}
842
843% -------------------------------------------------------------------------------------------------------------
844% Introduction
845% -------------------------------------------------------------------------------------------------------------
846\subsubsection{Introduction}
847
848Several important aspects of the ocean circulation are influenced by bottom topography.
849Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers, sills and
850channels that strongly constrain the path of water masses, but more subtle effects exist.
851For example, the topographic $\beta$-effect is usually larger than the planetary one along continental slopes.
852Topographic Rossby waves can be excited and can interact with the mean current.
853In the $z$-coordinate system presented in the previous section (\autoref{sec:PE_zco}),
854$z$-surfaces are geopotential surfaces.
855The bottom topography is discretised by steps.
856This often leads to a misrepresentation of a gradually sloping bottom and to
857large localized depth gradients associated with large localized vertical velocities.
858The response to such a velocity field often leads to numerical dispersion effects.
859One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of
860a full step one \cite{pacanowski.gnanadesikan_MWR98}.
861Another solution is to introduce a terrain-following coordinate system (hereafter $s$-coordinate).
862
863The $s$-coordinate avoids the discretisation error in the depth field since the layers of
864computation are gradually adjusted with depth to the ocean bottom.
865Relatively small topographic features as well as  gentle, large-scale slopes of the sea floor in the deep ocean,
866which would be ignored in typical $z$-model applications with the largest grid spacing at greatest depths,
867can easily be represented (with relatively low vertical resolution).
868A terrain-following model (hereafter $s$-model) also facilitates the modelling of the boundary layer flows over
869a large depth range, which in the framework of the $z$-model would require high vertical resolution over
870the whole depth range.
871Moreover, with a $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as
872the only boundaries of the domain (no more lateral boundary condition to specify).
873Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean,
874it has strong limitations as soon as stratification is introduced.
875The main two problems come from the truncation error in the horizontal pressure gradient and
876a possibly increased diapycnal diffusion.
877The horizontal pressure force in $s$-coordinate consists of two terms (see \autoref{apdx:A}),
878
879\begin{equation}
880  \label{eq:PE_p_sco}
881  \nabla p |_z = \nabla p |_s - \frac{1}{e_3} \pd[p]{s} \nabla z |_s
882\end{equation}
883
884The second term in \autoref{eq:PE_p_sco} depends on the tilt of the coordinate surface and
885leads to a truncation error that is not present in a $z$-model.
886In the special case of a $\sigma$-coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),
887\citet{haney_JPO91} and \citet{beckmann.haidvogel_JPO93} have given estimates of the magnitude of this truncation error.
888It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state,
889and the finite difference scheme.
890This error limits the possible topographic slopes that a model can handle at
891a given horizontal and vertical resolution.
892This is a severe restriction for large-scale applications using realistic bottom topography.
893The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive.
894This problem can be at least partially overcome by mixing $s$-coordinate and
895step-like representation of bottom topography \citep{gerdes_JGR93*a,gerdes_JGR93*b,madec.delecluse.ea_JPO96}.
896However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for
897a realistic bottom topography:
898an envelope topography is defined in $s$-coordinate on which a full or
899partial step bottom topography is then applied in order to adjust the model depth to the observed one
900(see \autoref{sec:DOM_zgr}.
901
902For numerical reasons a minimum of diffusion is required along the coordinate surfaces of
903any finite difference model.
904It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces.
905This is the case for a $z$-model as well as for a $s$-model.
906However, density varies more strongly on $s$-surfaces than on horizontal surfaces in regions of
907large topographic slopes, implying larger diapycnal diffusion in a $s$-model than in a $z$-model.
908Whereas such a diapycnal diffusion in a $z$-model tends to weaken horizontal density (pressure) gradients and thus
909the horizontal circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.
910For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification.
911Spurious diffusion along $s$-surfaces will induce a bump of isoneutral surfaces over the topography,
912and thus will generate there a baroclinic eddy.
913In contrast, the ocean will stay at rest in a $z$-model.
914As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below
915the strongly stratified portion of the water column (\ie\ the main thermocline) \citep{madec.delecluse.ea_JPO96}.
916An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces
917(see \autoref{subsec:PE_ldf}).
918Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,
919strongly exceeding the stability limit of such a operator when it is discretized (see \autoref{chap:LDF}).
920
921The $s$-coordinates introduced here \citep{lott.madec.ea_OM90,madec.delecluse.ea_JPO96} differ mainly in two aspects from
922similar models:
923it allows a representation of bottom topography with mixed full or partial step-like/terrain following topography;
924It also offers a completely general transformation, $s=s(i,j,z)$ for the vertical coordinate.
925
926% -------------------------------------------------------------------------------------------------------------
927% Curvilinear z-tilde coordinate System
928% -------------------------------------------------------------------------------------------------------------
929\subsection{\texorpdfstring{Curvilinear \ztilde-coordinate}{}}
930\label{subsec:PE_zco_tilde}
931
932The \ztilde -coordinate has been developed by \citet{leclair.madec_OM11}.
933It is available in \NEMO\ since the version 3.4 and is more robust in version 4.0 than previously.
934Nevertheless, it is currently not robust enough to be used in all possible configurations.
935Its use is therefore not recommended.
936
937\newpage
938
939% ================================================================
940% Subgrid Scale Physics
941% ================================================================
942\section{Subgrid scale physics}
943\label{sec:PE_zdf_ldf}
944
945The hydrostatic primitive equations describe the behaviour of a geophysical fluid at space and time scales larger than
946a few kilometres in the horizontal, a few meters in the vertical and a few minutes.
947They are usually solved at larger scales: the specified grid spacing and time step of the numerical model.
948The effects of smaller scale motions (coming from the advective terms in the Navier-Stokes equations)
949must be represented entirely in terms of large-scale patterns to close the equations.
950These effects appear in the equations as the divergence of turbulent fluxes
951(\ie\ fluxes associated with the mean correlation of small scale perturbations).
952Assuming a turbulent closure hypothesis is equivalent to choose a formulation for these fluxes.
953It is usually called the subgrid scale physics.
954It must be emphasized that this is the weakest part of the primitive equations,
955but also one of the most important for long-term simulations as
956small scale processes \textit{in fine} balance the surface input of kinetic energy and heat.
957
958The control exerted by gravity on the flow induces a strong anisotropy between the lateral and vertical motions.
959Therefore subgrid-scale physics \textbf{D}$^{\vect U}$, $D^{S}$ and $D^{T}$  in
960\autoref{eq:PE_dyn}, \autoref{eq:PE_tra_T} and \autoref{eq:PE_tra_S} are divided into
961a lateral part \textbf{D}$^{l \vect U}$, $D^{l S}$ and $D^{l T}$ and
962a vertical part \textbf{D}$^{v \vect U}$, $D^{v S}$ and $D^{v T}$.
963The formulation of these terms and their underlying physics are briefly discussed in the next two subsections.
964
965% -------------------------------------------------------------------------------------------------------------
966% Vertical Subgrid Scale Physics
967% -------------------------------------------------------------------------------------------------------------
968\subsection{Vertical subgrid scale physics}
969\label{subsec:PE_zdf}
970
971The model resolution is always larger than the scale at which the major sources of vertical turbulence occur
972(shear instability, internal wave breaking...).
973Turbulent motions are thus never explicitly solved, even partially, but always parameterized.
974The vertical turbulent fluxes are assumed to depend linearly on the gradients of large-scale quantities
975(for example, the turbulent heat flux is given by $\overline{T' w'} = -A^{v T} \partial_z \overline T$,
976where $A^{v T}$ is an eddy coefficient).
977This formulation is analogous to that of molecular diffusion and dissipation.
978This is quite clearly a necessary compromise: considering only the molecular viscosity acting on
979large scale severely underestimates the role of turbulent diffusion and dissipation,
980while an accurate consideration of the details of turbulent motions is simply impractical.
981The resulting vertical momentum and tracer diffusive operators are of second order:
982\begin{equation}
983  \label{eq:PE_zdf}
984  \begin{gathered}
985    \vect D^{v \vect U} = \pd[]{z} \lt( A^{vm} \pd[\vect U_h]{z} \rt) \ , \\
986          D^{vT}       = \pd[]{z} \lt( A^{vT} \pd[T]{z}         \rt) \quad \text{and} \quad
987          D^{vS}       = \pd[]{z} \lt( A^{vT} \pd[S]{z}         \rt)
988  \end{gathered}
989\end{equation}
990where $A^{vm}$ and $A^{vT}$ are the vertical eddy viscosity and diffusivity coefficients, respectively.
991At the sea surface and at the bottom, turbulent fluxes of momentum, heat and salt must be specified
992(see \autoref{chap:SBC} and \autoref{chap:ZDF} and \autoref{sec:TRA_bbl}).
993All the vertical physics is embedded in the specification of the eddy coefficients.
994They can be assumed to be either constant, or function of the local fluid properties
995(\eg\ Richardson number, Brunt-Vais\"{a}l\"{a} frequency, distance from the boundary ...),
996or computed from a turbulent closure model.
997The choices available in \NEMO\ are discussed in \autoref{chap:ZDF}).
998
999% -------------------------------------------------------------------------------------------------------------
1000% Lateral Diffusive and Viscous Operators Formulation
1001% -------------------------------------------------------------------------------------------------------------
1002\subsection{Formulation of the lateral diffusive and viscous operators}
1003\label{subsec:PE_ldf}
1004
1005Lateral turbulence can be roughly divided into a mesoscale turbulence associated with eddies
1006(which can be solved explicitly if the resolution is sufficient since
1007their underlying physics are included in the primitive equations),
1008and a sub mesoscale turbulence which is never explicitly solved even partially, but always parameterized.
1009The formulation of lateral eddy fluxes depends on whether the mesoscale is below or above the grid-spacing
1010(\ie\ the model is eddy-resolving or not).
1011
1012In non-eddy-resolving configurations, the closure is similar to that used for the vertical physics.
1013The lateral turbulent fluxes are assumed to depend linearly on the lateral gradients of large-scale quantities.
1014The resulting lateral diffusive and dissipative operators are of second order.
1015Observations show that lateral mixing induced by mesoscale turbulence tends to be along isopycnal surfaces
1016(or more precisely neutral surfaces \cite{mcdougall_JPO87}) rather than across them.
1017As the slope of neutral surfaces is small in the ocean, a common approximation is to assume that
1018the `lateral' direction is the horizontal, \ie\ the lateral mixing is performed along geopotential surfaces.
1019This leads to a geopotential second order operator for lateral subgrid scale physics.
1020This assumption can be relaxed: the eddy-induced turbulent fluxes can be better approached by assuming that
1021they depend linearly on the gradients of large-scale quantities computed along neutral surfaces.
1022In such a case, the diffusive operator is an isoneutral second order operator and
1023it has components in the three space directions.
1024However,
1025both horizontal and isoneutral operators have no effect on mean (\ie\ large scale) potential energy whereas
1026potential energy is a main source of turbulence (through baroclinic instabilities).
1027\citet{gent.mcwilliams_JPO90} proposed a parameterisation of mesoscale eddy-induced turbulence which
1028associates an eddy-induced velocity to the isoneutral diffusion.
1029Its mean effect is to reduce the mean potential energy of the ocean.
1030This leads to a formulation of lateral subgrid-scale physics made up of an isoneutral second order operator and
1031an eddy induced advective part.
1032In all these lateral diffusive formulations,
1033the specification of the lateral eddy coefficients remains the problematic point as
1034there is no really satisfactory formulation of these coefficients as a function of large-scale features.
1035
1036In eddy-resolving configurations, a second order operator can be used,
1037but usually the more scale selective biharmonic operator is preferred as
1038the grid-spacing is usually not small enough compared to the scale of the eddies.
1039The role devoted to the subgrid-scale physics is to dissipate the energy that
1040cascades toward the grid scale and thus to ensure the stability of the model while
1041not interfering with the resolved mesoscale activity.
1042Another approach is becoming more and more popular:
1043instead of specifying explicitly a sub-grid scale term in the momentum and tracer time evolution equations,
1044one uses an advective scheme which is diffusive enough to maintain the model stability.
1045It must be emphasised that then, all the sub-grid scale physics is included in the formulation of
1046the advection scheme.
1047
1048All these parameterisations of subgrid scale physics have advantages and drawbacks.
1049They are not all available in \NEMO. For active tracers (temperature and salinity) the main ones are:
1050Laplacian and bilaplacian operators acting along geopotential or iso-neutral surfaces,
1051\citet{gent.mcwilliams_JPO90} parameterisation, and various slightly diffusive advection schemes.
1052For momentum, the main ones are: Laplacian and bilaplacian operators acting along geopotential surfaces,
1053and UBS advection schemes when flux form is chosen for the momentum advection.
1054
1055\subsubsection{Lateral laplacian tracer diffusive operator}
1056
1057The lateral Laplacian tracer diffusive operator is defined by (see \autoref{apdx:B}):
1058\begin{equation}
1059  \label{eq:PE_iso_tensor}
1060  D^{lT} = \nabla \vect . \lt( A^{lT} \; \Re \; \nabla T \rt) \quad \text{with} \quad
1061  \Re =
1062    \begin{pmatrix}
1063      1    & 0    & -r_1          \\
1064      0    & 1    & -r_2          \\
1065      -r_1 & -r_2 & r_1^2 + r_2^2 \\
1066    \end{pmatrix}
1067\end{equation}
1068where $r_1$ and $r_2$ are the slopes between the surface along which the diffusive operator acts and
1069the model level (\eg\ $z$- or $s$-surfaces).
1070Note that the formulation \autoref{eq:PE_iso_tensor} is exact for
1071the rotation between geopotential and $s$-surfaces,
1072while it is only an approximation for the rotation between isoneutral and $z$- or $s$-surfaces.
1073Indeed, in the latter case, two assumptions are made to simplify \autoref{eq:PE_iso_tensor} \citep{cox_OM87}.
1074First, the horizontal contribution of the dianeutral mixing is neglected since the ratio between iso and
1075dia-neutral diffusive coefficients is known to be several orders of magnitude smaller than unity.
1076Second, the two isoneutral directions of diffusion are assumed to be independent since
1077the slopes are generally less than $10^{-2}$ in the ocean (see \autoref{apdx:B}).
1078
1079For \textit{iso-level} diffusion, $r_1$ and $r_2 $ are zero.
1080$\Re$ reduces to the identity in the horizontal direction, no rotation is applied.
1081
1082For \textit{geopotential} diffusion,
1083$r_1$ and $r_2 $ are the slopes between the geopotential and computational surfaces:
1084they are equal to $\sigma_1$ and $\sigma_2$, respectively (see \autoref{eq:PE_sco_slope}).
1085
1086For \textit{isoneutral} diffusion $r_1$ and $r_2$ are the slopes between the isoneutral and computational surfaces.
1087Therefore, they are different quantities, but have similar expressions in $z$- and $s$-coordinates.
1088In $z$-coordinates:
1089\begin{equation}
1090  \label{eq:PE_iso_slopes}
1091  r_1 = \frac{e_3}{e_1} \lt( \pd[\rho]{i} \rt) \lt( \pd[\rho]{k} \rt)^{-1} \quad
1092  r_2 = \frac{e_3}{e_2} \lt( \pd[\rho]{j} \rt) \lt( \pd[\rho]{k} \rt)^{-1}
1093\end{equation}
1094while in $s$-coordinates $\pd[]{k}$ is replaced by $\pd[]{s}$.
1095
1096\subsubsection{Eddy induced velocity}
1097
1098When the \textit{eddy induced velocity} parametrisation (eiv) \citep{gent.mcwilliams_JPO90} is used,
1099an additional tracer advection is introduced in combination with the isoneutral diffusion of tracers:
1100\[
1101  % \label{eq:PE_iso+eiv}
1102  D^{lT} = \nabla \cdot \lt( A^{lT} \; \Re \; \nabla T \rt) + \nabla \cdot \lt( \vect U^\ast \, T \rt)
1103\]
1104where $ \vect U^\ast = \lt( u^\ast,v^\ast,w^\ast \rt)$ is a non-divergent,
1105eddy-induced transport velocity. This velocity field is defined by:
1106\begin{gather}
1107  % \label{eq:PE_eiv}
1108  u^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_1 \rt) \\
1109  v^\ast =   \frac{1}{e_3}            \pd[]{k} \lt( A^{eiv} \;        \tilde{r}_2 \rt) \\
1110  w^\ast = - \frac{1}{e_1 e_2} \lt[   \pd[]{i} \lt( A^{eiv} \; e_2 \, \tilde{r}_1 \rt)
1111                                     + \pd[]{j} \lt( A^{eiv} \; e_1 \, \tilde{r}_2 \rt) \rt]
1112\end{gather}
1113where $A^{eiv}$ is the eddy induced velocity coefficient
1114(or equivalently the isoneutral thickness diffusivity coefficient),
1115and $\tilde r_1$ and $\tilde r_2$ are the slopes between isoneutral and \textit{geopotential} surfaces.
1116Their values are thus independent of the vertical coordinate, but their expression depends on the coordinate:
1117\begin{align}
1118  \label{eq:PE_slopes_eiv}
1119  \tilde{r}_n =
1120    \begin{cases}
1121      r_n            & \text{in $z$-coordinate}                \\
1122      r_n + \sigma_n & \text{in \zstar- and $s$-coordinates}
1123    \end{cases}
1124  \quad \text{where~} n = 1, 2
1125\end{align}
1126
1127The normal component of the eddy induced velocity is zero at all the boundaries.
1128This can be achieved in a model by tapering either the eddy coefficient or the slopes to zero in the vicinity of
1129the boundaries.
1130The latter strategy is used in \NEMO\ (cf. \autoref{chap:LDF}).
1131
1132\subsubsection{Lateral bilaplacian tracer diffusive operator}
1133
1134The lateral bilaplacian tracer diffusive operator is defined by:
1135\[
1136  % \label{eq:PE_bilapT}
1137  D^{lT}= - \Delta \; (\Delta T) \quad \text{where} \quad
1138  \Delta \bullet = \nabla \lt( \sqrt{B^{lT}} \; \Re \; \nabla \bullet \rt)
1139\]
1140It is the Laplacian operator given by \autoref{eq:PE_iso_tensor} applied twice with
1141the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1142
1143\subsubsection{Lateral Laplacian momentum diffusive operator}
1144
1145The Laplacian momentum diffusive operator along $z$- or $s$-surfaces is found by
1146applying \autoref{eq:PE_lap_vector} to the horizontal velocity vector (see \autoref{apdx:B}):
1147\begin{align*}
1148  % \label{eq:PE_lapU}
1149  \vect D^{l \vect U} &=   \nabla_h        \big( A^{lm}    \chi             \big)
1150                         - \nabla_h \times \big( A^{lm} \, \zeta \; \vect k \big) \\
1151                      &= \lt(   \frac{1}{e_1}     \pd[ \lt( A^{lm}    \chi      \rt) ]{i} \rt.
1152                              - \frac{1}{e_2 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{j} ,
1153                                \frac{1}{e_2}     \pd[ \lt( A^{lm}    \chi      \rt) ]{j}
1154                         \lt. + \frac{1}{e_1 e_3} \pd[ \lt( A^{lm} \; e_3 \zeta \rt) ]{i} \rt)
1155\end{align*}
1156
1157Such a formulation ensures a complete separation between the vorticity and horizontal divergence fields
1158(see \autoref{apdx:C}).
1159Unfortunately, it is only available in \textit{iso-level} direction.
1160When a rotation is required
1161(\ie\ geopotential diffusion in $s$-coordinates or isoneutral diffusion in both $z$- and $s$-coordinates),
1162the $u$ and $v$-fields are considered as independent scalar fields, so that the diffusive operator is given by:
1163\begin{gather*}
1164  % \label{eq:PE_lapU_iso}
1165    D_u^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla u \rt) \\
1166    D_v^{l \vect U} = \nabla . \lt( A^{lm} \; \Re \; \nabla v \rt)
1167\end{gather*}
1168where $\Re$ is given by \autoref{eq:PE_iso_tensor}.
1169It is the same expression as those used for diffusive operator on tracers.
1170It must be emphasised that such a formulation is only exact in a Cartesian coordinate system,
1171\ie\ on a $f$- or $\beta$-plane, not on the sphere.
1172It is also a very good approximation in vicinity of the Equator in
1173a geographical coordinate system \citep{lengaigne.madec.ea_JGR03}.
1174
1175\subsubsection{Lateral bilaplacian momentum diffusive operator}
1176
1177As for tracers, the bilaplacian order momentum diffusive operator is a re-entering Laplacian operator with
1178the harmonic eddy diffusion coefficient set to the square root of the biharmonic one.
1179Nevertheless it is currently not available in the iso-neutral case.
1180
1181\biblio
1182
1183\pindex
1184
1185\end{document}
Note: See TracBrowser for help on using the repository browser.