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 NEMO/trunk/doc/latex/NEMO/subfiles – NEMO

source: NEMO/trunk/doc/latex/NEMO/subfiles/chap_ZDF.tex @ 10405

Last change on this file since 10405 was 10354, checked in by nicolasmartin, 5 years ago

Vast edition of LaTeX subfiles to improve the readability by cutting sentences in a more suitable way
Every sentence begins in a new line and if necessary is splitted around 110 characters lenght for side-by-side visualisation,
this setting may not be adequate for everyone but something has to be set.
The punctuation was the primer trigger for the cutting process, otherwise subordinators and coordinators, in order to mostly keep a meaning for each line

File size: 79.5 KB
Line 
1\documentclass[../tex_main/NEMO_manual]{subfiles}
2\begin{document}
3% ================================================================
4% Chapter  Vertical Ocean Physics (ZDF)
5% ================================================================
6\chapter{Vertical Ocean Physics (ZDF)}
7\label{chap:ZDF}
8\minitoc
9
10%gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN.
11
12
13\newpage
14$\ $\newline    % force a new ligne
15
16
17% ================================================================
18% Vertical Mixing
19% ================================================================
20\section{Vertical mixing}
21\label{sec:ZDF_zdf}
22
23The discrete form of the ocean subgrid scale physics has been presented in
24\autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}.
25At the surface and bottom boundaries, the turbulent fluxes of momentum, heat and salt have to be defined.
26At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}),
27while at the bottom they are set to zero for heat and salt,
28unless a geothermal flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl} defined,
29see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum
30(see \autoref{sec:ZDF_bfr}).
31
32In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and
33diffusivity coefficients, $A_u^{vm}$ , $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$- points,
34respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}).
35These coefficients can be assumed to be either constant, or a function of the local Richardson number,
36or computed from a turbulent closure model (either TKE or GLS formulation).
37The computation of these coefficients is initialized in the \mdl{zdfini} module and performed in
38the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} modules.
39The trends due to the vertical momentum and tracer diffusion, including the surface forcing,
40are computed and added to the general 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}\forcode{ = .true.}) or a backward time stepping scheme
43(\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing coefficients,
44and thus of the formulation used (see \autoref{chap:STP}).
45
46% -------------------------------------------------------------------------------------------------------------
47%        Constant
48% -------------------------------------------------------------------------------------------------------------
49\subsection{Constant (\protect\key{zdfcst})}
50\label{subsec:ZDF_cst}
51%--------------------------------------------namzdf---------------------------------------------------------
52
53\nlst{namzdf}
54%--------------------------------------------------------------------------------------------------------------
55
56Options are defined through the \ngn{namzdf} namelist variables.
57When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to
58constant values over the whole ocean.
59This is the crudest way to define the vertical ocean physics.
60It is recommended that this option is only used in process studies, not in basin scale simulations.
61Typical values used in this case are:
62\begin{align*} 
63A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\
64A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
65\end{align*}
66
67These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.
68In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity,
69that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and
70$\sim10^{-9}~m^2.s^{-1}$ for salinity.
71
72
73% -------------------------------------------------------------------------------------------------------------
74%        Richardson Number Dependent
75% -------------------------------------------------------------------------------------------------------------
76\subsection{Richardson number dependent (\protect\key{zdfric})}
77\label{subsec:ZDF_ric}
78
79%--------------------------------------------namric---------------------------------------------------------
80
81\nlst{namzdf_ric}
82%--------------------------------------------------------------------------------------------------------------
83
84When \key{zdfric} is defined, a local Richardson number dependent formulation for the vertical momentum and
85tracer eddy coefficients is set through the \ngn{namzdf\_ric} namelist variables.
86The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.
87\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures.
88The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to
89a dependency between the vertical eddy coefficients and the local Richardson number
90($i.e.$ the ratio of stratification to vertical shear).
91Following \citet{Pacanowski_Philander_JPO81}, the following formulation has been implemented:
92\begin{equation} \label{eq:zdfric}
93   \left\{      \begin{aligned}
94         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
95         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
96   \end{aligned}    \right.
97\end{equation}
98where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number,
99$N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
100$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the constant case
101(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that
102can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$.
103The last three values can be modified by setting the \np{rn\_avmri}, \np{rn\_alp} and
104\np{nn\_ric} namelist parameters, respectively.
105
106A simple mixing-layer model to transfer and dissipate the atmospheric forcings
107(wind-stress and buoyancy fluxes) can be activated setting the \np{ln\_mldw}\forcode{ = .true.} in the namelist.
108
109In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and
110the 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}
114h_{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 parameter.
117
118In this similarity height relationship, the turbulent friction velocity:
119\begin{equation}
120u^{*} = \sqrt \frac {|\tau|} {\rho_o}
121\end{equation}
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 (\protect\key{zdftke})}
131\label{subsec:ZDF_tke}
132
133%--------------------------------------------namzdf_tke--------------------------------------------------
134
135\nlst{namzdf_tke}
136%--------------------------------------------------------------------------------------------------------------
137
138The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on
139a prognostic equation for $\bar{e}$, the turbulent kinetic energy,
140and a closure assumption for the turbulent length scales.
141This turbulent closure model has been developed by \citet{Bougeault1989} in the atmospheric case,
142adapted by \citet{Gaspar1990} for the oceanic case, and embedded in OPA, the ancestor of NEMO,
143by \citet{Blanke1993} for equatorial Atlantic simulations.
144Since then, significant modifications have been introduced by \citet{Madec1998} in both the implementation and
145the formulation of the mixing length scale.
146The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear,
147its destruction through stratification, its vertical diffusion, 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 \autoref{subsec: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 and diffusivity coefficients.
166The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with
167vertical 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 of the local Richardson number, $R_i$:
170\begin{align*} \label{eq:prt}
171P_{rt} = \begin{cases}
172                    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}  \\
173                    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}  \\
174                    \ \ 10 &      \text{if $\ 2 \leq R_i$} 
175            \end{cases}
176\end{align*}
177Options are defined through the  \ngn{namzdfy\_tke} namelist variables.
178The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable.
179
180At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as
181$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} namelist parameter.
182The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}), however a much larger value can be used when
183taking into account the surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}).
184The bottom value of TKE is assumed to be equal to the value of the level just above.
185The time integration of the $\bar{e}$ equation may formally lead to negative values because
186the numerical scheme does not ensure its positivity.
187To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} namelist parameter).
188Following \citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$.
189This allows the subsequent formulations to match that of \citet{Gargett1984} for the diffusion in
190the thermocline and deep ocean :  $K_\rho = 10^{-3} / N$.
191In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with
192too weak vertical diffusion.
193They must be specified at least larger than the molecular values, and are set through \np{rn\_avm0} and
194\np{rn\_avt0} (namzdf namelist, see \autoref{subsec:ZDF_cst}).
195
196\subsubsection{Turbulent length scale}
197For computational efficiency, the original formulation of the turbulent length scales proposed by
198\citet{Gaspar1990} has been simplified.
199Four formulations are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist parameter.
200The first two are based on the following first order approximation \citep{Blanke1993}:
201\begin{equation} \label{eq:tke_mxl0_1}
202l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N
203\end{equation}
204which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency.
205The resulting length scale is bounded by the distance to the surface or to the bottom
206(\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor (\np{nn\_mxl}\forcode{ = 1}).
207\citet{Blanke1993} notice that this simplification has two major drawbacks:
208it makes no sense for locally unstable stratification and the computation no longer uses all
209the information contained in the vertical density profile.
210To overcome these drawbacks, \citet{Madec1998} introduces the \np{nn\_mxl}\forcode{ = 2..3} cases,
211which add an extra assumption concerning the vertical gradient of the computed length scale.
212So, the length scales are first evaluated as in \autoref{eq:tke_mxl0_1} and then bounded such that:
213\begin{equation} \label{eq:tke_mxl_constraint}
214\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
215\qquad \text{with }\  l =  l_k = l_\epsilon
216\end{equation}
217\autoref{eq:tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than
218the variations of depth.
219It provides a better approximation of the \citet{Gaspar1990} formulation while being much less
220time consuming.
221In particular, it allows the length scale to be limited not only by the distance to the surface or
222to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as
223the thermocline (\autoref{fig:mixing_length}).
224In order to impose the \autoref{eq:tke_mxl_constraint} constraint, we introduce two additional length scales:
225$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and
226evaluate the dissipation and mixing length scales as
227(and note that here we use numerical indexing):
228%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
229\begin{figure}[!t] \begin{center}
230\includegraphics[width=1.00\textwidth]{Fig_mixing_length}
231\caption{ \protect\label{fig:mixing_length} 
232Illustration of the mixing length computation. }
233\end{center} 
234\end{figure}
235%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
236\begin{equation} \label{eq:tke_mxl2}
237\begin{aligned}
238 l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right)
239    \quad &\text{ from $k=1$ to $jpk$ }\ \\
240 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)}  \right)   
241    \quad &\text{ from $k=jpk$ to $1$ }\ \\
242\end{aligned}
243\end{equation}
244where $l^{(k)}$ is computed using \autoref{eq:tke_mxl0_1}, $i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$.
245
246In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same value:
247$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nn\_mxl}\forcode{ = 3} case,
248the dissipation and mixing turbulent length scales are give as in \citet{Gaspar1990}:
249\begin{equation} \label{eq:tke_mxl_gaspar}
250\begin{aligned}
251& l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }    \\
252& l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
253\end{aligned}
254\end{equation}
255
256At the ocean surface, a non zero length scale is set through the  \np{rn\_mxl0} namelist parameter.
257Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and
258$z_o$ the roughness parameter of the surface.
259Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}.
260In the ocean interior a minimum length scale is set to recover the molecular viscosity when
261$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ).
262
263
264\subsubsection{Surface wave breaking parameterization}
265%-----------------------------------------------------------------------%
266Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified to
267include the effect of surface wave breaking energetics.
268This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow.
269The \citet{Mellor_Blumberg_JPO04} modifications acts on surface length scale and TKE values and
270air-sea drag coefficient.
271The latter concerns the bulk formulea and is not discussed here.
272
273Following \citet{Craig_Banner_JPO94}, the boundary condition on surface TKE value is :
274\begin{equation}  \label{eq:ZDF_Esbc}
275\bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
276\end{equation}
277where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality which depends on the ''wave age'',
278ranging from 57 for mature waves to 146 for younger waves \citep{Mellor_Blumberg_JPO04}.
279The boundary condition on the turbulent length scale follows the Charnock's relation:
280\begin{equation} \label{eq:ZDF_Lsbc}
281l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
282\end{equation}
283where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant.
284\citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by
285\citet{Stacey_JPO99} citing observation evidence, and
286$\alpha_{CB} = 100$ the Craig and Banner's value.
287As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,
288with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}\forcode{ = 67.83} corresponds
289to $\alpha_{CB} = 100$.
290Further setting  \np{ln\_mxl0} to true applies \autoref{eq:ZDF_Lsbc} as surface boundary condition on length scale,
291with $\beta$ hard coded to the Stacey's value.
292Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on
293surface $\bar{e}$ value.
294
295
296\subsubsection{Langmuir cells}
297%--------------------------------------%
298Langmuir circulations (LC) can be described as ordered large-scale vertical motions in
299the surface layer of the oceans.
300Although LC have nothing to do with convection, the circulation pattern is rather similar to
301so-called convective rolls in the atmospheric boundary layer.
302The detailed physics behind LC is described in, for example, \citet{Craik_Leibovich_JFM76}.
303The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and
304wind drift currents.
305
306Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by
307\citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure.
308The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in
309an extra source terms of TKE, $P_{LC}$.
310The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled by setting \np{ln\_lc} to
311\forcode{.true.} in the namtke namelist.
312 
313By making an analogy with the characteristic convective velocity scale ($e.g.$, \citet{D'Alessio_al_JPO98}),
314$P_{LC}$ is assumed to be :
315\begin{equation}
316P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
317\end{equation}
318where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
319With no information about the wave field, $w_{LC}$ is assumed to be proportional to
320the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
321\footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity may be expressed as
322  $u_s =  0.016 \,|U_{10m}|$.
323  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of
324  $1.5~10^{-3}$ give the expression used of $u_s$ as a function of the module of surface stress}.
325For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at
326a finite depth $H_{LC}$ (which is often close to the mixed layer depth),
327and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).
328The resulting expression for $w_{LC}$ is :
329\begin{equation}
330w_{LC}  = \begin{cases}
331                   c_{LC} \,u_s \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
332                   0                             &      \text{otherwise} 
333                 \end{cases}
334\end{equation}
335where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise to fit LES data.
336The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second.
337The value of $c_{LC}$ is set through the \np{rn\_lc} namelist parameter,
338having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.
339
340The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
341$H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by
342converting its kinetic energy to potential energy, according to
343\begin{equation}
344- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} u_s^2
345\end{equation}
346
347
348\subsubsection{Mixing just below the mixed layer}
349%--------------------------------------------------------------%
350
351Vertical mixing parameterizations commonly used in ocean general circulation models tend to
352produce mixed-layer depths that are too shallow during summer months and windy conditions.
353This bias is particularly acute over the Southern Ocean.
354To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{Rodgers_2014}.
355The parameterization is an empirical one, $i.e.$ not derived from theoretical considerations,
356but rather is meant to account for observed processes that affect the density structure of
357the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme
358($i.e.$ near-inertial oscillations and ocean swells and waves).
359
360When using this parameterization ($i.e.$ when \np{nn\_etau}\forcode{ = 1}),
361the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations,
362swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition,
363plus a depth depend one given by:
364\begin{equation}  \label{eq:ZDF_Ehtau}
365S = (1-f_i) \; f_r \; e_s \; e^{-z / h_\tau} 
366\end{equation}
367where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that
368penetrate in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of
369the penetration, and $f_i$ is the ice concentration
370(no penetration if $f_i=1$, that is if the ocean is entirely covered by sea-ice).
371The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter.
372The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0}) or
373a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes
374(\np{nn\_etau}\forcode{ = 1}).
375
376Note that two other option existe, \np{nn\_etau}\forcode{ = 2..3}.
377They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer,
378or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrate the ocean.
379Those two options are obsolescent features introduced for test purposes.
380They will be removed in the next release.
381
382
383
384% from Burchard et al OM 2008 :
385% the most critical process not reproduced by statistical turbulence models is the activity of
386% internal waves and their interaction with turbulence. After the Reynolds decomposition,
387% internal waves are in principle included in the RANS equations, but later partially
388% excluded by the hydrostatic assumption and the model resolution.
389% Thus far, the representation of internal wave mixing in ocean models has been relatively crude
390% (e.g. Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002).
391
392
393
394% -------------------------------------------------------------------------------------------------------------
395%        TKE discretization considerations
396% -------------------------------------------------------------------------------------------------------------
397\subsection{TKE discretization considerations (\protect\key{zdftke})}
398\label{subsec:ZDF_tke_ene}
399
400%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
401\begin{figure}[!t]   \begin{center}
402\includegraphics[width=1.00\textwidth]{Fig_ZDF_TKE_time_scheme}
403\caption{ \protect\label{fig:TKE_time_scheme} 
404Illustration of the TKE time integration and its links to the momentum and tracer time integration. }
405\end{center} 
406\end{figure}
407%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
408
409The production of turbulence by vertical shear (the first term of the right hand side of
410\autoref{eq:zdftke_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion
411(first line in \autoref{eq:PE_zdf}).
412To do so a special care have to be taken for both the time and space discretization of
413the TKE equation \citep{Burchard_OM02,Marsaleix_al_OM08}.
414
415Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows how
416the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with
417the one-level forward time stepping of TKE equation.
418With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to
419the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and
420summing the result vertically:   
421\begin{equation} \label{eq:energ1}
422\begin{split}
423\int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
424&= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}         
425 - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
426\end{split}
427\end{equation}
428Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$,
429known at time $t$ (\autoref{fig:TKE_time_scheme}), as it is required when using the TKE scheme
430(see \autoref{sec:STP_forward_imp}).
431The first term of the right hand side of \autoref{eq:energ1} represents the kinetic energy transfer at
432the surface (atmospheric forcing) and at the bottom (friction effect).
433The second term is always negative.
434It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$.
435\autoref{eq:energ1} implies that, to be energetically consistent,
436the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
437${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$
438(and not by the more straightforward $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
439
440A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
441(second term of the right hand side of \autoref{eq:zdftke_e}).
442This term must balance the input of potential energy resulting from vertical mixing.
443The rate of change of potential energy (in 1D for the demonstration) due vertical mixing is obtained by
444multiplying vertical density diffusion tendency by $g\,z$ and and summing the result vertically:
445\begin{equation} \label{eq:energ2}
446\begin{split}
447\int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
448&= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta} 
449   - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
450&= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
451+ \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
452\end{split}
453\end{equation}
454where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
455The first term of the right hand side of \autoref{eq:energ2} is always zero because
456there is no diffusive flux through the ocean surface and bottom).
457The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
458Therefore \autoref{eq:energ1} implies that, to be energetically consistent,
459the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e}, the TKE equation.
460
461Let us now address the space discretization issue.
462The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in
463the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:cell}).
464A space averaging is thus required to obtain the shear TKE production term.
465By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of eddy coefficient by
466the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
467Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into account.
468
469The above energetic considerations leads to the following final discrete form for the TKE equation:
470\begin{equation} \label{eq:zdftke_ene}
471\begin{split}
472\frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv 
473\Biggl\{ \Biggr.
474  &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} } 
475                                                                              \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
476+&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} } 
477                                                                               \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j} 
478\Biggr. \Biggr\}   \\
479%
480- &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
481%
482+&\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]   \\
483%
484- &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
485\end{split}
486\end{equation}
487where the last two terms in \autoref{eq:zdftke_ene} (vertical diffusion and Kolmogorov dissipation)
488are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}).
489Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible.
490The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as
491they all appear in the right hand side of \autoref{eq:zdftke_ene}.
492For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
493
494% -------------------------------------------------------------------------------------------------------------
495%        GLS Generic Length Scale Scheme
496% -------------------------------------------------------------------------------------------------------------
497\subsection{GLS: Generic Length Scale (\protect\key{zdfgls})}
498\label{subsec:ZDF_gls}
499
500%--------------------------------------------namzdf_gls---------------------------------------------------------
501
502\nlst{namzdf_gls}
503%--------------------------------------------------------------------------------------------------------------
504
505The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations:
506one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale,
507$\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}.
508This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
509where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of
510well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982}, $k$-$\epsilon$ \citep{Rodi_1987},
511$k$-$\omega$ \citep{Wilcox_1988} among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).
512The GLS scheme is given by the following set of equations:
513\begin{equation} \label{eq:zdfgls_e}
514\frac{\partial \bar{e}}{\partial t} =
515\frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
516                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
517-K_\rho \,N^2
518+\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
519- \epsilon
520\end{equation}
521
522\begin{equation} \label{eq:zdfgls_psi}
523   \begin{split}
524\frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
525\frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
526                                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right]
527- C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
528&+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
529                                  \;\frac{\partial \psi}{\partial k}} \right]\;
530   \end{split}
531\end{equation}
532
533\begin{equation} \label{eq:zdfgls_kz}
534   \begin{split}
535         K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
536         K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
537   \end{split}
538\end{equation}
539
540\begin{equation} \label{eq:zdfgls_eps}
541{\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
542\end{equation}
543where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and
544$\epsilon$ the dissipation rate.
545The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of
546the choice of the turbulence model.
547Four different turbulent models are pre-defined (Tab.\autoref{tab:GLS}).
548They are made available through the \np{nn\_clo} namelist parameter.
549
550%--------------------------------------------------TABLE--------------------------------------------------
551\begin{table}[htbp]  \begin{center}
552%\begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
553\begin{tabular}{ccccc}
554                         &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
555%                        & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\ 
556\hline  \hline 
557\np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
558\hline 
559$( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
560$\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
561$\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
562$C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
563$C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
564$C_3$              &      1.           &     1.              &      1.                &       1.           \\
565$F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
566\hline
567\hline
568\end{tabular}
569\caption{   \protect\label{tab:GLS} 
570  Set of predefined GLS parameters, or equivalently predefined turbulence models available with
571  \protect\key{zdfgls} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\ngn{namzdf\_gls}.}
572\end{center}   \end{table}
573%--------------------------------------------------------------------------------------------------------------
574
575In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of
576the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length) value near physical boundaries
577(logarithmic boundary layer law).
578$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{Galperin_al_JAS88},
579or by \citet{Kantha_Clayson_1994} or one of the two functions suggested by \citet{Canuto_2001}
580(\np{nn\_stab\_func}\forcode{ = 0..3}, resp.).
581The value of $C_{0\mu}$ depends of the choice of the stability function.
582
583The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or
584Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp.
585As for TKE closure, the wave effect on the mixing is considered when
586\np{ln\_crban}\forcode{ = .true.} \citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}.
587The \np{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and
588\np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.
589
590The $\psi$ equation is known to fail in stably stratified flows, and for this reason
591almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
592With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$.
593A value of $c_{lim} = 0.53$ is often used \citep{Galperin_al_JAS88}.
594\cite{Umlauf_Burchard_CSR05} show that the value of the clipping factor is of crucial importance for
595the entrainment depth predicted in stably stratified situations,
596and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes.
597The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.},
598and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value.
599
600The time and space discretization of the GLS equations follows the same energetic consideration as for
601the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{Burchard_OM02}.
602Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}.
603
604% -------------------------------------------------------------------------------------------------------------
605%        OSM OSMOSIS BL Scheme
606% -------------------------------------------------------------------------------------------------------------
607\subsection{OSM: OSMOSIS boundary layer scheme (\protect\key{zdfosm})}
608\label{subsec:ZDF_osm}
609
610%--------------------------------------------namzdf_osm---------------------------------------------------------
611
612\nlst{namzdf_osm}
613%--------------------------------------------------------------------------------------------------------------
614
615The OSMOSIS turbulent closure scheme is based on......   TBC
616
617% ================================================================
618% Convection
619% ================================================================
620\section{Convection}
621\label{sec:ZDF_conv}
622
623%--------------------------------------------namzdf--------------------------------------------------------
624
625\nlst{namzdf}
626%--------------------------------------------------------------------------------------------------------------
627
628Static instabilities (i.e. light potential densities under heavy ones) may occur at particular ocean grid points.
629In nature, convective processes quickly re-establish the static stability of the water column.
630These processes have been removed from the model via the hydrostatic assumption so they must be parameterized.
631Three parameterisations are available to deal with convective processes:
632a non-penetrative convective adjustment or an enhanced vertical diffusion,
633or/and the use of a turbulent closure scheme.
634
635% -------------------------------------------------------------------------------------------------------------
636%       Non-Penetrative Convective Adjustment
637% -------------------------------------------------------------------------------------------------------------
638\subsection[Non-penetrative convective adjmt (\protect\np{ln\_tranpc}\forcode{ = .true.})]
639            {Non-penetrative convective adjustment (\protect\np{ln\_tranpc}\forcode{ = .true.})}
640\label{subsec:ZDF_npc}
641
642%--------------------------------------------namzdf--------------------------------------------------------
643
644\nlst{namzdf}
645%--------------------------------------------------------------------------------------------------------------
646
647%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
648\begin{figure}[!htb]    \begin{center}
649\includegraphics[width=0.90\textwidth]{Fig_npc}
650\caption{  \protect\label{fig:npc} 
651  Example of an unstable density profile treated by the non penetrative convective adjustment algorithm.
652  $1^{st}$ step: the initial profile is checked from the surface to the bottom.
653  It is found to be unstable between levels 3 and 4.
654  They are mixed.
655  The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed.
656  The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed.
657  The $1^{st}$ step ends since the density profile is then stable below the level 3.
658  $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step:
659  levels 2 to 5 are mixed.
660  The new density profile is checked.
661  It is found stable: end of algorithm.}
662\end{center}   \end{figure}
663%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
664
665Options are defined through the \ngn{namzdf} namelist variables.
666The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}\forcode{ = .true.}.
667It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of
668the water column, but only until the density structure becomes neutrally stable
669($i.e.$ until the mixed portion of the water column has \textit{exactly} the density of the water just below)
670\citep{Madec_al_JPO91}.
671The associated algorithm is an iterative process used in the following way (\autoref{fig:npc}):
672starting from the top of the ocean, the first instability is found.
673Assume in the following that the instability is located between levels $k$ and $k+1$.
674The temperature and salinity in the two levels are vertically mixed, conserving the heat and salt contents of
675the water column.
676The new density is then computed by a linear approximation.
677If the new density profile is still unstable between levels $k+1$ and $k+2$,
678levels $k$, $k+1$ and $k+2$ are then mixed.
679This process is repeated until stability is established below the level $k$
680(the mixing process can go down to the ocean bottom).
681The algorithm is repeated to check if the density profile between level $k-1$ and $k$ is unstable and/or
682if there is no deeper instability.
683
684This algorithm is significantly different from mixing statically unstable levels two by two.
685The latter procedure cannot converge with a finite number of iterations for some vertical profiles while
686the algorithm used in \NEMO converges for any profile in a number of iterations which is less than
687the number of vertical levels.
688This property is of paramount importance as pointed out by \citet{Killworth1989}:
689it avoids the existence of permanent and unrealistic static instabilities at the sea surface.
690This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in
691the north-western Mediterranean Sea \citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}.
692
693The current implementation has been modified in order to deal with any non linear equation of seawater
694(L. Brodeau, personnal communication).
695Two main differences have been introduced compared to the original algorithm:
696$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency
697(not the the difference in potential density);
698$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in
699the same way their temperature and salinity has been mixed.
700These two modifications allow the algorithm to perform properly and accurately with TEOS10 or EOS-80 without
701having to recompute the expansion coefficients at each mixing iteration.
702
703% -------------------------------------------------------------------------------------------------------------
704%       Enhanced Vertical Diffusion
705% -------------------------------------------------------------------------------------------------------------
706\subsection{Enhanced vertical diffusion (\protect\np{ln\_zdfevd}\forcode{ = .true.})}
707\label{subsec:ZDF_evd}
708
709%--------------------------------------------namzdf--------------------------------------------------------
710
711\nlst{namzdf}
712%--------------------------------------------------------------------------------------------------------------
713
714Options are defined through the  \ngn{namzdf} namelist variables.
715The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}.
716In this case, the vertical eddy mixing coefficients are assigned very large values
717(a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable
718($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar_PhD97, Lazar_al_JPO99}.
719This is done either on tracers only (\np{nn\_evdm}\forcode{ = 0}) or
720on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}).
721
722In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np{nn\_evdm}\forcode{ = 1},
723the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to
724the namelist parameter \np{rn\_avevd}.
725A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$.
726This parameterisation of convective processes is less time consuming than
727the convective adjustment algorithm presented above when mixing both tracers and
728momentum in the case of static instabilities.
729It requires the use of an implicit time stepping on vertical diffusion terms
730(i.e. \np{ln\_zdfexp}\forcode{ = .false.}).
731
732Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$.
733This removes a potential source of divergence of odd and even time step in
734a leapfrog environment \citep{Leclair_PhD2010} (see \autoref{sec:STP_mLF}).
735
736% -------------------------------------------------------------------------------------------------------------
737%       Turbulent Closure Scheme
738% -------------------------------------------------------------------------------------------------------------
739\subsection[Turbulent closure scheme (\protect\key{zdf}\{tke,gls,osm\})]{Turbulent Closure Scheme (\protect\key{zdftke}, \protect\key{zdfgls} or \protect\key{zdfosm})}
740\label{subsec:ZDF_tcs}
741
742The turbulent closure scheme presented in \autoref{subsec:ZDF_tke} and \autoref{subsec:ZDF_gls}
743(\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically unstable density profiles.
744In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in
745\autoref{eq:zdftke_e} or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative.
746It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring $A_u^{vm} {and}\;A_v^{vm}$
747(up to $1\;m^2s^{-1}$).
748These large values restore the static stability of the water column in a way similar to that of
749the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}).
750However, in the vicinity of the sea surface (first ocean layer), the eddy coefficients computed by
751the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
752because the mixing length scale is bounded by the distance to the sea surface.
753It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme,
754$i.e.$ setting the \np{ln\_zdfnpc} namelist parameter to true and
755defining the turbulent closure CPP key all together.
756
757The KPP turbulent closure scheme already includes enhanced vertical diffusion in the case of convection,
758as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp},
759therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the KPP scheme.
760% gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
761
762% ================================================================
763% Double Diffusion Mixing
764% ================================================================
765\section{Double diffusion mixing (\protect\key{zdfddm})}
766\label{sec:ZDF_ddm}
767
768%-------------------------------------------namzdf_ddm-------------------------------------------------
769%
770%\nlst{namzdf_ddm}
771%--------------------------------------------------------------------------------------------------------------
772
773Options are defined through the  \ngn{namzdf\_ddm} namelist variables.
774Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa.
775The former condition leads to salt fingering and the latter to diffusive convection.
776Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean.
777\citet{Merryfield1999} include a parameterisation of such phenomena in a global ocean model and show that
778it leads to relatively minor changes in circulation but exerts significant regional influences on
779temperature and salinity.
780This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key.
781
782Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
783\begin{align*} % \label{eq:zdfddm_Kz}
784    &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT}  \\
785    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
786\end{align*}
787where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
788and $o$ by processes other than double diffusion.
789The rates of double-diffusive mixing depend on the buoyancy ratio
790$R_\rho = \alpha \partial_z T / \beta \partial_z S$, where $\alpha$ and $\beta$ are coefficients of
791thermal expansion and saline contraction (see \autoref{subsec:TRA_eos}).
792To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt
793(1981):
794\begin{align} \label{eq:zdfddm_f}
795A_f^{vS} &=    \begin{cases}
796   \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
797   0                              &\text{otherwise} 
798            \end{cases}   
799\\           \label{eq:zdfddm_f_T}
800A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
801\end{align}
802
803%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
804\begin{figure}[!t]   \begin{center}
805\includegraphics[width=0.99\textwidth]{Fig_zdfddm}
806\caption{  \protect\label{fig:zdfddm}
807  From \citet{Merryfield1999} :
808  (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in regions of salt fingering.
809  Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$;
810  (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in regions of diffusive convection.
811  Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation.
812  The latter is not implemented in \NEMO. }
813\end{center}    \end{figure}
814%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
815
816The factor 0.7 in \autoref{eq:zdfddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of
817buoyancy flux of heat to buoyancy flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}).
818Following  \citet{Merryfield1999}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
819
820To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by
821Federov (1988) is used:
822\begin{align}  \label{eq:zdfddm_d}
823A_d^{vT} &=    \begin{cases}
824   1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
825                           &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
826   0                       &\text{otherwise} 
827            \end{cases}   
828\\          \label{eq:zdfddm_d_S}
829A_d^{vS} &=    \begin{cases}
830   A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right)
831                           &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
832   A_d^{vT} \ 0.15 \ R_\rho
833                           &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
834   0                       &\text{otherwise} 
835            \end{cases}   
836\end{align}
837
838The dependencies of \autoref{eq:zdfddm_f} to \autoref{eq:zdfddm_d_S} on $R_\rho$ are illustrated in
839\autoref{fig:zdfddm}.
840Implementing this requires computing $R_\rho$ at each grid point on every time step.
841This is done in \mdl{eosbn2} at the same time as $N^2$ is computed.
842This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive).
843
844% ================================================================
845% Bottom Friction
846% ================================================================
847\section{Bottom and top friction (\protect\mdl{zdfbfr})}
848\label{sec:ZDF_bfr}
849
850%--------------------------------------------nambfr--------------------------------------------------------
851%
852%\nlst{nambfr}
853%--------------------------------------------------------------------------------------------------------------
854
855Options to define the top and bottom friction are defined through the \ngn{nambfr} namelist variables.
856The bottom friction represents the friction generated by the bathymetry.
857The top friction represents the friction generated by the ice shelf/ocean interface.
858As the friction processes at the top and bottom are treated in similar way,
859only the bottom friction is described in detail below.
860
861
862Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as
863a condition on the vertical diffusive flux.
864For the bottom boundary layer, one has:
865\begin{equation} \label{eq:zdfbfr_flux}
866A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
867\end{equation}
868where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside
869the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean).
870How ${\cal F}_h^{\textbf U}$ influences the interior depends on the vertical resolution of the model near
871the bottom relative to the Ekman layer depth.
872For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m,
873one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$
874(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$).
875With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
876When the vertical mixing coefficient is this small, using a flux condition is equivalent to
877entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or
878bottom model layer.
879To illustrate this, consider the equation for $u$ at $k$, the last ocean level:
880\begin{equation} \label{eq:zdfbfr_flux2}
881\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}}
882\end{equation}
883If the bottom layer thickness is 200~m, the Ekman transport will be distributed over that depth.
884On the other hand, if the vertical resolution is high (1~m or less) and a turbulent closure model is used,
885the turbulent Ekman layer will be represented explicitly by the model.
886However, the logarithmic layer is never represented in current primitive equation model applications:
887it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
888Two choices are available in \NEMO: a linear and a quadratic bottom friction.
889Note that in both cases, the rotation between the interior velocity and the bottom friction is neglected in
890the present release of \NEMO.
891
892In the code, the bottom friction is imposed by adding the trend due to the bottom friction to
893the general momentum trend in \mdl{dynbfr}.
894For the time-split surface pressure gradient algorithm, the momentum trend due to
895the barotropic component needs to be handled separately.
896For this purpose it is convenient to compute and store coefficients which can be simply combined with
897bottom velocities and geometric values to provide the momentum trend due to bottom friction.
898These coefficients are computed in \mdl{zdfbfr} and generally take the form $c_b^{\textbf U}$ where:
899\begin{equation} \label{eq:zdfbfr_bdef}
900\frac{\partial {\textbf U_h}}{\partial t} =
901  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
902\end{equation}
903where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
904
905% -------------------------------------------------------------------------------------------------------------
906%       Linear Bottom Friction
907% -------------------------------------------------------------------------------------------------------------
908\subsection{Linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 0..1})}
909\label{subsec:ZDF_bfr_linear}
910
911The linear bottom friction parameterisation (including the special case of a free-slip condition) assumes that
912the bottom friction is proportional to the interior velocity (i.e. the velocity of the last model level):
913\begin{equation} \label{eq:zdfbfr_linear}
914{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
915\end{equation}
916where $r$ is a friction coefficient expressed in ms$^{-1}$.
917This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,
918and setting $r = H / \tau$, where $H$ is the ocean depth.
919Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}.
920A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models.
921One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$
922(\citet{Gill1982}, Eq. 9.6.6).
923For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$,
924and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
925This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days.
926It can be changed by specifying \np{rn\_bfri1} (namelist parameter).
927
928For the linear friction case the coefficients defined in the general expression \autoref{eq:zdfbfr_bdef} are:
929\begin{equation} \label{eq:zdfbfr_linbfr_b}
930\begin{split}
931 c_b^u &= - r\\
932 c_b^v &= - r\\
933\end{split}
934\end{equation}
935When \np{nn\_botfr}\forcode{ = 1}, the value of $r$ used is \np{rn\_bfri1}.
936Setting \np{nn\_botfr}\forcode{ = 0} is equivalent to setting $r=0$ and
937leads to a free-slip bottom boundary condition.
938These values are assigned in \mdl{zdfbfr}.
939From v3.2 onwards there is support for local enhancement of these values via an externally defined 2D mask array
940(\np{ln\_bfr2d}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file.
941The mask values should vary from 0 to 1.
942Locations with a non-zero mask value will have the friction coefficient increased by
943$mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri1}.
944
945% -------------------------------------------------------------------------------------------------------------
946%       Non-Linear Bottom Friction
947% -------------------------------------------------------------------------------------------------------------
948\subsection{Non-linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 2})}
949\label{subsec:ZDF_bfr_nonlinear}
950
951The non-linear bottom friction parameterisation assumes that the bottom friction is quadratic:
952\begin{equation} \label{eq:zdfbfr_nonlinear}
953{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
954}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
955\end{equation}
956where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy due to tides,
957internal waves breaking and other short time scale currents.
958A typical value of the drag coefficient is $C_D = 10^{-3} $.
959As an example, the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and
960$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} uses $C_D = 1.4\;10^{-3}$ and
961$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
962The CME choices have been set as default values (\np{rn\_bfri2} and \np{rn\_bfeb2} namelist parameters).
963
964As for the linear case, the bottom friction is imposed in the code by adding the trend due to
965the bottom friction to the general momentum trend in \mdl{dynbfr}.
966For the non-linear friction case the terms computed in \mdl{zdfbfr} are:
967\begin{equation} \label{eq:zdfbfr_nonlinbfr}
968\begin{split}
969 c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\
970 c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\
971\end{split}
972\end{equation}
973
974The coefficients that control the strength of the non-linear bottom friction are initialised as namelist parameters:
975$C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}.
976Note for applications which treat tides explicitly a low or even zero value of \np{rn\_bfeb2} is recommended.
977From v3.2 onwards a local enhancement of $C_D$ is possible via an externally defined 2D mask array
978(\np{ln\_bfr2d}\forcode{ = .true.}).
979This works in the same way as for the linear bottom friction case with non-zero masked locations increased by
980$mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri2}.
981
982% -------------------------------------------------------------------------------------------------------------
983%       Bottom Friction Log-layer
984% -------------------------------------------------------------------------------------------------------------
985\subsection[Log-layer btm frict enhncmnt (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})]
986            {Log-layer bottom friction enhancement (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})}
987\label{subsec:ZDF_bfr_loglayer}
988
989In the non-linear bottom friction case, the drag coefficient, $C_D$, can be optionally enhanced using
990a "law of the wall" scaling.
991If  \np{ln\_loglayer} = .true., $C_D$ is no longer constant but is related to the thickness of
992the last wet layer in each column by:
993\begin{equation}
994C_D = \left ( {\kappa \over {\rm log}\left ( 0.5e_{3t}/rn\_bfrz0 \right ) } \right )^2
995\end{equation}
996
997\noindent where $\kappa$ is the von-Karman constant and \np{rn\_bfrz0} is a roughness length provided via
998the namelist.
999
1000For stability, the drag coefficient is bounded such that it is kept greater or equal to
1001the base \np{rn\_bfri2} value and it is not allowed to exceed the value of an additional namelist parameter:
1002\np{rn\_bfri2\_max}, i.e.:
1003\begin{equation}
1004rn\_bfri2 \leq C_D \leq rn\_bfri2\_max
1005\end{equation}
1006
1007\noindent Note also that a log-layer enhancement can also be applied to the top boundary friction if
1008under ice-shelf cavities are in use (\np{ln\_isfcav}\forcode{ = .true.}).
1009In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}.
1010
1011% -------------------------------------------------------------------------------------------------------------
1012%       Bottom Friction stability
1013% -------------------------------------------------------------------------------------------------------------
1014\subsection{Bottom friction stability considerations}
1015\label{subsec:ZDF_bfr_stability}
1016
1017Some care needs to exercised over the choice of parameters to ensure that the implementation of
1018bottom friction does not induce numerical instability.
1019For the purposes of stability analysis, an approximation to \autoref{eq:zdfbfr_flux2} is:
1020\begin{equation} \label{eq:Eqn_bfrstab}
1021\begin{split}
1022 \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
1023               &= -\frac{ru}{e_{3u}}\;2\rdt\\
1024\end{split}
1025\end{equation}
1026\noindent where linear bottom friction and a leapfrog timestep have been assumed.
1027To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have:
1028\begin{equation}
1029 |\Delta u| < \;|u|
1030\end{equation}
1031\noindent which, using \autoref{eq:Eqn_bfrstab}, gives:
1032\begin{equation}
1033r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
1034\end{equation}
1035This same inequality can also be derived in the non-linear bottom friction case if
1036a velocity of 1 m.s$^{-1}$ is assumed.
1037Alternatively, this criterion can be rearranged to suggest a minimum bottom box thickness to ensure stability:
1038\begin{equation}
1039e_{3u} > 2\;r\;\rdt
1040\end{equation}
1041\noindent which it may be necessary to impose if partial steps are being used.
1042For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m.
1043For most applications, with physically sensible parameters these restrictions should not be of concern.
1044But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.
1045To ensure stability limits are imposed on the bottom friction coefficients both
1046during initialisation and at each time step.
1047Checks at initialisation are made in \mdl{zdfbfr} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1048The number of breaches of the stability criterion are reported as well as
1049the minimum and maximum values that have been set.
1050The criterion is also checked at each time step, using the actual velocity, in \mdl{dynbfr}.
1051Values of the bottom friction coefficient are reduced as necessary to ensure stability;
1052these changes are not reported.
1053
1054Limits on the bottom friction coefficient are not imposed if the user has elected to
1055handle the bottom friction implicitly (see \autoref{subsec:ZDF_bfr_imp}).
1056The number of potential breaches of the explicit stability criterion are still reported for information purposes.
1057
1058% -------------------------------------------------------------------------------------------------------------
1059%       Implicit Bottom Friction
1060% -------------------------------------------------------------------------------------------------------------
1061\subsection{Implicit bottom friction (\protect\np{ln\_bfrimp}\forcode{ = .true.})}
1062\label{subsec:ZDF_bfr_imp}
1063
1064An optional implicit form of bottom friction has been implemented to improve model stability.
1065We recommend this option for shelf sea and coastal ocean applications, especially for split-explicit time splitting.
1066This option can be invoked by setting \np{ln\_bfrimp} to \forcode{.true.} in the \textit{nambfr} namelist.
1067This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \textit{namzdf} namelist.
1068
1069This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp},
1070the bottom boundary condition is implemented implicitly.
1071
1072\begin{equation} \label{eq:dynzdf_bfr}
1073\left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{mbk}
1074    = \binom{c_{b}^{u}u^{n+1}_{mbk}}{c_{b}^{v}v^{n+1}_{mbk}}
1075\end{equation}
1076
1077where $mbk$ is the layer number of the bottom wet layer.
1078Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so, it is implicit.
1079
1080If split-explicit time splitting is used, care must be taken to avoid the double counting of the bottom friction in
1081the 2-D barotropic momentum equations.
1082As NEMO only updates the barotropic pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation,
1083we need to remove the bottom friction induced by these two terms which has been included in the 3-D momentum trend
1084and update it with the latest value.
1085On the other hand, the bottom friction contributed by the other terms
1086(e.g. the advection term, viscosity term) has been included in the 3-D momentum equations and
1087should not be added in the 2-D barotropic mode.
1088
1089The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the following:
1090
1091\begin{equation} \label{eq:dynspg_ts_bfr1}
1092\frac{\textbf{U}_{med}-\textbf{U}^{m-1}}{2\Delta t}=-g\nabla\eta-f\textbf{k}\times\textbf{U}^{m}+c_{b}
1093\left(\textbf{U}_{med}-\textbf{U}^{m-1}\right)
1094\end{equation}
1095\begin{equation} \label{eq:dynspg_ts_bfr2}
1096\frac{\textbf{U}^{m+1}-\textbf{U}_{med}}{2\Delta t}=\textbf{T}+
1097\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{U}^{'}\right)-
10982\Delta t_{bc}c_{b}\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{u}_{b}\right)
1099\end{equation}
1100
1101where $\textbf{T}$ is the vertical integrated 3-D momentum trend.
1102We assume the leap-frog time-stepping is used here.
1103$\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step.
1104$c_{b}$ is the friction coefficient.
1105$\eta$ is the sea surface level calculated in the barotropic loops while $\eta^{'}$ is the sea surface level used in
1106the 3-D baroclinic mode.
1107$\textbf{u}_{b}$ is the bottom layer horizontal velocity.
1108
1109
1110
1111
1112% -------------------------------------------------------------------------------------------------------------
1113%       Bottom Friction with split-explicit time splitting
1114% -------------------------------------------------------------------------------------------------------------
1115\subsection[Bottom friction w/ split-explicit time splitting (\protect\np{ln\_bfrimp})]
1116            {Bottom friction with split-explicit time splitting (\protect\np{ln\_bfrimp})}
1117\label{subsec:ZDF_bfr_ts}
1118
1119When calculating the momentum trend due to bottom friction in \mdl{dynbfr},
1120the bottom velocity at the before time step is used.
1121This velocity includes both the baroclinic and barotropic components which is appropriate when
1122using either the explicit or filtered surface pressure gradient algorithms
1123(\key{dynspg\_exp} or \key{dynspg\_flt}).
1124Extra attention is required, however, when using split-explicit time stepping (\key{dynspg\_ts}).
1125In this case the free surface equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro},
1126while the three dimensional prognostic variables are solved with the longer time step of \np{rn\_rdt} seconds.
1127The trend in the barotropic momentum due to bottom friction appropriate to this method is that given by
1128the selected parameterisation ($i.e.$ linear or non-linear bottom friction) computed with
1129the evolving velocities at each barotropic timestep.
1130
1131In the case of non-linear bottom friction, we have elected to partially linearise the problem by
1132keeping the coefficients fixed throughout the barotropic time-stepping to those computed in
1133\mdl{zdfbfr} using the now timestep.
1134This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to:
1135
1136\begin{enumerate}
1137\item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before barotropic velocity to
1138  the bottom friction component of the vertically integrated momentum trend.
1139  Note the same stability check that is carried out on the bottom friction coefficient in \mdl{dynbfr} has to
1140  be applied here to ensure that the trend removed matches that which was added in \mdl{dynbfr}.
1141\item At each barotropic step, compute the contribution of the current barotropic velocity to
1142  the trend due to bottom friction.
1143  Add this contribution to the vertically integrated momentum trend.
1144  This contribution is handled implicitly which eliminates the need to impose a stability criteria on
1145  the values of the bottom friction coefficient within the barotropic loop.
1146\end{enumerate}
1147
1148Note that the use of an implicit formulation within the barotropic loop for the bottom friction trend means that
1149any limiting of the bottom friction coefficient in \mdl{dynbfr} does not adversely affect the solution when
1150using split-explicit time splitting.
1151This is because the major contribution to bottom friction is likely to come from the barotropic component which
1152uses the unrestricted value of the coefficient.
1153However, if the limiting is thought to be having a major effect
1154(a more likely prospect in coastal and shelf seas applications) then
1155the fully implicit form of the bottom friction should be used (see \autoref{subsec:ZDF_bfr_imp})
1156which can be selected by setting \np{ln\_bfrimp} $=$ \forcode{.true.}.
1157
1158Otherwise, the implicit formulation takes the form:
1159\begin{equation} \label{eq:zdfbfr_implicitts}
1160 \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ] 
1161\end{equation}
1162where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),
1163$c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and
1164$RHS$ represents all the components to the vertically integrated momentum trend except for
1165that due to bottom friction.
1166
1167
1168
1169
1170% ================================================================
1171% Tidal Mixing
1172% ================================================================
1173\section{Tidal mixing (\protect\key{zdftmx})}
1174\label{sec:ZDF_tmx}
1175
1176%--------------------------------------------namzdf_tmx--------------------------------------------------
1177%
1178%\nlst{namzdf_tmx}
1179%--------------------------------------------------------------------------------------------------------------
1180
1181
1182% -------------------------------------------------------------------------------------------------------------
1183%        Bottom intensified tidal mixing
1184% -------------------------------------------------------------------------------------------------------------
1185\subsection{Bottom intensified tidal mixing}
1186\label{subsec:ZDF_tmx_bottom}
1187
1188Options are defined through the  \ngn{namzdf\_tmx} namelist variables.
1189The parameterization of tidal mixing follows the general formulation for the vertical eddy diffusivity proposed by
1190\citet{St_Laurent_al_GRL02} and first introduced in an OGCM by \citep{Simmons_al_OM04}.
1191In this formulation an additional vertical diffusivity resulting from internal tide breaking,
1192$A^{vT}_{tides}$ is expressed as a function of $E(x,y)$,
1193the energy transfer from barotropic tides to baroclinic tides:
1194\begin{equation} \label{eq:Ktides}
1195A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 }
1196\end{equation}
1197where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
1198$\rho$ the density, $q$ the tidal dissipation efficiency, and $F(z)$ the vertical structure function.
1199
1200The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) and
1201is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).
1202The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter)
1203represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally,
1204with the remaining $1-q$ radiating away as low mode internal waves and
1205contributing to the background internal wave field.
1206A value of $q=1/3$ is typically used \citet{St_Laurent_al_GRL02}.
1207The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical.
1208It is implemented as a simple exponential decaying upward away from the bottom,
1209with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter,
1210with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},
1211\begin{equation} \label{eq:Fz}
1212F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) }
1213\end{equation}
1214and is normalized so that vertical integral over the water column is unity.
1215
1216The associated vertical viscosity is calculated from the vertical diffusivity assuming a Prandtl number of 1,
1217$i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$.
1218In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity is capped at $300\,cm^2/s$ and
1219impose a lower limit on $N^2$ of \np{rn\_n2min} usually set to $10^{-8} s^{-2}$.
1220These bounds are usually rarely encountered.
1221
1222The internal wave energy map, $E(x, y)$ in \autoref{eq:Ktides}, is derived from a barotropic model of
1223the tides utilizing a parameterization of the conversion of barotropic tidal energy into internal waves.
1224The essential goal of the parameterization is to represent the momentum exchange between the barotropic tides and
1225the unrepresented internal waves induced by the tidal flow over rough topography in a stratified ocean.
1226In the current version of \NEMO, the map is built from the output of
1227the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}.
1228This model provides the dissipation associated with internal wave energy for the M2 and K1 tides component
1229(\autoref{fig:ZDF_M2_K1_tmx}).
1230The S2 dissipation is simply approximated as being $1/4$ of the M2 one.
1231The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$.
1232Its global mean value is $1.1$ TW,
1233in agreement with independent estimates \citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.
1234
1235%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1236\begin{figure}[!t]   \begin{center}
1237\includegraphics[width=0.90\textwidth]{Fig_ZDF_M2_K1_tmx}
1238\caption{  \protect\label{fig:ZDF_M2_K1_tmx} 
1239(a) M2 and (b) K1 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). }
1240\end{center}   \end{figure}
1241%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
1242 
1243% -------------------------------------------------------------------------------------------------------------
1244%        Indonesian area specific treatment
1245% -------------------------------------------------------------------------------------------------------------
1246\subsection{Indonesian area specific treatment (\protect\np{ln\_zdftmx\_itf})}
1247\label{subsec:ZDF_tmx_itf}
1248
1249When the Indonesian Through Flow (ITF) area is included in the model domain,
1250a specific treatment of tidal induced mixing in this area can be used.
1251It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide an input NetCDF file,
1252\ifile{mask\_itf}, which contains a mask array defining the ITF area where the specific treatment is applied.
1253
1254When \np{ln\_tmx\_itf}\forcode{ = .true.}, the two key parameters $q$ and $F(z)$ are adjusted following
1255the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}:
1256
1257First, the Indonesian archipelago is a complex geographic region with a series of
1258large, deep, semi-enclosed basins connected via numerous narrow straits.
1259Once generated, internal tides remain confined within this semi-enclosed area and hardly radiate away.
1260Therefore all the internal tides energy is consumed within this area.
1261So it is assumed that $q = 1$, $i.e.$ all the energy generated is available for mixing.
1262Note that for test purposed, the ITF tidal dissipation efficiency is a namelist parameter (\np{rn\_tfe\_itf}).
1263A value of $1$ or close to is this recommended for this parameter.
1264
1265Second, the vertical structure function, $F(z)$, is no more associated with a bottom intensification of the mixing,
1266but with a maximum of energy available within the thermocline.
1267\citet{Koch-Larrouy_al_GRL07} have suggested that the vertical distribution of
1268the energy dissipation proportional to $N^2$ below the core of the thermocline and to $N$ above.
1269The resulting $F(z)$ is:
1270\begin{equation} \label{eq:Fz_itf}
1271F(i,j,k) \sim     \left\{ \begin{aligned}
1272\frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\
1273\frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$}
1274                      \end{aligned} \right.
1275\end{equation}
1276
1277Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,
1278which agrees with the independent estimates inferred from observations.
1279Introduced in a regional OGCM, the parameterization improves the water mass characteristics in
1280the different Indonesian seas, suggesting that the horizontal and vertical distributions of
1281the mixing are adequately prescribed \citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}.
1282Note also that such a parameterisation has a significant impact on the behaviour of
1283global coupled GCMs \citep{Koch-Larrouy_al_CD10}.
1284
1285
1286% ================================================================
1287% Internal wave-driven mixing
1288% ================================================================
1289\section{Internal wave-driven mixing (\protect\key{zdftmx\_new})}
1290\label{sec:ZDF_tmx_new}
1291
1292%--------------------------------------------namzdf_tmx_new------------------------------------------
1293%
1294%\nlst{namzdf_tmx_new}
1295%--------------------------------------------------------------------------------------------------------------
1296
1297The parameterization of mixing induced by breaking internal waves is a generalization of
1298the approach originally proposed by \citet{St_Laurent_al_GRL02}.
1299A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed,
1300and the resulting diffusivity is obtained as
1301\begin{equation} \label{eq:Kwave}
1302A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 }
1303\end{equation}
1304where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of
1305the energy available for mixing.
1306If the \np{ln\_mevar} namelist parameter is set to false, the mixing efficiency is taken as constant and
1307equal to 1/6 \citep{Osborn_JPO80}.
1308In the opposite (recommended) case, $R_f$ is instead a function of
1309the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$,
1310with $\nu$ the molecular viscosity of seawater, following the model of \cite{Bouffard_Boegman_DAO2013} and
1311the implementation of \cite{de_lavergne_JPO2016_efficiency}.
1312Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when
1313the mixing efficiency is constant.
1314
1315In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary
1316as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to true, a recommended choice.
1317This parameterization of differential mixing, due to \cite{Jackson_Rehmann_JPO2014},
1318is implemented as in \cite{de_lavergne_JPO2016_efficiency}.
1319
1320The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$,
1321is constructed from three static maps of column-integrated internal wave energy dissipation,
1322$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures
1323(de Lavergne et al., in prep):
1324\begin{align*}
1325F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\
1326F_{pyc}(i,j,k) &\propto N^{n\_p}\\
1327F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} }
1328\end{align*} 
1329In the above formula, $h_{ab}$ denotes the height above bottom,
1330$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by
1331\begin{equation*}
1332h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; ,
1333\end{equation*}
1334The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_tmx\_new} namelist)
1335controls the stratification-dependence of the pycnocline-intensified dissipation.
1336It can take values of 1 (recommended) or 2.
1337Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of
1338the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps.
1339$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and
1340$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of
1341the abyssal hill topography \citep{Goff_JGR2010} and the latitude.
1342
1343% ================================================================
1344
1345
1346
1347\end{document}
Note: See TracBrowser for help on using the repository browser.