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_ZDF.tex in trunk/NEMO/DOC/BETA/Chapters – NEMO

source: trunk/NEMO/DOC/BETA/Chapters/Chap_ZDF.tex @ 705

Last change on this file since 705 was 705, checked in by smasson, 17 years ago

Historical date: first beta version of the documentation, see ticket:1

  • Property svn:executable set to *
File size: 33.5 KB
Line 
1% ================================================================
2% Chapter Ñ Vertical Ocean Physics (ZDF)
3% ================================================================
4\chapter{Vertical Ocean Physics (ZDF)}
5\label{ZDF}
6\minitoc
7
8%gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN.
9
10% ================================================================
11% Vertical Mixing
12% ================================================================
13\section{Vertical Mixing}
14\label{ZDF_zdf}
15
16The discrete form of the ocean subgrid scale physics has been presented in \S\ref{TRA_zdf} and \S\ref{DYN_zdf}. At the surface and bottom boundaries, the turbulent fluxes of momentum, heat and salt have to be defined. At the surface they are prescribed from the surface forcing (see Chap.~\ref{SBC}), while at the bottom they are set to zero for heat and salt, unless a geothermal flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl} defined, see \S\ref{TRA_bbc}), and specified through a bottom friction parameterization for momentum (see \S\ref{ZDF_bfr}).
17
18In this section we briefly discuss the various choices offered to compute
19the vertical eddy viscosity and diffusivity coefficients, $A_u^{vm}$ , $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$-points, respectively (see \S\ref{TRA_zdf} and \S\ref{DYN_zdf}). These coefficients can be assumed to be either constant, or a function of the local Richardson number, or computed from a turbulent closure model (either TKE or KPP formulation). The computation of these coefficients is initialized in \mdl{zdfini} module and performed in \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfkpp} modules. The trends due to the vertical momentum and tracer diffusion, including the surface forcing,
20are computed and added to the general trend in \mdl{dynzdf} and \mdl{trazdf} modules, respectively. These trends can be computed using either a forward time scheme (cpp variable
21\np{np\_zdfexp} or a backward time scheme (default option) depending on the magnitude of the mixing coefficients used, and thus of the formulation used (see \S\ref{DOM_nxt}).
22
23% -------------------------------------------------------------------------------------------------------------
24%        Constant
25% -------------------------------------------------------------------------------------------------------------
26\subsection{Constant (\key{zdfcst})}
27\label{ZDF_cst}
28%--------------------------------------------namzdf---------------------------------------------------------
29\namdisplay{namzdf}
30%--------------------------------------------------------------------------------------------------------------
31
32When the \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to
33constant values over the whole ocean. This is the crudest way to define the
34vertical ocean physics. It is recommended to use this option only in process
35studies, not in basin scale simulation. Typical values used in this case
36are:
37\begin{align*} 
38A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\
39\\
40A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
41\end{align*}
42
43These values are set through \np{avm0} and \np{avt0} namelist parameters. In any case, do not use
44values smaller that those associated to the molecular viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and $\sim10^{-9}~m^2.s^{-1}$ for salinity.
45
46
47% -------------------------------------------------------------------------------------------------------------
48%        Richardson Number Dependent
49% -------------------------------------------------------------------------------------------------------------
50\subsection{Richardson Number Dependent (\key{zdfric})}
51\label{ZDF_ric}
52
53%--------------------------------------------namric---------------------------------------------------------
54\namdisplay{namric}
55%--------------------------------------------------------------------------------------------------------------
56
57When \key{zdfric} is defined, a local Richardson number dependent formulation of the vertical momentum and tracer eddy coefficients is set. The vertical mixing coefficients are diagnosed from the large scale variables computed by the model (order 0.5 closure scheme). \textit{In situ} measurements allow to link vertical turbulent activity to large scale ocean structures. The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to a dependency between the vertical turbulent eddy coefficients and the local Richardson number ($i.e.$ ratio of stratification over vertical shear). Following \citet{PacPhil1981}, the following formulation has been implemented:
58\begin{equation} \label{Eq_zdfric}
59   \left\{      \begin{aligned}
60         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
61         \\
62         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
63   \end{aligned}    \right.
64\end{equation}
65where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number, $N$ is the local brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}), $A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in constant case (see \S\ref{ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1} $ is the maximum value that can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$. The last three coefficients can be modified by setting \np{avmri}, \np{alp} and \np{nric} namelist parameter, respectively.
66
67% -------------------------------------------------------------------------------------------------------------
68%        TKE Turbulent Closure Scheme
69% -------------------------------------------------------------------------------------------------------------
70\subsection{TKE Turbulent Closure Scheme (\key{zdftke})}
71\label{ZDF_tke}
72
73%--------------------------------------------namtke---------------------------------------------------------
74\namdisplay{nam_tke}
75%--------------------------------------------------------------------------------------------------------------
76
77The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on a prognostic equation for $\bar {e}$, the turbulent kinetic energy, and a closure assumption for the turbulent length scales. This turbulent closure model has been developed by \citet{Bougeault1989} in atmospheric cases, adapted by \citet{Gaspar1990} for oceanic cases, and embedded in OPA by \citet{Blanke1993} for equatorial Atlantic simulations. Since then, significant modifications have been introduced by \citet{Madec1998} in both the implementation and the formulation of the mixing length scale. The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear, its destruction through stratification, its vertical diffusion and its dissipation of \citet{Kolmogorov1942} type:
78\begin{equation} \label{Eq_zdftke_e}
79\frac{\partial \bar{e}}{\partial t} =
80\frac{A^{vm}}{e_3 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
81                     +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
82-A^{vT}\,N^2
83+\frac{1}{e_3} \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
84            \;\frac{\partial \bar{e}}{\partial k}} \right]
85- c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
86\end{equation}
87\begin{equation} \label{Eq_zdftke_kz}
88   \begin{split}
89         A^{vm} &= C_k\  l_k\  \sqrt {\bar{e}}
90    \\
91         A^{vT} &= A^{vm} / P_{rt}
92   \end{split}
93\end{equation}
94where $N$ designates the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
95$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing turbulent
96length scales, $P_{rt} $ is the Prandtl number. The constants $C_k = \sqrt {2} /2$ and
97$C_\epsilon = 0.1$ are designed to deal with vertical mixing at any depth \citep{Gaspar1990}. They are set through namelist parameter \np{ediff} and \np{ediss}. $P_{rt} $ can be
98set to unity or, following \citet{Blanke1993}, be a function of the
99local Richardson number, $R_i $:
100\begin{align*} \label{Eq_prt}
101P_{rt} = \begin{cases}
102                    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}  \\
103                    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}  \\
104                    \ \ 10 &      \text{if $\ 2 \leq R_i$} 
105            \end{cases}
106\end{align*}
107Note that a horizontal Shapiro filter can be optionally applied to $R_i$. Nevertheless it is an obsolescent option that is notrecommanded.  The choice of $P_{rt} $ is controlled by \np{npdl} namelist parameter.
108
109For computational efficiency, the original formulation of the turbulent length scales proposed by \citet{Gaspar1990} has been simplified. Four formulations are proposed, the choice of which is controlled by \np{nmxl} namelist parameter. The first two are based on the following first order approximation \citep{Blanke1993}:
110\begin{equation} \label{Eq_tke_mxl0_1}
111l_k = l_\epsilon = \sqrt {2 \bar e} / N
112\end{equation}
113which is obtained in a stable stratified region with constant values of the brunt-Vais\"{a}l\"{a} frequency. The resulting turbulent length scale is bounded by the distance to the surface or to the bottom (\np{nmxl}=0) or by the local vertical scale factor (\np{nmxl}=1). \citet{Blanke1993} notice that this simplification has two major drawbacks: it has no sense for local unstable
114stratification and the computation no longer uses the whole information contained in the vertical density profile. To overcome this drawbacks, \citet{Madec1998} introduces the \np{nmxl}=2 or 3 cases, which add of an hypothesis on the vertical gradient of the computed length scale. So, the length scales are first evaluated as in \eqref{Eq_tke_mxl0_1} and then bounded such that:
115\begin{equation} \label{Eq_tke_mxl_constraint}
116\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
117\qquad \text{with }\  l =  l_k = l_\epsilon
118\end{equation}
119
120\eqref{Eq_tke_mxl_constraint} means that the vertical variations of the length scale cannot be
121larger than the variations of depth. It provides a better approximation of the \citet{Gaspar1990} formulation while being much less time consuming. In particular, it allows the length scale to be limited not only by the distance to the surface or to the ocean bottom but also by the distance to a
122strongly stratified portion of the water column such as the thermocline (Fig.~\ref{Fig_mixing_length}). In order to imposed \eqref{Eq_tke_mxl_constraint} constraint, we introduce two additonnal length scal: $l_{up}$ and $l_{dwn}$, the upward and downward length scale, and evaluate the dissipation and mixing turbulent length scales as (caution here we use the numerical indexation):
123%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
124\begin{figure}[!t] \label{Fig_mixing_length}  \begin{center}
125\includegraphics[width=1.00\textwidth]{./Figures/Fig_mixing_length.pdf}
126\caption {Illustration of the mixing length computation. }
127\end{center} 
128\end{figure}
129%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
130\begin{equation} \label{Eq_tke_mxl2}
131\begin{aligned}
132 l_{up  }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3T}^{(k)} \right)
133    \quad &\text{ from $k=1$ to $jpk$ }\ \\
134 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3T}^{(k-1)}  \right)   
135    \quad &\text{ from $k=jpk$ to $1$ }\ \\
136\end{aligned}
137\end{equation}
138where $l^{(k)}$ is compute using \eqref{Eq_tke_mxl0_1}, $i.e.$ $l^{(k)} = \sqrt {2 \bar e^{(k)} / N^{(k)} }$.
139
140In the \np{nmxl}=2 case, the dissipation and mixing turbulent length scales take a same value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nmxl}=2 case, the dissipation and mixing turbulent length scales are give as in \citet{Gaspar1990}:
141\begin{equation} \label{Eq_tke_mxl_gaspar}
142\begin{aligned}
143& l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }    \\
144& l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
145\end{aligned}
146\end{equation}
147
148At the sea surface the value of $\bar{e}$ is prescribed from the wind
149stress field: $\bar{e}=ebb\;\left| \tau \right|$ ($ebb=60$ by default)
150with a minimal threshold of $emin0=10^{-4}~m^2.s^{-2}$ (namelist
151parameters). Its bottom value is assumed to be equal to the value of the level just above. The time
152integration of the $\bar{e}$ equation may formally lead to negative values
153because the numerical scheme does not ensure the positivity. To overcome
154this problem, a cut-off in the minimum value of $\bar{e}$ is used. Following
155\citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. This allows the subsequent formulations to match
156\citet{Gargett1984} one for the diffusion in the thermocline and deep ocean
157$(A^{vT} = 10^{-3} / N)$. In addition, a
158cut-off is applied on $A^{vm}$ and $A^{vT}$ to avoid numerical
159instabilities associated with too weak vertical diffusion. They must be
160specified at least larger than the molecular values, and are set through
161\textit{avm0} and \textit{avt0} (\textbf{namelist} parameters).
162
163% -------------------------------------------------------------------------------------------------------------
164%        K Profile Parametrisation (KPP)
165% -------------------------------------------------------------------------------------------------------------
166\subsection{K Profile Parametrisation (KPP) (\key{zdfkpp}) }
167\label{ZDF_kpp}
168
169%--------------------------------------------namkpp--------------------------------------------------------
170\namdisplay{namkpp}
171%--------------------------------------------------------------------------------------------------------------
172
173The KKP scheme has been implemented by J. Chanut ...
174\colorbox{yellow}{Add a description of KPP here.}
175
176
177% ================================================================
178% Convection
179% ================================================================
180\section{Convection}
181\label{ZDF_conv}
182
183%--------------------------------------------namzdf--------------------------------------------------------
184\namdisplay{namzdf}
185%--------------------------------------------------------------------------------------------------------------
186
187Static instabilities (i.e. light potential densities under heavy ones) may
188occur at particular ocean grid points. In nature, convective processes
189quickly re-establish the static stability of the water column. These
190processes have been removed from the model via the hydrostatic assumption:
191they must be parameterized. Three parameterisations are available to deal
192with convective processes: either a non-penetrative convective adjustment or
193an enhanced vertical diffusion, or/and the use of a turbulent closure
194scheme.
195
196% -------------------------------------------------------------------------------------------------------------
197%       Non-Penetrative Convective Adjustment
198% -------------------------------------------------------------------------------------------------------------
199\subsection{Non-Penetrative Convective Adjustment (\np{ln\_tranpc}=T) }
200\label{ZDF_npc}
201
202%--------------------------------------------namnpc--------------------------------------------------------
203\namdisplay{namnpc}
204%--------------------------------------------------------------------------------------------------------------
205
206
207%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
208\begin{figure}[!ht] \label{Fig_npc}    \begin{center}
209\includegraphics[width=0.90\textwidth]{./Figures/Fig_npc.pdf}
210\caption {Example of an unstable density profile treated by the non penetrative convective adjustment algorithm. $1^{st}$ step: the initial profile is checked from the surface to the bottom. It is found to be unstable between levels 3 and 4. They are mixed. The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. The $1^{st}$ step ends since the density profile is then stable below the level 3. $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: levels 2 to 5 are mixed. The new density profile is checked. It is found stable: end of algorithm.}
211\end{center}   \end{figure}
212%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
213
214The non-penetrative convective adjustment algorithm is used when \np{ln\_zdfnpc}=T. It is applied at each \np{nnpc1} time step and mixes downwards instantaneously the statically unstable portion of the water column, but only until the density structure becomes neutrally stable ($i.e.$ until the mixed portion of the water column has \textit{exactly} the density of the water just below) \citep{Madec1991a}. This algorithm is an iterative process used in the following way
215(Fig. \ref{Fig_npc}): going from the top of the ocean towards the bottom, the first
216instability is searched. Assume in the following that the instability is
217located between levels $k$ and $k+1$. The two levels are vertically mixed, for
218potential temperature and salinity, conserving the heat and salt contents of
219the water column. The new density is then computed by a linear
220approximation. If the new density profile is still unstable between levels
221$k+1$ and $k+2$, levels $k$, $k+1$ and $k+2$ are then mixed. This process is repeated until
222stability is established below the level $k$ (the mixing process can go down to
223the ocean bottom). The algorithm is repeated to check if the density profile
224between level $k-1$ and $k$ is unstable and/or if there is no deeper instability.
225
226This algorithm is significantly different from mixing two by two statically unstable levels. The latter procedure cannot converge with a finite number
227of iterations for some vertical profiles while the algorithm used in OPA
228converges for any profile in a number of iterations less than the number of
229vertical levels. This property is of paramount importance as pointed out by
230\citet{Killworth1989}: it avoids the existence of permanent and unrealistic
231static instabilities at the sea surface. This non-penetrative convective
232algorithm has been proved successful in studying the deep water formation in
233the north-western Mediterranean Sea \citep{Madec1991a, Madec1991b, Madec1991c}.
234
235Note that in this algorithm the potential density referenced to the sea
236surface is used to check whether the density profile is stable or not.
237Moreover, the mixing in potential density is assumed to be linear. This
238assures the convergence of the algorithm even when the equation of state is
239non-linear. Small static instabilities can thus persist due to cabbeling:
240they will be treated at the next time step. Moreover, temperature and
241salinity, and thus density, are mixed, but the corresponding velocity fields
242remain unchanged. When using a Richardson dependent eddy viscosity, the
243mixing of momentum is done through the vertical diffusion: after a static
244adjustment, the Richardson number is zero and thus the eddy viscosity
245coefficient is at a maximum. When this algorithm is used with constant
246vertical eddy viscosity, spurious solution can occur as the vertical
247momentum diffusion remains small even after a static adjustment. In that
248latter case, we recommend to add momentum mixing in a manner that mimics the
249mixing in temperature and salinity \citep{Speich1992, Speich1996}.
250
251%AMT This presentation fails to mention the big drawback of this scheme: many water masses of the world ocean, especially AABW, are unstable when represented in surface-referenced potential density sigma_0. The scheme erroneously mixes them up.
252
253% -------------------------------------------------------------------------------------------------------------
254%       Enhanced Vertical Diffusion
255% -------------------------------------------------------------------------------------------------------------
256\subsection{Enhanced Vertical Diffusion (\np{ln\_zdfevd}=T)}
257\label{ZDF_evd}
258
259%--------------------------------------------namzdf--------------------------------------------------------
260\namdisplay{namzdf}
261%--------------------------------------------------------------------------------------------------------------
262
263The enhanced vertical diffusion parameterization is used when \np{ln\_zdfevd} is
264defined. In this case, the vertical eddy mixing coefficients are assigned to be very large (a typical value is $1\;m^2s^{-1})$ in regions where the stratification is unstable (i.e. when the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar1997, Lazar1999}. This is done either on tracers only (\np{n\_evdm}=0) or on both momentum and
265tracers (\np{n\_evdm}=1) mixing coefficients.
266
267In practice, when $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$ are set to a large value,
268\np{avevd}, and  if \np{n\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm} $ . Typical value for $avevd$ is in-between 1 and $100~m^2.s^{-1}$. This parameterisation of convective processes is less time consuming than the convective adjustment algorithm presented above when mixing both tracers and momentum in case of static instabilities. It requires the use of an implicit time stepping on vertical diffusion terms (i.e. \np{ln\_zdfexp}=F).
269
270% -------------------------------------------------------------------------------------------------------------
271%       Turbulent Closure Scheme
272% -------------------------------------------------------------------------------------------------------------
273\subsection{Turbulent Closure Scheme (\key{zdftke})}
274\label{ZDF_tcs}
275
276The turbulent closure scheme presented in \S\ref{ZDF_tke} and used when the
277\key{zdftke} is defined allows, in theory, to deal with
278statically unstable density profiles. In such a
279case, the term of destruction of turbulent kinetic energy through
280stratification in \eqref{Eq_zdftke_e} becomes a source term as $N^2$ is negative. It
281results large values of both $A_T^{vT}$ and the four neighbouring$A_u^{vm}
282{and}\;A_v^{vm} $ (up to $1\;m^2s^{-1})$ that are able to restore the
283static stability of the water column in a way similar to that of the
284enhanced vertical diffusion parameterization (\S\ref{ZDF_evd}). Nevertheless, the
285eddy coefficients computed by the turbulent scheme do usually not exceed
286$10^{-2}m.s^{-1}$ in the vicinity of the sea surface (first ocean layer) due to
287the bound of the turbulent length scale by the distance to the sea surface
288(see {\S}VI.7-c). It can thus be useful to combine the enhanced vertical
289diffusion with the turbulent closure, i.e. defining \np{np\_zdfevd} and
290\key{zdftke} CPP variables all together.
291
292The KPP scheme includes enhanced vertical diffusion in the case of convection, as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, therefore \np{np\_zdfevd} should not be used with the KPP scheme. %gm%  + one word on non local flux with KPP scheme
293
294
295% ================================================================
296% Double Diffusion Mixing
297% ================================================================
298\section{Double Diffusion Mixing (\key{zdfddm})}
299\label{ZDF_ddm}
300
301%-------------------------------------------namddm--------------------------------------------------------
302\namdisplay{namddm}
303%--------------------------------------------------------------------------------------------------------------
304
305Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa. The former condition leads to salt fingering and the latter to diffusive convection. Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the oceans.  \citet{Merryfield1999} include a parameterization of such phenomena in a global ocean model and show that it leads to relatively minor changes in circulation but exerts significant regional influences on temperature and salinity.
306
307Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
308\begin{align*} % \label{Eq_zdfddm_Kz}
309    &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT}  \\
310    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
311\end{align*}
312where subscript $f$ represents mixing by salt fingering,
313$d$ by diffusive convection, and $o$ by processes other than
314double diffusion. The rates of double-diffusive mixing depend on buoyancy ratio $R_\rho = \alpha \partial_z T / \partial_z S$, where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline contraction (see \S\ref{TRA_eos}. To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt (1981):
315\begin{align} \label{Eq_zdfddm_f}
316A_f^{vS} &=    \begin{cases}
317   \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
318   0                              &\text{otherwise} 
319            \end{cases}   
320\\
321A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
322\end{align}
323
324%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
325\begin{figure}[!t] \label{Fig_zdfddm}  \begin{center}
326\includegraphics[width=0.99\textwidth]{./Figures/Fig_zdfddm.pdf}
327\caption {From \citet{Merryfield1999} : (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in regions of salt fingering. Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$ ; (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vT}$ for temperature and salt in regions of diffusive convection. Heavy curves denote the Federov parameterization and thin curves the Kelley parameterization. The latter is not implemented in \NEMO }
328\end{center}    \end{figure}
329%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
330
331The factor 0.7 in \eqref{Eq_zdfddm_f} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of buoyancy fluxes due to transport of heat and salt (e.g., McDougall and Taylor 1984). Following  \citet{Merryfield1999}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
332
333To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:
334\begin{align} \label{Eq_zdfddm_d}
335A_d^{vT} &=    \begin{cases}
336   1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
337                           &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
338   0                       &\text{otherwise} 
339            \end{cases}   
340\\
341A_d^{vS} &=    \begin{cases}
342   A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right)
343                           &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
344   A_d^{vT} \ 0.15 \ R_\rho
345                           &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
346   0                       &\text{otherwise} 
347            \end{cases}   
348\end{align}
349
350The dependences of \eqref{Eq_zdfddm_f} to \eqref{Eq_zdfddm_d} on $R_\rho$ are illustrated in Fig.~\ref{Fig_zdfddm}. Implementing this requires computing $R_\rho$ at each
351grid point and time step. This is done in \mdl{eosbn2} at the same time as $N^2$ is computed. This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive).
352
353
354% ================================================================
355% Bottom Friction
356% ================================================================
357\section{Bottom Friction}
358\label{ZDF_bfr}
359
360%--------------------------------------------nambfr--------------------------------------------------------
361\namdisplay{nambfr}
362%--------------------------------------------------------------------------------------------------------------
363
364Both surface momentum flux (wind stress) and the bottom momentum flux
365(bottom friction) enter the equations as a condition on the vertical
366diffusive flux. For the bottom boundary layer, one has:
367\begin{equation} \label{Eq_zdfbfr_flux}
368A^{vm} \left( \partial \textbf{U}_h / \partial z \right) = \textbf{F}_h
369\end{equation}
370where $\textbf{F}_h$ is supposed to represent the horizontal momentum flux
371outside the logarithmic turbulent boundary layer (thickness of the order of
3721~m in the ocean). How $\textbf{F}_h$ influences the interior depends on the
373vertical resolution of the model near the bottom relative to the Ekman layer
374depth. For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m, one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$ (for a Coriolis
375frequency $f = 10^{-4}$~m$^2$s$^{-1}$). With a background diffusion coefficient
376$A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m. When the vertical mixing coefficient is this small, using a flux condition is equivalent to entering the
377viscous forces (either wind stress or bottom friction) as a body force over
378the depth of the top or bottom model layer. To illustrate this, consider the
379equation for $u$ at $k$, the last ocean level:
380\begin{equation} \label{Eq_zdfbfr_flux2}
381\frac{\partial u \; (k)}{\partial t} = \frac{1}{e_{3u}} \left[ A^{vm} \; (k) \frac{U \; (k-1) - U \; (k)}{e_{3uw} \; (k-1)} - F_u \right] \approx - \frac{F_u}{e_{3u}}
382\end{equation}
383For example, if the bottom layer thickness is 200~m, the Ekman transport will be
384distributed over that depth. On the other hand, if the vertical resolution
385is high (1~m or less) and a turbulent closure model is used, the turbulent
386Ekman layer will be represented explicitly by the model. However, the
387logarithmic layer is never represented in current primitive equation model
388applications: it is \emph{necessary} to parameterize the flux $\textbf{F}_h $. Two
389choices are available in OPA: a linear and a quadratic bottom friction. Note
390that in both cases, the rotation between the interior velocity and the
391bottom friction is neglected in the present release of OPA.
392
393% -------------------------------------------------------------------------------------------------------------
394%       Linear Bottom Friction
395% -------------------------------------------------------------------------------------------------------------
396\subsection{Linear Bottom Friction}
397\label{ZDF_bfr_linear}
398
399%-------------------------------------------nambfr--------------------------------------------------------
400\begin{flushright}
401(\textbf{namelist} !nbotfr :\textit{ nbotfr = 0, = 1 or = 3})
402\end{flushright}
403
404The linear bottom friction parameterization assumes that the bottom friction is proportional to the interior velocity (i.e. the velocity of the last model level):
405\begin{equation} \label{Eq_zdfbfr_linear}
406\textbf{F}_h = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \textbf{U}_h^b
407\end{equation}
408where $\textbf{U}_h^b$ is the horizontal velocity vector of the bottom
409ocean layer and $r$ a friction coefficient expressed in m.s$^{-1}$. This
410coefficient is generally estimated by setting a typical decay time $\tau $ in the
411deep ocean, $r = H / \tau$. Commonly accepted values of $\tau$ are of the
412order of 100 to 200 days \citep{Weatherly1984}. A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ 
413corresponding to 115 days is usually used in quasi-geostrophic models. One may
414consider the linear friction as an approximation of quadratic friction,
415$r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982}, Eq. 9.6.6). With a drag coefficient $C_D = 0.002$, a typical value of tidal currents $U_{av} =0.1$~m.s$^{-1}$,
416and assuming an ocean depth $H = 4000$~m, the resulting friction
417coefficient is $r = 4\;10^{-4}$~m.s$^{-1}$. This is the default value used in
418OPA. It corresponds to a decay time scale of 115~days. It can be changed by
419specifying \np{bfric1} (namelist parameter).
420
421In the code, the bottom friction is specified by updating the value of the
422vertical eddy coefficient at the bottom level. Indeed, the discrete
423formulation of (\ref{Eq_zdfbfr_linear}) at the last ocean $T-$level, using the fact that
424$\textbf {U}_h =0$ inside the bottom, leads to
425\begin{equation} \label{Eq_zdfbfr_linKz}
426\begin{split}
427A_u^{vm} &= r\;e_{3uw}\\
428A_v^{vm} &= r\;e_{3uw}\\
429\end{split}
430\end{equation}
431
432Such an update is done in \mdl{zdfbfr} when \np{nbotfr}=1 and the value of $r$ used is
433\np{bfric1}. Setting \np{nbotfr}=3 is equivalent to set $r=0$ and leads to a free-slip bottom boundary condition,
434while setting \np{nbotfr}=0 imposes $r=2\;A_{vb}^{\rm {\bf U}} $, where $A_{vb}^{\rm {\bf U}} $ is the
435background vertical eddy coefficient: a no-slip boundary condition is used.
436Note that this latter choice generally leads to an underestimation of the
437bottom friction: for a deepest level thickness of $200~m$ and $A_{vb}^{\rm {\bf U}}
438=10^{-4}$m$^2$.s$^{-1}$, the friction coefficient is only $r=10^{-6}$m.s$^{-1}$.
439
440
441% -------------------------------------------------------------------------------------------------------------
442%       Non-Linear Bottom Friction
443% -------------------------------------------------------------------------------------------------------------
444\subsection{Non-Linear Bottom Friction}
445\label{ZDF_bfr_nonlinear}
446
447\begin{center}
448(\textbf{namelist} !nbotfr : \textit{nbotfr = 2})
449\end{center}
450
451The non-linear bottom friction parameterization assumes that the bottom
452friction is quadratic:
453\begin{equation} \label{Eq_zdfbfr_nonlinear}
454\textbf {F}_h = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
455}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
456\end{equation}
457
458with $\textbf{U}_h^b = (u_b\;,\;v_b)$ the horizontal interior velocity (i.e. the horizontal velocity of the bottom ocean layer), $C_D$ a drag coefficient, and $e_b $ a bottom turbulent kinetic energy due to tides, internal waves breaking and other short time scale currents. A typical value of the drag coefficient is $C_D = 10^{-3} $. As an example, the CME experiment \citep{Treguier1992} uses $C_D = 10^{-3}$ and $e_b = 2.5\;10^{-3}$m$^2$.s$^{-2}$, while the FRAM experiment \citep{Killworth1992} uses $e_b =0$ 
459and $e_b =2.5\;\;10^{-3}$m$^2$.s$^{-2}$. The FRAM choices have been set as
460default value (\np{bfric2} and \np{bfeb2} namelist parameters).
461
462As for the linear case, the bottom friction is specified in the code by
463updating the value of the vertical eddy coefficient at the bottom level:
464
465\begin{equation} \label{Eq_zdfbfr_nonlinKz}
466\begin{split}
467A_u^{vm} &=C_D\; e_{3uw} \left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^
468{1/2}\\
469A_v^{vm} &=C_D\; e_{3uw} \left[  \left(\bar{\bar{u}}^{i+1,j}\right)^2 + v^2 + e_b \right]^{1/2}\\
470\end{split}
471\end{equation}
472
473This update is done in \mdl{zdfbfr}. The coefficients that control the strength of the
474non-linear bottom friction are initialized as namelist parameters: ($C_D$= \np{bfri2}, and $e_b$ =\np{bfeb2}).
475
476% ================================================================
Note: See TracBrowser for help on using the repository browser.