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 branches/2012/dev_MERGE_2012/DOC/TexFiles/Chapters – NEMO

source: branches/2012/dev_MERGE_2012/DOC/TexFiles/Chapters/Chap_ZDF.tex @ 3764

Last change on this file since 3764 was 3764, checked in by smasson, 11 years ago

dev_MERGE_2012: report bugfixes done in the trunk from r3555 to r3763 into dev_MERGE_2012

File size: 72.0 KB
Line 
1% ================================================================
2% Chapter  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\newpage
12$\ $\newline    % force a new ligne
13
14
15% ================================================================
16% Vertical Mixing
17% ================================================================
18\section{Vertical Mixing}
19\label{ZDF_zdf}
20
21The discrete form of the ocean subgrid scale physics has been presented in
22\S\ref{TRA_zdf} and \S\ref{DYN_zdf}. At the surface and bottom boundaries,
23the turbulent fluxes of momentum, heat and salt have to be defined. At the
24surface they are prescribed from the surface forcing (see Chap.~\ref{SBC}),
25while at the bottom they are set to zero for heat and salt, unless a geothermal
26flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl} 
27defined, see \S\ref{TRA_bbc}), and specified through a bottom friction
28parameterisation for momentum (see \S\ref{ZDF_bfr}).
29
30In this section we briefly discuss the various choices offered to compute
31the vertical eddy viscosity and diffusivity coefficients, $A_u^{vm}$ ,
32$A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$-
33points, respectively (see \S\ref{TRA_zdf} and \S\ref{DYN_zdf}). These
34coefficients can be assumed to be either constant, or a function of the local
35Richardson number, or computed from a turbulent closure model (either
36TKE or KPP formulation). The computation of these coefficients is initialized
37in the \mdl{zdfini} module and performed in the \mdl{zdfric}, \mdl{zdftke} or
38\mdl{zdfkpp} modules. The trends due to the vertical momentum and tracer
39diffusion, including the surface forcing, are computed and added to the
40general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.
41These trends can be computed using either a forward time stepping scheme
42(namelist parameter \np{ln\_zdfexp}=true) or a backward time stepping
43scheme (\np{ln\_zdfexp}=false) depending on the magnitude of the mixing
44coefficients, and thus of the formulation used (see \S\ref{STP}).
45
46% -------------------------------------------------------------------------------------------------------------
47%        Constant
48% -------------------------------------------------------------------------------------------------------------
49\subsection{Constant (\key{zdfcst})}
50\label{ZDF_cst}
51%--------------------------------------------namzdf---------------------------------------------------------
52\namdisplay{namzdf}
53%--------------------------------------------------------------------------------------------------------------
54
55When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients
56are set to constant values over the whole ocean. This is the crudest way to define
57the vertical ocean physics. It is recommended that this option is only used in
58process studies, not in basin scale simulations. Typical values used in this case are:
59\begin{align*} 
60A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\
61A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
62\end{align*}
63
64These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.
65In all cases, do not use values smaller that those associated with the molecular
66viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum,
67$\sim10^{-7}~m^2.s^{-1}$ for temperature and $\sim10^{-9}~m^2.s^{-1}$ for salinity.
68
69
70% -------------------------------------------------------------------------------------------------------------
71%        Richardson Number Dependent
72% -------------------------------------------------------------------------------------------------------------
73\subsection{Richardson Number Dependent (\key{zdfric})}
74\label{ZDF_ric}
75
76%--------------------------------------------namric---------------------------------------------------------
77\namdisplay{namzdf_ric}
78%--------------------------------------------------------------------------------------------------------------
79
80When \key{zdfric} is defined, a local Richardson number dependent formulation
81for the vertical momentum and tracer eddy coefficients is set. The vertical mixing
82coefficients are diagnosed from the large scale variables computed by the model.
83\textit{In situ} measurements have been used to link vertical turbulent activity to
84large scale ocean structures. The hypothesis of a mixing mainly maintained by the
85growth of Kelvin-Helmholtz like instabilities leads to a dependency between the
86vertical eddy coefficients and the local Richardson number ($i.e.$ the
87ratio of stratification to vertical shear). Following \citet{Pacanowski_Philander_JPO81}, the following
88formulation has been implemented:
89\begin{equation} \label{Eq_zdfric}
90   \left\{      \begin{aligned}
91         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
92         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
93   \end{aligned}    \right.
94\end{equation}
95where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson
96number, $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
97$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the
98constant case (see \S\ref{ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ 
99is the maximum value that can be reached by the coefficient when $Ri\leq 0$,
100$a=5$ and $n=2$. The last three values can be modified by setting the
101\np{rn\_avmri}, \np{rn\_alp} and \np{nn\_ric} namelist parameters, respectively.
102
103A simple mixing-layer model to transfer and dissipate the atmospheric
104 forcings (wind-stress and buoyancy fluxes) can be activated setting
105the \np{ln\_mldw} =.true. in the namelist.
106
107In this case, the local depth of turbulent wind-mixing or "Ekman depth"
108 $h_{e}(x,y,t)$ is evaluated and the vertical eddy coefficients prescribed within this layer.
109
110This depth is assumed proportional to the "depth of frictional influence" that is limited by rotation:
111\begin{equation}
112         h_{e} = Ek \frac {u^{*}} {f_{0}}    \\
113\end{equation}
114where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis
115parameter.
116
117In this similarity height relationship, the turbulent friction velocity:
118\begin{equation}
119         u^{*} = \sqrt \frac {|\tau|} {\rho_o}     \\
120\end{equation}
121
122is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$.
123The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}.
124Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to
125the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{Lermusiaux2001}.
126
127% -------------------------------------------------------------------------------------------------------------
128%        TKE Turbulent Closure Scheme
129% -------------------------------------------------------------------------------------------------------------
130\subsection{TKE Turbulent Closure Scheme (\key{zdftke})}
131\label{ZDF_tke}
132
133%--------------------------------------------namzdf_tke--------------------------------------------------
134\namdisplay{namzdf_tke}
135%--------------------------------------------------------------------------------------------------------------
136
137The vertical eddy viscosity and diffusivity coefficients are computed from a TKE
138turbulent closure model based on a prognostic equation for $\bar{e}$, the turbulent
139kinetic energy, and a closure assumption for the turbulent length scales. This
140turbulent closure model has been developed by \citet{Bougeault1989} in the
141atmospheric case, adapted by \citet{Gaspar1990} for the oceanic case, and
142embedded in OPA, the ancestor of NEMO, by \citet{Blanke1993} for equatorial Atlantic
143simulations. Since then, significant modifications have been introduced by
144\citet{Madec1998} in both the implementation and the formulation of the mixing
145length scale. The time evolution of $\bar{e}$ is the result of the production of
146$\bar{e}$ through vertical shear, its destruction through stratification, its vertical
147diffusion, and its dissipation of \citet{Kolmogorov1942} type:
148\begin{equation} \label{Eq_zdftke_e}
149\frac{\partial \bar{e}}{\partial t} =
150\frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
151                    +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
152-K_\rho\,N^2
153+\frac{1}{e_3} \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
154            \;\frac{\partial \bar{e}}{\partial k}} \right]
155- c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
156\end{equation}
157\begin{equation} \label{Eq_zdftke_kz}
158   \begin{split}
159         K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }     \\
160         K_\rho &= A^{vm} / P_{rt}
161   \end{split}
162\end{equation}
163where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
164$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,
165$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity
166and diffusivity coefficients. The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ 
167$\approx 0.7$ are designed to deal with vertical mixing at any depth \citep{Gaspar1990}.
168They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}.
169$P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function
170of the local Richardson number, $R_i$:
171\begin{align*} \label{Eq_prt}
172P_{rt} = \begin{cases}
173                    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}  \\
174                    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}  \\
175                    \ \ 10 &      \text{if $\ 2 \leq R_i$} 
176            \end{cases}
177\end{align*}
178The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist parameter.
179
180At the sea surface, the value of $\bar{e}$ is prescribed from the wind
181stress field as $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} 
182namelist parameter. The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}),
183however a much larger value can be used when taking into account the
184surface wave breaking (see below Eq. \eqref{ZDF_Esbc}).
185The bottom value of TKE is assumed to be equal to the value of the level just above.
186The time integration of the $\bar{e}$ equation may formally lead to negative values
187because the numerical scheme does not ensure its positivity. To overcome this
188problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} 
189namelist parameter). Following \citet{Gaspar1990}, the cut-off value is set
190to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. This allows the subsequent formulations
191to match that of \citet{Gargett1984} for the diffusion in the thermocline and
192deep ocean :  $K_\rho = 10^{-3} / N$.
193In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical
194instabilities associated with too weak vertical diffusion. They must be
195specified at least larger than the molecular values, and are set through
196\np{rn\_avm0} and \np{rn\_avt0} (namzdf namelist, see \S\ref{ZDF_cst}).
197
198\subsubsection{Turbulent length scale}
199For computational efficiency, the original formulation of the turbulent length
200scales proposed by \citet{Gaspar1990} has been simplified. Four formulations
201are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist
202parameter. The first two are based on the following first order approximation
203\citep{Blanke1993}:
204\begin{equation} \label{Eq_tke_mxl0_1}
205l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N
206\end{equation}
207which is valid in a stable stratified region with constant values of the Brunt-
208Vais\"{a}l\"{a} frequency. The resulting length scale is bounded by the distance
209to the surface or to the bottom (\np{nn\_mxl} = 0) or by the local vertical scale factor
210(\np{nn\_mxl} = 1). \citet{Blanke1993} notice that this simplification has two major
211drawbacks: it makes no sense for locally unstable stratification and the
212computation no longer uses all the information contained in the vertical density
213profile. To overcome these drawbacks, \citet{Madec1998} introduces the
214\np{nn\_mxl} = 2 or 3 cases, which add an extra assumption concerning the vertical
215gradient of the computed length scale. So, the length scales are first evaluated
216as in \eqref{Eq_tke_mxl0_1} and then bounded such that:
217\begin{equation} \label{Eq_tke_mxl_constraint}
218\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
219\qquad \text{with }\  l =  l_k = l_\epsilon
220\end{equation}
221\eqref{Eq_tke_mxl_constraint} means that the vertical variations of the length
222scale cannot be larger than the variations of depth. It provides a better
223approximation of the \citet{Gaspar1990} formulation while being much less
224time consuming. In particular, it allows the length scale to be limited not only
225by the distance to the surface or to the ocean bottom but also by the distance
226to a strongly stratified portion of the water column such as the thermocline
227(Fig.~\ref{Fig_mixing_length}). In order to impose the \eqref{Eq_tke_mxl_constraint} 
228constraint, we introduce two additional length scales: $l_{up}$ and $l_{dwn}$,
229the upward and downward length scales, and evaluate the dissipation and
230mixing length scales as (and note that here we use numerical indexing):
231%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
232\begin{figure}[!t] \begin{center}
233\includegraphics[width=1.00\textwidth]{./TexFiles/Figures/Fig_mixing_length.pdf}
234\caption{ \label{Fig_mixing_length} 
235Illustration of the mixing length computation. }
236\end{center} 
237\end{figure}
238%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
239\begin{equation} \label{Eq_tke_mxl2}
240\begin{aligned}
241 l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right)
242    \quad &\text{ from $k=1$ to $jpk$ }\ \\
243 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)}  \right)   
244    \quad &\text{ from $k=jpk$ to $1$ }\ \\
245\end{aligned}
246\end{equation}
247where $l^{(k)}$ is computed using \eqref{Eq_tke_mxl0_1},
248$i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$.
249
250In the \np{nn\_mxl}~=~2 case, the dissipation and mixing length scales take the same
251value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the
252\np{nn\_mxl}~=~3 case, the dissipation and mixing turbulent length scales are give
253as in \citet{Gaspar1990}:
254\begin{equation} \label{Eq_tke_mxl_gaspar}
255\begin{aligned}
256& l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }    \\
257& l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
258\end{aligned}
259\end{equation}
260
261At the ocean surface, a non zero length scale is set through the  \np{rn\_lmin0} namelist
262parameter. Usually the surface scale is given by $l_o = \kappa \,z_o$ 
263where $\kappa = 0.4$ is von Karman's constant and $z_o$ the roughness
264parameter of the surface. Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94} 
265leads to a 0.04~m, the default value of \np{rn\_lsurf}. In the ocean interior
266a minimum length scale is set to recover the molecular viscosity when $\bar{e}$ 
267reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ).
268
269
270\subsubsection{Surface wave breaking parameterization}
271%-----------------------------------------------------------------------%
272Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified
273to include the effect of surface wave breaking energetics. This results in a reduction of summertime
274surface temperature when the mixed layer is relatively shallow. The \citet{Mellor_Blumberg_JPO04} 
275modifications acts on surface length scale and TKE values and air-sea drag coefficient.
276The latter concerns the bulk formulea and is not discussed here.
277
278Following \citet{Craig_Banner_JPO94}, the boundary condition on surface TKE value is :
279\begin{equation}  \label{ZDF_Esbc}
280\bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
281\end{equation}
282where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality
283which depends on the ''wave age'', ranging from 57 for mature waves to 146 for
284younger waves \citep{Mellor_Blumberg_JPO04}.
285The boundary condition on the turbulent length scale follows the Charnock's relation:
286\begin{equation} \label{ZDF_Lsbc}
287l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
288\end{equation}
289where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant.
290\citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by \citet{Stacey_JPO99}
291citing observation evidence, and $\alpha_{CB} = 100$ the Craig and Banner's value.
292As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,
293with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}~=~67.83 corresponds
294to $\alpha_{CB} = 100$. further setting  \np{ln\_lsurf} to true applies \eqref{ZDF_Lsbc} 
295as surface boundary condition on length scale, with $\beta$ hard coded to the Stacet's value.
296Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters)
297is applied on surface $\bar{e}$ value.
298
299
300\subsubsection{Langmuir cells}
301%--------------------------------------%
302Langmuir circulations (LC) can be described as ordered large-scale vertical motions
303in the surface layer of the oceans. Although LC have nothing to do with convection,
304the circulation pattern is rather similar to so-called convective rolls in the atmospheric
305boundary layer. The detailed physics behind LC is described in, for example,
306\citet{Craik_Leibovich_JFM76}. The prevailing explanation is that LC arise from
307a nonlinear interaction between the Stokes drift and wind drift currents.
308
309Here we introduced in the TKE turbulent closure the simple parameterization of
310Langmuir circulations proposed by \citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure.
311The parameterization, tuned against large-eddy simulation, includes the whole effect
312of LC in an extra source terms of TKE, $P_{LC}$.
313The presence of $P_{LC}$ in \eqref{Eq_zdftke_e}, the TKE equation, is controlled
314by setting \np{ln\_lc} to \textit{true} in the namtke namelist.
315 
316By making an analogy with the characteristic convective velocity scale
317($e.g.$, \citet{D'Alessio_al_JPO98}), $P_{LC}$ is assumed to be :
318\begin{equation}
319P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
320\end{equation}
321where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
322With no information about the wave field, $w_{LC}$ is assumed to be proportional to
323the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
324\footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity
325may be expressed as $u_s =  0.016 \,|U_{10m}|$. Assuming an air density of
326$\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of $1.5~10^{-3}$ give the expression
327used of $u_s$ as a function of the module of surface stress}.
328For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as
329at a finite depth $H_{LC}$ (which is often close to the mixed layer depth), and simply
330varies as a sine function in between (a first-order profile for the Langmuir cell structures).
331The resulting expression for $w_{LC}$ is :
332\begin{equation}
333w_{LC}  = \begin{cases}
334                   c_{LC} \,u_s \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
335                   0                             &      \text{otherwise} 
336                 \end{cases}
337\end{equation}
338where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise
339to fit LES data. The chosen value yields maximum vertical velocities $w_{LC}$ of the order
340of a few centimeters per second. The value of $c_{LC}$ is set through the \np{rn\_lc} 
341namelist parameter, having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.
342
343The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
344$H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift
345can reach on its own by converting its kinetic energy to potential energy, according to
346\begin{equation}
347- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} u_s^2
348\end{equation}
349
350
351\subsubsection{Mixing just below the mixed layer}
352%--------------------------------------------------------------%
353
354To be add here a description of "penetration of TKE" and the associated namelist parameters
355 \np{nn\_etau}, \np{rn\_efr} and \np{nn\_htau}.
356
357% from Burchard et al OM 2008 :
358% the most critical process not reproduced by statistical turbulence models is the activity of internal waves and their interaction with turbulence. After the Reynolds decomposition, internal waves are in principle included in the RANS equations, but later partially excluded by the hydrostatic assumption and the model resolution. Thus far, the representation of internal wave mixing in ocean models has been relatively crude (e.g. Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002).
359
360
361
362% -------------------------------------------------------------------------------------------------------------
363%        TKE discretization considerations
364% -------------------------------------------------------------------------------------------------------------
365\subsection{TKE discretization considerations (\key{zdftke})}
366\label{ZDF_tke_ene}
367
368%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
369\begin{figure}[!t]   \begin{center}
370\includegraphics[width=1.00\textwidth]{./TexFiles/Figures/Fig_ZDF_TKE_time_scheme.pdf}
371\caption{ \label{Fig_TKE_time_scheme} 
372Illustration of the TKE time integration and its links to the momentum and tracer time integration. }
373\end{center} 
374\end{figure}
375%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
376
377The production of turbulence by vertical shear (the first term of the right hand side
378of \eqref{Eq_zdftke_e}) should balance the loss of kinetic energy associated with
379the vertical momentum diffusion (first line in \eqref{Eq_PE_zdf}). To do so a special care
380have to be taken for both the time and space discretization of the TKE equation
381\citep{Burchard_OM02,Marsaleix_al_OM08}.
382
383Let us first address the time stepping issue. Fig.~\ref{Fig_TKE_time_scheme} shows
384how the two-level Leap-Frog time stepping of the momentum and tracer equations interplays
385with the one-level forward time stepping of TKE equation. With this framework, the total loss
386of kinetic energy (in 1D for the demonstration) due to the vertical momentum diffusion is
387obtained by multiplying this quantity by $u^t$ and summing the result vertically:   
388\begin{equation} \label{Eq_energ1}
389\begin{split}
390\int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
391&= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}         
392 - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
393\end{split}
394\end{equation}
395Here, the vertical diffusion of momentum is discretized backward in time
396with a coefficient, $K_m$, known at time $t$ (Fig.~\ref{Fig_TKE_time_scheme}),
397as it is required when using the TKE scheme (see \S\ref{STP_forward_imp}).
398The first term of the right hand side of \eqref{Eq_energ1} represents the kinetic energy
399transfer at the surface (atmospheric forcing) and at the bottom (friction effect).
400The second term is always negative. It is the dissipation rate of kinetic energy,
401and thus minus the shear production rate of $\bar{e}$. \eqref{Eq_energ1} 
402implies that, to be energetically consistent, the production rate of $\bar{e}$ 
403used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
404${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ (and not by the more straightforward
405$K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
406
407A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
408(second term of the right hand side of \eqref{Eq_zdftke_e}). This term
409must balance the input of potential energy resulting from vertical mixing.
410The rate of change of potential energy (in 1D for the demonstration) due vertical
411mixing is obtained by multiplying vertical density diffusion
412tendency by $g\,z$ and and summing the result vertically:
413\begin{equation} \label{Eq_energ2}
414\begin{split}
415\int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
416&= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta} 
417   - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
418&= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
419+ \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
420\end{split}
421\end{equation}
422where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
423The first term of the right hand side of \eqref{Eq_energ2}  is always zero
424because there is no diffusive flux through the ocean surface and bottom).
425The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
426Therefore \eqref{Eq_energ1} implies that, to be energetically consistent, the product
427${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \eqref{Eq_zdftke_e}, the TKE equation.
428
429Let us now address the space discretization issue.
430The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity
431components are in the centre of the side faces of a $t$-box in staggered C-grid
432(Fig.\ref{Fig_cell}). A space averaging is thus required to obtain the shear TKE production term.
433By redoing the \eqref{Eq_energ1} in the 3D case, it can be shown that the product of
434eddy coefficient by the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
435Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into
436account.
437
438The above energetic considerations leads to
439the following final discrete form for the TKE equation:
440\begin{equation} \label{Eq_zdftke_ene}
441\begin{split}
442\frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv 
443\Biggl\{ \Biggr.
444  &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} } 
445                                                                              \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
446+&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} } 
447                                                                               \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j} 
448\Biggr. \Biggr\}   \\
449%
450- &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
451%
452+&\frac{1}{{e_3w}^{t+\rdt}}  \;\delta_{k+1/2} \left[   {K_m}^{t-\rdt} \,\frac{\delta_{k}[(\bar{e})^{t+\rdt}]} {{e_3w}^{t+\rdt}}   \right]   \\
453%
454- &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
455\end{split}
456\end{equation}
457where the last two terms in \eqref{Eq_zdftke_ene} (vertical diffusion and Kolmogorov dissipation)
458are time stepped using a backward scheme (see\S\ref{STP_forward_imp}).
459Note that the Kolmogorov term has been linearized in time in order to render
460the implicit computation possible. The restart of the TKE scheme
461requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as they all appear in
462the right hand side of \eqref{Eq_zdftke_ene}. For the latter, it is in fact
463the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
464
465% -------------------------------------------------------------------------------------------------------------
466%        GLS Generic Length Scale Scheme
467% -------------------------------------------------------------------------------------------------------------
468\subsection{GLS Generic Length Scale (\key{zdfgls})}
469\label{ZDF_gls}
470
471%--------------------------------------------namzdf_gls---------------------------------------------------------
472\namdisplay{namzdf_gls}
473%--------------------------------------------------------------------------------------------------------------
474
475The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on
476two prognostic equations: one for the turbulent kinetic energy $\bar {e}$, and another
477for the generic length scale, $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}.
478This later variable is defined as : $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
479where the triplet $(p, m, n)$ value given in Tab.\ref{Tab_GLS} allows to recover
480a number of well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982},
481$k$-$\epsilon$ \citep{Rodi_1987}, $k$-$\omega$ \citep{Wilcox_1988} 
482among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).
483The GLS scheme is given by the following set of equations:
484\begin{equation} \label{Eq_zdfgls_e}
485\frac{\partial \bar{e}}{\partial t} =
486\frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
487                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
488-K_\rho \,N^2
489+\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
490- \epsilon
491\end{equation}
492
493\begin{equation} \label{Eq_zdfgls_psi}
494   \begin{split}
495\frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
496\frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
497                                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
498- C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
499&+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
500                                  \;\frac{\partial \psi}{\partial k}} \right]\;
501   \end{split}
502\end{equation}
503
504\begin{equation} \label{Eq_zdfgls_kz}
505   \begin{split}
506         K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
507         K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
508   \end{split}
509\end{equation}
510
511\begin{equation} \label{Eq_zdfgls_eps}
512{\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
513\end{equation}
514where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2})
515and $\epsilon$ the dissipation rate.
516The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$)
517depends of the choice of the turbulence model. Four different turbulent models are pre-defined
518(Tab.\ref{Tab_GLS}). They are made available through the \np{nn\_clo} namelist parameter.
519
520%--------------------------------------------------TABLE--------------------------------------------------
521\begin{table}[htbp]  \begin{center}
522%\begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
523\begin{tabular}{ccccc}
524                         &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
525%                        & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\ 
526\hline  \hline 
527\np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
528\hline 
529$( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
530$\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
531$\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
532$C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
533$C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
534$C_3$              &      1.           &     1.              &      1.                &       1.           \\
535$F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
536\hline
537\hline
538\end{tabular}
539\caption{   \label{Tab_GLS} 
540Set of predefined GLS parameters, or equivalently predefined turbulence models available
541with \key{zdfgls} and controlled by the \np{nn\_clos} namelist parameter.}
542\end{center}   \end{table}
543%--------------------------------------------------------------------------------------------------------------
544
545In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force
546the convergence of the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length)
547value near physical boundaries (logarithmic boundary layer law). $C_{\mu}$ and $C_{\mu'}$ 
548are calculated from stability function proposed by \citet{Galperin_al_JAS88}, or by \citet{Kantha_Clayson_1994} 
549or one of the two functions suggested by \citet{Canuto_2001}  (\np{nn\_stab\_func} = 0, 1, 2 or 3, resp.}).
550The value of $C_{0\mu}$ depends of the choice of the stability function.
551
552The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated
553thanks to Dirichlet or Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp.
554As for TKE closure , the wave effect on the mixing is considered when \np{ln\_crban}~=~true
555\citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. The \np{rn\_crban} namelist parameter
556is $\alpha_{CB}$ in \eqref{ZDF_Esbc} and \np{rn\_charn} provides the value of $\beta$ in \eqref{ZDF_Lsbc}.
557
558The $\psi$ equation is known to fail in stably stratified flows, and for this reason
559almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
560With this clipping, the maximum permissible length scale is determined by
561$l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. A value of $c_{lim} = 0.53$ is often used
562\citep{Galperin_al_JAS88}. \cite{Umlauf_Burchard_CSR05} show that the value of
563the clipping factor is of crucial importance for the entrainment depth predicted in
564stably stratified situations, and that its value has to be chosen in accordance
565with the algebraic model for the turbulent fluxes. The clipping is only activated
566if \np{ln\_length\_lim}=true, and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value.
567
568The time and space discretization of the GLS equations follows the same energetic
569consideration as for the TKE case described in \S\ref{ZDF_tke_ene}  \citep{Burchard_OM02}.
570Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}.
571
572% -------------------------------------------------------------------------------------------------------------
573%        K Profile Parametrisation (KPP)
574% -------------------------------------------------------------------------------------------------------------
575\subsection{K Profile Parametrisation (KPP) (\key{zdfkpp}) }
576\label{ZDF_kpp}
577
578%--------------------------------------------namkpp--------------------------------------------------------
579\namdisplay{namzdf_kpp}
580%--------------------------------------------------------------------------------------------------------------
581
582The KKP scheme has been implemented by J. Chanut ...
583
584\colorbox{yellow}{Add a description of KPP here.}
585
586
587% ================================================================
588% Convection
589% ================================================================
590\section{Convection}
591\label{ZDF_conv}
592
593%--------------------------------------------namzdf--------------------------------------------------------
594\namdisplay{namzdf}
595%--------------------------------------------------------------------------------------------------------------
596
597Static instabilities (i.e. light potential densities under heavy ones) may
598occur at particular ocean grid points. In nature, convective processes
599quickly re-establish the static stability of the water column. These
600processes have been removed from the model via the hydrostatic
601assumption so they must be parameterized. Three parameterisations
602are available to deal with convective processes: a non-penetrative
603convective adjustment or an enhanced vertical diffusion, or/and the
604use of a turbulent closure scheme.
605
606% -------------------------------------------------------------------------------------------------------------
607%       Non-Penetrative Convective Adjustment
608% -------------------------------------------------------------------------------------------------------------
609\subsection   [Non-Penetrative Convective Adjustment (\np{ln\_tranpc}) ]
610         {Non-Penetrative Convective Adjustment (\np{ln\_tranpc}=.true.) }
611\label{ZDF_npc}
612
613%--------------------------------------------namzdf--------------------------------------------------------
614\namdisplay{namzdf}
615%--------------------------------------------------------------------------------------------------------------
616
617%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
618\begin{figure}[!htb]    \begin{center}
619\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_npc.pdf}
620\caption{  \label{Fig_npc} 
621Example of an unstable density profile treated by the non penetrative
622convective adjustment algorithm. $1^{st}$ step: the initial profile is checked from
623the surface to the bottom. It is found to be unstable between levels 3 and 4.
624They are mixed. The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5
625are mixed. The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are
626mixed. The $1^{st}$ step ends since the density profile is then stable below
627the level 3. $2^{nd}$ step: the new $\rho$ profile is checked following the same
628procedure as in $1^{st}$ step: levels 2 to 5 are mixed. The new density profile
629is checked. It is found stable: end of algorithm.}
630\end{center}   \end{figure}
631%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
632
633The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}=true.
634It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously
635the statically unstable portion of the water column, but only until the density
636structure becomes neutrally stable ($i.e.$ until the mixed portion of the water
637column has \textit{exactly} the density of the water just below) \citep{Madec_al_JPO91}.
638The associated algorithm is an iterative process used in the following way
639(Fig. \ref{Fig_npc}): starting from the top of the ocean, the first instability is
640found. Assume in the following that the instability is located between levels
641$k$ and $k+1$. The potential temperature and salinity in the two levels are
642vertically mixed, conserving the heat and salt contents of the water column.
643The new density is then computed by a linear approximation. If the new
644density profile is still unstable between levels $k+1$ and $k+2$, levels $k$,
645$k+1$ and $k+2$ are then mixed. This process is repeated until stability is
646established below the level $k$ (the mixing process can go down to the
647ocean bottom). The algorithm is repeated to check if the density profile
648between level $k-1$ and $k$ is unstable and/or if there is no deeper instability.
649
650This algorithm is significantly different from mixing statically unstable levels
651two by two. The latter procedure cannot converge with a finite number
652of iterations for some vertical profiles while the algorithm used in \NEMO 
653converges for any profile in a number of iterations which is less than the
654number of vertical levels. This property is of paramount importance as
655pointed out by \citet{Killworth1989}: it avoids the existence of permanent
656and unrealistic static instabilities at the sea surface. This non-penetrative
657convective algorithm has been proved successful in studies of the deep
658water formation in the north-western Mediterranean Sea
659\citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}.
660
661Note that in the current implementation of this algorithm presents several
662limitations. First, potential density referenced to the sea surface is used to
663check whether the density profile is stable or not. This is a strong
664simplification which leads to large errors for realistic ocean simulations.
665Indeed, many water masses of the world ocean, especially Antarctic Bottom
666Water, are unstable when represented in surface-referenced potential density.
667The scheme will erroneously mix them up. Second, the mixing of potential
668density is assumed to be linear. This assures the convergence of the algorithm
669even when the equation of state is non-linear. Small static instabilities can thus
670persist due to cabbeling: they will be treated at the next time step.
671Third, temperature and salinity, and thus density, are mixed, but the
672corresponding velocity fields remain unchanged. When using a Richardson
673Number dependent eddy viscosity, the mixing of momentum is done through
674the vertical diffusion: after a static adjustment, the Richardson Number is zero
675and thus the eddy viscosity coefficient is at a maximum. When this convective
676adjustment algorithm is used with constant vertical eddy viscosity, spurious
677solutions can occur since the vertical momentum diffusion remains small even
678after a static adjustment. In that case, we recommend the addition of momentum
679mixing in a manner that mimics the mixing in temperature and salinity
680\citep{Speich_PhD92, Speich_al_JPO96}.
681
682% -------------------------------------------------------------------------------------------------------------
683%       Enhanced Vertical Diffusion
684% -------------------------------------------------------------------------------------------------------------
685\subsection   [Enhanced Vertical Diffusion (\np{ln\_zdfevd})]
686         {Enhanced Vertical Diffusion (\np{ln\_zdfevd}=true)}
687\label{ZDF_evd}
688
689%--------------------------------------------namzdf--------------------------------------------------------
690\namdisplay{namzdf}
691%--------------------------------------------------------------------------------------------------------------
692
693The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}=true.
694In this case, the vertical eddy mixing coefficients are assigned very large values
695(a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable
696($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative)
697\citep{Lazar_PhD97, Lazar_al_JPO99}. This is done either on tracers only
698(\np{nn\_evdm}=0) or on both momentum and tracers (\np{nn\_evdm}=1).
699
700In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and
701if \np{nn\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ 
702values also, are set equal to the namelist parameter \np{rn\_avevd}. A typical value
703for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterisation of
704convective processes is less time consuming than the convective adjustment
705algorithm presented above when mixing both tracers and momentum in the
706case of static instabilities. It requires the use of an implicit time stepping on
707vertical diffusion terms (i.e. \np{ln\_zdfexp}=false).
708
709Note that the stability test is performed on both \textit{before} and \textit{now} 
710values of $N^2$. This removes a potential source of divergence of odd and
711even time step in a leapfrog environment \citep{Leclair_PhD2010} (see \S\ref{STP_mLF}).
712
713% -------------------------------------------------------------------------------------------------------------
714%       Turbulent Closure Scheme
715% -------------------------------------------------------------------------------------------------------------
716\subsection{Turbulent Closure Scheme (\key{zdftke} or \key{zdfgls})}
717\label{ZDF_tcs}
718
719The turbulent closure scheme presented in \S\ref{ZDF_tke} and \S\ref{ZDF_gls} 
720(\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically
721unstable density profiles. In such a case, the term corresponding to the
722destruction of turbulent kinetic energy through stratification in \eqref{Eq_zdftke_e} 
723or \eqref{Eq_zdfgls_e} becomes a source term, since $N^2$ is negative.
724It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring
725$A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1})$. These large values
726restore the static stability of the water column in a way similar to that of the
727enhanced vertical diffusion parameterisation (\S\ref{ZDF_evd}). However,
728in the vicinity of the sea surface (first ocean layer), the eddy coefficients
729computed by the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
730because the mixing length scale is bounded by the distance to the sea surface.
731It can thus be useful to combine the enhanced vertical
732diffusion with the turbulent closure scheme, $i.e.$ setting the \np{ln\_zdfnpc} 
733namelist parameter to true and defining the turbulent closure CPP key all together.
734
735The KPP turbulent closure scheme already includes enhanced vertical diffusion
736in the case of convection, as governed by the variables $bvsqcon$ and $difcon$ 
737found in \mdl{zdfkpp}, therefore \np{ln\_zdfevd}=false should be used with the KPP
738scheme. %gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
739
740% ================================================================
741% Double Diffusion Mixing
742% ================================================================
743\section  [Double Diffusion Mixing (\key{zdfddm})]
744      {Double Diffusion Mixing (\key{zdfddm})}
745\label{ZDF_ddm}
746
747%-------------------------------------------namzdf_ddm-------------------------------------------------
748\namdisplay{namzdf_ddm}
749%--------------------------------------------------------------------------------------------------------------
750
751Double diffusion occurs when relatively warm, salty water overlies cooler, fresher
752water, or vice versa. The former condition leads to salt fingering and the latter
753to diffusive convection. Double-diffusive phenomena contribute to diapycnal
754mixing in extensive regions of the ocean.  \citet{Merryfield1999} include a
755parameterisation of such phenomena in a global ocean model and show that
756it leads to relatively minor changes in circulation but exerts significant regional
757influences on temperature and salinity. This parameterisation has been
758introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key.
759
760Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
761\begin{align*} % \label{Eq_zdfddm_Kz}
762    &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT}  \\
763    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
764\end{align*}
765where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
766and $o$ by processes other than double diffusion. The rates of double-diffusive
767mixing depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,
768where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline
769contraction (see \S\ref{TRA_eos}). To represent mixing of $S$ and $T$ by salt
770fingering, we adopt the diapycnal diffusivities suggested by Schmitt (1981):
771\begin{align} \label{Eq_zdfddm_f}
772A_f^{vS} &=    \begin{cases}
773   \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
774   0                              &\text{otherwise} 
775            \end{cases}   
776\\           \label{Eq_zdfddm_f_T}
777A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
778\end{align}
779
780%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
781\begin{figure}[!t]   \begin{center}
782\includegraphics[width=0.99\textwidth]{./TexFiles/Figures/Fig_zdfddm.pdf}
783\caption{  \label{Fig_zdfddm}
784From \citet{Merryfield1999} : (a) Diapycnal diffusivities $A_f^{vT}$ 
785and $A_f^{vS}$ for temperature and salt in regions of salt fingering. Heavy
786curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves
787$A^{\ast v} = 10^{-4}~m^2.s^{-1}$ ; (b) diapycnal diffusivities $A_d^{vT}$ and
788$A_d^{vS}$ for temperature and salt in regions of diffusive convection. Heavy
789curves denote the Federov parameterisation and thin curves the Kelley
790parameterisation. The latter is not implemented in \NEMO. }
791\end{center}    \end{figure}
792%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
793
794The factor 0.7 in \eqref{Eq_zdfddm_f_T} reflects the measured ratio
795$\alpha F_T /\beta F_S \approx  0.7$ of buoyancy flux of heat to buoyancy
796flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}). Following  \citet{Merryfield1999},
797we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
798
799To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:
800\begin{align}  \label{Eq_zdfddm_d}
801A_d^{vT} &=    \begin{cases}
802   1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
803                           &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
804   0                       &\text{otherwise} 
805            \end{cases}   
806\\          \label{Eq_zdfddm_d_S}
807A_d^{vS} &=    \begin{cases}
808   A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right)
809                           &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
810   A_d^{vT} \ 0.15 \ R_\rho
811                           &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
812   0                       &\text{otherwise} 
813            \end{cases}   
814\end{align}
815
816The dependencies of \eqref{Eq_zdfddm_f} to \eqref{Eq_zdfddm_d_S} on $R_\rho$ 
817are illustrated in Fig.~\ref{Fig_zdfddm}. Implementing this requires computing
818$R_\rho$ at each grid point on every time step. This is done in \mdl{eosbn2} at the
819same time as $N^2$ is computed. This avoids duplication in the computation of
820$\alpha$ and $\beta$ (which is usually quite expensive).
821
822% ================================================================
823% Bottom Friction
824% ================================================================
825\section  [Bottom Friction (\textit{zdfbfr})]   {Bottom Friction (\mdl{zdfbfr} module)}
826\label{ZDF_bfr}
827
828%--------------------------------------------nambfr--------------------------------------------------------
829\namdisplay{nambfr}
830%--------------------------------------------------------------------------------------------------------------
831
832Both the surface momentum flux (wind stress) and the bottom momentum
833flux (bottom friction) enter the equations as a condition on the vertical
834diffusive flux. For the bottom boundary layer, one has:
835\begin{equation} \label{Eq_zdfbfr_flux}
836A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
837\end{equation}
838where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum
839outside the logarithmic turbulent boundary layer (thickness of the order of
8401~m in the ocean). How ${\cal F}_h^{\textbf U}$ influences the interior depends on the
841vertical resolution of the model near the bottom relative to the Ekman layer
842depth. For example, in order to obtain an Ekman layer depth
843$d = \sqrt{2\;A^{vm}} / f = 50$~m, one needs a vertical diffusion coefficient
844$A^{vm} = 0.125$~m$^2$s$^{-1}$ (for a Coriolis frequency
845$f = 10^{-4}$~m$^2$s$^{-1}$). With a background diffusion coefficient
846$A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
847When the vertical mixing coefficient is this small, using a flux condition is
848equivalent to entering the viscous forces (either wind stress or bottom friction)
849as a body force over the depth of the top or bottom model layer. To illustrate
850this, consider the equation for $u$ at $k$, the last ocean level:
851\begin{equation} \label{Eq_zdfbfr_flux2}
852\frac{\partial u_k}{\partial t} = \frac{1}{e_{3u}} \left[ \frac{A_{uw}^{vm}}{e_{3uw}} \delta_{k+1/2}\;[u] - {\cal F}^u_h \right] \approx - \frac{{\cal F}^u_{h}}{e_{3u}}
853\end{equation}
854If the bottom layer thickness is 200~m, the Ekman transport will
855be distributed over that depth. On the other hand, if the vertical resolution
856is high (1~m or less) and a turbulent closure model is used, the turbulent
857Ekman layer will be represented explicitly by the model. However, the
858logarithmic layer is never represented in current primitive equation model
859applications: it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
860Two choices are available in \NEMO: a linear and a quadratic bottom friction.
861Note that in both cases, the rotation between the interior velocity and the
862bottom friction is neglected in the present release of \NEMO.
863
864In the code, the bottom friction is imposed by adding the trend due to the bottom
865friction to the general momentum trend in \mdl{dynbfr}. For the time-split surface
866pressure gradient algorithm, the momentum trend due to the barotropic component
867needs to be handled separately. For this purpose it is convenient to compute and
868store coefficients which can be simply combined with bottom velocities and geometric
869values to provide the momentum trend due to bottom friction.
870These coefficients are computed in \mdl{zdfbfr} and generally take the form
871$c_b^{\textbf U}$ where:
872\begin{equation} \label{Eq_zdfbfr_bdef}
873\frac{\partial {\textbf U_h}}{\partial t} =
874  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
875\end{equation}
876where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
877
878% -------------------------------------------------------------------------------------------------------------
879%       Linear Bottom Friction
880% -------------------------------------------------------------------------------------------------------------
881\subsection{Linear Bottom Friction (\np{nn\_botfr} = 0 or 1) }
882\label{ZDF_bfr_linear}
883
884The linear bottom friction parameterisation (including the special case
885of a free-slip condition) assumes that the bottom friction
886is proportional to the interior velocity (i.e. the velocity of the last
887model level):
888\begin{equation} \label{Eq_zdfbfr_linear}
889{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
890\end{equation}
891where $r$ is a friction coefficient expressed in ms$^{-1}$.
892This coefficient is generally estimated by setting a typical decay time
893$\tau$ in the deep ocean,
894and setting $r = H / \tau$, where $H$ is the ocean depth. Commonly accepted
895values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}.
896A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used
897in quasi-geostrophic models. One may consider the linear friction as an
898approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982},
899Eq. 9.6.6). For example, with a drag coefficient $C_D = 0.002$, a typical speed
900of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, and assuming an ocean depth
901$H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
902This is the default value used in \NEMO. It corresponds to a decay time scale
903of 115~days. It can be changed by specifying \np{rn\_bfric1} (namelist parameter).
904
905For the linear friction case the coefficients defined in the general
906expression \eqref{Eq_zdfbfr_bdef} are:
907\begin{equation} \label{Eq_zdfbfr_linbfr_b}
908\begin{split}
909 c_b^u &= - r\\
910 c_b^v &= - r\\
911\end{split}
912\end{equation}
913When \np{nn\_botfr}=1, the value of $r$ used is \np{rn\_bfric1}.
914Setting \np{nn\_botfr}=0 is equivalent to setting $r=0$ and leads to a free-slip
915bottom boundary condition. These values are assigned in \mdl{zdfbfr}.
916From v3.2 onwards there is support for local enhancement of these values
917via an externally defined 2D mask array (\np{ln\_bfr2d}=true) given
918in the \ifile{bfr\_coef} input NetCDF file. The mask values should vary from 0 to 1.
919Locations with a non-zero mask value will have the friction coefficient increased
920by $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfric1}.
921
922% -------------------------------------------------------------------------------------------------------------
923%       Non-Linear Bottom Friction
924% -------------------------------------------------------------------------------------------------------------
925\subsection{Non-Linear Bottom Friction (\np{nn\_botfr} = 2)}
926\label{ZDF_bfr_nonlinear}
927
928The non-linear bottom friction parameterisation assumes that the bottom
929friction is quadratic:
930\begin{equation} \label{Eq_zdfbfr_nonlinear}
931{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
932}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
933\end{equation}
934where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy
935due to tides, internal waves breaking and other short time scale currents.
936A typical value of the drag coefficient is $C_D = 10^{-3} $. As an example,
937the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and
938$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} 
939uses $C_D = 1.4\;10^{-3}$ and $e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
940The CME choices have been set as default values (\np{rn\_bfric2} and \np{rn\_bfeb2} 
941namelist parameters).
942
943As for the linear case, the bottom friction is imposed in the code by
944adding the trend due to the bottom friction to the general momentum trend
945in \mdl{dynbfr}.
946For the non-linear friction case the terms
947computed in \mdl{zdfbfr}  are:
948\begin{equation} \label{Eq_zdfbfr_nonlinbfr}
949\begin{split}
950 c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\
951 c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\
952\end{split}
953\end{equation}
954
955The coefficients that control the strength of the non-linear bottom friction are
956initialised as namelist parameters: $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}.
957Note for applications which treat tides explicitly a low or even zero value of
958\np{rn\_bfeb2} is recommended. From v3.2 onwards a local enhancement of $C_D$ 
959is possible via an externally defined 2D mask array (\np{ln\_bfr2d}=true).
960See previous section for details.
961
962% -------------------------------------------------------------------------------------------------------------
963%       Bottom Friction stability
964% -------------------------------------------------------------------------------------------------------------
965\subsection{Bottom Friction stability considerations}
966\label{ZDF_bfr_stability}
967
968Some care needs to exercised over the choice of parameters to ensure that the
969implementation of bottom friction does not induce numerical instability. For
970the purposes of stability analysis, an approximation to \eqref{Eq_zdfbfr_flux2}
971is:
972\begin{equation} \label{Eqn_bfrstab}
973\begin{split}
974 \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
975               &= -\frac{ru}{e_{3u}}\;2\rdt\\
976\end{split}
977\end{equation}
978\noindent where linear bottom friction and a leapfrog timestep have been assumed.
979To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have:
980\begin{equation}
981 |\Delta u| < \;|u|
982\end{equation}
983\noindent which, using \eqref{Eqn_bfrstab}, gives:
984\begin{equation}
985r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
986\end{equation}
987This same inequality can also be derived in the non-linear bottom friction case
988if a velocity of 1 m.s$^{-1}$ is assumed. Alternatively, this criterion can be
989rearranged to suggest a minimum bottom box thickness to ensure stability:
990\begin{equation}
991e_{3u} > 2\;r\;\rdt
992\end{equation}
993\noindent which it may be necessary to impose if partial steps are being used.
994For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then
995$e_{3u}$ should be greater than 3.6 m. For most applications, with physically
996sensible parameters these restrictions should not be of concern. But
997caution may be necessary if attempts are made to locally enhance the bottom
998friction parameters.
999To ensure stability limits are imposed on the bottom friction coefficients both during
1000initialisation and at each time step. Checks at initialisation are made in \mdl{zdfbfr} 
1001(assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1002The number of breaches of the stability criterion are reported as well as the minimum
1003and maximum values that have been set. The criterion is also checked at each time step,
1004using the actual velocity, in \mdl{dynbfr}. Values of the bottom friction coefficient are
1005reduced as necessary to ensure stability; these changes are not reported.
1006
1007Limits on the bottom friction coefficient are not imposed if the user has elected to
1008handle the bottom friction implicitly (see \S\ref{ZDF_bfr_imp}). The number of potential
1009breaches of the explicit stability criterion are still reported for information purposes.
1010
1011% -------------------------------------------------------------------------------------------------------------
1012%       Implicit Bottom Friction
1013% -------------------------------------------------------------------------------------------------------------
1014\subsection{Implicit Bottom Friction (\np{ln\_bfrimp}$=$\textit{T})}
1015\label{ZDF_bfr_imp}
1016
1017An optional implicit form of bottom friction has been implemented to improve
1018model stability. We recommend this option for shelf sea and coastal ocean applications, especially
1019for split-explicit time splitting. This option can be invoked by setting \np{ln\_bfrimp} 
1020to \textit{true} in the \textit{nambfr} namelist. This option requires \np{ln\_zdfexp} to be \textit{false} 
1021in the \textit{namzdf} namelist.
1022
1023This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, the
1024bottom boundary condition is implemented implicitly.
1025
1026\begin{equation} \label{Eq_dynzdf_bfr}
1027\left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{mbk}
1028    = \binom{c_{b}^{u}u^{n+1}_{mbk}}{c_{b}^{v}v^{n+1}_{mbk}}
1029\end{equation}
1030
1031where $mbk$ is the layer number of the bottom wet layer. superscript $n+1$ means the velocity used in the
1032friction formula is to be calculated, so, it is implicit.
1033
1034If split-explicit time splitting is used, care must be taken to avoid the double counting of
1035the bottom friction in the 2-D barotropic momentum equations. As NEMO only updates the barotropic
1036pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, we need to remove
1037the bottom friction induced by these two terms which has been included in the 3-D momentum trend
1038and update it with the latest value. On the other hand, the bottom friction contributed by the
1039other terms (e.g. the advection term, viscosity term) has been included in the 3-D momentum equations
1040and should not be added in the 2-D barotropic mode.
1041
1042The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the
1043following:
1044
1045\begin{equation} \label{Eq_dynspg_ts_bfr1}
1046\frac{\textbf{U}_{med}-\textbf{U}^{m-1}}{2\Delta t}=-g\nabla\eta-f\textbf{k}\times\textbf{U}^{m}+c_{b}
1047\left(\textbf{U}_{med}-\textbf{U}^{m-1}\right)
1048\end{equation}
1049\begin{equation} \label{Eq_dynspg_ts_bfr2}
1050\frac{\textbf{U}^{m+1}-\textbf{U}_{med}}{2\Delta t}=\textbf{T}+
1051\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{U}^{'}\right)-
10522\Delta t_{bc}c_{b}\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{u}_{b}\right)
1053\end{equation}
1054
1055where $\textbf{T}$ is the vertical integrated 3-D momentum trend. We assume the leap-frog time-stepping
1056is used here. $\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step.
1057 $c_{b}$ is the friction coefficient. $\eta$ is the sea surface level calculated in the barotropic loops
1058while $\eta^{'}$ is the sea surface level used in the 3-D baroclinic mode. $\textbf{u}_{b}$ is the bottom
1059layer horizontal velocity.
1060
1061
1062
1063
1064% -------------------------------------------------------------------------------------------------------------
1065%       Bottom Friction with split-explicit time splitting
1066% -------------------------------------------------------------------------------------------------------------
1067\subsection{Bottom Friction with split-explicit time splitting (\np{ln\_bfrimp}$=$\textit{F})}
1068\label{ZDF_bfr_ts}
1069
1070When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, the
1071bottom velocity at the before time step is used. This velocity includes both the
1072baroclinic and barotropic components which is appropriate when using either the
1073explicit or filtered surface pressure gradient algorithms (\key{dynspg\_exp} or
1074{\key{dynspg\_flt}). Extra attention is required, however, when using
1075split-explicit time stepping (\key{dynspg\_ts}). In this case the free surface
1076equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, while the three
1077dimensional prognostic variables are solved with the longer time step
1078of \np{rn\_rdt} seconds. The trend in the barotropic momentum due to bottom
1079friction appropriate to this method is that given by the selected parameterisation
1080($i.e.$ linear or non-linear bottom friction) computed with the evolving velocities
1081at each barotropic timestep.
1082
1083In the case of non-linear bottom friction, we have elected to partially linearise
1084the problem by keeping the coefficients fixed throughout the barotropic
1085time-stepping to those computed in \mdl{zdfbfr} using the now timestep.
1086This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to:
1087
1088\begin{enumerate}
1089\item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before
1090barotropic velocity to the bottom friction component of the vertically
1091integrated momentum trend. Note the same stability check that is carried out
1092on the bottom friction coefficient in \mdl{dynbfr} has to be applied here to
1093ensure that the trend removed matches that which was added in \mdl{dynbfr}.
1094\item At each barotropic step, compute the contribution of the current barotropic
1095velocity to the trend due to bottom friction. Add this contribution to the
1096vertically integrated momentum trend. This contribution is handled implicitly which
1097eliminates the need to impose a stability criteria on the values of the bottom friction
1098coefficient within the barotropic loop.
1099\end{enumerate}
1100
1101Note that the use of an implicit formulation within the barotropic loop
1102for the bottom friction trend means that any limiting of the bottom friction coefficient
1103in \mdl{dynbfr} does not adversely affect the solution when using split-explicit time
1104splitting. This is because the major contribution to bottom friction is likely to come from
1105the barotropic component which uses the unrestricted value of the coefficient. However, if the
1106limiting is thought to be having a major effect (a more likely prospect in coastal and shelf seas
1107applications) then the fully implicit form of the bottom friction should be used (see \S\ref{ZDF_bfr_imp} )
1108which can be selected by setting \np{ln\_bfrimp} $=$ \textit{true}.
1109
1110Otherwise, the implicit formulation takes the form:
1111\begin{equation} \label{Eq_zdfbfr_implicitts}
1112 \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ] 
1113\end{equation}
1114where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),
1115$c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and $RHS$ represents
1116all the components to the vertically integrated momentum trend except for that due to bottom friction.
1117
1118
1119
1120
1121% ================================================================
1122% Tidal Mixing
1123% ================================================================
1124\section{Tidal Mixing (\key{zdftmx})}
1125\label{ZDF_tmx}
1126
1127%--------------------------------------------namzdf_tmx--------------------------------------------------
1128\namdisplay{namzdf_tmx}
1129%--------------------------------------------------------------------------------------------------------------
1130
1131
1132% -------------------------------------------------------------------------------------------------------------
1133%        Bottom intensified tidal mixing
1134% -------------------------------------------------------------------------------------------------------------
1135\subsection{Bottom intensified tidal mixing}
1136\label{ZDF_tmx_bottom}
1137
1138The parameterization of tidal mixing follows the general formulation for
1139the vertical eddy diffusivity proposed by \citet{St_Laurent_al_GRL02} and
1140first introduced in an OGCM by \citep{Simmons_al_OM04}.
1141In this formulation an additional vertical diffusivity resulting from internal tide breaking,
1142$A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, the energy transfer from barotropic
1143tides to baroclinic tides :
1144\begin{equation} \label{Eq_Ktides}
1145A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 }
1146\end{equation}
1147where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency
1148(see \S\ref{TRA_bn2}), $\rho$ the density, $q$ the tidal dissipation efficiency,
1149and $F(z)$ the vertical structure function.
1150
1151The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter)
1152and is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).
1153The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter)
1154represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally,
1155with the remaining $1-q$ radiating away as low mode internal waves and
1156contributing to the background internal wave field. A value of $q=1/3$ is typically used 
1157\citet{St_Laurent_al_GRL02}.
1158The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical.
1159It is implemented as a simple exponential decaying upward away from the bottom,
1160with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},
1161\begin{equation} \label{Eq_Fz}
1162F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) }
1163\end{equation}
1164and is normalized so that vertical integral over the water column is unity.
1165
1166The associated vertical viscosity is calculated from the vertical
1167diffusivity assuming a Prandtl number of 1, $i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$.
1168In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity
1169is capped at $300\,cm^2/s$ and impose a lower limit on $N^2$ of \np{rn\_n2min} 
1170usually set to $10^{-8} s^{-2}$. These bounds are usually rarely encountered.
1171
1172The internal wave energy map, $E(x, y)$ in \eqref{Eq_Ktides}, is derived
1173from a barotropic model of the tides utilizing a parameterization of the
1174conversion of barotropic tidal energy into internal waves.
1175The essential goal of the parameterization is to represent the momentum
1176exchange between the barotropic tides and the unrepresented internal waves
1177induced by the tidal flow over rough topography in a stratified ocean.
1178In the current version of \NEMO, the map is built from the output of
1179the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}.
1180This model provides the dissipation associated with internal wave energy for the M2 and K1
1181tides component (Fig.~\ref{Fig_ZDF_M2_K1_tmx}). The S2 dissipation is simply approximated
1182as being $1/4$ of the M2 one. The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$.
1183Its global mean value is $1.1$ TW, in agreement with independent estimates
1184\citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.
1185
1186%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1187\begin{figure}[!t]   \begin{center}
1188\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_ZDF_M2_K1_tmx.pdf}
1189\caption{  \label{Fig_ZDF_M2_K1_tmx} 
1190(a) M2 and (b) K1 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). }
1191\end{center}   \end{figure}
1192%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1193 
1194% -------------------------------------------------------------------------------------------------------------
1195%        Indonesian area specific treatment
1196% -------------------------------------------------------------------------------------------------------------
1197\subsection{Indonesian area specific treatment (\np{ln\_zdftmx\_itf})}
1198\label{ZDF_tmx_itf}
1199
1200When the Indonesian Through Flow (ITF) area is included in the model domain,
1201a specific treatment of tidal induced mixing in this area can be used.
1202It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide
1203an input NetCDF file, \ifile{mask\_itf}, which contains a mask array defining the ITF area
1204where the specific treatment is applied.
1205
1206When \np{ln\_tmx\_itf}=true, the two key parameters $q$ and $F(z)$ are adjusted following
1207the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}:
1208
1209First, the Indonesian archipelago is a complex geographic region
1210with a series of large, deep, semi-enclosed basins connected via
1211numerous narrow straits. Once generated, internal tides remain
1212confined within this semi-enclosed area and hardly radiate away.
1213Therefore all the internal tides energy is consumed within this area.
1214So it is assumed that $q = 1$, $i.e.$ all the energy generated is available for mixing.
1215Note that for test purposed, the ITF tidal dissipation efficiency is a
1216namelist parameter (\np{rn\_tfe\_itf}). A value of $1$ or close to is
1217this recommended for this parameter.
1218
1219Second, the vertical structure function, $F(z)$, is no more associated
1220with a bottom intensification of the mixing, but with a maximum of
1221energy available within the thermocline. \citet{Koch-Larrouy_al_GRL07} 
1222have suggested that the vertical distribution of the energy dissipation
1223proportional to $N^2$ below the core of the thermocline and to $N$ above.
1224The resulting $F(z)$ is:
1225\begin{equation} \label{Eq_Fz_itf}
1226F(i,j,k) \sim     \left\{ \begin{aligned}
1227\frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\
1228\frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$}
1229                      \end{aligned} \right.
1230\end{equation}
1231
1232Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,
1233which agrees with the independent estimates inferred from observations.
1234Introduced in a regional OGCM, the parameterization improves the water mass
1235characteristics in the different Indonesian seas, suggesting that the horizontal
1236and vertical distributions of the mixing are adequately prescribed
1237\citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}.
1238Note also that such a parameterisation has a significant impact on the behaviour
1239of global coupled GCMs \citep{Koch-Larrouy_al_CD10}.
1240
1241
1242% ================================================================
Note: See TracBrowser for help on using the repository browser.