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/2014/dev_r4642_WavesWG/DOC/TexFiles/Chapters – NEMO

source: branches/2014/dev_r4642_WavesWG/DOC/TexFiles/Chapters/Chap_ZDF.tex @ 4644

Last change on this file since 4644 was 4644, checked in by acc, 10 years ago

Branch 2014/dev_r4642_WavesWG #1323. Import of surface wave components from the 2013/dev_ECMWF_waves branch + a few compatability changes and some mislaid documentation

File size: 76.1 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
55Options are defined through the  \ngn{namzdf} namelist variables.
56When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients
57are set to constant values over the whole ocean. This is the crudest way to define
58the vertical ocean physics. It is recommended that this option is only used in
59process studies, not in basin scale simulations. Typical values used in this case are:
60\begin{align*} 
61A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\
62A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
63\end{align*}
64
65These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.
66In all cases, do not use values smaller that those associated with the molecular
67viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum,
68$\sim10^{-7}~m^2.s^{-1}$ for temperature and $\sim10^{-9}~m^2.s^{-1}$ for salinity.
69
70
71% -------------------------------------------------------------------------------------------------------------
72%        Richardson Number Dependent
73% -------------------------------------------------------------------------------------------------------------
74\subsection{Richardson Number Dependent (\key{zdfric})}
75\label{ZDF_ric}
76
77%--------------------------------------------namric---------------------------------------------------------
78\namdisplay{namzdf_ric}
79%--------------------------------------------------------------------------------------------------------------
80
81When \key{zdfric} is defined, a local Richardson number dependent formulation
82for the vertical momentum and tracer eddy coefficients is set through the  \ngn{namzdf\_ric} 
83namelist variables.The vertical mixing
84coefficients are diagnosed from the large scale variables computed by the model.
85\textit{In situ} measurements have been used to link vertical turbulent activity to
86large scale ocean structures. The hypothesis of a mixing mainly maintained by the
87growth of Kelvin-Helmholtz like instabilities leads to a dependency between the
88vertical eddy coefficients and the local Richardson number ($i.e.$ the
89ratio of stratification to vertical shear). Following \citet{Pacanowski_Philander_JPO81}, the following
90formulation has been implemented:
91\begin{equation} \label{Eq_zdfric}
92   \left\{      \begin{aligned}
93         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
94         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
95   \end{aligned}    \right.
96\end{equation}
97where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson
98number, $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
99$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the
100constant case (see \S\ref{ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ 
101is the maximum value that can be reached by the coefficient when $Ri\leq 0$,
102$a=5$ and $n=2$. The last three values can be modified by setting the
103\np{rn\_avmri}, \np{rn\_alp} and \np{nn\_ric} namelist parameters, respectively.
104
105A simple mixing-layer model to transfer and dissipate the atmospheric
106 forcings (wind-stress and buoyancy fluxes) can be activated setting
107the \np{ln\_mldw} =.true. in the namelist.
108
109In this case, the local depth of turbulent wind-mixing or "Ekman depth"
110 $h_{e}(x,y,t)$ is evaluated and the vertical eddy coefficients prescribed within this layer.
111
112This depth is assumed proportional to the "depth of frictional influence" that is limited by rotation:
113\begin{equation}
114         h_{e} = Ek \frac {u^{*}} {f_{0}}    \\
115\end{equation}
116where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis
117parameter.
118
119In this similarity height relationship, the turbulent friction velocity:
120\begin{equation}
121         u^{*} = \sqrt \frac {|\tau|} {\rho_o}     \\
122\end{equation}
123
124is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$.
125The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}.
126Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to
127the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{Lermusiaux2001}.
128
129% -------------------------------------------------------------------------------------------------------------
130%        TKE Turbulent Closure Scheme
131% -------------------------------------------------------------------------------------------------------------
132\subsection{TKE Turbulent Closure Scheme (\key{zdftke})}
133\label{ZDF_tke}
134
135%--------------------------------------------namzdf_tke--------------------------------------------------
136\namdisplay{namzdf_tke}
137%--------------------------------------------------------------------------------------------------------------
138
139The vertical eddy viscosity and diffusivity coefficients are computed from a TKE
140turbulent closure model based on a prognostic equation for $\bar{e}$, the turbulent
141kinetic energy, and a closure assumption for the turbulent length scales. This
142turbulent closure model has been developed by \citet{Bougeault1989} in the
143atmospheric case, adapted by \citet{Gaspar1990} for the oceanic case, and
144embedded in OPA, the ancestor of NEMO, by \citet{Blanke1993} for equatorial Atlantic
145simulations. Since then, significant modifications have been introduced by
146\citet{Madec1998} in both the implementation and the formulation of the mixing
147length scale. The time evolution of $\bar{e}$ is the result of the production of
148$\bar{e}$ through vertical shear, its destruction through stratification, its vertical
149diffusion, and its dissipation of \citet{Kolmogorov1942} type:
150\begin{equation} \label{Eq_zdftke_e}
151\frac{\partial \bar{e}}{\partial t} =
152\frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
153                    +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
154-K_\rho\,N^2
155+\frac{1}{e_3} \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
156            \;\frac{\partial \bar{e}}{\partial k}} \right]
157- c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
158\end{equation}
159\begin{equation} \label{Eq_zdftke_kz}
160   \begin{split}
161         K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }     \\
162         K_\rho &= A^{vm} / P_{rt}
163   \end{split}
164\end{equation}
165where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
166$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,
167$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity
168and diffusivity coefficients. The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ 
169$\approx 0.7$ are designed to deal with vertical mixing at any depth \citep{Gaspar1990}.
170They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}.
171$P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function
172of the local Richardson number, $R_i$:
173\begin{align*} \label{Eq_prt}
174P_{rt} = \begin{cases}
175                    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}  \\
176                    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}  \\
177                    \ \ 10 &      \text{if $\ 2 \leq R_i$} 
178            \end{cases}
179\end{align*}
180Options are defined through the  \ngn{namzdfy\_tke} namelist variables.
181The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable.
182
183At the sea surface, the value of $\bar{e}$ is prescribed from the wind
184stress field as $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} 
185namelist parameter. The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}),
186however a much larger value can be used when taking into account the
187surface wave breaking (see below Eq. \eqref{ZDF_Esbc}).
188The bottom value of TKE is assumed to be equal to the value of the level just above.
189The time integration of the $\bar{e}$ equation may formally lead to negative values
190because the numerical scheme does not ensure its positivity. To overcome this
191problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} 
192namelist parameter). Following \citet{Gaspar1990}, the cut-off value is set
193to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. This allows the subsequent formulations
194to match that of \citet{Gargett1984} for the diffusion in the thermocline and
195deep ocean :  $K_\rho = 10^{-3} / N$.
196In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical
197instabilities associated with too weak vertical diffusion. They must be
198specified at least larger than the molecular values, and are set through
199\np{rn\_avm0} and \np{rn\_avt0} (\ngn{namzdf} namelist, see \S\ref{ZDF_cst}).
200
201\subsubsection{Turbulent length scale}
202For computational efficiency, the original formulation of the turbulent length
203scales proposed by \citet{Gaspar1990} has been simplified. Four formulations
204are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist
205parameter. The first two are based on the following first order approximation
206\citep{Blanke1993}:
207\begin{equation} \label{Eq_tke_mxl0_1}
208l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N
209\end{equation}
210which is valid in a stable stratified region with constant values of the Brunt-
211Vais\"{a}l\"{a} frequency. The resulting length scale is bounded by the distance
212to the surface or to the bottom (\np{nn\_mxl} = 0) or by the local vertical scale factor
213(\np{nn\_mxl} = 1). \citet{Blanke1993} notice that this simplification has two major
214drawbacks: it makes no sense for locally unstable stratification and the
215computation no longer uses all the information contained in the vertical density
216profile. To overcome these drawbacks, \citet{Madec1998} introduces the
217\np{nn\_mxl} = 2 or 3 cases, which add an extra assumption concerning the vertical
218gradient of the computed length scale. So, the length scales are first evaluated
219as in \eqref{Eq_tke_mxl0_1} and then bounded such that:
220\begin{equation} \label{Eq_tke_mxl_constraint}
221\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
222\qquad \text{with }\  l =  l_k = l_\epsilon
223\end{equation}
224\eqref{Eq_tke_mxl_constraint} means that the vertical variations of the length
225scale cannot be larger than the variations of depth. It provides a better
226approximation of the \citet{Gaspar1990} formulation while being much less
227time consuming. In particular, it allows the length scale to be limited not only
228by the distance to the surface or to the ocean bottom but also by the distance
229to a strongly stratified portion of the water column such as the thermocline
230(Fig.~\ref{Fig_mixing_length}). In order to impose the \eqref{Eq_tke_mxl_constraint} 
231constraint, we introduce two additional length scales: $l_{up}$ and $l_{dwn}$,
232the upward and downward length scales, and evaluate the dissipation and
233mixing length scales as (and note that here we use numerical indexing):
234%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
235\begin{figure}[!t] \begin{center}
236\includegraphics[width=1.00\textwidth]{./TexFiles/Figures/Fig_mixing_length.pdf}
237\caption{ \label{Fig_mixing_length} 
238Illustration of the mixing length computation. }
239\end{center} 
240\end{figure}
241%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
242\begin{equation} \label{Eq_tke_mxl2}
243\begin{aligned}
244 l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right)
245    \quad &\text{ from $k=1$ to $jpk$ }\ \\
246 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)}  \right)   
247    \quad &\text{ from $k=jpk$ to $1$ }\ \\
248\end{aligned}
249\end{equation}
250where $l^{(k)}$ is computed using \eqref{Eq_tke_mxl0_1},
251$i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$.
252
253In the \np{nn\_mxl}~=~2 case, the dissipation and mixing length scales take the same
254value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the
255\np{nn\_mxl}~=~3 case, the dissipation and mixing turbulent length scales are give
256as in \citet{Gaspar1990}:
257\begin{equation} \label{Eq_tke_mxl_gaspar}
258\begin{aligned}
259& l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }    \\
260& l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
261\end{aligned}
262\end{equation}
263
264At the ocean surface, a non zero length scale is set through the
265\np{rn\_lmin0} namelist parameter. Usually the surface scale is given
266by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant
267and $z_o$ the roughness parameter of the surface. Assuming $z_o=0.1$~m
268\citep{Craig_Banner_JPO94} leads to a 0.04~m, the default value of
269\np{rn\_lsurf}. In the ocean interior a minimum length scale is set to
270recover the molecular viscosity when $\bar{e}$ reach its minimum value
271($1\times 10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$).
272
273
274\subsubsection{Surface wave breaking parameterization}
275%-----------------------------------------------------------------------%
276Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified
277to include the effect of surface wave breaking energetics. This results in a reduction of summertime
278surface temperature when the mixed layer is relatively shallow. The \citet{Mellor_Blumberg_JPO04} 
279modifications acts on surface length scale and TKE values and air-sea drag coefficient.
280The latter concerns the bulk formulea and is not discussed here.
281
282Following \citet{Craig_Banner_JPO94}, the boundary condition on surface TKE value is :
283\begin{equation}  \label{ZDF_Esbc}
284\bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
285\end{equation}
286where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of
287proportionality which depends on the ''wave age'', ranging from 57 for mature
288waves to 146 for younger waves \citep{Mellor_Blumberg_JPO04}.  The boundary
289condition on the turbulent length scale follows Charnock's relation:
290\begin{equation} \label{ZDF_Lsbc}
291l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
292\end{equation}
293where $\kappa=0.40$ is the von Karman constant, and $\beta$ is Charnock's
294constant.  \citet{Mellor_Blumberg_JPO04} suggest $\beta = 2\times10^{5}$ the value
295chosen by \citet{Stacey_JPO99} citing observation evidence, and $\alpha_{CB}
296= 100$ the Craig and Banner's value.  As the surface boundary condition
297on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with
298$e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}~=~67.83
299corresponds to $\alpha_{CB} = 100$. further setting  \np{ln\_lsurf} to true
300applies \eqref{ZDF_Lsbc} as surface boundary condition on length scale, with
301$\beta$ hard coded to Stacey's value.  Note that a minimal threshold
302of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on
303surface $\bar{e}$ value.
304
305\subsubsection{Surface wave breaking flux from a wave model \np{ln\_wavetke}}
306%-----------------------------------------------------------------------%
307The constant of proportionality $\alpha_{CB}$ in Eq~\eqref{ZDF_Esbc} relates
308the water-side friction velocity $w_*$ to the turbulent energy flux as follows,
309\begin{equation}
310   \Phi_\mathrm{oc} = \rho_\mathrm{w} \alpha_\mathrm{CB} w_*^3.
311   \label{Eq_ZDF_alpha}
312\end{equation}
313The default option in NEMO \eqref{ZDF_Esbc} is to assume $\alpha_{CB}
314= 100$ as explained in the previous section.
315However, the energy flux can be computed from the dissipation source term
316of a wave model \citep{Janssen_AH04,Janssen_JGR12,Janssen_ECMWF13},
317\begin{equation}
318   \Phi_\mathrm{oc} = \Phi_\mathrm{in} - \rho_\mathrm{w}g \int_0^{2\pi} 
319                      \int_0^{\infty} (S_\mathrm{in} + S_\mathrm{ds})\, 
320                        d\omega d\theta.
321   \label{Eq_ZDF_phioc}
322\end{equation}
323Assuming high-frequency equilibrium and ignoring the direct turbulent energy
324flux from the atmosphere to the ocean we get
325\begin{equation}
326   \Phi_\mathrm{oc} = -\rho_\mathrm{w}g \int_0^{2\pi} 
327                      \int_0^{\omega_\mathrm{c}} S_\mathrm{ds}\, 
328                        d\omega d\theta = -\rho_\mathrm{a} m u_*^3.
329   \label{Eq_ZDF_m}
330\end{equation}
331Here, $m \approx -\sqrt{\rho_\mathrm{a}/\rho_\mathrm{w}} \alpha_\mathrm{CB}$
332is the energy flux \emph{from} the waves (thus always negative) normalized by
333the air friction velocity $u_*$. It
334is archived as PHIOC (GRIB parameter 212, table 140) in ERA-Interim and also
335operationally by ECMWF.  The namelist parameter \np{ln\_wavetke} controls
336the wave TKE flux.  We assume that the flux has been converted to physical
337units following \eqref{Eq_ZDF_m} before ingested by NEMO.
338
339NEMO computes the upper boundary condition following
340\citet{Mellor_Blumberg_JPO04}, see \eqref{ZDF_Esbc}.  Since $e$ varies
341rapidly with depth, we want to weight the surface value $\overline{e}_o$
342by the thickness of the uppermost level to attain a value representative
343for the turbulence level of the uppermost level,
344\begin{equation}
345   \overline{e}_1 = \frac{\overline{e}_o}{L} \int_{-L}^{0} e(z) \,dz.
346   \label{Eq_ZDF_eavg}
347\end{equation}
348Here $L = \Delta z_1/2$ is the depth of the $T$-point of the uppermost level.
349This adjustment is crucial with model configurations with a thick uppermost
350level, e.g. ORCA1L42.  If we assume, as \citet{Mellor_Blumberg_JPO04} do, that in the
351wave-affected layer the roughness length can be set to a constant which we
352choose to be $z_\mathrm{w} = 0.5H_\mathrm{s}$ and that in this near-surface
353region diffusion balances dissipation, the TKE equation attains the simple
354exponential solution \citep{Mellor_Blumberg_JPO04}
355\begin{equation}
356  e(z) = \overline{e}_o \exp(2\lambda z/3).
357   \label{Eq_ZDF_phimb}
358\end{equation}
359Here, the length scale $\lambda^{-1}$ is sea-state dependent, see Eq (10) by
360\citet{Mellor_Blumberg_JPO04},
361\begin{equation}
362   \lambda = [3/(S_q B_1 \kappa^2)]^{1/2}z_\mathrm{w}^{-1} \approx
363   \frac{2.38}{z_\mathrm{w}}.
364\end{equation}
365We have assumed $S_q=0.2$ and $B=16.6$ \citep{Mellor_Yamada_1982}, as
366used in NEMO.  For a wave height of 2.5 m, which is close to the global
367mean, $\lambda^{-1} \approx 0.5\, \mathrm{m}$.  Integrating \eqref{Eq_ZDF_phimb}
368is straightforward, and the average TKE in \eqref{Eq_ZDF_eavg} becomes
369\begin{equation}
370  \overline{e}_1 = \overline{e}_o \frac{3}{2\lambda L} \left[1 -
371  \exp(-2\lambda L/3)\right].
372\end{equation}
373The wave model energy flux is controlled by \np{ln\_wavetke} in namelist
374\ngn{namsbc}.
375
376
377\subsubsection{Langmuir cells}
378%--------------------------------------%
379Langmuir circulations (LC) can be described as ordered large-scale vertical motions
380in the surface layer of the oceans. Although LC have nothing to do with convection,
381the circulation pattern is rather similar to so-called convective rolls in the atmospheric
382boundary layer. The detailed physics behind LC is described in, for example,
383\citet{Craik_Leibovich_JFM76}. The prevailing explanation is that LC arise from
384a nonlinear interaction between the Stokes drift and wind drift currents.
385
386Here we introduced in the TKE turbulent closure the simple parameterization of
387Langmuir circulations proposed by \citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure.
388The parameterization, tuned against large-eddy simulation, includes the whole effect
389of LC in an extra source terms of TKE, $P_{LC}$.
390The presence of $P_{LC}$ in \eqref{Eq_zdftke_e}, the TKE equation, is controlled
391by setting \np{ln\_lc} to \textit{true} in the namtke namelist.
392 
393By making an analogy with the characteristic convective velocity scale
394($e.g.$, \citet{D'Alessio_al_JPO98}), $P_{LC}$ is assumed to be:
395\begin{equation}
396P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
397\end{equation}
398where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
399With no information about the wave field, $w_{LC}$ is assumed to be proportional to
400the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
401\footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity
402may be expressed as $u_s =  0.016 \,|U_{10m}|$. Assuming an air density of
403$\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of $1.5~10^{-3}$ give the expression
404used of $u_s$ as a function of the module of surface stress}.
405For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as
406at a finite depth $H_{LC}$ (which is often close to the mixed layer depth), and simply
407varies as a sine function in between (a first-order profile for the Langmuir cell structures).
408The resulting expression for $w_{LC}$ is :
409\begin{equation}
410w_{LC}  = \begin{cases}
411                   c_{LC} \,u_s \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
412                   0                             &      \text{otherwise} 
413                 \end{cases}
414\end{equation}
415where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise
416to fit LES data. The chosen value yields maximum vertical velocities $w_{LC}$ of the order
417of a few centimeters per second. The value of $c_{LC}$ is set through the \np{rn\_lc} 
418namelist parameter, having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.
419
420The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
421$H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift
422can reach on its own by converting its kinetic energy to potential energy, according to
423\begin{equation}
424- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} u_s^2
425\end{equation}
426
427
428\subsubsection{Mixing just below the mixed layer}
429%--------------------------------------------------------------%
430
431To be add here a description of "penetration of TKE" and the associated namelist parameters
432 \np{nn\_etau}, \np{rn\_efr} and \np{nn\_htau}.
433
434% from Burchard et al OM 2008 :
435% 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).
436
437
438
439% -------------------------------------------------------------------------------------------------------------
440%        TKE discretization considerations
441% -------------------------------------------------------------------------------------------------------------
442\subsection{TKE discretization considerations (\key{zdftke})}
443\label{ZDF_tke_ene}
444
445%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
446\begin{figure}[!t]   \begin{center}
447\includegraphics[width=1.00\textwidth]{./TexFiles/Figures/Fig_ZDF_TKE_time_scheme.pdf}
448\caption{ \label{Fig_TKE_time_scheme} 
449Illustration of the TKE time integration and its links to the momentum and tracer time integration. }
450\end{center} 
451\end{figure}
452%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
453
454The production of turbulence by vertical shear (the first term of the right hand side
455of \eqref{Eq_zdftke_e}) should balance the loss of kinetic energy associated with
456the vertical momentum diffusion (first line in \eqref{Eq_PE_zdf}). To do so a special care
457have to be taken for both the time and space discretization of the TKE equation
458\citep{Burchard_OM02,Marsaleix_al_OM08}.
459
460Let us first address the time stepping issue. Fig.~\ref{Fig_TKE_time_scheme} shows
461how the two-level Leap-Frog time stepping of the momentum and tracer equations interplays
462with the one-level forward time stepping of TKE equation. With this framework, the total loss
463of kinetic energy (in 1D for the demonstration) due to the vertical momentum diffusion is
464obtained by multiplying this quantity by $u^t$ and summing the result vertically:   
465\begin{equation} \label{Eq_energ1}
466\begin{split}
467\int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
468&= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}         
469 - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
470\end{split}
471\end{equation}
472Here, the vertical diffusion of momentum is discretized backward in time
473with a coefficient, $K_m$, known at time $t$ (Fig.~\ref{Fig_TKE_time_scheme}),
474as it is required when using the TKE scheme (see \S\ref{STP_forward_imp}).
475The first term of the right hand side of \eqref{Eq_energ1} represents the kinetic energy
476transfer at the surface (atmospheric forcing) and at the bottom (friction effect).
477The second term is always negative. It is the dissipation rate of kinetic energy,
478and thus minus the shear production rate of $\bar{e}$. \eqref{Eq_energ1} 
479implies that, to be energetically consistent, the production rate of $\bar{e}$ 
480used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
481${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ (and not by the more straightforward
482$K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
483
484A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
485(second term of the right hand side of \eqref{Eq_zdftke_e}). This term
486must balance the input of potential energy resulting from vertical mixing.
487The rate of change of potential energy (in 1D for the demonstration) due vertical
488mixing is obtained by multiplying vertical density diffusion
489tendency by $g\,z$ and and summing the result vertically:
490\begin{equation} \label{Eq_energ2}
491\begin{split}
492\int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
493&= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta} 
494   - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
495&= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
496+ \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
497\end{split}
498\end{equation}
499where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
500The first term of the right hand side of \eqref{Eq_energ2}  is always zero
501because there is no diffusive flux through the ocean surface and bottom).
502The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
503Therefore \eqref{Eq_energ1} implies that, to be energetically consistent, the product
504${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \eqref{Eq_zdftke_e}, the TKE equation.
505
506Let us now address the space discretization issue.
507The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity
508components are in the centre of the side faces of a $t$-box in staggered C-grid
509(Fig.\ref{Fig_cell}). A space averaging is thus required to obtain the shear TKE production term.
510By redoing the \eqref{Eq_energ1} in the 3D case, it can be shown that the product of
511eddy coefficient by the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
512Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into
513account.
514
515The above energetic considerations leads to
516the following final discrete form for the TKE equation:
517\begin{equation} \label{Eq_zdftke_ene}
518\begin{split}
519\frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv 
520\Biggl\{ \Biggr.
521  &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} } 
522                                                                              \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
523+&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} } 
524                                                                               \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j} 
525\Biggr. \Biggr\}   \\
526%
527- &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
528%
529+&\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]   \\
530%
531- &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
532\end{split}
533\end{equation}
534where the last two terms in \eqref{Eq_zdftke_ene} (vertical diffusion and Kolmogorov dissipation)
535are time stepped using a backward scheme (see\S\ref{STP_forward_imp}).
536Note that the Kolmogorov term has been linearized in time in order to render
537the implicit computation possible. The restart of the TKE scheme
538requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as they all appear in
539the right hand side of \eqref{Eq_zdftke_ene}. For the latter, it is in fact
540the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
541
542% -------------------------------------------------------------------------------------------------------------
543%        GLS Generic Length Scale Scheme
544% -------------------------------------------------------------------------------------------------------------
545\subsection{GLS Generic Length Scale (\key{zdfgls})}
546\label{ZDF_gls}
547
548%--------------------------------------------namzdf_gls---------------------------------------------------------
549\namdisplay{namzdf_gls}
550%--------------------------------------------------------------------------------------------------------------
551
552The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on
553two prognostic equations: one for the turbulent kinetic energy $\bar {e}$, and another
554for the generic length scale, $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}.
555This later variable is defined as : $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
556where the triplet $(p, m, n)$ value given in Tab.\ref{Tab_GLS} allows to recover
557a number of well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982},
558$k$-$\epsilon$ \citep{Rodi_1987}, $k$-$\omega$ \citep{Wilcox_1988} 
559among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).
560The GLS scheme is given by the following set of equations:
561\begin{equation} \label{Eq_zdfgls_e}
562\frac{\partial \bar{e}}{\partial t} =
563\frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
564                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
565-K_\rho \,N^2
566+\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
567- \epsilon
568\end{equation}
569
570\begin{equation} \label{Eq_zdfgls_psi}
571   \begin{split}
572\frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
573\frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
574                                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
575- C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
576&+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
577                                  \;\frac{\partial \psi}{\partial k}} \right]\;
578   \end{split}
579\end{equation}
580
581\begin{equation} \label{Eq_zdfgls_kz}
582   \begin{split}
583         K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
584         K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
585   \end{split}
586\end{equation}
587
588\begin{equation} \label{Eq_zdfgls_eps}
589{\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
590\end{equation}
591where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2})
592and $\epsilon$ the dissipation rate.
593The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$)
594depends of the choice of the turbulence model. Four different turbulent models are pre-defined
595(Tab.\ref{Tab_GLS}). They are made available through the \np{nn\_clo} namelist parameter.
596
597%--------------------------------------------------TABLE--------------------------------------------------
598\begin{table}[htbp]  \begin{center}
599%\begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
600\begin{tabular}{ccccc}
601                         &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
602%                        & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\ 
603\hline  \hline 
604\np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
605\hline 
606$( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
607$\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
608$\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
609$C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
610$C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
611$C_3$              &      1.           &     1.              &      1.                &       1.           \\
612$F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
613\hline
614\hline
615\end{tabular}
616\caption{   \label{Tab_GLS} 
617Set of predefined GLS parameters, or equivalently predefined turbulence models available
618with \key{zdfgls} and controlled by the \np{nn\_clos} namelist variable in \ngn{namzdf\_gls} .}
619\end{center}   \end{table}
620%--------------------------------------------------------------------------------------------------------------
621
622In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force
623the convergence of the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length)
624value near physical boundaries (logarithmic boundary layer law). $C_{\mu}$ and $C_{\mu'}$ 
625are calculated from stability function proposed by \citet{Galperin_al_JAS88}, or by \citet{Kantha_Clayson_1994} 
626or one of the two functions suggested by \citet{Canuto_2001}  (\np{nn\_stab\_func} = 0, 1, 2 or 3, resp.}).
627The value of $C_{0\mu}$ depends of the choice of the stability function.
628
629The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated
630thanks to Dirichlet or Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp.
631As for TKE closure , the wave effect on the mixing is considered when \np{ln\_crban}~=~true
632\citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. The \np{rn\_crban} namelist parameter
633is $\alpha_{CB}$ in \eqref{ZDF_Esbc} and \np{rn\_charn} provides the value of $\beta$ in \eqref{ZDF_Lsbc}.
634
635The $\psi$ equation is known to fail in stably stratified flows, and for this reason
636almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
637With this clipping, the maximum permissible length scale is determined by
638$l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. A value of $c_{lim} = 0.53$ is often used
639\citep{Galperin_al_JAS88}. \cite{Umlauf_Burchard_CSR05} show that the value of
640the clipping factor is of crucial importance for the entrainment depth predicted in
641stably stratified situations, and that its value has to be chosen in accordance
642with the algebraic model for the turbulent fluxes. The clipping is only activated
643if \np{ln\_length\_lim}=true, and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value.
644
645The time and space discretization of the GLS equations follows the same energetic
646consideration as for the TKE case described in \S\ref{ZDF_tke_ene}  \citep{Burchard_OM02}.
647Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}.
648
649% -------------------------------------------------------------------------------------------------------------
650%        K Profile Parametrisation (KPP)
651% -------------------------------------------------------------------------------------------------------------
652\subsection{K Profile Parametrisation (KPP) (\key{zdfkpp}) }
653\label{ZDF_kpp}
654
655%--------------------------------------------namkpp--------------------------------------------------------
656\namdisplay{namzdf_kpp}
657%--------------------------------------------------------------------------------------------------------------
658
659The KKP scheme has been implemented by J. Chanut ...
660Options are defined through the  \ngn{namzdf\_kpp} namelist variables.
661
662\colorbox{yellow}{Add a description of KPP here.}
663
664
665% ================================================================
666% Convection
667% ================================================================
668\section{Convection}
669\label{ZDF_conv}
670
671%--------------------------------------------namzdf--------------------------------------------------------
672\namdisplay{namzdf}
673%--------------------------------------------------------------------------------------------------------------
674
675Static instabilities (i.e. light potential densities under heavy ones) may
676occur at particular ocean grid points. In nature, convective processes
677quickly re-establish the static stability of the water column. These
678processes have been removed from the model via the hydrostatic
679assumption so they must be parameterized. Three parameterisations
680are available to deal with convective processes: a non-penetrative
681convective adjustment or an enhanced vertical diffusion, or/and the
682use of a turbulent closure scheme.
683
684% -------------------------------------------------------------------------------------------------------------
685%       Non-Penetrative Convective Adjustment
686% -------------------------------------------------------------------------------------------------------------
687\subsection   [Non-Penetrative Convective Adjustment (\np{ln\_tranpc}) ]
688         {Non-Penetrative Convective Adjustment (\np{ln\_tranpc}=.true.) }
689\label{ZDF_npc}
690
691%--------------------------------------------namzdf--------------------------------------------------------
692\namdisplay{namzdf}
693%--------------------------------------------------------------------------------------------------------------
694
695%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
696\begin{figure}[!htb]    \begin{center}
697\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_npc.pdf}
698\caption{  \label{Fig_npc} 
699Example of an unstable density profile treated by the non penetrative
700convective adjustment algorithm. $1^{st}$ step: the initial profile is checked from
701the surface to the bottom. It is found to be unstable between levels 3 and 4.
702They are mixed. The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5
703are mixed. The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are
704mixed. The $1^{st}$ step ends since the density profile is then stable below
705the level 3. $2^{nd}$ step: the new $\rho$ profile is checked following the same
706procedure as in $1^{st}$ step: levels 2 to 5 are mixed. The new density profile
707is checked. It is found stable: end of algorithm.}
708\end{center}   \end{figure}
709%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
710
711Options are defined through the  \ngn{namzdf} namelist variables.
712The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}=true.
713It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously
714the statically unstable portion of the water column, but only until the density
715structure becomes neutrally stable ($i.e.$ until the mixed portion of the water
716column has \textit{exactly} the density of the water just below) \citep{Madec_al_JPO91}.
717The associated algorithm is an iterative process used in the following way
718(Fig. \ref{Fig_npc}): starting from the top of the ocean, the first instability is
719found. Assume in the following that the instability is located between levels
720$k$ and $k+1$. The potential temperature and salinity in the two levels are
721vertically mixed, conserving the heat and salt contents of the water column.
722The new density is then computed by a linear approximation. If the new
723density profile is still unstable between levels $k+1$ and $k+2$, levels $k$,
724$k+1$ and $k+2$ are then mixed. This process is repeated until stability is
725established below the level $k$ (the mixing process can go down to the
726ocean bottom). The algorithm is repeated to check if the density profile
727between level $k-1$ and $k$ is unstable and/or if there is no deeper instability.
728
729This algorithm is significantly different from mixing statically unstable levels
730two by two. The latter procedure cannot converge with a finite number
731of iterations for some vertical profiles while the algorithm used in \NEMO 
732converges for any profile in a number of iterations which is less than the
733number of vertical levels. This property is of paramount importance as
734pointed out by \citet{Killworth1989}: it avoids the existence of permanent
735and unrealistic static instabilities at the sea surface. This non-penetrative
736convective algorithm has been proved successful in studies of the deep
737water formation in the north-western Mediterranean Sea
738\citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}.
739
740Note that in the current implementation of this algorithm presents several
741limitations. First, potential density referenced to the sea surface is used to
742check whether the density profile is stable or not. This is a strong
743simplification which leads to large errors for realistic ocean simulations.
744Indeed, many water masses of the world ocean, especially Antarctic Bottom
745Water, are unstable when represented in surface-referenced potential density.
746The scheme will erroneously mix them up. Second, the mixing of potential
747density is assumed to be linear. This assures the convergence of the algorithm
748even when the equation of state is non-linear. Small static instabilities can thus
749persist due to cabbeling: they will be treated at the next time step.
750Third, temperature and salinity, and thus density, are mixed, but the
751corresponding velocity fields remain unchanged. When using a Richardson
752Number dependent eddy viscosity, the mixing of momentum is done through
753the vertical diffusion: after a static adjustment, the Richardson Number is zero
754and thus the eddy viscosity coefficient is at a maximum. When this convective
755adjustment algorithm is used with constant vertical eddy viscosity, spurious
756solutions can occur since the vertical momentum diffusion remains small even
757after a static adjustment. In that case, we recommend the addition of momentum
758mixing in a manner that mimics the mixing in temperature and salinity
759\citep{Speich_PhD92, Speich_al_JPO96}.
760
761% -------------------------------------------------------------------------------------------------------------
762%       Enhanced Vertical Diffusion
763% -------------------------------------------------------------------------------------------------------------
764\subsection   [Enhanced Vertical Diffusion (\np{ln\_zdfevd})]
765         {Enhanced Vertical Diffusion (\np{ln\_zdfevd}=true)}
766\label{ZDF_evd}
767
768%--------------------------------------------namzdf--------------------------------------------------------
769\namdisplay{namzdf}
770%--------------------------------------------------------------------------------------------------------------
771
772Options are defined through the  \ngn{namzdf} namelist variables.
773The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}=true.
774In this case, the vertical eddy mixing coefficients are assigned very large values
775(a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable
776($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative)
777\citep{Lazar_PhD97, Lazar_al_JPO99}. This is done either on tracers only
778(\np{nn\_evdm}=0) or on both momentum and tracers (\np{nn\_evdm}=1).
779
780In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and
781if \np{nn\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ 
782values also, are set equal to the namelist parameter \np{rn\_avevd}. A typical value
783for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterisation of
784convective processes is less time consuming than the convective adjustment
785algorithm presented above when mixing both tracers and momentum in the
786case of static instabilities. It requires the use of an implicit time stepping on
787vertical diffusion terms (i.e. \np{ln\_zdfexp}=false).
788
789Note that the stability test is performed on both \textit{before} and \textit{now} 
790values of $N^2$. This removes a potential source of divergence of odd and
791even time step in a leapfrog environment \citep{Leclair_PhD2010} (see \S\ref{STP_mLF}).
792
793% -------------------------------------------------------------------------------------------------------------
794%       Turbulent Closure Scheme
795% -------------------------------------------------------------------------------------------------------------
796\subsection{Turbulent Closure Scheme (\key{zdftke} or \key{zdfgls})}
797\label{ZDF_tcs}
798
799The turbulent closure scheme presented in \S\ref{ZDF_tke} and \S\ref{ZDF_gls} 
800(\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically
801unstable density profiles. In such a case, the term corresponding to the
802destruction of turbulent kinetic energy through stratification in \eqref{Eq_zdftke_e} 
803or \eqref{Eq_zdfgls_e} becomes a source term, since $N^2$ is negative.
804It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring
805$A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1})$. These large values
806restore the static stability of the water column in a way similar to that of the
807enhanced vertical diffusion parameterisation (\S\ref{ZDF_evd}). However,
808in the vicinity of the sea surface (first ocean layer), the eddy coefficients
809computed by the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
810because the mixing length scale is bounded by the distance to the sea surface.
811It can thus be useful to combine the enhanced vertical
812diffusion with the turbulent closure scheme, $i.e.$ setting the \np{ln\_zdfnpc} 
813namelist parameter to true and defining the turbulent closure CPP key all together.
814
815The KPP turbulent closure scheme already includes enhanced vertical diffusion
816in the case of convection, as governed by the variables $bvsqcon$ and $difcon$ 
817found in \mdl{zdfkpp}, therefore \np{ln\_zdfevd}=false should be used with the KPP
818scheme. %gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
819
820% ================================================================
821% Double Diffusion Mixing
822% ================================================================
823\section  [Double Diffusion Mixing (\key{zdfddm})]
824      {Double Diffusion Mixing (\key{zdfddm})}
825\label{ZDF_ddm}
826
827%-------------------------------------------namzdf_ddm-------------------------------------------------
828\namdisplay{namzdf_ddm}
829%--------------------------------------------------------------------------------------------------------------
830
831Options are defined through the  \ngn{namzdf\_ddm} namelist variables.
832Double diffusion occurs when relatively warm, salty water overlies cooler, fresher
833water, or vice versa. The former condition leads to salt fingering and the latter
834to diffusive convection. Double-diffusive phenomena contribute to diapycnal
835mixing in extensive regions of the ocean.  \citet{Merryfield1999} include a
836parameterisation of such phenomena in a global ocean model and show that
837it leads to relatively minor changes in circulation but exerts significant regional
838influences on temperature and salinity. This parameterisation has been
839introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key.
840
841Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
842\begin{align*} % \label{Eq_zdfddm_Kz}
843    &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT}  \\
844    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
845\end{align*}
846where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
847and $o$ by processes other than double diffusion. The rates of double-diffusive
848mixing depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,
849where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline
850contraction (see \S\ref{TRA_eos}). To represent mixing of $S$ and $T$ by salt
851fingering, we adopt the diapycnal diffusivities suggested by Schmitt (1981):
852\begin{align} \label{Eq_zdfddm_f}
853A_f^{vS} &=    \begin{cases}
854   \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
855   0                              &\text{otherwise} 
856            \end{cases}   
857\\           \label{Eq_zdfddm_f_T}
858A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
859\end{align}
860
861%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
862\begin{figure}[!t]   \begin{center}
863\includegraphics[width=0.99\textwidth]{./TexFiles/Figures/Fig_zdfddm.pdf}
864\caption{  \label{Fig_zdfddm}
865From \citet{Merryfield1999} : (a) Diapycnal diffusivities $A_f^{vT}$ 
866and $A_f^{vS}$ for temperature and salt in regions of salt fingering. Heavy
867curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves
868$A^{\ast v} = 10^{-4}~m^2.s^{-1}$ ; (b) diapycnal diffusivities $A_d^{vT}$ and
869$A_d^{vS}$ for temperature and salt in regions of diffusive convection. Heavy
870curves denote the Federov parameterisation and thin curves the Kelley
871parameterisation. The latter is not implemented in \NEMO. }
872\end{center}    \end{figure}
873%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
874
875The factor 0.7 in \eqref{Eq_zdfddm_f_T} reflects the measured ratio
876$\alpha F_T /\beta F_S \approx  0.7$ of buoyancy flux of heat to buoyancy
877flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}). Following  \citet{Merryfield1999},
878we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
879
880To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:
881\begin{align}  \label{Eq_zdfddm_d}
882A_d^{vT} &=    \begin{cases}
883   1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
884                           &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
885   0                       &\text{otherwise} 
886            \end{cases}   
887\\          \label{Eq_zdfddm_d_S}
888A_d^{vS} &=    \begin{cases}
889   A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right)
890                           &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
891   A_d^{vT} \ 0.15 \ R_\rho
892                           &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
893   0                       &\text{otherwise} 
894            \end{cases}   
895\end{align}
896
897The dependencies of \eqref{Eq_zdfddm_f} to \eqref{Eq_zdfddm_d_S} on $R_\rho$ 
898are illustrated in Fig.~\ref{Fig_zdfddm}. Implementing this requires computing
899$R_\rho$ at each grid point on every time step. This is done in \mdl{eosbn2} at the
900same time as $N^2$ is computed. This avoids duplication in the computation of
901$\alpha$ and $\beta$ (which is usually quite expensive).
902
903% ================================================================
904% Bottom Friction
905% ================================================================
906\section  [Bottom Friction (\textit{zdfbfr})]   {Bottom Friction (\mdl{zdfbfr} module)}
907\label{ZDF_bfr}
908
909%--------------------------------------------nambfr--------------------------------------------------------
910\namdisplay{nambfr}
911%--------------------------------------------------------------------------------------------------------------
912
913Options are defined through the  \ngn{nambfr} namelist variables.
914Both the surface momentum flux (wind stress) and the bottom momentum
915flux (bottom friction) enter the equations as a condition on the vertical
916diffusive flux. For the bottom boundary layer, one has:
917\begin{equation} \label{Eq_zdfbfr_flux}
918A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
919\end{equation}
920where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum
921outside the logarithmic turbulent boundary layer (thickness of the order of
9221~m in the ocean). How ${\cal F}_h^{\textbf U}$ influences the interior depends on the
923vertical resolution of the model near the bottom relative to the Ekman layer
924depth. For example, in order to obtain an Ekman layer depth
925$d = \sqrt{2\;A^{vm}} / f = 50$~m, one needs a vertical diffusion coefficient
926$A^{vm} = 0.125$~m$^2$s$^{-1}$ (for a Coriolis frequency
927$f = 10^{-4}$~m$^2$s$^{-1}$). With a background diffusion coefficient
928$A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
929When the vertical mixing coefficient is this small, using a flux condition is
930equivalent to entering the viscous forces (either wind stress or bottom friction)
931as a body force over the depth of the top or bottom model layer. To illustrate
932this, consider the equation for $u$ at $k$, the last ocean level:
933\begin{equation} \label{Eq_zdfbfr_flux2}
934\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}}
935\end{equation}
936If the bottom layer thickness is 200~m, the Ekman transport will
937be distributed over that depth. On the other hand, if the vertical resolution
938is high (1~m or less) and a turbulent closure model is used, the turbulent
939Ekman layer will be represented explicitly by the model. However, the
940logarithmic layer is never represented in current primitive equation model
941applications: it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
942Two choices are available in \NEMO: a linear and a quadratic bottom friction.
943Note that in both cases, the rotation between the interior velocity and the
944bottom friction is neglected in the present release of \NEMO.
945
946In the code, the bottom friction is imposed by adding the trend due to the bottom
947friction to the general momentum trend in \mdl{dynbfr}. For the time-split surface
948pressure gradient algorithm, the momentum trend due to the barotropic component
949needs to be handled separately. For this purpose it is convenient to compute and
950store coefficients which can be simply combined with bottom velocities and geometric
951values to provide the momentum trend due to bottom friction.
952These coefficients are computed in \mdl{zdfbfr} and generally take the form
953$c_b^{\textbf U}$ where:
954\begin{equation} \label{Eq_zdfbfr_bdef}
955\frac{\partial {\textbf U_h}}{\partial t} =
956  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
957\end{equation}
958where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
959
960% -------------------------------------------------------------------------------------------------------------
961%       Linear Bottom Friction
962% -------------------------------------------------------------------------------------------------------------
963\subsection{Linear Bottom Friction (\np{nn\_botfr} = 0 or 1) }
964\label{ZDF_bfr_linear}
965
966The linear bottom friction parameterisation (including the special case
967of a free-slip condition) assumes that the bottom friction
968is proportional to the interior velocity (i.e. the velocity of the last
969model level):
970\begin{equation} \label{Eq_zdfbfr_linear}
971{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
972\end{equation}
973where $r$ is a friction coefficient expressed in ms$^{-1}$.
974This coefficient is generally estimated by setting a typical decay time
975$\tau$ in the deep ocean,
976and setting $r = H / \tau$, where $H$ is the ocean depth. Commonly accepted
977values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}.
978A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used
979in quasi-geostrophic models. One may consider the linear friction as an
980approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982},
981Eq. 9.6.6). For example, with a drag coefficient $C_D = 0.002$, a typical speed
982of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, and assuming an ocean depth
983$H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
984This is the default value used in \NEMO. It corresponds to a decay time scale
985of 115~days. It can be changed by specifying \np{rn\_bfric1} (namelist parameter).
986
987For the linear friction case the coefficients defined in the general
988expression \eqref{Eq_zdfbfr_bdef} are:
989\begin{equation} \label{Eq_zdfbfr_linbfr_b}
990\begin{split}
991 c_b^u &= - r\\
992 c_b^v &= - r\\
993\end{split}
994\end{equation}
995When \np{nn\_botfr}=1, the value of $r$ used is \np{rn\_bfric1}.
996Setting \np{nn\_botfr}=0 is equivalent to setting $r=0$ and leads to a free-slip
997bottom boundary condition. These values are assigned in \mdl{zdfbfr}.
998From v3.2 onwards there is support for local enhancement of these values
999via an externally defined 2D mask array (\np{ln\_bfr2d}=true) given
1000in the \ifile{bfr\_coef} input NetCDF file. The mask values should vary from 0 to 1.
1001Locations with a non-zero mask value will have the friction coefficient increased
1002by $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfric1}.
1003
1004% -------------------------------------------------------------------------------------------------------------
1005%       Non-Linear Bottom Friction
1006% -------------------------------------------------------------------------------------------------------------
1007\subsection{Non-Linear Bottom Friction (\np{nn\_botfr} = 2)}
1008\label{ZDF_bfr_nonlinear}
1009
1010The non-linear bottom friction parameterisation assumes that the bottom
1011friction is quadratic:
1012\begin{equation} \label{Eq_zdfbfr_nonlinear}
1013{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
1014}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
1015\end{equation}
1016where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy
1017due to tides, internal waves breaking and other short time scale currents.
1018A typical value of the drag coefficient is $C_D = 10^{-3} $. As an example,
1019the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and
1020$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} 
1021uses $C_D = 1.4\;10^{-3}$ and $e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
1022The CME choices have been set as default values (\np{rn\_bfric2} and \np{rn\_bfeb2} 
1023namelist parameters).
1024
1025As for the linear case, the bottom friction is imposed in the code by
1026adding the trend due to the bottom friction to the general momentum trend
1027in \mdl{dynbfr}.
1028For the non-linear friction case the terms
1029computed in \mdl{zdfbfr}  are:
1030\begin{equation} \label{Eq_zdfbfr_nonlinbfr}
1031\begin{split}
1032 c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\
1033 c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\
1034\end{split}
1035\end{equation}
1036
1037The coefficients that control the strength of the non-linear bottom friction are
1038initialised as namelist parameters: $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}.
1039Note for applications which treat tides explicitly a low or even zero value of
1040\np{rn\_bfeb2} is recommended. From v3.2 onwards a local enhancement of $C_D$ 
1041is possible via an externally defined 2D mask array (\np{ln\_bfr2d}=true).
1042See previous section for details.
1043
1044% -------------------------------------------------------------------------------------------------------------
1045%       Bottom Friction stability
1046% -------------------------------------------------------------------------------------------------------------
1047\subsection{Bottom Friction stability considerations}
1048\label{ZDF_bfr_stability}
1049
1050Some care needs to exercised over the choice of parameters to ensure that the
1051implementation of bottom friction does not induce numerical instability. For
1052the purposes of stability analysis, an approximation to \eqref{Eq_zdfbfr_flux2}
1053is:
1054\begin{equation} \label{Eqn_bfrstab}
1055\begin{split}
1056 \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
1057               &= -\frac{ru}{e_{3u}}\;2\rdt\\
1058\end{split}
1059\end{equation}
1060\noindent where linear bottom friction and a leapfrog timestep have been assumed.
1061To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have:
1062\begin{equation}
1063 |\Delta u| < \;|u|
1064\end{equation}
1065\noindent which, using \eqref{Eqn_bfrstab}, gives:
1066\begin{equation}
1067r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
1068\end{equation}
1069This same inequality can also be derived in the non-linear bottom friction case
1070if a velocity of 1 m.s$^{-1}$ is assumed. Alternatively, this criterion can be
1071rearranged to suggest a minimum bottom box thickness to ensure stability:
1072\begin{equation}
1073e_{3u} > 2\;r\;\rdt
1074\end{equation}
1075\noindent which it may be necessary to impose if partial steps are being used.
1076For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then
1077$e_{3u}$ should be greater than 3.6 m. For most applications, with physically
1078sensible parameters these restrictions should not be of concern. But
1079caution may be necessary if attempts are made to locally enhance the bottom
1080friction parameters.
1081To ensure stability limits are imposed on the bottom friction coefficients both during
1082initialisation and at each time step. Checks at initialisation are made in \mdl{zdfbfr} 
1083(assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1084The number of breaches of the stability criterion are reported as well as the minimum
1085and maximum values that have been set. The criterion is also checked at each time step,
1086using the actual velocity, in \mdl{dynbfr}. Values of the bottom friction coefficient are
1087reduced as necessary to ensure stability; these changes are not reported.
1088
1089Limits on the bottom friction coefficient are not imposed if the user has elected to
1090handle the bottom friction implicitly (see \S\ref{ZDF_bfr_imp}). The number of potential
1091breaches of the explicit stability criterion are still reported for information purposes.
1092
1093% -------------------------------------------------------------------------------------------------------------
1094%       Implicit Bottom Friction
1095% -------------------------------------------------------------------------------------------------------------
1096\subsection{Implicit Bottom Friction (\np{ln\_bfrimp}$=$\textit{T})}
1097\label{ZDF_bfr_imp}
1098
1099An optional implicit form of bottom friction has been implemented to improve
1100model stability. We recommend this option for shelf sea and coastal ocean applications, especially
1101for split-explicit time splitting. This option can be invoked by setting \np{ln\_bfrimp} 
1102to \textit{true} in the \textit{nambfr} namelist. This option requires \np{ln\_zdfexp} to be \textit{false} 
1103in the \textit{namzdf} namelist.
1104
1105This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, the
1106bottom boundary condition is implemented implicitly.
1107
1108\begin{equation} \label{Eq_dynzdf_bfr}
1109\left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{mbk}
1110    = \binom{c_{b}^{u}u^{n+1}_{mbk}}{c_{b}^{v}v^{n+1}_{mbk}}
1111\end{equation}
1112
1113where $mbk$ is the layer number of the bottom wet layer. superscript $n+1$ means the velocity used in the
1114friction formula is to be calculated, so, it is implicit.
1115
1116If split-explicit time splitting is used, care must be taken to avoid the double counting of
1117the bottom friction in the 2-D barotropic momentum equations. As NEMO only updates the barotropic
1118pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, we need to remove
1119the bottom friction induced by these two terms which has been included in the 3-D momentum trend
1120and update it with the latest value. On the other hand, the bottom friction contributed by the
1121other terms (e.g. the advection term, viscosity term) has been included in the 3-D momentum equations
1122and should not be added in the 2-D barotropic mode.
1123
1124The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the
1125following:
1126
1127\begin{equation} \label{Eq_dynspg_ts_bfr1}
1128\frac{\textbf{U}_{med}-\textbf{U}^{m-1}}{2\Delta t}=-g\nabla\eta-f\textbf{k}\times\textbf{U}^{m}+c_{b}
1129\left(\textbf{U}_{med}-\textbf{U}^{m-1}\right)
1130\end{equation}
1131\begin{equation} \label{Eq_dynspg_ts_bfr2}
1132\frac{\textbf{U}^{m+1}-\textbf{U}_{med}}{2\Delta t}=\textbf{T}+
1133\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{U}^{'}\right)-
11342\Delta t_{bc}c_{b}\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{u}_{b}\right)
1135\end{equation}
1136
1137where $\textbf{T}$ is the vertical integrated 3-D momentum trend. We assume the leap-frog time-stepping
1138is used here. $\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step.
1139 $c_{b}$ is the friction coefficient. $\eta$ is the sea surface level calculated in the barotropic loops
1140while $\eta^{'}$ is the sea surface level used in the 3-D baroclinic mode. $\textbf{u}_{b}$ is the bottom
1141layer horizontal velocity.
1142
1143
1144
1145
1146% -------------------------------------------------------------------------------------------------------------
1147%       Bottom Friction with split-explicit time splitting
1148% -------------------------------------------------------------------------------------------------------------
1149\subsection{Bottom Friction with split-explicit time splitting (\np{ln\_bfrimp}$=$\textit{F})}
1150\label{ZDF_bfr_ts}
1151
1152When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, the
1153bottom velocity at the before time step is used. This velocity includes both the
1154baroclinic and barotropic components which is appropriate when using either the
1155explicit or filtered surface pressure gradient algorithms (\key{dynspg\_exp} or
1156{\key{dynspg\_flt}). Extra attention is required, however, when using
1157split-explicit time stepping (\key{dynspg\_ts}). In this case the free surface
1158equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, while the three
1159dimensional prognostic variables are solved with the longer time step
1160of \np{rn\_rdt} seconds. The trend in the barotropic momentum due to bottom
1161friction appropriate to this method is that given by the selected parameterisation
1162($i.e.$ linear or non-linear bottom friction) computed with the evolving velocities
1163at each barotropic timestep.
1164
1165In the case of non-linear bottom friction, we have elected to partially linearise
1166the problem by keeping the coefficients fixed throughout the barotropic
1167time-stepping to those computed in \mdl{zdfbfr} using the now timestep.
1168This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to:
1169
1170\begin{enumerate}
1171\item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before
1172barotropic velocity to the bottom friction component of the vertically
1173integrated momentum trend. Note the same stability check that is carried out
1174on the bottom friction coefficient in \mdl{dynbfr} has to be applied here to
1175ensure that the trend removed matches that which was added in \mdl{dynbfr}.
1176\item At each barotropic step, compute the contribution of the current barotropic
1177velocity to the trend due to bottom friction. Add this contribution to the
1178vertically integrated momentum trend. This contribution is handled implicitly which
1179eliminates the need to impose a stability criteria on the values of the bottom friction
1180coefficient within the barotropic loop.
1181\end{enumerate}
1182
1183Note that the use of an implicit formulation within the barotropic loop
1184for the bottom friction trend means that any limiting of the bottom friction coefficient
1185in \mdl{dynbfr} does not adversely affect the solution when using split-explicit time
1186splitting. This is because the major contribution to bottom friction is likely to come from
1187the barotropic component which uses the unrestricted value of the coefficient. However, if the
1188limiting is thought to be having a major effect (a more likely prospect in coastal and shelf seas
1189applications) then the fully implicit form of the bottom friction should be used (see \S\ref{ZDF_bfr_imp} )
1190which can be selected by setting \np{ln\_bfrimp} $=$ \textit{true}.
1191
1192Otherwise, the implicit formulation takes the form:
1193\begin{equation} \label{Eq_zdfbfr_implicitts}
1194 \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ] 
1195\end{equation}
1196where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),
1197$c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and $RHS$ represents
1198all the components to the vertically integrated momentum trend except for that due to bottom friction.
1199
1200
1201
1202
1203% ================================================================
1204% Tidal Mixing
1205% ================================================================
1206\section{Tidal Mixing (\key{zdftmx})}
1207\label{ZDF_tmx}
1208
1209%--------------------------------------------namzdf_tmx--------------------------------------------------
1210\namdisplay{namzdf_tmx}
1211%--------------------------------------------------------------------------------------------------------------
1212
1213
1214% -------------------------------------------------------------------------------------------------------------
1215%        Bottom intensified tidal mixing
1216% -------------------------------------------------------------------------------------------------------------
1217\subsection{Bottom intensified tidal mixing}
1218\label{ZDF_tmx_bottom}
1219
1220Options are defined through the  \ngn{namzdf\_tmx} namelist variables.
1221The parameterization of tidal mixing follows the general formulation for
1222the vertical eddy diffusivity proposed by \citet{St_Laurent_al_GRL02} and
1223first introduced in an OGCM by \citep{Simmons_al_OM04}.
1224In this formulation an additional vertical diffusivity resulting from internal tide breaking,
1225$A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, the energy transfer from barotropic
1226tides to baroclinic tides :
1227\begin{equation} \label{Eq_Ktides}
1228A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 }
1229\end{equation}
1230where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency
1231(see \S\ref{TRA_bn2}), $\rho$ the density, $q$ the tidal dissipation efficiency,
1232and $F(z)$ the vertical structure function.
1233
1234The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter)
1235and is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).
1236The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter)
1237represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally,
1238with the remaining $1-q$ radiating away as low mode internal waves and
1239contributing to the background internal wave field. A value of $q=1/3$ is typically used 
1240\citet{St_Laurent_al_GRL02}.
1241The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical.
1242It is implemented as a simple exponential decaying upward away from the bottom,
1243with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},
1244\begin{equation} \label{Eq_Fz}
1245F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) }
1246\end{equation}
1247and is normalized so that vertical integral over the water column is unity.
1248
1249The associated vertical viscosity is calculated from the vertical
1250diffusivity assuming a Prandtl number of 1, $i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$.
1251In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity
1252is capped at $300\,cm^2/s$ and impose a lower limit on $N^2$ of \np{rn\_n2min} 
1253usually set to $10^{-8} s^{-2}$. These bounds are usually rarely encountered.
1254
1255The internal wave energy map, $E(x, y)$ in \eqref{Eq_Ktides}, is derived
1256from a barotropic model of the tides utilizing a parameterization of the
1257conversion of barotropic tidal energy into internal waves.
1258The essential goal of the parameterization is to represent the momentum
1259exchange between the barotropic tides and the unrepresented internal waves
1260induced by the tidal flow over rough topography in a stratified ocean.
1261In the current version of \NEMO, the map is built from the output of
1262the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}.
1263This model provides the dissipation associated with internal wave energy for the M2 and K1
1264tides component (Fig.~\ref{Fig_ZDF_M2_K1_tmx}). The S2 dissipation is simply approximated
1265as being $1/4$ of the M2 one. The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$.
1266Its global mean value is $1.1$ TW, in agreement with independent estimates
1267\citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.
1268
1269%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1270\begin{figure}[!t]   \begin{center}
1271\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_ZDF_M2_K1_tmx.pdf}
1272\caption{  \label{Fig_ZDF_M2_K1_tmx} 
1273(a) M2 and (b) K1 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). }
1274\end{center}   \end{figure}
1275%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1276 
1277% -------------------------------------------------------------------------------------------------------------
1278%        Indonesian area specific treatment
1279% -------------------------------------------------------------------------------------------------------------
1280\subsection{Indonesian area specific treatment (\np{ln\_zdftmx\_itf})}
1281\label{ZDF_tmx_itf}
1282
1283When the Indonesian Through Flow (ITF) area is included in the model domain,
1284a specific treatment of tidal induced mixing in this area can be used.
1285It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide
1286an input NetCDF file, \ifile{mask\_itf}, which contains a mask array defining the ITF area
1287where the specific treatment is applied.
1288
1289When \np{ln\_tmx\_itf}=true, the two key parameters $q$ and $F(z)$ are adjusted following
1290the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}:
1291
1292First, the Indonesian archipelago is a complex geographic region
1293with a series of large, deep, semi-enclosed basins connected via
1294numerous narrow straits. Once generated, internal tides remain
1295confined within this semi-enclosed area and hardly radiate away.
1296Therefore all the internal tides energy is consumed within this area.
1297So it is assumed that $q = 1$, $i.e.$ all the energy generated is available for mixing.
1298Note that for test purposed, the ITF tidal dissipation efficiency is a
1299namelist parameter (\np{rn\_tfe\_itf}). A value of $1$ or close to is
1300this recommended for this parameter.
1301
1302Second, the vertical structure function, $F(z)$, is no more associated
1303with a bottom intensification of the mixing, but with a maximum of
1304energy available within the thermocline. \citet{Koch-Larrouy_al_GRL07} 
1305have suggested that the vertical distribution of the energy dissipation
1306proportional to $N^2$ below the core of the thermocline and to $N$ above.
1307The resulting $F(z)$ is:
1308\begin{equation} \label{Eq_Fz_itf}
1309F(i,j,k) \sim     \left\{ \begin{aligned}
1310\frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\
1311\frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$}
1312                      \end{aligned} \right.
1313\end{equation}
1314
1315Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,
1316which agrees with the independent estimates inferred from observations.
1317Introduced in a regional OGCM, the parameterization improves the water mass
1318characteristics in the different Indonesian seas, suggesting that the horizontal
1319and vertical distributions of the mixing are adequately prescribed
1320\citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}.
1321Note also that such a parameterisation has a significant impact on the behaviour
1322of global coupled GCMs \citep{Koch-Larrouy_al_CD10}.
1323
1324
1325% ================================================================
Note: See TracBrowser for help on using the repository browser.