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 @ 11597

Last change on this file since 11597 was 11597, checked in by nicolasmartin, 5 years ago

Continuation of coding rules application
Recovery of some sections deleted by the previous commit

File size: 83.1 KB
Line 
1\documentclass[../main/NEMO_manual]{subfiles}
2
3%% Custom aliases
4\newcommand{\cf}{\ensuremath{C\kern-0.14em f}}
5
6\begin{document}
7\chapter{Vertical Ocean Physics (ZDF)}
8\label{chap:ZDF}
9
10\chaptertoc
11
12%gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN.
13
14%% =================================================================================================
15\section{Vertical mixing}
16\label{sec:ZDF}
17
18The discrete form of the ocean subgrid scale physics has been presented in
19\autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}.
20At the surface and bottom boundaries, the turbulent fluxes of momentum, heat and salt have to be defined.
21At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}),
22while at the bottom they are set to zero for heat and salt,
23unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie\ \np{ln_trabbc}{ln\_trabbc} defined,
24see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum
25(see \autoref{sec:ZDF_drg}).
26
27In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and
28diffusivity coefficients, $A_u^{vm}$ , $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$- points,
29respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}).
30These coefficients can be assumed to be either constant, or a function of the local Richardson number,
31or computed from a turbulent closure model (either TKE or GLS or OSMOSIS formulation).
32The computation of these coefficients is initialized in the \mdl{zdfphy} module and performed in
33the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} or \mdl{zdfosm} modules.
34The trends due to the vertical momentum and tracer diffusion, including the surface forcing,
35are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.
36%These trends can be computed using either a forward time stepping scheme
37%(namelist parameter \np[=.true.]{ln_zdfexp}{ln\_zdfexp}) or a backward time stepping scheme
38%(\np[=.false.]{ln_zdfexp}{ln\_zdfexp}) depending on the magnitude of the mixing coefficients,
39%and thus of the formulation used (see \autoref{chap:TD}).
40
41
42\begin{listing}
43  \nlst{namzdf}
44  \caption{\forcode{&namzdf}}
45  \label{lst:namzdf}
46\end{listing}
47
48%% =================================================================================================
49\subsection[Constant (\forcode{ln_zdfcst})]{Constant (\protect\np{ln_zdfcst}{ln\_zdfcst})}
50\label{subsec:ZDF_cst}
51
52Options are defined through the \nam{zdf}{zdf} namelist variables.
53When \np{ln_zdfcst}{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to
54constant values over the whole ocean.
55This is the crudest way to define the vertical ocean physics.
56It is recommended to use this option only in process studies, not in basin scale simulations.
57Typical values used in this case are:
58\begin{align*}
59  A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}   \\
60  A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
61\end{align*}
62
63These values are set through the \np{rn_avm0}{rn\_avm0} and \np{rn_avt0}{rn\_avt0} namelist parameters.
64In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity,
65that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and
66$\sim10^{-9}~m^2.s^{-1}$ for salinity.
67
68%% =================================================================================================
69\subsection[Richardson number dependent (\forcode{ln_zdfric})]{Richardson number dependent (\protect\np{ln_zdfric}{ln\_zdfric})}
70\label{subsec:ZDF_ric}
71
72
73\begin{listing}
74  \nlst{namzdf_ric}
75  \caption{\forcode{&namzdf_ric}}
76  \label{lst:namzdf_ric}
77\end{listing}
78
79When \np[=.true.]{ln_zdfric}{ln\_zdfric}, a local Richardson number dependent formulation for the vertical momentum and
80tracer eddy coefficients is set through the \nam{zdf_ric}{zdf\_ric} namelist variables.
81The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.
82\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures.
83The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to
84a dependency between the vertical eddy coefficients and the local Richardson number
85(\ie\ the ratio of stratification to vertical shear).
86Following \citet{pacanowski.philander_JPO81}, the following formulation has been implemented:
87\[
88  % \label{eq:ZDF_ric}
89  \left\{
90    \begin{aligned}
91      A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
92      A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
93    \end{aligned}
94  \right.
95\]
96where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number,
97$N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
98$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the constant case
99(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that
100can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$.
101The last three values can be modified by setting the \np{rn_avmri}{rn\_avmri}, \np{rn_alp}{rn\_alp} and
102\np{nn_ric}{nn\_ric} namelist parameters, respectively.
103
104A simple mixing-layer model to transfer and dissipate the atmospheric forcings
105(wind-stress and buoyancy fluxes) can be activated setting the \np[=.true.]{ln_mldw}{ln\_mldw} in the namelist.
106
107In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and
108the vertical eddy coefficients prescribed within this layer.
109
110This depth is assumed proportional to the "depth of frictional influence" that is limited by rotation:
111\[
112  h_{e} = Ek \frac {u^{*}} {f_{0}}
113\]
114where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis parameter.
115
116In this similarity height relationship, the turbulent friction velocity:
117\[
118  u^{*} = \sqrt \frac {|\tau|} {\rho_o}
119\]
120is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$.
121The final $h_{e}$ is further constrained by the adjustable bounds \np{rn_mldmin}{rn\_mldmin} and \np{rn_mldmax}{rn\_mldmax}.
122Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to
123the empirical values \np{rn_wtmix}{rn\_wtmix} and \np{rn_wvmix}{rn\_wvmix} \citep{lermusiaux_JMS01}.
124
125%% =================================================================================================
126\subsection[TKE turbulent closure scheme (\forcode{ln_zdftke})]{TKE turbulent closure scheme (\protect\np{ln_zdftke}{ln\_zdftke})}
127\label{subsec:ZDF_tke}
128
129\begin{listing}
130  \nlst{namzdf_tke}
131  \caption{\forcode{&namzdf_tke}}
132  \label{lst:namzdf_tke}
133\end{listing}
134
135The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on
136a prognostic equation for $\bar{e}$, the turbulent kinetic energy,
137and a closure assumption for the turbulent length scales.
138This turbulent closure model has been developed by \citet{bougeault.lacarrere_MWR89} in the atmospheric case,
139adapted by \citet{gaspar.gregoris.ea_JGR90} for the oceanic case, and embedded in OPA, the ancestor of \NEMO,
140by \citet{blanke.delecluse_JPO93} for equatorial Atlantic simulations.
141Since then, significant modifications have been introduced by \citet{madec.delecluse.ea_NPM98} in both the implementation and
142the formulation of the mixing length scale.
143The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear,
144its destruction through stratification, its vertical diffusion, and its dissipation of \citet{kolmogorov_IANS42} type:
145\begin{equation}
146  \label{eq:ZDF_tke_e}
147  \frac{\partial \bar{e}}{\partial t} =
148  \frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
149      +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
150  -K_\rho\,N^2
151  +\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
152      \;\frac{\partial \bar{e}}{\partial k}} \right]
153  - c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
154\end{equation}
155\[
156  % \label{eq:ZDF_tke_kz}
157  \begin{split}
158    K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }    \\
159    K_\rho &= A^{vm} / P_{rt}
160  \end{split}
161\]
162where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
163$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,
164$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity and diffusivity coefficients.
165The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with
166vertical mixing at any depth \citep{gaspar.gregoris.ea_JGR90}.
167They are set through namelist parameters \np{nn_ediff}{nn\_ediff} and \np{nn_ediss}{nn\_ediss}.
168$P_{rt}$ can be set to unity or, following \citet{blanke.delecluse_JPO93}, be a function of the local Richardson number, $R_i$:
169\begin{align*}
170  % \label{eq:ZDF_prt}
171  P_{rt} =
172  \begin{cases}
173    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}   \\
174    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}   \\
175    \ \ 10 &      \text{if $\ 2 \leq R_i$}
176  \end{cases}
177\end{align*}
178The choice of $P_{rt}$ is controlled by the \np{nn_pdl}{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}{rn\_ebb} namelist parameter.
182The default value of $e_{bb}$ is 3.75. \citep{gaspar.gregoris.ea_JGR90}), 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}{rn\_emin} namelist parameter).
188Following \citet{gaspar.gregoris.ea_JGR90}, 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{gargett_JMR84} 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}{rn\_avm0} and
194\np{rn_avt0}{rn\_avt0} (\nam{zdf}{zdf} namelist, see \autoref{subsec:ZDF_cst}).
195
196%% =================================================================================================
197\subsubsection{Turbulent length scale}
198
199For computational efficiency, the original formulation of the turbulent length scales proposed by
200\citet{gaspar.gregoris.ea_JGR90} has been simplified.
201Four formulations are proposed, the choice of which is controlled by the \np{nn_mxl}{nn\_mxl} namelist parameter.
202The first two are based on the following first order approximation \citep{blanke.delecluse_JPO93}:
203\begin{equation}
204  \label{eq:ZDF_tke_mxl0_1}
205  l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N
206\end{equation}
207which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency.
208The resulting length scale is bounded by the distance to the surface or to the bottom
209(\np[=0]{nn_mxl}{nn\_mxl}) or by the local vertical scale factor (\np[=1]{nn_mxl}{nn\_mxl}).
210\citet{blanke.delecluse_JPO93} notice that this simplification has two major drawbacks:
211it makes no sense for locally unstable stratification and the computation no longer uses all
212the information contained in the vertical density profile.
213To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np[=2, 3]{nn_mxl}{nn\_mxl} cases,
214which add an extra assumption concerning the vertical gradient of the computed length scale.
215So, the length scales are first evaluated as in \autoref{eq:ZDF_tke_mxl0_1} and then bounded such that:
216\begin{equation}
217  \label{eq:ZDF_tke_mxl_constraint}
218  \frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
219  \qquad \text{with }\  l =  l_k = l_\epsilon
220\end{equation}
221\autoref{eq:ZDF_tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than
222the variations of depth.
223It provides a better approximation of the \citet{gaspar.gregoris.ea_JGR90} formulation while being much less
224time consuming.
225In particular, it allows the length scale to be limited not only by the distance to the surface or
226to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as
227the thermocline (\autoref{fig:ZDF_mixing_length}).
228In order to impose the \autoref{eq:ZDF_tke_mxl_constraint} constraint, we introduce two additional length scales:
229$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and
230evaluate the dissipation and mixing length scales as
231(and note that here we use numerical indexing):
232\begin{figure}[!t]
233  \centering
234  \includegraphics[width=0.66\textwidth]{Fig_mixing_length}
235  \caption[Mixing length computation]{Illustration of the mixing length computation}
236  \label{fig:ZDF_mixing_length}
237\end{figure}
238\[
239  % \label{eq:ZDF_tke_mxl2}
240  \begin{aligned}
241    l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right)
242    \quad &\text{ from $k=1$ to $jpk$ }\ \\
243    l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)\right)
244    \quad &\text{ from $k=jpk$ to $1$ }\ \\
245  \end{aligned}
246\]
247where $l^{(k)}$ is computed using \autoref{eq:ZDF_tke_mxl0_1}, \ie\ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$.
248
249In the \np[=2]{nn_mxl}{nn\_mxl} case, the dissipation and mixing length scales take the same value:
250$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np[=3]{nn_mxl}{nn\_mxl} case,
251the dissipation and mixing turbulent length scales are give as in \citet{gaspar.gregoris.ea_JGR90}:
252\[
253  % \label{eq:ZDF_tke_mxl_gaspar}
254  \begin{aligned}
255    & l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }   \\
256    & l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
257  \end{aligned}
258\]
259
260At the ocean surface, a non zero length scale is set through the  \np{rn_mxl0}{rn\_mxl0} namelist parameter.
261Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and
262$z_o$ the roughness parameter of the surface.
263Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn_mxl0}{rn\_mxl0}.
264In the ocean interior a minimum length scale is set to recover the molecular viscosity when
265$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ).
266
267%% =================================================================================================
268\subsubsection{Surface wave breaking parameterization}
269
270Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to
271include the effect of surface wave breaking energetics.
272This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow.
273The \citet{mellor.blumberg_JPO04} modifications acts on surface length scale and TKE values and
274air-sea drag coefficient.
275The latter concerns the bulk formulae and is not discussed here.
276
277Following \citet{craig.banner_JPO94}, the boundary condition on surface TKE value is :
278\begin{equation}
279  \label{eq:ZDF_Esbc}
280  \bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
281\end{equation}
282where $\alpha_{CB}$ is the \citet{craig.banner_JPO94} constant of proportionality which depends on the ''wave age'',
283ranging from 57 for mature waves to 146 for younger waves \citep{mellor.blumberg_JPO04}.
284The boundary condition on the turbulent length scale follows the Charnock's relation:
285\begin{equation}
286  \label{eq:ZDF_Lsbc}
287  l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
288\end{equation}
289where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant.
290\citet{mellor.blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by
291\citet{stacey_JPO99} citing observation evidence, and
292$\alpha_{CB} = 100$ the Craig and Banner's value.
293As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,
294with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter, setting \np[=67.83]{rn_ebb}{rn\_ebb} corresponds
295to $\alpha_{CB} = 100$.
296Further setting  \np[=.true.]{ln_mxl0}{ln\_mxl0},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale,
297with $\beta$ hard coded to the Stacey's value.
298Note that a minimal threshold of \np{rn_emin0}{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the
299surface $\bar{e}$ value.
300
301%% =================================================================================================
302\subsubsection{Langmuir cells}
303
304Langmuir circulations (LC) can be described as ordered large-scale vertical motions in
305the surface layer of the oceans.
306Although LC have nothing to do with convection, the circulation pattern is rather similar to
307so-called convective rolls in the atmospheric boundary layer.
308The detailed physics behind LC is described in, for example, \citet{craik.leibovich_JFM76}.
309The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and
310wind drift currents.
311
312Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by
313\citep{axell_JGR02} for a $k-\epsilon$ turbulent closure.
314The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in
315an extra source term of TKE, $P_{LC}$.
316The presence of $P_{LC}$ in \autoref{eq:ZDF_tke_e}, the TKE equation, is controlled by setting \np{ln_lc}{ln\_lc} to
317\forcode{.true.} in the \nam{zdf_tke}{zdf\_tke} namelist.
318
319By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}),
320$P_{LC}$ is assumed to be :
321\[
322P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
323\]
324where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
325With no information about the wave field, $w_{LC}$ is assumed to be proportional to
326the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
327\footnote{Following \citet{li.garrett_JMR93}, the surface Stoke drift velocity may be expressed as
328  $u_s =  0.016 \,|U_{10m}|$.
329  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of
330  $1.5~10^{-3}$ give the expression used of $u_s$ as a function of the module of surface stress
331}.
332For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at
333a finite depth $H_{LC}$ (which is often close to the mixed layer depth),
334and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).
335The resulting expression for $w_{LC}$ is :
336\[
337  w_{LC}  =
338  \begin{cases}
339    c_{LC} \,u_s \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
340    0                             &      \text{otherwise}
341  \end{cases}
342\]
343where $c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data.
344The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second.
345The value of $c_{LC}$ is set through the \np{rn_lc}{rn\_lc} namelist parameter,
346having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}.
347
348The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
349$H_{LC}$ is the depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by
350converting its kinetic energy to potential energy, according to
351\[
352- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} u_s^2
353\]
354
355%% =================================================================================================
356\subsubsection{Mixing just below the mixed layer}
357
358Vertical mixing parameterizations commonly used in ocean general circulation models tend to
359produce mixed-layer depths that are too shallow during summer months and windy conditions.
360This bias is particularly acute over the Southern Ocean.
361To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{rodgers.aumont.ea_B14}.
362The parameterization is an empirical one, \ie\ not derived from theoretical considerations,
363but rather is meant to account for observed processes that affect the density structure of
364the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme
365(\ie\ near-inertial oscillations and ocean swells and waves).
366
367When using this parameterization (\ie\ when \np[=1]{nn_etau}{nn\_etau}),
368the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations,
369swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition,
370plus a depth depend one given by:
371\begin{equation}
372  \label{eq:ZDF_Ehtau}
373  S = (1-f_i) \; f_r \; e_s \; e^{-z / h_\tau}
374\end{equation}
375where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that
376penetrates in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of
377the penetration, and $f_i$ is the ice concentration
378(no penetration if $f_i=1$, \ie\ if the ocean is entirely covered by sea-ice).
379The value of $f_r$, usually a few percents, is specified through \np{rn_efr}{rn\_efr} namelist parameter.
380The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np[=0]{nn_etau}{nn\_etau}) or
381a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes
382(\np[=1]{nn_etau}{nn\_etau}).
383
384Note that two other option exist, \np[=2, 3]{nn_etau}{nn\_etau}.
385They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer,
386or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean.
387Those two options are obsolescent features introduced for test purposes.
388They will be removed in the next release.
389
390% This should be explain better below what this rn_eice parameter is meant for:
391In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn_eice}{rn\_eice} namelist parameter.
392This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration
393is greater than 25\%.
394
395% from Burchard et al OM 2008 :
396% the most critical process not reproduced by statistical turbulence models is the activity of
397% internal waves and their interaction with turbulence. After the Reynolds decomposition,
398% internal waves are in principle included in the RANS equations, but later partially
399% excluded by the hydrostatic assumption and the model resolution.
400% Thus far, the representation of internal wave mixing in ocean models has been relatively crude
401% (\eg\ Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002).
402
403%% =================================================================================================
404\subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls})]{GLS: Generic Length Scale (\protect\np{ln_zdfgls}{ln\_zdfgls})}
405\label{subsec:ZDF_gls}
406
407
408\begin{listing}
409  \nlst{namzdf_gls}
410  \caption{\forcode{&namzdf_gls}}
411  \label{lst:namzdf_gls}
412\end{listing}
413
414The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations:
415one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale,
416$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}.
417This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
418where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:ZDF_GLS} allows to recover a number of
419well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87},
420$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}).
421The GLS scheme is given by the following set of equations:
422\begin{equation}
423  \label{eq:ZDF_gls_e}
424  \frac{\partial \bar{e}}{\partial t} =
425  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
426      +\left( \frac{\partial v}{\partial k} \right)^2} \right]
427  -K_\rho \,N^2
428  +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
429  - \epsilon
430\end{equation}
431
432\[
433  % \label{eq:ZDF_gls_psi}
434  \begin{split}
435    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
436      \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
437          +\left( \frac{\partial v}{\partial k} \right)^2} \right]
438      - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
439    &+\frac{1}{e_3\;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
440        \;\frac{\partial \psi}{\partial k}} \right]\;
441  \end{split}
442\]
443
444\[
445  % \label{eq:ZDF_gls_kz}
446  \begin{split}
447    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
448    K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
449  \end{split}
450\]
451
452\[
453  % \label{eq:ZDF_gls_eps}
454  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
455\]
456where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and
457$\epsilon$ the dissipation rate.
458The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of
459the choice of the turbulence model.
460Four different turbulent models are pre-defined (\autoref{tab:ZDF_GLS}).
461They are made available through the \np{nn_clo}{nn\_clo} namelist parameter.
462
463\begin{table}[htbp]
464  \centering
465  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
466  \begin{tabular}{ccccc}
467    &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\
468    % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\
469    \hline
470    \hline
471    \np{nn_clo}{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\
472    \hline
473    $( p , n , m )$         &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
474    $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
475    $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
476    $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
477    $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
478    $C_3$              &      1.           &     1.              &      1.                &       1.           \\
479    $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
480    \hline
481    \hline
482  \end{tabular}
483  \caption[Set of predefined GLS parameters or equivalently predefined turbulence models available]{
484    Set of predefined GLS parameters, or equivalently predefined turbulence models available with
485    \protect\np[=.true.]{ln_zdfgls}{ln\_zdfgls} and controlled by
486    the \protect\np{nn_clos}{nn\_clos} namelist variable in \protect\nam{zdf_gls}{zdf\_gls}.}
487  \label{tab:ZDF_GLS}
488\end{table}
489
490In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of
491the mixing length towards $\kappa z_b$ ($\kappa$ is the Von Karman constant and $z_b$ the rugosity length scale) value near physical boundaries
492(logarithmic boundary layer law).
493$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88},
494or by \citet{kantha.clayson_JGR94} or one of the two functions suggested by \citet{canuto.howard.ea_JPO01}
495(\np[=0, 3]{nn_stab_func}{nn\_stab\_func}, resp.).
496The value of $C_{0\mu}$ depends on the choice of the stability function.
497
498The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or
499Neumann condition through \np{nn_bc_surf}{nn\_bc\_surf} and \np{nn_bc_bot}{nn\_bc\_bot}, resp.
500As for TKE closure, the wave effect on the mixing is considered when
501\np[ > 0.]{rn_crban}{rn\_crban} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}.
502The \np{rn_crban}{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and
503\np{rn_charn}{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.
504
505The $\psi$ equation is known to fail in stably stratified flows, and for this reason
506almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
507With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$.
508A value of $c_{lim} = 0.53$ is often used \citep{galperin.kantha.ea_JAS88}.
509\cite{umlauf.burchard_CSR05} show that the value of the clipping factor is of crucial importance for
510the entrainment depth predicted in stably stratified situations,
511and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes.
512The clipping is only activated if \np[=.true.]{ln_length_lim}{ln\_length\_lim},
513and the $c_{lim}$ is set to the \np{rn_clim_galp}{rn\_clim\_galp} value.
514
515The time and space discretization of the GLS equations follows the same energetic consideration as for
516the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{burchard_OM02}.
517Evaluation of the 4 GLS turbulent closure schemes can be found in \citet{warner.sherwood.ea_OM05} in ROMS model and
518 in \citet{reffray.guillaume.ea_GMD15} for the \NEMO\ model.
519
520%% =================================================================================================
521\subsection[OSM: OSMosis boundary layer scheme (\forcode{ln_zdfosm})]{OSM: OSMosis boundary layer scheme (\protect\np{ln_zdfosm}{ln\_zdfosm})}
522\label{subsec:ZDF_osm}
523
524\begin{listing}
525  \nlst{namzdf_osm}
526  \caption{\forcode{&namzdf_osm}}
527  \label{lst:namzdf_osm}
528\end{listing}
529
530The OSMOSIS turbulent closure scheme is based on......   TBC
531
532%% =================================================================================================
533\subsection[ Discrete energy conservation for TKE and GLS schemes]{Discrete energy conservation for TKE and GLS schemes}
534\label{subsec:ZDF_tke_ene}
535
536\begin{figure}[!t]
537  \centering
538  \includegraphics[width=0.66\textwidth]{Fig_ZDF_TKE_time_scheme}
539  \caption[Subgrid kinetic energy integration in GLS and TKE schemes]{
540    Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and
541    its links to the momentum and tracer time integration.}
542  \label{fig:ZDF_TKE_time_scheme}
543\end{figure}
544
545The production of turbulence by vertical shear (the first term of the right hand side of
546\autoref{eq:ZDF_tke_e}) and  \autoref{eq:ZDF_gls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion
547(first line in \autoref{eq:MB_zdf}).
548To do so a special care has to be taken for both the time and space discretization of
549the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}.
550
551Let us first address the time stepping issue. \autoref{fig:ZDF_TKE_time_scheme} shows how
552the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with
553the one-level forward time stepping of the equation for $\bar{e}$.
554With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to
555the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and
556summing the result vertically:
557\begin{equation}
558  \label{eq:ZDF_energ1}
559  \begin{split}
560    \int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
561    &= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}
562    - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
563  \end{split}
564\end{equation}
565Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$,
566known at time $t$ (\autoref{fig:ZDF_TKE_time_scheme}), as it is required when using the TKE scheme
567(see \autoref{sec:TD_forward_imp}).
568The first term of the right hand side of \autoref{eq:ZDF_energ1} represents the kinetic energy transfer at
569the surface (atmospheric forcing) and at the bottom (friction effect).
570The second term is always negative.
571It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$.
572\autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
573the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
574${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$
575(and not by the more straightforward $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
576
577A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
578(second term of the right hand side of \autoref{eq:ZDF_tke_e} and \autoref{eq:ZDF_gls_e}).
579This term must balance the input of potential energy resulting from vertical mixing.
580The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by
581multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically:
582\begin{equation}
583  \label{eq:ZDF_energ2}
584  \begin{split}
585    \int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
586    &= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta}
587    - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
588    &= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
589    + \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
590  \end{split}
591\end{equation}
592where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
593The first term of the right hand side of \autoref{eq:ZDF_energ2} is always zero because
594there is no diffusive flux through the ocean surface and bottom).
595The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
596Therefore \autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
597the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:ZDF_tke_e} and  \autoref{eq:ZDF_gls_e}.
598
599Let us now address the space discretization issue.
600The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in
601the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:DOM_cell}).
602A space averaging is thus required to obtain the shear TKE production term.
603By redoing the \autoref{eq:ZDF_energ1} in the 3D case, it can be shown that the product of eddy coefficient by
604the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
605Furthermore, the time variation of $e_3$ has be taken into account.
606
607The above energetic considerations leads to the following final discrete form for the TKE equation:
608\begin{equation}
609  \label{eq:ZDF_tke_ene}
610  \begin{split}
611    \frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv
612    \Biggl\{ \Biggr.
613    &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} }
614        \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
615    +&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} }
616        \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j}
617    \Biggr. \Biggr\}   \\
618    %
619    - &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
620    %
621    +&\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]   \\
622    %
623    - &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
624  \end{split}
625\end{equation}
626where the last two terms in \autoref{eq:ZDF_tke_ene} (vertical diffusion and Kolmogorov dissipation)
627are time stepped using a backward scheme (see\autoref{sec:TD_forward_imp}).
628Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible.
629%The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as
630%they all appear in the right hand side of \autoref{eq:ZDF_tke_ene}.
631%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
632
633%% =================================================================================================
634\section{Convection}
635\label{sec:ZDF_conv}
636
637Static instabilities (\ie\ light potential densities under heavy ones) may occur at particular ocean grid points.
638In nature, convective processes quickly re-establish the static stability of the water column.
639These processes have been removed from the model via the hydrostatic assumption so they must be parameterized.
640Three parameterisations are available to deal with convective processes:
641a non-penetrative convective adjustment or an enhanced vertical diffusion,
642or/and the use of a turbulent closure scheme.
643
644%% =================================================================================================
645\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc})]{Non-penetrative convective adjustment (\protect\np{ln_tranpc}{ln\_tranpc})}
646\label{subsec:ZDF_npc}
647
648\begin{figure}[!htb]
649  \centering
650  \includegraphics[width=0.66\textwidth]{Fig_npc}
651  \caption[Unstable density profile treated by the non penetrative convective adjustment algorithm]{
652    Example of an unstable density profile treated by
653    the non penetrative convective adjustment algorithm.
654    $1^{st}$ step: the initial profile is checked from the surface to the bottom.
655    It is found to be unstable between levels 3 and 4.
656    They are mixed.
657    The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed.
658    The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed.
659    The $1^{st}$ step ends since the density profile is then stable below the level 3.
660    $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step:
661    levels 2 to 5 are mixed.
662    The new density profile is checked.
663    It is found stable: end of algorithm.}
664  \label{fig:ZDF_npc}
665\end{figure}
666
667Options are defined through the \nam{zdf}{zdf} namelist variables.
668The non-penetrative convective adjustment is used when \np[=.true.]{ln_zdfnpc}{ln\_zdfnpc}.
669It is applied at each \np{nn_npc}{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of
670the water column, but only until the density structure becomes neutrally stable
671(\ie\ until the mixed portion of the water column has \textit{exactly} the density of the water just below)
672\citep{madec.delecluse.ea_JPO91}.
673The associated algorithm is an iterative process used in the following way (\autoref{fig:ZDF_npc}):
674starting from the top of the ocean, the first instability is found.
675Assume in the following that the instability is located between levels $k$ and $k+1$.
676The temperature and salinity in the two levels are vertically mixed, conserving the heat and salt contents of
677the water column.
678The new density is then computed by a linear approximation.
679If the new density profile is still unstable between levels $k+1$ and $k+2$,
680levels $k$, $k+1$ and $k+2$ are then mixed.
681This process is repeated until stability is established below the level $k$
682(the mixing process can go down to the ocean bottom).
683The algorithm is repeated to check if the density profile between level $k-1$ and $k$ is unstable and/or
684if there is no deeper instability.
685
686This algorithm is significantly different from mixing statically unstable levels two by two.
687The latter procedure cannot converge with a finite number of iterations for some vertical profiles while
688the algorithm used in \NEMO\ converges for any profile in a number of iterations which is less than
689the number of vertical levels.
690This property is of paramount importance as pointed out by \citet{killworth_iprc89}:
691it avoids the existence of permanent and unrealistic static instabilities at the sea surface.
692This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in
693the north-western Mediterranean Sea \citep{madec.delecluse.ea_JPO91, madec.chartier.ea_DAO91, madec.crepon_iprc91}.
694
695The current implementation has been modified in order to deal with any non linear equation of seawater
696(L. Brodeau, personnal communication).
697Two main differences have been introduced compared to the original algorithm:
698$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency
699(not the difference in potential density);
700$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in
701the same way their temperature and salinity has been mixed.
702These two modifications allow the algorithm to perform properly and accurately with TEOS10 or EOS-80 without
703having to recompute the expansion coefficients at each mixing iteration.
704
705%% =================================================================================================
706\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd})]{Enhanced vertical diffusion (\protect\np{ln_zdfevd}{ln\_zdfevd})}
707\label{subsec:ZDF_evd}
708
709Options are defined through the  \nam{zdf}{zdf} namelist variables.
710The enhanced vertical diffusion parameterisation is used when \np[=.true.]{ln_zdfevd}{ln\_zdfevd}.
711In this case, the vertical eddy mixing coefficients are assigned very large values
712in regions where the stratification is unstable
713(\ie\ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}.
714This is done either on tracers only (\np[=0]{nn_evdm}{nn\_evdm}) or
715on both momentum and tracers (\np[=1]{nn_evdm}{nn\_evdm}).
716
717In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np[=1]{nn_evdm}{nn\_evdm},
718the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to
719the namelist parameter \np{rn_avevd}{rn\_avevd}.
720A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$.
721This parameterisation of convective processes is less time consuming than
722the convective adjustment algorithm presented above when mixing both tracers and
723momentum in the case of static instabilities.
724
725Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$.
726This removes a potential source of divergence of odd and even time step in
727a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:TD_mLF}).
728
729%% =================================================================================================
730\subsection[Handling convection with turbulent closure schemes (\forcode{ln_zdf_}\{\forcode{tke,gls,osm}\})]{Handling convection with turbulent closure schemes (\forcode{ln_zdf{tke,gls,osm}})}
731\label{subsec:ZDF_tcs}
732
733The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and
734\autoref{subsec:ZDF_osm} (\ie\ \np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} or \np{ln_zdfosm}{ln\_zdfosm} defined) deal, in theory,
735with statically unstable density profiles.
736In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in
737\autoref{eq:ZDF_tke_e} or \autoref{eq:ZDF_gls_e} becomes a source term, since $N^2$ is negative.
738It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at
739velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$).
740These large values restore the static stability of the water column in a way similar to that of
741the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}).
742However, in the vicinity of the sea surface (first ocean layer), the eddy coefficients computed by
743the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
744because the mixing length scale is bounded by the distance to the sea surface.
745It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme,
746\ie\ setting the \np{ln_zdfnpc}{ln\_zdfnpc} namelist parameter to true and
747defining the turbulent closure (\np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} = \forcode{.true.}) all together.
748
749The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection,
750%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp},
751therefore \np[=.false.]{ln_zdfevd}{ln\_zdfevd} should be used with the OSMOSIS scheme.
752% gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
753
754%% =================================================================================================
755\section[Double diffusion mixing (\forcode{ln_zdfddm})]{Double diffusion mixing (\protect\np{ln_zdfddm}{ln\_zdfddm})}
756\label{subsec:ZDF_ddm}
757
758%
759%\nlst{namzdf_ddm}
760
761This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter
762\np{ln_zdfddm}{ln\_zdfddm} in \nam{zdf}{zdf}.
763Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa.
764The former condition leads to salt fingering and the latter to diffusive convection.
765Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean.
766\citet{merryfield.holloway.ea_JPO99} include a parameterisation of such phenomena in a global ocean model and show that
767it leads to relatively minor changes in circulation but exerts significant regional influences on
768temperature and salinity.
769
770Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
771\begin{align*}
772  % \label{eq:ZDF_ddm_Kz}
773  &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT} \\
774  &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
775\end{align*}
776where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
777and $o$ by processes other than double diffusion.
778The rates of double-diffusive mixing depend on the buoyancy ratio
779$R_\rho = \alpha \partial_z T / \beta \partial_z S$, where $\alpha$ and $\beta$ are coefficients of
780thermal expansion and saline contraction (see \autoref{subsec:TRA_eos}).
781To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt
782(1981):
783\begin{align}
784  \label{eq:ZDF_ddm_f}
785  A_f^{vS} &=
786             \begin{cases}
787               \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
788               0                              &\text{otherwise}
789             \end{cases}
790  \\         \label{eq:ZDF_ddm_f_T}
791  A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho
792\end{align}
793
794\begin{figure}[!t]
795  \centering
796  \includegraphics[width=0.66\textwidth]{Fig_zdfddm}
797  \caption[Diapycnal diffusivities for temperature and salt in regions of salt fingering and
798  diffusive convection]{
799    From \citet{merryfield.holloway.ea_JPO99}:
800    (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in
801    regions of salt fingering.
802    Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and
803    thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$;
804    (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in
805    regions of diffusive convection.
806    Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation.
807    The latter is not implemented in \NEMO.}
808  \label{fig:ZDF_ddm}
809\end{figure}
810
811The factor 0.7 in \autoref{eq:ZDF_ddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of
812buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}).
813Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
814
815To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by
816Federov (1988) is used:
817\begin{align}
818  % \label{eq:ZDF_ddm_d}
819  A_d^{vT} &=
820             \begin{cases}
821               1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
822               &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
823               0                       &\text{otherwise}
824             \end{cases}
825                                       \nonumber \\
826  \label{eq:ZDF_ddm_d_S}
827  A_d^{vS} &=
828             \begin{cases}
829               A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right) &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
830               A_d^{vT} \ 0.15 \ R_\rho               &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
831               0                       &\text{otherwise}
832             \end{cases}
833\end{align}
834
835The dependencies of \autoref{eq:ZDF_ddm_f} to \autoref{eq:ZDF_ddm_d_S} on $R_\rho$ are illustrated in
836\autoref{fig:ZDF_ddm}.
837Implementing this requires computing $R_\rho$ at each grid point on every time step.
838This is done in \mdl{eosbn2} at the same time as $N^2$ is computed.
839This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive).
840
841%% =================================================================================================
842\section[Bottom and top friction (\textit{zdfdrg.F90})]{Bottom and top friction (\protect\mdl{zdfdrg})}
843\label{sec:ZDF_drg}
844
845%
846\begin{listing}
847  \nlst{namdrg}
848  \caption{\forcode{&namdrg}}
849  \label{lst:namdrg}
850\end{listing}
851\begin{listing}
852  \nlst{namdrg_top}
853  \caption{\forcode{&namdrg_top}}
854  \label{lst:namdrg_top}
855\end{listing}
856\begin{listing}
857  \nlst{namdrg_bot}
858  \caption{\forcode{&namdrg_bot}}
859  \label{lst:namdrg_bot}
860\end{listing}
861
862
863Options to define the top and bottom friction are defined through the \nam{drg}{drg} namelist variables.
864The bottom friction represents the friction generated by the bathymetry.
865The top friction represents the friction generated by the ice shelf/ocean interface.
866As the friction processes at the top and the bottom are treated in and identical way,
867the description below considers mostly the bottom friction case, if not stated otherwise.
868
869Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as
870a condition on the vertical diffusive flux.
871For the bottom boundary layer, one has:
872 \[
873   % \label{eq:ZDF_bfr_flux}
874   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
875 \]
876where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside
877the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean).
878How ${\cal F}_h^{\textbf U}$ influences the interior depends on the vertical resolution of the model near
879the bottom relative to the Ekman layer depth.
880For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m,
881one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$
882(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$).
883With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
884When the vertical mixing coefficient is this small, using a flux condition is equivalent to
885entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or
886bottom model layer.
887To illustrate this, consider the equation for $u$ at $k$, the last ocean level:
888\begin{equation}
889  \label{eq:ZDF_drg_flux2}
890  \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}}
891\end{equation}
892If the bottom layer thickness is 200~m, the Ekman transport will be distributed over that depth.
893On the other hand, if the vertical resolution is high (1~m or less) and a turbulent closure model is used,
894the turbulent Ekman layer will be represented explicitly by the model.
895However, the logarithmic layer is never represented in current primitive equation model applications:
896it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
897Two choices are available in \NEMO: a linear and a quadratic bottom friction.
898Note that in both cases, the rotation between the interior velocity and the bottom friction is neglected in
899the present release of \NEMO.
900
901In the code, the bottom friction is imposed by adding the trend due to the bottom friction to
902 the general momentum trend in \mdl{dynzdf}.
903For the time-split surface pressure gradient algorithm, the momentum trend due to
904the barotropic component needs to be handled separately.
905For this purpose it is convenient to compute and store coefficients which can be simply combined with
906bottom velocities and geometric values to provide the momentum trend due to bottom friction.
907 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where:
908\begin{equation}
909  \label{eq:ZDF_bfr_bdef}
910  \frac{\partial {\textbf U_h}}{\partial t} =
911  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
912\end{equation}
913where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
914Note than from \NEMO\ 4.0, drag coefficients are only computed at cell centers (\ie\ at T-points) and refer to as $c_b^T$ in the following. These are then linearly interpolated in space to get $c_b^\textbf{U}$ at velocity points.
915
916%% =================================================================================================
917\subsection[Linear top/bottom friction (\forcode{ln_lin})]{Linear top/bottom friction (\protect\np{ln_lin}{ln\_lin})}
918\label{subsec:ZDF_drg_linear}
919
920The linear friction parameterisation (including the special case of a free-slip condition) assumes that
921the friction is proportional to the interior velocity (\ie\ the velocity of the first/last model level):
922\[
923  % \label{eq:ZDF_bfr_linear}
924  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
925\]
926where $r$ is a friction coefficient expressed in $m s^{-1}$.
927This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,
928and setting $r = H / \tau$, where $H$ is the ocean depth.
929Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{weatherly_JMR84}.
930A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models.
931One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$
932(\citet{gill_bk82}, Eq. 9.6.6).
933For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$,
934and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
935This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days.
936It can be changed by specifying \np{rn_Uc0}{rn\_Uc0} (namelist parameter).
937
938 For the linear friction case the drag coefficient used in the general expression \autoref{eq:ZDF_bfr_bdef} is:
939\[
940  % \label{eq:ZDF_bfr_linbfr_b}
941    c_b^T = - r
942\]
943When \np[=.true.]{ln_lin}{ln\_lin}, the value of $r$ used is \np{rn_Uc0}{rn\_Uc0}*\np{rn_Cd0}{rn\_Cd0}.
944Setting \np[=.true.]{ln_OFF}{ln\_OFF} (and \forcode{ln_lin=.true.}) is equivalent to setting $r=0$ and leads to a free-slip boundary condition.
945
946These values are assigned in \mdl{zdfdrg}.
947Note that there is support for local enhancement of these values via an externally defined 2D mask array
948(\np[=.true.]{ln_boost}{ln\_boost}) given in the \ifile{bfr\_coef} input NetCDF file.
949The mask values should vary from 0 to 1.
950Locations with a non-zero mask value will have the friction coefficient increased by
951$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
952
953%% =================================================================================================
954\subsection[Non-linear top/bottom friction (\forcode{ln_non_lin})]{Non-linear top/bottom friction (\protect\np{ln_non_lin}{ln\_non\_lin})}
955\label{subsec:ZDF_drg_nonlinear}
956
957The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic:
958\[
959  % \label{eq:ZDF_drg_nonlinear}
960  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
961  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
962\]
963where $C_D$ is a drag coefficient, and $e_b $ a top/bottom turbulent kinetic energy due to tides,
964internal waves breaking and other short time scale currents.
965A typical value of the drag coefficient is $C_D = 10^{-3} $.
966As an example, the CME experiment \citep{treguier_JGR92} uses $C_D = 10^{-3}$ and
967$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and
968$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
969The CME choices have been set as default values (\np{rn_Cd0}{rn\_Cd0} and \np{rn_ke0}{rn\_ke0} namelist parameters).
970
971As for the linear case, the friction is imposed in the code by adding the trend due to
972the friction to the general momentum trend in \mdl{dynzdf}.
973For the non-linear friction case the term computed in \mdl{zdfdrg} is:
974\[
975  % \label{eq:ZDF_drg_nonlinbfr}
976    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2}
977\]
978
979The coefficients that control the strength of the non-linear friction are initialised as namelist parameters:
980$C_D$= \np{rn_Cd0}{rn\_Cd0}, and $e_b$ =\np{rn_bfeb2}{rn\_bfeb2}.
981Note that for applications which consider tides explicitly, a low or even zero value of \np{rn_bfeb2}{rn\_bfeb2} is recommended. A local enhancement of $C_D$ is again possible via an externally defined 2D mask array
982(\np[=.true.]{ln_boost}{ln\_boost}).
983This works in the same way as for the linear friction case with non-zero masked locations increased by
984$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
985
986%% =================================================================================================
987\subsection[Log-layer top/bottom friction (\forcode{ln_loglayer})]{Log-layer top/bottom friction (\protect\np{ln_loglayer}{ln\_loglayer})}
988\label{subsec:ZDF_drg_loglayer}
989
990In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using
991a "law of the wall" scaling. This assumes that the model vertical resolution can capture the logarithmic layer which typically occur for layers thinner than 1 m or so.
992If  \np[=.true.]{ln_loglayer}{ln\_loglayer}, $C_D$ is no longer constant but is related to the distance to the wall (or equivalently to the half of the top/bottom layer thickness):
993\[
994  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2
995\]
996
997\noindent where $\kappa$ is the von-Karman constant and \np{rn_z0}{rn\_z0} is a roughness length provided via the namelist.
998
999The drag coefficient is bounded such that it is kept greater or equal to
1000the base \np{rn_Cd0}{rn\_Cd0} value which occurs where layer thicknesses become large and presumably logarithmic layers are not resolved at all. For stability reason, it is also not allowed to exceed the value of an additional namelist parameter:
1001\np{rn_Cdmax}{rn\_Cdmax}, \ie
1002\[
1003  rn\_Cd0 \leq C_D \leq rn\_Cdmax
1004\]
1005
1006\noindent The log-layer enhancement can also be applied to the top boundary friction if
1007under ice-shelf cavities are activated (\np[=.true.]{ln_isfcav}{ln\_isfcav}).
1008%In this case, the relevant namelist parameters are \np{rn_tfrz0}{rn\_tfrz0}, \np{rn_tfri2}{rn\_tfri2} and \np{rn_tfri2_max}{rn\_tfri2\_max}.
1009
1010%% =================================================================================================
1011\subsection[Explicit top/bottom friction (\forcode{ln_drgimp=.false.})]{Explicit top/bottom friction (\protect\np[=.false.]{ln_drgimp}{ln\_drgimp})}
1012\label{subsec:ZDF_drg_stability}
1013
1014Setting \np[=.false.]{ln_drgimp}{ln\_drgimp} means that bottom friction is treated explicitly in time, which has the advantage of simplifying the interaction with the split-explicit free surface (see \autoref{subsec:ZDF_drg_ts}). The latter does indeed require the knowledge of bottom stresses in the course of the barotropic sub-iteration, which becomes less straightforward in the implicit case. In the explicit case, top/bottom stresses can be computed using \textit{before} velocities and inserted in the overall momentum tendency budget. This reads:
1015
1016At the top (below an ice shelf cavity):
1017\[
1018  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1019  = c_{t}^{\textbf{U}}\textbf{u}^{n-1}_{t}
1020\]
1021
1022At the bottom (above the sea floor):
1023\[
1024  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1025  = c_{b}^{\textbf{U}}\textbf{u}^{n-1}_{b}
1026\]
1027
1028Since this is conditionally stable, some care needs to exercised over the choice of parameters to ensure that the implementation of explicit top/bottom friction does not induce numerical instability.
1029For the purposes of stability analysis, an approximation to \autoref{eq:ZDF_drg_flux2} is:
1030\begin{equation}
1031  \label{eq:ZDF_Eqn_drgstab}
1032  \begin{split}
1033    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
1034    &= -\frac{ru}{e_{3u}}\;2\rdt\\
1035  \end{split}
1036\end{equation}
1037\noindent where linear friction and a leapfrog timestep have been assumed.
1038To ensure that the friction cannot reverse the direction of flow it is necessary to have:
1039\[
1040  |\Delta u| < \;|u|
1041\]
1042\noindent which, using \autoref{eq:ZDF_Eqn_drgstab}, gives:
1043\[
1044  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
1045\]
1046This same inequality can also be derived in the non-linear bottom friction case if
1047a velocity of 1 m.s$^{-1}$ is assumed.
1048Alternatively, this criterion can be rearranged to suggest a minimum bottom box thickness to ensure stability:
1049\[
1050  e_{3u} > 2\;r\;\rdt
1051\]
1052\noindent which it may be necessary to impose if partial steps are being used.
1053For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m.
1054For most applications, with physically sensible parameters these restrictions should not be of concern.
1055But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.
1056To ensure stability limits are imposed on the top/bottom friction coefficients both
1057during initialisation and at each time step.
1058Checks at initialisation are made in \mdl{zdfdrg} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1059The number of breaches of the stability criterion are reported as well as
1060the minimum and maximum values that have been set.
1061The criterion is also checked at each time step, using the actual velocity, in \mdl{dynzdf}.
1062Values of the friction coefficient are reduced as necessary to ensure stability;
1063these changes are not reported.
1064
1065Limits on the top/bottom friction coefficient are not imposed if the user has elected to
1066handle the friction implicitly (see \autoref{subsec:ZDF_drg_imp}).
1067The number of potential breaches of the explicit stability criterion are still reported for information purposes.
1068
1069%% =================================================================================================
1070\subsection[Implicit top/bottom friction (\forcode{ln_drgimp=.true.})]{Implicit top/bottom friction (\protect\np[=.true.]{ln_drgimp}{ln\_drgimp})}
1071\label{subsec:ZDF_drg_imp}
1072
1073An optional implicit form of bottom friction has been implemented to improve model stability.
1074We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting.
1075This option can be invoked by setting \np{ln_drgimp}{ln\_drgimp} to \forcode{.true.} in the \nam{drg}{drg} namelist.
1076%This option requires \np{ln_zdfexp}{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf}{zdf} namelist.
1077
1078This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step:
1079
1080At the top (below an ice shelf cavity):
1081\[
1082  % \label{eq:ZDF_dynZDF__drg_top}
1083  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1084  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t}
1085\]
1086
1087At the bottom (above the sea floor):
1088\[
1089  % \label{eq:ZDF_dynZDF__drg_bot}
1090  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1091  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b}
1092\]
1093
1094where $t$ and $b$ refers to top and bottom layers respectively.
1095Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit.
1096
1097%% =================================================================================================
1098\subsection[Bottom friction with split-explicit free surface]{Bottom friction with split-explicit free surface}
1099\label{subsec:ZDF_drg_ts}
1100
1101With split-explicit free surface, the sub-stepping of barotropic equations needs the knowledge of top/bottom stresses. An obvious way to satisfy this is to take them as constant over the course of the barotropic integration and equal to the value used to update the baroclinic momentum trend. Provided \np[=.false.]{ln_drgimp}{ln\_drgimp} and a centred or \textit{leap-frog} like integration of barotropic equations is used (\ie\ \forcode{ln_bt_fw=.false.}, cf \autoref{subsec:DYN_spg_ts}), this does ensure that barotropic and baroclinic dynamics feel the same stresses during one leapfrog time step. However, if \np[=.true.]{ln_drgimp}{ln\_drgimp},  stresses depend on the \textit{after} value of the velocities which themselves depend on the barotropic iteration result. This cyclic dependency makes difficult obtaining consistent stresses in 2d and 3d dynamics. Part of this mismatch is then removed when setting the final barotropic component of 3d velocities to the time splitting estimate. This last step can be seen as a necessary evil but should be minimized since it interferes with the adjustment to the boundary conditions.
1102
1103The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO\ is as follows:
1104\begin{enumerate}
1105\item To extend the stability of the barotropic sub-stepping, bottom stresses are refreshed at each sub-iteration. The baroclinic part of the flow entering the stresses is frozen at the initial time of the barotropic iteration. In case of non-linear friction, the drag coefficient is also constant.
1106\item In case of an implicit drag, specific computations are performed in \mdl{dynzdf} which renders the overall scheme mixed explicit/implicit: the barotropic components of 3d velocities are removed before seeking for the implicit vertical diffusion result. Top/bottom stresses due to the barotropic components are explicitly accounted for thanks to the updated values of barotropic velocities. Then the implicit solution of 3d velocities is obtained. Lastly, the residual barotropic component is replaced by the time split estimate.
1107\end{enumerate}
1108
1109Note that other strategies are possible, like considering vertical diffusion step in advance, \ie\ prior barotropic integration.
1110
1111%% =================================================================================================
1112\section[Internal wave-driven mixing (\forcode{ln_zdfiwm})]{Internal wave-driven mixing (\protect\np{ln_zdfiwm}{ln\_zdfiwm})}
1113\label{subsec:ZDF_tmx_new}
1114
1115%
1116\begin{listing}
1117  \nlst{namzdf_iwm}
1118  \caption{\forcode{&namzdf_iwm}}
1119  \label{lst:namzdf_iwm}
1120\end{listing}
1121
1122The parameterization of mixing induced by breaking internal waves is a generalization of
1123the approach originally proposed by \citet{st-laurent.simmons.ea_GRL02}.
1124A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed,
1125and the resulting diffusivity is obtained as
1126\[
1127  % \label{eq:ZDF_Kwave}
1128  A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 }
1129\]
1130where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of
1131the energy available for mixing.
1132If the \np{ln_mevar}{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and
1133equal to 1/6 \citep{osborn_JPO80}.
1134In the opposite (recommended) case, $R_f$ is instead a function of
1135the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$,
1136with $\nu$ the molecular viscosity of seawater, following the model of \cite{bouffard.boegman_DAO13} and
1137the implementation of \cite{de-lavergne.madec.ea_JPO16}.
1138Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when
1139the mixing efficiency is constant.
1140
1141In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary
1142as a function of $Re_b$ by setting the \np{ln_tsdiff}{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice.
1143This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14},
1144is implemented as in \cite{de-lavergne.madec.ea_JPO16}.
1145
1146The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$,
1147is constructed from three static maps of column-integrated internal wave energy dissipation,
1148$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures:
1149
1150\begin{align*}
1151  F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\
1152  F_{pyc}(i,j,k) &\propto N^{n_p}\\
1153  F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} }
1154\end{align*}
1155In the above formula, $h_{ab}$ denotes the height above bottom,
1156$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by
1157\[
1158  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; ,
1159\]
1160The $n_p$ parameter (given by \np{nn_zpyc}{nn\_zpyc} in \nam{zdf_iwm}{zdf\_iwm} namelist)
1161controls the stratification-dependence of the pycnocline-intensified dissipation.
1162It can take values of $1$ (recommended) or $2$.
1163Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of
1164the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps.
1165$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and
1166$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of
1167the abyssal hill topography \citep{goff_JGR10} and the latitude.
1168%
1169% Jc: input files names ?
1170
1171%% =================================================================================================
1172\section[Surface wave-induced mixing (\forcode{ln_zdfswm})]{Surface wave-induced mixing (\protect\np{ln_zdfswm}{ln\_zdfswm})}
1173\label{subsec:ZDF_swm}
1174
1175Surface waves produce an enhanced mixing through wave-turbulence interaction.
1176In addition to breaking waves induced turbulence (\autoref{subsec:ZDF_tke}),
1177the influence of non-breaking waves can be accounted introducing
1178wave-induced viscosity and diffusivity as a function of the wave number spectrum.
1179Following \citet{qiao.yuan.ea_OD10}, a formulation of wave-induced mixing coefficient
1180is provided  as a function of wave amplitude, Stokes Drift and wave-number:
1181
1182\begin{equation}
1183  \label{eq:ZDF_Bv}
1184  B_{v} = \alpha {A} {U}_{st} {exp(3kz)}
1185\end{equation}
1186
1187Where $B_{v}$ is the wave-induced mixing coefficient, $A$ is the wave amplitude,
1188${U}_{st}$ is the Stokes Drift velocity, $k$ is the wave number and $\alpha$
1189is a constant which should be determined by observations or
1190numerical experiments and is set to be 1.
1191
1192The coefficient $B_{v}$ is then directly added to the vertical viscosity
1193and diffusivity coefficients.
1194
1195In order to account for this contribution set: \forcode{ln_zdfswm=.true.},
1196then wave interaction has to be activated through \forcode{ln_wave=.true.},
1197the Stokes Drift can be evaluated by setting \forcode{ln_sdw=.true.}
1198(see \autoref{subsec:SBC_wave_sdw})
1199and the needed wave fields can be provided either in forcing or coupled mode
1200(for more information on wave parameters and settings see \autoref{sec:SBC_wave})
1201
1202%% =================================================================================================
1203\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp})]{Adaptive-implicit vertical advection(\protect\np{ln_zad_Aimp}{ln\_zad\_Aimp})}
1204\label{subsec:ZDF_aimp}
1205
1206The adaptive-implicit vertical advection option in NEMO is based on the work of
1207\citep{shchepetkin_OM15}.  In common with most ocean models, the timestep used with NEMO
1208needs to satisfy multiple criteria associated with different physical processes in order
1209to maintain numerical stability. \citep{shchepetkin_OM15} pointed out that the vertical
1210CFL criterion is commonly the most limiting. \citep{lemarie.debreu.ea_OM15} examined the
1211constraints for a range of time and space discretizations and provide the CFL stability
1212criteria for a range of advection schemes. The values for the Leap-Frog with Robert
1213asselin filter time-stepping (as used in NEMO) are reproduced in
1214\autoref{tab:ZDF_zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these
1215restrictions but at the cost of large dispersive errors and, possibly, large numerical
1216viscosity. The adaptive-implicit vertical advection option provides a targetted use of the
1217implicit scheme only when and where potential breaches of the vertical CFL condition
1218occur. In many practical applications these events may occur remote from the main area of
1219interest or due to short-lived conditions such that the extra numerical diffusion or
1220viscosity does not greatly affect the overall solution. With such applications, setting:
1221\forcode{ln_zad_Aimp=.true.} should allow much longer model timesteps to be used whilst
1222retaining the accuracy of the high order explicit schemes over most of the domain.
1223
1224\begin{table}[htbp]
1225  \centering
1226  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}}
1227  \begin{tabular}{r|ccc}
1228    \hline
1229    spatial discretization  & 2$^nd$ order centered & 3$^rd$ order upwind & 4$^th$ order compact \\
1230    advective CFL criterion &                 0.904 &              0.472  &                0.522 \\
1231    \hline
1232  \end{tabular}
1233  \caption[Advective CFL criteria for the leapfrog with Robert Asselin filter time-stepping]{
1234    The advective CFL criteria for a range of spatial discretizations for
1235    the leapfrog with Robert Asselin filter time-stepping
1236    ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}.}
1237  \label{tab:ZDF_zad_Aimp_CFLcrit}
1238\end{table}
1239
1240In particular, the advection scheme remains explicit everywhere except where and when
1241local vertical velocities exceed a threshold set just below the explicit stability limit.
1242Once the threshold is reached a tapered transition towards an implicit scheme is used by
1243partitioning the vertical velocity into a part that can be treated explicitly and any
1244excess that must be treated implicitly. The partitioning is achieved via a Courant-number
1245dependent weighting algorithm as described in \citep{shchepetkin_OM15}.
1246
1247The local cell Courant number ($Cu$) used for this partitioning is:
1248
1249\begin{equation}
1250  \label{eq:ZDF_Eqn_zad_Aimp_Courant}
1251  \begin{split}
1252    Cu &= {2 \rdt \over e^n_{3t_{ijk}}} \bigg (\big [ \texttt{Max}(w^n_{ijk},0.0) - \texttt{Min}(w^n_{ijk+1},0.0) \big ]    \\
1253       &\phantom{=} +\big [ \texttt{Max}(e_{{2_u}ij}e^n_{{3_{u}}ijk}u^n_{ijk},0.0) - \texttt{Min}(e_{{2_u}i-1j}e^n_{{3_{u}}i-1jk}u^n_{i-1jk},0.0) \big ]
1254                     \big / e_{{1_t}ij}e_{{2_t}ij}            \\
1255       &\phantom{=} +\big [ \texttt{Max}(e_{{1_v}ij}e^n_{{3_{v}}ijk}v^n_{ijk},0.0) - \texttt{Min}(e_{{1_v}ij-1}e^n_{{3_{v}}ij-1k}v^n_{ij-1k},0.0) \big ]
1256                     \big / e_{{1_t}ij}e_{{2_t}ij} \bigg )    \\
1257  \end{split}
1258\end{equation}
1259
1260\noindent and the tapering algorithm follows \citep{shchepetkin_OM15} as:
1261
1262\begin{align}
1263  \label{eq:ZDF_Eqn_zad_Aimp_partition}
1264Cu_{min} &= 0.15 \nonumber \\
1265Cu_{max} &= 0.3  \nonumber \\
1266Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\
1267Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\
1268\cf &=
1269     \begin{cases}
1270        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\
1271        (Cu - Cu_{min})^2 / (Fcu +  (Cu - Cu_{min})^2)             &\text{else if $Cu < Cu_{cut}$} \\
1272        (Cu - Cu_{max}) / Cu                                       &\text{else}
1273     \end{cases}
1274\end{align}
1275
1276\begin{figure}[!t]
1277  \centering
1278  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_coeff}
1279  \caption[Partitioning coefficient used to partition vertical velocities into parts]{
1280    The value of the partitioning coefficient (\cf) used to partition vertical velocities into
1281    parts to be treated implicitly and explicitly for a range of typical Courant numbers
1282    (\forcode{ln_zad_Aimp=.true.}).}
1283  \label{fig:ZDF_zad_Aimp_coeff}
1284\end{figure}
1285
1286\noindent The partitioning coefficient is used to determine the part of the vertical
1287velocity that must be handled implicitly ($w_i$) and to subtract this from the total
1288vertical velocity ($w_n$) to leave that which can continue to be handled explicitly:
1289
1290\begin{align}
1291  \label{eq:ZDF_Eqn_zad_Aimp_partition2}
1292    w_{i_{ijk}} &= \cf_{ijk} w_{n_{ijk}}     \nonumber \\
1293    w_{n_{ijk}} &= (1-\cf_{ijk}) w_{n_{ijk}}
1294\end{align}
1295
1296\noindent Note that the coefficient is such that the treatment is never fully implicit;
1297the three cases from \autoref{eq:ZDF_Eqn_zad_Aimp_partition} can be considered as:
1298fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the
1299coefficient (\cf) varies as shown in \autoref{fig:ZDF_zad_Aimp_coeff}. Note with these values
1300the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly
1301implicit' is 0.45 which is just below the stability limited given in
1302\autoref{tab:ZDF_zad_Aimp_CFLcrit}  for a 3rd order scheme.
1303
1304The $w_i$ component is added to the implicit solvers for the vertical mixing in
1305\mdl{dynzdf} and \mdl{trazdf} in a similar way to \citep{shchepetkin_OM15}.  This is
1306sufficient for the flux-limited advection scheme (\forcode{ln_traadv_mus}) but further
1307intervention is required when using the flux-corrected scheme (\forcode{ln_traadv_fct}).
1308For these schemes the implicit upstream fluxes must be added to both the monotonic guess
1309and to the higher order solution when calculating the antidiffusive fluxes. The implicit
1310vertical fluxes are then removed since they are added by the implicit solver later on.
1311
1312The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be
1313used in a wide range of simulations. The following test simulation, however, does illustrate
1314the potential benefits and will hopefully encourage further testing and feedback from users:
1315
1316\begin{figure}[!t]
1317  \centering
1318  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_overflow_frames}
1319  \caption[OVERFLOW: time-series of temperature vertical cross-sections]{
1320    A time-series of temperature vertical cross-sections for the OVERFLOW test case.
1321    These results are for the default settings with \forcode{nn_rdt=10.0} and
1322    without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}).}
1323  \label{fig:ZDF_zad_Aimp_overflow_frames}
1324\end{figure}
1325
1326%% =================================================================================================
1327\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case}
1328
1329The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case}
1330provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case
1331by only a few extra physics choices namely:
1332
1333\begin{verbatim}
1334     ln_dynldf_OFF = .false.
1335     ln_dynldf_lap = .true.
1336     ln_dynldf_hor = .true.
1337     ln_zdfnpc     = .true.
1338     ln_traadv_fct = .true.
1339        nn_fct_h   =  2
1340        nn_fct_v   =  2
1341\end{verbatim}
1342
1343\noindent which were chosen to provide a slightly more stable and less noisy solution. The
1344result when using the default value of \forcode{nn_rdt=10.} without adaptive-implicit
1345vertical velocity is illustrated in \autoref{fig:ZDF_zad_Aimp_overflow_frames}. The mass of
1346cold water, initially sitting on the shelf, moves down the slope and forms a
1347bottom-trapped, dense plume. Even with these extra physics choices the model is close to
1348stability limits and attempts with \forcode{nn_rdt=30.} will fail after about 5.5 hours
1349with excessively high horizontal velocities. This time-scale corresponds with the time the
1350plume reaches the steepest part of the topography and, although detected as a horizontal
1351CFL breach, the instability originates from a breach of the vertical CFL limit. This is a good
1352candidate, therefore, for use of the adaptive-implicit vertical advection scheme.
1353
1354The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps
1355are shown in \autoref{fig:ZDF_zad_Aimp_overflow_all_rdt} (together with the equivalent
1356frames from the base run).  In this simple example the use of the adaptive-implicit
1357vertcal advection scheme has enabled a 12x increase in the model timestep without
1358significantly altering the solution (although at this extreme the plume is more diffuse
1359and has not travelled so far).  Notably, the solution with and without the scheme is
1360slightly different even with \forcode{nn_rdt=10.}; suggesting that the base run was
1361close enough to instability to trigger the scheme despite completing successfully.
1362To assist in diagnosing how active the scheme is, in both location and time, the 3D
1363implicit and explicit components of the vertical velocity are available via XIOS as
1364\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient
1365(\cf) is also available as \texttt{wi\_cff}. For a quick oversight of
1366the schemes activity the global maximum values of the absolute implicit component
1367of the vertical velocity and the partitioning coefficient are written to the netCDF
1368version of the run statistics file (\texttt{run.stat.nc}) if this is active (see
1369\autoref{sec:MISC_opt} for activation details).
1370
1371\autoref{fig:ZDF_zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for
1372the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is
1373active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the
1374test case is close to stability limits even with this value. At the larger timesteps, the
1375vertical velocity is treated mostly implicitly at some location throughout the run. The
1376oscillatory nature of this measure appears to be linked to the progress of the plume front
1377as each cusp is associated with the location of the maximum shifting to the adjacent cell.
1378This is illustrated in \autoref{fig:ZDF_zad_Aimp_maxCf_loc} where the i- and k- locations of the
1379maximum have been overlaid for the base run case.
1380
1381\medskip
1382\noindent Only limited tests have been performed in more realistic configurations. In the
1383ORCA2\_ICE\_PISCES reference configuration the scheme does activate and passes
1384restartability and reproducibility tests but it is unable to improve the model's stability
1385enough to allow an increase in the model time-step. A view of the time-series of maximum
1386partitioning coefficient (not shown here)  suggests that the default time-step of 5400s is
1387already pushing at stability limits, especially in the initial start-up phase. The
1388time-series does not, however, exhibit any of the 'cuspiness' found with the overflow
1389tests.
1390
1391\medskip
1392\noindent A short test with an eORCA1 configuration promises more since a test using a
1393time-step of 3600s remains stable with \forcode{ln_zad_Aimp=.true.} whereas the
1394time-step is limited to 2700s without.
1395
1396\begin{figure}[!t]
1397  \centering
1398  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_overflow_all_rdt}
1399  \caption[OVERFLOW: sample temperature vertical cross-sections from mid- and end-run]{
1400    Sample temperature vertical cross-sections from mid- and end-run using
1401    different values for \forcode{nn_rdt} and with or without adaptive implicit vertical advection.
1402    Without the adaptive implicit vertical advection
1403    only the run with the shortest timestep is able to run to completion.
1404    Note also that the colour-scale has been chosen to confirm that
1405    temperatures remain within the original range of 10$^o$ to 20$^o$.}
1406  \label{fig:ZDF_zad_Aimp_overflow_all_rdt}
1407\end{figure}
1408
1409\begin{figure}[!t]
1410  \centering
1411  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_maxCf}
1412  \caption[OVERFLOW: maximum partitioning coefficient during a series of test runs]{
1413    The maximum partitioning coefficient during a series of test runs with
1414    increasing model timestep length.
1415    At the larger timesteps,
1416    the vertical velocity is treated mostly implicitly at some location throughout the run.}
1417  \label{fig:ZDF_zad_Aimp_maxCf}
1418\end{figure}
1419
1420\begin{figure}[!t]
1421  \centering
1422  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_maxCf_loc}
1423  \caption[OVERFLOW: maximum partitioning coefficient for the case overlaid]{
1424    The maximum partitioning coefficient for the \forcode{nn_rdt=10.0} case overlaid with
1425    information on the gridcell i- and k-locations of the maximum value.}
1426  \label{fig:ZDF_zad_Aimp_maxCf_loc}
1427\end{figure}
1428
1429\onlyinsubfile{\input{../../global/epilogue}}
1430
1431\end{document}
Note: See TracBrowser for help on using the repository browser.