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/branches/2020/dev_r14116_HPC-04_mcastril_Mixed_Precision_implementation_final/doc/latex/NEMO/subfiles – NEMO

source: NEMO/branches/2020/dev_r14116_HPC-04_mcastril_Mixed_Precision_implementation_final/doc/latex/NEMO/subfiles/chap_ZDF.tex @ 14644

Last change on this file since 14644 was 14644, checked in by sparonuz, 3 years ago

Merge trunk -r14642:HEAD

File size: 97.6 KB
Line 
1\documentclass[../main/NEMO_manual]{subfiles}
2
3\begin{document}
4
5\chapter{Vertical Ocean Physics (ZDF)}
6\label{chap:ZDF}
7
8\chaptertoc
9
10\paragraph{Changes record} ~\\
11
12{\footnotesize
13  \begin{tabularx}{\textwidth}{l||X|X}
14    Release & Author(s) & Modifications \\
15    \hline
16    {\em   4.0} & {\em ...} & {\em ...} \\
17    {\em   3.6} & {\em ...} & {\em ...} \\
18    {\em   3.4} & {\em ...} & {\em ...} \\
19    {\em <=3.4} & {\em ...} & {\em ...}
20  \end{tabularx}
21}
22
23\clearpage
24
25\cmtgm{ Add here a small introduction to ZDF and naming of the different physics
26(similar to what have been written for TRA and DYN).}
27
28%% =================================================================================================
29\section{Vertical mixing}
30\label{sec:ZDF}
31
32The discrete form of the ocean subgrid scale physics has been presented in
33\autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}.
34At the surface and bottom boundaries, the turbulent fluxes of momentum, heat and salt have to be defined.
35At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}),
36while at the bottom they are set to zero for heat and salt,
37unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie\ \np{ln_trabbc}{ln\_trabbc} defined,
38see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum
39(see \autoref{sec:ZDF_drg}).
40
41In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and
42diffusivity coefficients, $A_u^{vm}$ , $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$- points,
43respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}).
44These coefficients can be assumed to be either constant, or a function of the local Richardson number,
45or computed from a turbulent closure model (either TKE or GLS or OSMOSIS formulation).
46The computation of these coefficients is initialized in the \mdl{zdfphy} module and performed in
47the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} or \mdl{zdfosm} modules.
48The trends due to the vertical momentum and tracer diffusion, including the surface forcing,
49are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.
50%These trends can be computed using either a forward time stepping scheme
51%(namelist parameter \np[=.true.]{ln_zdfexp}{ln\_zdfexp}) or a backward time stepping scheme
52%(\np[=.false.]{ln_zdfexp}{ln\_zdfexp}) depending on the magnitude of the mixing coefficients,
53%and thus of the formulation used (see \autoref{chap:TD}).
54
55\begin{listing}
56  \nlst{namzdf}
57  \caption{\forcode{&namzdf}}
58  \label{lst:namzdf}
59\end{listing}
60
61%% =================================================================================================
62\subsection[Constant (\forcode{ln_zdfcst})]{Constant (\protect\np{ln_zdfcst}{ln\_zdfcst})}
63\label{subsec:ZDF_cst}
64
65Options are defined through the \nam{zdf}{zdf} namelist variables.
66When \np{ln_zdfcst}{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to
67constant values over the whole ocean.
68This is the crudest way to define the vertical ocean physics.
69It is recommended to use this option only in process studies, not in basin scale simulations.
70Typical values used in this case are:
71\begin{align*}
72  A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}   \\
73  A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
74\end{align*}
75
76These values are set through the \np{rn_avm0}{rn\_avm0} and \np{rn_avt0}{rn\_avt0} namelist parameters.
77In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity,
78that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and
79$\sim10^{-9}~m^2.s^{-1}$ for salinity.
80
81%% =================================================================================================
82\subsection[Richardson number dependent (\forcode{ln_zdfric})]{Richardson number dependent (\protect\np{ln_zdfric}{ln\_zdfric})}
83\label{subsec:ZDF_ric}
84
85\begin{listing}
86  \nlst{namzdf_ric}
87  \caption{\forcode{&namzdf_ric}}
88  \label{lst:namzdf_ric}
89\end{listing}
90
91When \np[=.true.]{ln_zdfric}{ln\_zdfric}, a local Richardson number dependent formulation for the vertical momentum and
92tracer eddy coefficients is set through the \nam{zdf_ric}{zdf\_ric} namelist variables.
93The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.
94\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures.
95The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to
96a dependency between the vertical eddy coefficients and the local Richardson number
97(\ie\ the ratio of stratification to vertical shear).
98Following \citet{pacanowski.philander_JPO81}, the following formulation has been implemented:
99\[
100  % \label{eq:ZDF_ric}
101  \left\{
102    \begin{aligned}
103      A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
104      A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
105    \end{aligned}
106  \right.
107\]
108where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number,
109$N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
110$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the constant case
111(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that
112can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$.
113The last three values can be modified by setting the \np{rn_avmri}{rn\_avmri}, \np{rn_alp}{rn\_alp} and
114\np{nn_ric}{nn\_ric} namelist parameters, respectively.
115
116A simple mixing-layer model to transfer and dissipate the atmospheric forcings
117(wind-stress and buoyancy fluxes) can be activated setting the \np[=.true.]{ln_mldw}{ln\_mldw} in the namelist.
118
119In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and
120the vertical eddy coefficients prescribed within this layer.
121
122This depth is assumed proportional to the "depth of frictional influence" that is limited by rotation:
123\[
124  h_{e} = Ek \frac {u^{*}} {f_{0}}
125\]
126where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis parameter.
127
128In this similarity height relationship, the turbulent friction velocity:
129\[
130  u^{*} = \sqrt \frac {|\tau|} {\rho_o}
131\]
132is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$.
133The final $h_{e}$ is further constrained by the adjustable bounds \np{rn_mldmin}{rn\_mldmin} and \np{rn_mldmax}{rn\_mldmax}.
134Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to
135the empirical values \np{rn_wtmix}{rn\_wtmix} and \np{rn_wvmix}{rn\_wvmix} \citep{lermusiaux_JMS01}.
136
137%% =================================================================================================
138\subsection[TKE turbulent closure scheme (\forcode{ln_zdftke})]{TKE turbulent closure scheme (\protect\np{ln_zdftke}{ln\_zdftke})}
139\label{subsec:ZDF_tke}
140
141\begin{listing}
142  \nlst{namzdf_tke}
143  \caption{\forcode{&namzdf_tke}}
144  \label{lst:namzdf_tke}
145\end{listing}
146
147The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on
148a prognostic equation for $\bar{e}$, the turbulent kinetic energy,
149and a closure assumption for the turbulent length scales.
150This turbulent closure model has been developed by \citet{bougeault.lacarrere_MWR89} in the atmospheric case,
151adapted by \citet{gaspar.gregoris.ea_JGR90} for the oceanic case, and embedded in OPA, the ancestor of \NEMO,
152by \citet{blanke.delecluse_JPO93} for equatorial Atlantic simulations.
153Since then, significant modifications have been introduced by \citet{madec.delecluse.ea_NPM98} in both the implementation and
154the formulation of the mixing length scale.
155The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear,
156its destruction through stratification, its vertical diffusion, and its dissipation of \citet{kolmogorov_IANS42} type:
157\begin{equation}
158  \label{eq:ZDF_tke_e}
159  \frac{\partial \bar{e}}{\partial t} =
160  \frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
161      +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
162  -K_\rho\,N^2
163  +\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
164      \;\frac{\partial \bar{e}}{\partial k}} \right]
165  - c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
166\end{equation}
167\[
168  % \label{eq:ZDF_tke_kz}
169  \begin{split}
170    K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }    \\
171    K_\rho &= A^{vm} / P_{rt}
172  \end{split}
173\]
174where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),
175$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,
176$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity and diffusivity coefficients.
177The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with
178vertical mixing at any depth \citep{gaspar.gregoris.ea_JGR90}.
179They are set through namelist parameters \np{nn_ediff}{nn\_ediff} and \np{nn_ediss}{nn\_ediss}.
180$P_{rt}$ can be set to unity or, following \citet{blanke.delecluse_JPO93}, be a function of the local Richardson number, $R_i$:
181\begin{align*}
182  % \label{eq:ZDF_prt}
183  P_{rt} =
184  \begin{cases}
185    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}   \\
186    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}   \\
187    \ \ 10 &      \text{if $\ 2 \leq R_i$}
188  \end{cases}
189\end{align*}
190The choice of $P_{rt}$ is controlled by the \np{nn_pdl}{nn\_pdl} namelist variable.
191
192At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as
193$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter.
194The default value of $e_{bb}$ is 3.75. \citep{gaspar.gregoris.ea_JGR90}), however a much larger value can be used when
195taking into account the surface wave breaking (see below \autoref{eq:ZDF_Esbc}).
196The bottom value of TKE is assumed to be equal to the value of the level just above.
197The time integration of the $\bar{e}$ equation may formally lead to negative values because
198the numerical scheme does not ensure its positivity.
199To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn_emin}{rn\_emin} namelist parameter).
200Following \citet{gaspar.gregoris.ea_JGR90}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$.
201This allows the subsequent formulations to match that of \citet{gargett_JMR84} for the diffusion in
202the thermocline and deep ocean :  $K_\rho = 10^{-3} / N$.
203In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with
204too weak vertical diffusion.
205They must be specified at least larger than the molecular values, and are set through \np{rn_avm0}{rn\_avm0} and
206\np{rn_avt0}{rn\_avt0} (\nam{zdf}{zdf} namelist, see \autoref{subsec:ZDF_cst}).
207
208%% =================================================================================================
209\subsubsection{Turbulent length scale}
210
211For computational efficiency, the original formulation of the turbulent length scales proposed by
212\citet{gaspar.gregoris.ea_JGR90} has been simplified.
213Four formulations are proposed, the choice of which is controlled by the \np{nn_mxl}{nn\_mxl} namelist parameter.
214The first two are based on the following first order approximation \citep{blanke.delecluse_JPO93}:
215\begin{equation}
216  \label{eq:ZDF_tke_mxl0_1}
217  l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N
218\end{equation}
219which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency.
220The resulting length scale is bounded by the distance to the surface or to the bottom
221(\np[=0]{nn_mxl}{nn\_mxl}) or by the local vertical scale factor (\np[=1]{nn_mxl}{nn\_mxl}).
222\citet{blanke.delecluse_JPO93} notice that this simplification has two major drawbacks:
223it makes no sense for locally unstable stratification and the computation no longer uses all
224the information contained in the vertical density profile.
225To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np[=2, 3]{nn_mxl}{nn\_mxl} cases,
226which add an extra assumption concerning the vertical gradient of the computed length scale.
227So, the length scales are first evaluated as in \autoref{eq:ZDF_tke_mxl0_1} and then bounded such that:
228\begin{equation}
229  \label{eq:ZDF_tke_mxl_constraint}
230  \frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
231  \qquad \text{with }\  l =  l_k = l_\epsilon
232\end{equation}
233\autoref{eq:ZDF_tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than
234the variations of depth.
235It provides a better approximation of the \citet{gaspar.gregoris.ea_JGR90} formulation while being much less
236time consuming.
237In particular, it allows the length scale to be limited not only by the distance to the surface or
238to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as
239the thermocline (\autoref{fig:ZDF_mixing_length}).
240In order to impose the \autoref{eq:ZDF_tke_mxl_constraint} constraint, we introduce two additional length scales:
241$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and
242evaluate the dissipation and mixing length scales as
243(and note that here we use numerical indexing):
244\begin{figure}[!t]
245  \centering
246  \includegraphics[width=0.66\textwidth]{ZDF_mixing_length}
247  \caption[Mixing length computation]{Illustration of the mixing length computation}
248  \label{fig:ZDF_mixing_length}
249\end{figure}
250\[
251  % \label{eq:ZDF_tke_mxl2}
252  \begin{aligned}
253    l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right)
254    \quad &\text{ from $k=1$ to $jpk$ }\ \\
255    l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)\right)
256    \quad &\text{ from $k=jpk$ to $1$ }\ \\
257  \end{aligned}
258\]
259where $l^{(k)}$ is computed using \autoref{eq:ZDF_tke_mxl0_1}, \ie\ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$.
260
261In the \np[=2]{nn_mxl}{nn\_mxl} case, the dissipation and mixing length scales take the same value:
262$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np[=3]{nn_mxl}{nn\_mxl} case,
263the dissipation and mixing turbulent length scales are give as in \citet{gaspar.gregoris.ea_JGR90}:
264\[
265  % \label{eq:ZDF_tke_mxl_gaspar}
266  \begin{aligned}
267    & l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }   \\
268    & l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
269  \end{aligned}
270\]
271
272At the ocean surface, a non zero length scale is set through the  \np{rn_mxl0}{rn\_mxl0} namelist parameter.
273Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and
274$z_o$ the roughness parameter of the surface.
275Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn_mxl0}{rn\_mxl0}.
276In the ocean interior a minimum length scale is set to recover the molecular viscosity when
277$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ).
278
279%% =================================================================================================
280\subsubsection{Surface wave breaking parameterization}
281
282Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to
283include the effect of surface wave breaking energetics.
284This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow.
285The \citet{mellor.blumberg_JPO04} modifications acts on surface length scale and TKE values and
286air-sea drag coefficient.
287The latter concerns the bulk formulae and is not discussed here.
288
289Following \citet{craig.banner_JPO94}, the boundary condition on surface TKE value is :
290\begin{equation}
291  \label{eq:ZDF_Esbc}
292  \bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
293\end{equation}
294where $\alpha_{CB}$ is the \citet{craig.banner_JPO94} constant of proportionality which depends on the ''wave age'',
295ranging from 57 for mature waves to 146 for younger waves \citep{mellor.blumberg_JPO04}.
296The boundary condition on the turbulent length scale follows the Charnock's relation:
297\begin{equation}
298  \label{eq:ZDF_Lsbc}
299  l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
300\end{equation}
301where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant.
302\citet{mellor.blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by
303\citet{stacey_JPO99} citing observation evidence, and
304$\alpha_{CB} = 100$ the Craig and Banner's value.
305As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,
306with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter, setting \np[=67.83]{rn_ebb}{rn\_ebb} corresponds
307to $\alpha_{CB} = 100$.
308Further setting  \np[=.true.]{ln_mxl0}{ln\_mxl0},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale,
309with $\beta$ hard coded to the Stacey's value.
310Note that a minimal threshold of \np{rn_emin0}{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the
311surface $\bar{e}$ value.
312
313%% =================================================================================================
314\subsubsection{Langmuir cells}
315
316Langmuir circulations (LC) can be described as ordered large-scale vertical motions in
317the surface layer of the oceans.
318Although LC have nothing to do with convection, the circulation pattern is rather similar to
319so-called convective rolls in the atmospheric boundary layer.
320The detailed physics behind LC is described in, for example, \citet{craik.leibovich_JFM76}.
321The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and
322wind drift currents.
323
324Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by
325\citep{axell_JGR02} for a $k-\epsilon$ turbulent closure.
326The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in
327an extra source term of TKE, $P_{LC}$.
328The presence of $P_{LC}$ in \autoref{eq:ZDF_tke_e}, the TKE equation, is controlled by setting \np{ln_lc}{ln\_lc} to
329\forcode{.true.} in the \nam{zdf_tke}{zdf\_tke} namelist.
330
331By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}),
332$P_{LC}$ is assumed to be :
333\[
334P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
335\]
336where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
337With no information about the wave field, $w_{LC}$ is assumed to be proportional to
338the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
339\footnote{Following \citet{li.garrett_JMR93}, the surface Stoke drift velocity may be expressed as
340  $u_s =  0.016 \,|U_{10m}|$.
341  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of
342  $1.5~10^{-3}$ give the expression used of $u_s$ as a function of the module of surface stress
343}.
344For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at
345a finite depth $H_{LC}$ (which is often close to the mixed layer depth),
346and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).
347The resulting expression for $w_{LC}$ is :
348\[
349  w_{LC}  =
350  \begin{cases}
351    c_{LC} \,u_s \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
352    0                             &      \text{otherwise}
353  \end{cases}
354\]
355where $c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data.
356The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second.
357The value of $c_{LC}$ is set through the \np{rn_lc}{rn\_lc} namelist parameter,
358having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}.
359
360The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
361$H_{LC}$ is the depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by
362converting its kinetic energy to potential energy, according to
363\[
364- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} u_s^2
365\]
366
367%% =================================================================================================
368\subsubsection{Mixing just below the mixed layer}
369
370Vertical mixing parameterizations commonly used in ocean general circulation models tend to
371produce mixed-layer depths that are too shallow during summer months and windy conditions.
372This bias is particularly acute over the Southern Ocean.
373To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{rodgers.aumont.ea_B14}.
374The parameterization is an empirical one, \ie\ not derived from theoretical considerations,
375but rather is meant to account for observed processes that affect the density structure of
376the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme
377(\ie\ near-inertial oscillations and ocean swells and waves).
378
379When using this parameterization (\ie\ when \np[=1]{nn_etau}{nn\_etau}),
380the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations,
381swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition,
382plus a depth depend one given by:
383\begin{equation}
384  \label{eq:ZDF_Ehtau}
385  S = (1-f_i) \; f_r \; e_s \; e^{-z / h_\tau}
386\end{equation}
387where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that
388penetrates in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of
389the penetration, and $f_i$ is the ice concentration
390(no penetration if $f_i=1$, \ie\ if the ocean is entirely covered by sea-ice).
391The value of $f_r$, usually a few percents, is specified through \np{rn_efr}{rn\_efr} namelist parameter.
392The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np[=0]{nn_etau}{nn\_etau}) or
393a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes
394(\np[=1]{nn_etau}{nn\_etau}).
395
396Note that two other option exist, \np[=2, 3]{nn_etau}{nn\_etau}.
397They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer,
398or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean.
399Those two options are obsolescent features introduced for test purposes.
400They will be removed in the next release.
401
402% This should be explain better below what this rn_eice parameter is meant for:
403In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn_eice}{rn\_eice} namelist parameter.
404This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration
405is greater than 25\%.
406
407% from Burchard et al OM 2008 :
408% the most critical process not reproduced by statistical turbulence models is the activity of
409% internal waves and their interaction with turbulence. After the Reynolds decomposition,
410% internal waves are in principle included in the RANS equations, but later partially
411% excluded by the hydrostatic assumption and the model resolution.
412% Thus far, the representation of internal wave mixing in ocean models has been relatively crude
413% (\eg\ Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002).
414
415%% =================================================================================================
416\subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls})]{GLS: Generic Length Scale (\protect\np{ln_zdfgls}{ln\_zdfgls})}
417\label{subsec:ZDF_gls}
418
419\begin{listing}
420  \nlst{namzdf_gls}
421  \caption{\forcode{&namzdf_gls}}
422  \label{lst:namzdf_gls}
423\end{listing}
424
425The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations:
426one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale,
427$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}.
428This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
429where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:ZDF_GLS} allows to recover a number of
430well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87},
431$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}).
432The GLS scheme is given by the following set of equations:
433\begin{equation}
434  \label{eq:ZDF_gls_e}
435  \frac{\partial \bar{e}}{\partial t} =
436  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
437      +\left( \frac{\partial v}{\partial k} \right)^2} \right]
438  -K_\rho \,N^2
439  +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
440  - \epsilon
441\end{equation}
442
443\[
444  % \label{eq:ZDF_gls_psi}
445  \begin{split}
446    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
447      \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
448          +\left( \frac{\partial v}{\partial k} \right)^2} \right]
449      - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
450    &+\frac{1}{e_3\;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
451        \;\frac{\partial \psi}{\partial k}} \right]\;
452  \end{split}
453\]
454
455\[
456  % \label{eq:ZDF_gls_kz}
457  \begin{split}
458    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
459    K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
460  \end{split}
461\]
462
463\[
464  % \label{eq:ZDF_gls_eps}
465  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
466\]
467where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and
468$\epsilon$ the dissipation rate.
469The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of
470the choice of the turbulence model.
471Four different turbulent models are pre-defined (\autoref{tab:ZDF_GLS}).
472They are made available through the \np{nn_clo}{nn\_clo} namelist parameter.
473
474\begin{table}[htbp]
475  \centering
476  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
477  \begin{tabular}{ccccc}
478    &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\
479    % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\
480    \hline
481    \hline
482    \np{nn_clo}{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\
483    \hline
484    $( p , n , m )$         &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
485    $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
486    $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
487    $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
488    $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
489    $C_3$              &      1.           &     1.              &      1.                &       1.           \\
490    $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
491    \hline
492    \hline
493  \end{tabular}
494  \caption[Set of predefined GLS parameters or equivalently predefined turbulence models available]{
495    Set of predefined GLS parameters, or equivalently predefined turbulence models available with
496    \protect\np[=.true.]{ln_zdfgls}{ln\_zdfgls} and controlled by
497    the \protect\np{nn_clos}{nn\_clos} namelist variable in \protect\nam{zdf_gls}{zdf\_gls}.}
498  \label{tab:ZDF_GLS}
499\end{table}
500
501In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of
502the mixing length towards $\kappa z_b$ ($\kappa$ is the Von Karman constant and $z_b$ the rugosity length scale) value near physical boundaries
503(logarithmic boundary layer law).
504$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88},
505or by \citet{kantha.clayson_JGR94} or one of the two functions suggested by \citet{canuto.howard.ea_JPO01}
506(\np[=0, 3]{nn_stab_func}{nn\_stab\_func}, resp.).
507The value of $C_{0\mu}$ depends on the choice of the stability function.
508
509The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or
510Neumann condition through \np{nn_bc_surf}{nn\_bc\_surf} and \np{nn_bc_bot}{nn\_bc\_bot}, resp.
511As for TKE closure, the wave effect on the mixing is considered when
512\np[ > 0.]{rn_crban}{rn\_crban} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}.
513The \np{rn_crban}{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and
514\np{rn_charn}{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.
515
516The $\psi$ equation is known to fail in stably stratified flows, and for this reason
517almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
518With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$.
519A value of $c_{lim} = 0.53$ is often used \citep{galperin.kantha.ea_JAS88}.
520\cite{umlauf.burchard_CSR05} show that the value of the clipping factor is of crucial importance for
521the entrainment depth predicted in stably stratified situations,
522and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes.
523The clipping is only activated if \np[=.true.]{ln_length_lim}{ln\_length\_lim},
524and the $c_{lim}$ is set to the \np{rn_clim_galp}{rn\_clim\_galp} value.
525
526The time and space discretization of the GLS equations follows the same energetic consideration as for
527the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{burchard_OM02}.
528Evaluation of the 4 GLS turbulent closure schemes can be found in \citet{warner.sherwood.ea_OM05} in ROMS model and
529 in \citet{reffray.bourdalle-badie.ea_GMD15} for the \NEMO\ model.
530
531% -------------------------------------------------------------------------------------------------------------
532%        OSM OSMOSIS BL Scheme
533% -------------------------------------------------------------------------------------------------------------
534\subsection[OSM: OSMOSIS boundary layer scheme (\forcode{ln_zdfosm = .true.})]
535{OSM: OSMOSIS boundary layer scheme (\protect\np{ln_zdfosm}{ln\_zdfosm})}
536\label{subsec:ZDF_osm}
537
538\begin{listing}
539  \nlst{namzdf_osm}
540  \caption{\forcode{&namzdf_osm}}
541  \label{lst:namzdf_osm}
542\end{listing}
543
544%--------------------------------------------------------------------------------------------------------------
545\paragraph{Namelist choices}
546Most of the namelist options refer to how to specify the Stokes
547surface drift and penetration depth. There are three options:
548\begin{description}
549  \item \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave} Default value in \texttt{namelist\_ref}. In this case the Stokes drift is
550      assumed to be parallel to the surface wind stress, with
551      magnitude consistent with a constant turbulent Langmuir number
552    $\mathrm{La}_t=$ \protect\np{rn_m_la} {rn\_m\_la} i.e.\
553    $u_{s0}=\tau/(\mathrm{La}_t^2\rho_0)$.  Default value of
554    \protect\np{rn_m_la}{rn\_m\_la} is 0.3. The Stokes penetration
555      depth $\delta = $ \protect\np{rn_osm_dstokes}{rn\_osm\_dstokes}; this has default value
556      of 5~m.
557
558  \item \protect\np[=1]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is
559      assumed to be parallel to the surface wind stress, with
560      magnitude as in the classical Pierson-Moskowitz wind-sea
561      spectrum.  Significant wave height and
562      wave-mean period taken from this spectrum are used to calculate the Stokes penetration
563      depth, following the approach set out in  \citet{breivik.janssen.ea_JPO14}.
564
565    \item \protect\np[=2]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is
566      taken from  ECMWF wave model output, though only the component parallel
567      to the wind stress is retained. Significant wave height and
568      wave-mean period from ECMWF wave model output are used to calculate the Stokes penetration
569      depth, again following \citet{breivik.janssen.ea_JPO14}.
570
571    \end{description}
572
573    Others refer to the treatment of diffusion and viscosity beneath
574    the surface boundary layer:
575\begin{description}
576   \item \protect\np{ln_kpprimix} {ln\_kpprimix}  Default is \np{.true.}. Switches on KPP-style Ri \#-dependent
577     mixing below the surface boundary layer. If this is set
578     \texttt{.true.}  the following variable settings are honoured:
579    \item \protect\np{rn_riinfty}{rn\_riinfty} Critical value of local Ri \# below which
580      shear instability increases vertical mixing from background value.
581    \item \protect\np{rn_difri} {rn\_difri} Maximum value of Ri \#-dependent mixing at $\mathrm{Ri}=0$.
582    \item \protect\np{ln_convmix}{ln\_convmix} If \texttt{.true.} then, where water column is unstable, specify
583       diffusivity equal to \protect\np{rn_dif_conv}{rn\_dif\_conv} (default value is 1 m~s$^{-2}$).
584 \end{description}
585 Diagnostic output is controlled by:
586  \begin{description}
587    \item \protect\np{ln_dia_osm}{ln\_dia\_osm} Default is \np{.false.}; allows XIOS output of OSMOSIS internal fields.
588  \end{description}
589Obsolete namelist parameters include:
590\begin{description}
591\item \protect\np{ln_use_osm_la}\np{ln\_use\_osm\_la} With \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave},
592  \protect\np{rn_osm_dstokes} {rn\_osm\_dstokes} is always used to specify the Stokes
593  penetration depth.
594\item \protect\np{nn_ave} {nn\_ave} Choice of averaging method for KPP-style Ri \#
595  mixing. Not taken account of.
596\item \protect\np{rn_osm_hbl0} {rn\_osm\_hbl0} Depth of initial boundary layer is now set
597  by a density criterion similar to that used in calculating \emph{hmlp} (output as \texttt{mldr10\_1}) in \mdl{zdfmxl}.
598\end{description}
599
600\subsubsection{Summary}
601Much of the time the turbulent motions in the ocean surface boundary
602layer (OSBL) are not given by
603classical shear turbulence. Instead they are in a regime known as
604`Langmuir turbulence',  dominated by an
605interaction between the currents and the Stokes drift of the surface waves \citep[e.g.][]{mcwilliams.sullivan.ea_JFM97}.
606This regime is characterised by strong vertical turbulent motion, and appears when the surface Stokes drift $u_{s0}$ is much greater than the friction velocity $u_{\ast}$. More specifically Langmuir turbulence is thought to be crucial where the turbulent Langmuir number $\mathrm{La}_{t}=(u_{\ast}/u_{s0}) > 0.4$.
607
608The OSMOSIS model is fundamentally based on results of Large Eddy
609Simulations (LES) of Langmuir turbulence and aims to fully describe
610this Langmuir regime. The description in this section is of necessity incomplete and further details are available in Grant. A (2019); in prep.
611
612The OSMOSIS turbulent closure scheme is a similarity-scale scheme in
613the same spirit as the K-profile
614parameterization (KPP) scheme of \citet{large.mcwilliams.ea_RG94}.
615A specified shape of diffusivity, scaled by the (OSBL) depth
616$h_{\mathrm{BL}}$ and a turbulent velocity scale, is imposed throughout the
617boundary layer
618$-h_{\mathrm{BL}}<z<\eta$. The turbulent closure model
619also includes fluxes of tracers and momentum that are``non-local'' (independent of the local property gradient).
620
621Rather than the OSBL
622depth being diagnosed in terms of a bulk Richardson number criterion,
623as in KPP, it is set by a prognostic equation that is informed by
624energy budget considerations reminiscent of the classical mixed layer
625models of \citet{kraus.turner_T67}.
626The model also includes an explicit parametrization of the structure
627of the pycnocline (the stratified region at the bottom of the OSBL).
628
629Presently, mixing below the OSBL is handled by the Richardson
630number-dependent mixing scheme used in \citet{large.mcwilliams.ea_RG94}.
631
632Convective parameterizations such as described in \autoref{sec:ZDF_conv}
633below should not be used with the OSMOSIS-OBL model: instabilities
634within the OSBL are part of the model, while instabilities below the
635ML are handled by the Ri \# dependent scheme.
636
637\subsubsection{Depth and velocity scales}
638The model supposes a boundary layer of thickness $h_{\mathrm{bl}}$ enclosing a well-mixed layer of thickness $h_{\mathrm{ml}}$ and a relatively thin pycnocline at the base of thickness $\Delta h$; \autoref{fig:OSBL_structure} shows typical (a) buoyancy structure and (b) turbulent buoyancy flux profile for the unstable boundary layer (losing buoyancy at the surface; e.g.\ cooling).
639\begin{figure}[!t]
640  \begin{center}
641    %\includegraphics[width=0.7\textwidth]{ZDF_OSM_structure_of_OSBL}
642    \caption{
643      \protect\label{fig:OSBL_structure}
644     The structure of the entraining  boundary layer. (a) Mean buoyancy profile. (b) Profile of the buoyancy flux.
645    }
646  \end{center}
647\end{figure}
648The pycnocline in the OSMOSIS scheme is assumed to have a finite thickness, and may include a number of model levels. This means that the OSMOSIS scheme must parametrize both the thickness of the pycnocline, and the turbulent fluxes within the pycnocline.
649
650Consideration of the power input by wind acting on the Stokes drift suggests that the Langmuir turbulence has velocity scale:
651\begin{equation}
652  \label{eq:ZDF_w_La}
653  w_{*L}= \left(u_*^2 u_{s\,0}\right)^{1/3};
654\end{equation}
655but at times the Stokes drift may be weak due to e.g.\ ice cover, short fetch, misalignment with the surface stress, etc.\ so  a composite velocity scale is assumed for the stable (warming) boundary layer:
656\begin{equation}
657  \label{eq:ZDF_composite-nu}
658  \nu_{\ast}= \left\{ u_*^3 \left[1-\exp(-.5 \mathrm{La}_t^2)\right]+w_{*L}^3\right\}^{1/3}.
659\end{equation}
660For the unstable boundary layer this is merged with the standard convective velocity scale $w_{*C}=\left(\overline{w^\prime b^\prime}_0 \,h_\mathrm{ml}\right)^{1/3}$, where $\overline{w^\prime b^\prime}_0$ is the upwards surface buoyancy flux, to give:
661\begin{equation}
662  \label{eq:ZDF_vel-scale-unstable}
663  \omega_* = \left(\nu_*^3 + 0.5 w_{*C}^3\right)^{1/3}.
664\end{equation}
665
666\subsubsection{The flux gradient model}
667The flux-gradient relationships used in the OSMOSIS scheme take the form:
668
669\begin{equation}
670  \label{eq:ZDF_flux-grad-gen}
671  \overline{w^\prime\chi^\prime}=-K\frac{\partial\overline{\chi}}{\partial z} + N_{\chi,s} +N_{\chi,b} +N_{\chi,t},
672\end{equation}
673
674where $\chi$ is a general variable and $N_{\chi,s}, N_{\chi,b} \mathrm{and} N_{\chi,t}$  are the non-gradient terms, and represent the effects of the different terms in the turbulent flux-budget on the transport of $\chi$. $N_{\chi,s}$ represents the effects that the Stokes shear has on the transport of $\chi$, $N_{\chi,b}$  the effect of buoyancy, and $N_{\chi,t}$ the effect of the turbulent transport.  The same general form for the flux-gradient relationship is used to parametrize the transports of momentum, heat and salinity.
675
676In terms of the non-dimensionalized depth variables
677
678\begin{equation}
679  \label{eq:ZDF_sigma}
680  \sigma_{\mathrm{ml}}= -z/h_{\mathrm{ml}}; \;\sigma_{\mathrm{bl}}= -z/h_{\mathrm{bl}},
681\end{equation}
682
683in unstable conditions the eddy diffusivity ($K_d$) and eddy viscosity ($K_\nu$) profiles are parametrized as:
684
685\begin{align}
686  \label{eq:ZDF_diff-unstable}
687  K_d=&0.8\, \omega_*\, h_{\mathrm{ml}} \, \sigma_{\mathrm{ml}} \left(1-\beta_d \sigma_{\mathrm{ml}}\right)^{3/2}
688  \\
689  \label{eq:ZDF_visc-unstable}
690  K_\nu =& 0.3\, \omega_* \,h_{\mathrm{ml}}\, \sigma_{\mathrm{ml}} \left(1-\beta_\nu \sigma_{\mathrm{ml}}\right)\left(1-\tfrac{1}{2}\sigma_{\mathrm{ml}}^2\right)
691\end{align}
692
693where $\beta_d$ and $\beta_\nu$ are parameters that are determined by matching \autoref{eq:ZDF_diff-unstable} and \autoref{eq:ZDF_visc-unstable} to the eddy diffusivity and viscosity at the base of the well-mixed layer, given by
694
695\begin{equation}
696  \label{eq:ZDF_diff-wml-base}
697  K_{d,\mathrm{ml}}=K_{\nu,\mathrm{ml}}=\,0.16\,\omega_* \Delta h.
698\end{equation}
699
700For stable conditions the eddy diffusivity/viscosity profiles are given by:
701
702\begin{align}
703  \label{eq:ZDF_diff-stable}
704  K_d= & 0.75\,\, \nu_*\, h_{\mathrm{ml}}\,\,  \exp\left[-2.8
705       \left(h_{\mathrm{bl}}/L_L\right)^2\right]\sigma_{\mathrm{ml}}
706       \left(1-\sigma_{\mathrm{ml}}\right)^{3/2} \\
707  \label{eq:ZDF_visc-stable}
708  K_\nu = & 0.375\,\,  \nu_*\, h_{\mathrm{ml}} \,\, \exp\left[-2.8 \left(h_{\mathrm{bl}}/L_L\right)^2\right] \sigma_{\mathrm{ml}} \left(1-\sigma_{\mathrm{ml}}\right)\left(1-\tfrac{1}{2}\sigma_{\mathrm{ml}}^2\right).
709\end{align}
710
711The shape of the eddy viscosity and diffusivity profiles is the same as the shape in the unstable OSBL. The eddy diffusivity/viscosity depends on the stability parameter $h_{\mathrm{bl}}/{L_L}$ where $ L_L$ is analogous to the Obukhov length, but for Langmuir turbulence:
712\begin{equation}
713  \label{eq:ZDF_L_L}
714  L_L=-w_{*L}^3/\left<\overline{w^\prime b^\prime}\right>_L,
715\end{equation}
716with the mean turbulent buoyancy flux averaged over the boundary layer given in terms of its surface value $\overline{w^\prime b^\prime}_0$ and (downwards) )solar irradiance $I(z)$ by
717\begin{equation}
718  \label{eq:ZDF_stable-av-buoy-flux}
719  \left<\overline{w^\prime b^\prime}\right>_L = \tfrac{1}{2} {\overline{w^\prime b^\prime}}_0-g\alpha_E\left[\tfrac{1}{2}(I(0)+I(-h))-\left<I\right>\right].
720\end{equation}
721
722In unstable conditions the eddy diffusivity and viscosity depend on stability through the velocity scale $\omega_*$, which depends on the two velocity scales $\nu_*$ and $w_{*C}$.
723
724Details of the non-gradient terms in \autoref{eq:ZDF_flux-grad-gen} and of the fluxes within the pycnocline $-h_{\mathrm{bl}}<z<h_{\mathrm{ml}}$ can be found in Grant (2019).
725
726\subsubsection{Evolution of the boundary layer depth}
727
728The prognostic equation for the depth of the neutral/unstable boundary layer is given by \iffalse \citep{grant+etal18?}, \fi
729
730\begin{equation}
731  \label{eq:ZDF_dhdt-unstable}
732%\frac{\partial h_\mathrm{bl}}{\partial t} + \mathbf{U}_b\cdot\nabla h_\mathrm{bl}= W_b - \frac{{\overline{w^\prime b^\prime}}_\mathrm{ent}}{\Delta B_\mathrm{bl}}
733   \frac{\partial h_\mathrm{bl}}{\partial t} = W_b - \frac{{\overline{w^\prime b^\prime}}_\mathrm{ent}}{\Delta B_\mathrm{bl}}
734\end{equation}
735where $h_\mathrm{bl}$ is the horizontally-varying depth of the OSBL,
736$\mathbf{U}_b$ and $W_b$ are the mean horizontal and vertical
737velocities at the base of the OSBL, ${\overline{w^\prime
738    b^\prime}}_\mathrm{ent}$ is the buoyancy flux due to entrainment
739and $\Delta B_\mathrm{bl}$ is the difference between the buoyancy
740averaged over the depth of the OSBL (i.e.\ including the ML and
741pycnocline) and the buoyancy just below the base of the OSBL. This
742equation for the case when the pycnocline has a finite thickness,
743based on the potential energy budget of the OSBL, is the leading term
744\iffalse \citep{grant+etal18?} \fi of a generalization of that used in mixed-layer
745models e.g.\ \citet{kraus.turner_T67}, in which the thickness of the pycnocline is taken to be zero.
746
747The entrainment flux for the combination of convective and Langmuir turbulence is given by
748\begin{equation}
749  \label{eq:ZDF_entrain-flux}
750  {\overline{w^\prime b^\prime}}_\mathrm{ent} = -\alpha_{\mathrm{B}} {\overline{w^\prime b^\prime}}_0 - \alpha_{\mathrm{S}} \frac{u_*^3}{h_{\mathrm{ml}}}
751  + G\left(\delta/h_{\mathrm{ml}} \right)\left[\alpha_{\mathrm{S}}e^{-1.5\, \mathrm{La}_t}-\alpha_{\mathrm{L}} \frac{w_{\mathrm{*L}}^3}{h_{\mathrm{ml}}}\right]
752\end{equation}
753where the factor $G\equiv 1 - \mathrm{e}^ {-25\delta/h_{\mathrm{bl}}}(1-4\delta/h_{\mathrm{bl}})$ models the lesser efficiency of Langmuir mixing when the boundary-layer depth is much greater than the Stokes depth, and $\alpha_{\mathrm{B}}$, $\alpha_{S}$  and $\alpha_{\mathrm{L}}$ depend on the ratio of the appropriate eddy turnover time to the inertial timescale $f^{-1}$. Results from the LES suggest $\alpha_{\mathrm{B}}=0.18 F(fh_{\mathrm{bl}}/w_{*C})$, $\alpha_{S}=0.15 F(fh_{\mathrm{bl}}/u_*$  and $\alpha_{\mathrm{L}}=0.035 F(fh_{\mathrm{bl}}/u_{*L})$, where $F(x)\equiv\tanh(x^{-1})^{0.69}$.
754
755For the stable boundary layer, the equation for the depth of the OSBL is:
756
757\begin{equation}
758  \label{eq:ZDF_dhdt-stable}
759\max\left(\Delta B_{bl},\frac{w_{*L}^2}{h_\mathrm{bl}}\right)\frac{\partial h_\mathrm{bl}}{\partial t} = \left(0.06 + 0.52\,\frac{ h_\mathrm{bl}}{L_L}\right) \frac{w_{*L}^3}{h_\mathrm{bl}} +\left<\overline{w^\prime b^\prime}\right>_L.
760\end{equation}
761
762\autoref{eq:ZDF_dhdt-unstable} always leads to the depth of the entraining OSBL increasing (ignoring the effect of the mean vertical motion), but the change in the thickness of the stable OSBL given by \autoref{eq:ZDF_dhdt-stable} can be positive or negative, depending on the magnitudes of $\left<\overline{w^\prime b^\prime}\right>_L$ and $h_\mathrm{bl}/L_L$. The rate at which the depth of the OSBL can decrease is limited by choosing an effective buoyancy $w_{*L}^2/h_\mathrm{bl}$, in place of $\Delta B_{bl}$ which will be $\approx 0$ for the collapsing OSBL.
763
764
765%% =================================================================================================
766\subsection[ Discrete energy conservation for TKE and GLS schemes]{Discrete energy conservation for TKE and GLS schemes}
767\label{subsec:ZDF_tke_ene}
768
769\begin{figure}[!t]
770  \centering
771  \includegraphics[width=0.66\textwidth]{ZDF_TKE_time_scheme}
772  \caption[Subgrid kinetic energy integration in GLS and TKE schemes]{
773    Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and
774    its links to the momentum and tracer time integration.}
775  \label{fig:ZDF_TKE_time_scheme}
776\end{figure}
777
778The production of turbulence by vertical shear (the first term of the right hand side of
779\autoref{eq:ZDF_tke_e}) and  \autoref{eq:ZDF_gls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion
780(first line in \autoref{eq:MB_zdf}).
781To do so a special care has to be taken for both the time and space discretization of
782the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}.
783
784Let us first address the time stepping issue. \autoref{fig:ZDF_TKE_time_scheme} shows how
785the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with
786the one-level forward time stepping of the equation for $\bar{e}$.
787With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to
788the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and
789summing the result vertically:
790\begin{equation}
791  \label{eq:ZDF_energ1}
792  \begin{split}
793    \int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
794    &= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}
795    - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
796  \end{split}
797\end{equation}
798Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$,
799known at time $t$ (\autoref{fig:ZDF_TKE_time_scheme}), as it is required when using the TKE scheme
800(see \autoref{sec:TD_forward_imp}).
801The first term of the right hand side of \autoref{eq:ZDF_energ1} represents the kinetic energy transfer at
802the surface (atmospheric forcing) and at the bottom (friction effect).
803The second term is always negative.
804It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$.
805\autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
806the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
807${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$
808(and not by the more straightforward $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
809
810A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
811(second term of the right hand side of \autoref{eq:ZDF_tke_e} and \autoref{eq:ZDF_gls_e}).
812This term must balance the input of potential energy resulting from vertical mixing.
813The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by
814multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically:
815\begin{equation}
816  \label{eq:ZDF_energ2}
817  \begin{split}
818    \int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
819    &= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta}
820    - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
821    &= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
822    + \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
823  \end{split}
824\end{equation}
825where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
826The first term of the right hand side of \autoref{eq:ZDF_energ2} is always zero because
827there is no diffusive flux through the ocean surface and bottom).
828The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
829Therefore \autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
830the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:ZDF_tke_e} and  \autoref{eq:ZDF_gls_e}.
831
832Let us now address the space discretization issue.
833The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in
834the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:DOM_cell}).
835A space averaging is thus required to obtain the shear TKE production term.
836By redoing the \autoref{eq:ZDF_energ1} in the 3D case, it can be shown that the product of eddy coefficient by
837the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
838Furthermore, the time variation of $e_3$ has be taken into account.
839
840The above energetic considerations leads to the following final discrete form for the TKE equation:
841\begin{equation}
842  \label{eq:ZDF_tke_ene}
843  \begin{split}
844    \frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv
845    \Biggl\{ \Biggr.
846    &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} }
847        \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
848    +&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} }
849        \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j}
850    \Biggr. \Biggr\}   \\
851    %
852    - &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
853    %
854    +&\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]   \\
855    %
856    - &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
857  \end{split}
858\end{equation}
859where the last two terms in \autoref{eq:ZDF_tke_ene} (vertical diffusion and Kolmogorov dissipation)
860are time stepped using a backward scheme (see\autoref{sec:TD_forward_imp}).
861Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible.
862%The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as
863%they all appear in the right hand side of \autoref{eq:ZDF_tke_ene}.
864%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
865
866%% =================================================================================================
867\section{Convection}
868\label{sec:ZDF_conv}
869
870Static instabilities (\ie\ light potential densities under heavy ones) may occur at particular ocean grid points.
871In nature, convective processes quickly re-establish the static stability of the water column.
872These processes have been removed from the model via the hydrostatic assumption so they must be parameterized.
873Three parameterisations are available to deal with convective processes:
874a non-penetrative convective adjustment or an enhanced vertical diffusion,
875or/and the use of a turbulent closure scheme.
876
877%% =================================================================================================
878\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc})]{Non-penetrative convective adjustment (\protect\np{ln_tranpc}{ln\_tranpc})}
879\label{subsec:ZDF_npc}
880
881\begin{figure}[!htb]
882  \centering
883  \includegraphics[width=0.66\textwidth]{ZDF_npc}
884  \caption[Unstable density profile treated by the non penetrative convective adjustment algorithm]{
885    Example of an unstable density profile treated by
886    the non penetrative convective adjustment algorithm.
887    $1^{st}$ step: the initial profile is checked from the surface to the bottom.
888    It is found to be unstable between levels 3 and 4.
889    They are mixed.
890    The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed.
891    The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed.
892    The $1^{st}$ step ends since the density profile is then stable below the level 3.
893    $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step:
894    levels 2 to 5 are mixed.
895    The new density profile is checked.
896    It is found stable: end of algorithm.}
897  \label{fig:ZDF_npc}
898\end{figure}
899
900Options are defined through the \nam{zdf}{zdf} namelist variables.
901The non-penetrative convective adjustment is used when \np[=.true.]{ln_zdfnpc}{ln\_zdfnpc}.
902It is applied at each \np{nn_npc}{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of
903the water column, but only until the density structure becomes neutrally stable
904(\ie\ until the mixed portion of the water column has \textit{exactly} the density of the water just below)
905\citep{madec.delecluse.ea_JPO91}.
906The associated algorithm is an iterative process used in the following way (\autoref{fig:ZDF_npc}):
907starting from the top of the ocean, the first instability is found.
908Assume in the following that the instability is located between levels $k$ and $k+1$.
909The temperature and salinity in the two levels are vertically mixed, conserving the heat and salt contents of
910the water column.
911The new density is then computed by a linear approximation.
912If the new density profile is still unstable between levels $k+1$ and $k+2$,
913levels $k$, $k+1$ and $k+2$ are then mixed.
914This process is repeated until stability is established below the level $k$
915(the mixing process can go down to the ocean bottom).
916The algorithm is repeated to check if the density profile between level $k-1$ and $k$ is unstable and/or
917if there is no deeper instability.
918
919This algorithm is significantly different from mixing statically unstable levels two by two.
920The latter procedure cannot converge with a finite number of iterations for some vertical profiles while
921the algorithm used in \NEMO\ converges for any profile in a number of iterations which is less than
922the number of vertical levels.
923This property is of paramount importance as pointed out by \citet{killworth_iprc89}:
924it avoids the existence of permanent and unrealistic static instabilities at the sea surface.
925This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in
926the north-western Mediterranean Sea \citep{madec.delecluse.ea_JPO91, madec.chartier.ea_DAO91, madec.crepon_iprc91}.
927
928The current implementation has been modified in order to deal with any non linear equation of seawater
929(L. Brodeau, personnal communication).
930Two main differences have been introduced compared to the original algorithm:
931$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency
932(not the difference in potential density);
933$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in
934the same way their temperature and salinity has been mixed.
935These two modifications allow the algorithm to perform properly and accurately with TEOS10 or EOS-80 without
936having to recompute the expansion coefficients at each mixing iteration.
937
938%% =================================================================================================
939\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd})]{Enhanced vertical diffusion (\protect\np{ln_zdfevd}{ln\_zdfevd})}
940\label{subsec:ZDF_evd}
941
942Options are defined through the  \nam{zdf}{zdf} namelist variables.
943The enhanced vertical diffusion parameterisation is used when \np[=.true.]{ln_zdfevd}{ln\_zdfevd}.
944In this case, the vertical eddy mixing coefficients are assigned very large values
945in regions where the stratification is unstable
946(\ie\ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}.
947This is done either on tracers only (\np[=0]{nn_evdm}{nn\_evdm}) or
948on both momentum and tracers (\np[=1]{nn_evdm}{nn\_evdm}).
949
950In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np[=1]{nn_evdm}{nn\_evdm},
951the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to
952the namelist parameter \np{rn_avevd}{rn\_avevd}.
953A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$.
954This parameterisation of convective processes is less time consuming than
955the convective adjustment algorithm presented above when mixing both tracers and
956momentum in the case of static instabilities.
957
958Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$.
959This removes a potential source of divergence of odd and even time step in
960a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:TD_mLF}).
961
962%% =================================================================================================
963\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}})}
964\label{subsec:ZDF_tcs}
965
966The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and
967\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,
968with statically unstable density profiles.
969In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in
970\autoref{eq:ZDF_tke_e} or \autoref{eq:ZDF_gls_e} becomes a source term, since $N^2$ is negative.
971It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at
972velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$).
973These large values restore the static stability of the water column in a way similar to that of
974the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}).
975However, in the vicinity of the sea surface (first ocean layer), the eddy coefficients computed by
976the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
977because the mixing length scale is bounded by the distance to the sea surface.
978It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme,
979\ie\ setting the \np{ln_zdfnpc}{ln\_zdfnpc} namelist parameter to true and
980defining the turbulent closure (\np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} = \forcode{.true.}) all together.
981
982The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection,
983%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp},
984therefore \np[=.false.]{ln_zdfevd}{ln\_zdfevd} should be used with the OSMOSIS scheme.
985% gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
986
987%% =================================================================================================
988\section[Double diffusion mixing (\forcode{ln_zdfddm})]{Double diffusion mixing (\protect\np{ln_zdfddm}{ln\_zdfddm})}
989\label{subsec:ZDF_ddm}
990
991%\nlst{namzdf_ddm}
992
993This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter
994\np{ln_zdfddm}{ln\_zdfddm} in \nam{zdf}{zdf}.
995Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa.
996The former condition leads to salt fingering and the latter to diffusive convection.
997Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean.
998\citet{merryfield.holloway.ea_JPO99} include a parameterisation of such phenomena in a global ocean model and show that
999it leads to relatively minor changes in circulation but exerts significant regional influences on
1000temperature and salinity.
1001
1002Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
1003\begin{align*}
1004  % \label{eq:ZDF_ddm_Kz}
1005  &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT} \\
1006  &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
1007\end{align*}
1008where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
1009and $o$ by processes other than double diffusion.
1010The rates of double-diffusive mixing depend on the buoyancy ratio
1011$R_\rho = \alpha \partial_z T / \beta \partial_z S$, where $\alpha$ and $\beta$ are coefficients of
1012thermal expansion and saline contraction (see \autoref{subsec:TRA_eos}).
1013To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt
1014(1981):
1015\begin{align}
1016  \label{eq:ZDF_ddm_f}
1017  A_f^{vS} &=
1018             \begin{cases}
1019               \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
1020               0                              &\text{otherwise}
1021             \end{cases}
1022  \\         \label{eq:ZDF_ddm_f_T}
1023  A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho
1024\end{align}
1025
1026\begin{figure}[!t]
1027  \centering
1028  \includegraphics[width=0.66\textwidth]{ZDF_ddm}
1029  \caption[Diapycnal diffusivities for temperature and salt in regions of salt fingering and
1030  diffusive convection]{
1031    From \citet{merryfield.holloway.ea_JPO99}:
1032    (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in
1033    regions of salt fingering.
1034    Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and
1035    thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$;
1036    (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in
1037    regions of diffusive convection.
1038    Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation.
1039    The latter is not implemented in \NEMO.}
1040  \label{fig:ZDF_ddm}
1041\end{figure}
1042
1043The factor 0.7 in \autoref{eq:ZDF_ddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of
1044buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}).
1045Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
1046
1047To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by
1048Federov (1988) is used:
1049\begin{align}
1050  % \label{eq:ZDF_ddm_d}
1051  A_d^{vT} &=
1052             \begin{cases}
1053               1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
1054               &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
1055               0                       &\text{otherwise}
1056             \end{cases}
1057                                       \nonumber \\
1058  \label{eq:ZDF_ddm_d_S}
1059  A_d^{vS} &=
1060             \begin{cases}
1061               A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right) &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
1062               A_d^{vT} \ 0.15 \ R_\rho               &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
1063               0                       &\text{otherwise}
1064             \end{cases}
1065\end{align}
1066
1067The dependencies of \autoref{eq:ZDF_ddm_f} to \autoref{eq:ZDF_ddm_d_S} on $R_\rho$ are illustrated in
1068\autoref{fig:ZDF_ddm}.
1069Implementing this requires computing $R_\rho$ at each grid point on every time step.
1070This is done in \mdl{eosbn2} at the same time as $N^2$ is computed.
1071This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive).
1072
1073%% =================================================================================================
1074\section[Bottom and top friction (\textit{zdfdrg.F90})]{Bottom and top friction (\protect\mdl{zdfdrg})}
1075\label{sec:ZDF_drg}
1076
1077\begin{listing}
1078  \nlst{namdrg}
1079  \caption{\forcode{&namdrg}}
1080  \label{lst:namdrg}
1081\end{listing}
1082
1083\begin{listing}
1084  \nlst{namdrg_top}
1085  \caption{\forcode{&namdrg_top}}
1086  \label{lst:namdrg_top}
1087\end{listing}
1088
1089\begin{listing}
1090  \nlst{namdrg_bot}
1091  \caption{\forcode{&namdrg_bot}}
1092  \label{lst:namdrg_bot}
1093\end{listing}
1094
1095Options to define the top and bottom friction are defined through the \nam{drg}{drg} namelist variables.
1096The bottom friction represents the friction generated by the bathymetry.
1097The top friction represents the friction generated by the ice shelf/ocean interface.
1098As the friction processes at the top and the bottom are treated in and identical way,
1099the description below considers mostly the bottom friction case, if not stated otherwise.
1100
1101Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as
1102a condition on the vertical diffusive flux.
1103For the bottom boundary layer, one has:
1104 \[
1105   % \label{eq:ZDF_bfr_flux}
1106   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
1107 \]
1108where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside
1109the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean).
1110How ${\cal F}_h^{\textbf U}$ influences the interior depends on the vertical resolution of the model near
1111the bottom relative to the Ekman layer depth.
1112For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m,
1113one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$
1114(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$).
1115With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
1116When the vertical mixing coefficient is this small, using a flux condition is equivalent to
1117entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or
1118bottom model layer.
1119To illustrate this, consider the equation for $u$ at $k$, the last ocean level:
1120\begin{equation}
1121  \label{eq:ZDF_drg_flux2}
1122  \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}}
1123\end{equation}
1124If the bottom layer thickness is 200~m, the Ekman transport will be distributed over that depth.
1125On the other hand, if the vertical resolution is high (1~m or less) and a turbulent closure model is used,
1126the turbulent Ekman layer will be represented explicitly by the model.
1127However, the logarithmic layer is never represented in current primitive equation model applications:
1128it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
1129Two choices are available in \NEMO: a linear and a quadratic bottom friction.
1130Note that in both cases, the rotation between the interior velocity and the bottom friction is neglected in
1131the present release of \NEMO.
1132
1133In the code, the bottom friction is imposed by adding the trend due to the bottom friction to
1134 the general momentum trend in \mdl{dynzdf}.
1135For the time-split surface pressure gradient algorithm, the momentum trend due to
1136the barotropic component needs to be handled separately.
1137For this purpose it is convenient to compute and store coefficients which can be simply combined with
1138bottom velocities and geometric values to provide the momentum trend due to bottom friction.
1139 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where:
1140\begin{equation}
1141  \label{eq:ZDF_bfr_bdef}
1142  \frac{\partial {\textbf U_h}}{\partial t} =
1143  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
1144\end{equation}
1145where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
1146Note 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.
1147
1148%% =================================================================================================
1149\subsection[Linear top/bottom friction (\forcode{ln_lin})]{Linear top/bottom friction (\protect\np{ln_lin}{ln\_lin})}
1150\label{subsec:ZDF_drg_linear}
1151
1152The linear friction parameterisation (including the special case of a free-slip condition) assumes that
1153the friction is proportional to the interior velocity (\ie\ the velocity of the first/last model level):
1154\[
1155  % \label{eq:ZDF_bfr_linear}
1156  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
1157\]
1158where $r$ is a friction coefficient expressed in $m s^{-1}$.
1159This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,
1160and setting $r = H / \tau$, where $H$ is the ocean depth.
1161Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{weatherly_JMR84}.
1162A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models.
1163One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$
1164(\citet{gill_bk82}, Eq. 9.6.6).
1165For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$,
1166and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
1167This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days.
1168It can be changed by specifying \np{rn_Uc0}{rn\_Uc0} (namelist parameter).
1169
1170 For the linear friction case the drag coefficient used in the general expression \autoref{eq:ZDF_bfr_bdef} is:
1171\[
1172  % \label{eq:ZDF_bfr_linbfr_b}
1173    c_b^T = - r
1174\]
1175When \np[=.true.]{ln_lin}{ln\_lin}, the value of $r$ used is \np{rn_Uc0}{rn\_Uc0}*\np{rn_Cd0}{rn\_Cd0}.
1176Setting \np[=.true.]{ln_drg_OFF}{ln\_drg\_OFF} (and \forcode{ln_lin=.true.}) is equivalent to setting $r=0$ and leads to a free-slip boundary condition.
1177
1178These values are assigned in \mdl{zdfdrg}.
1179Note that there is support for local enhancement of these values via an externally defined 2D mask array
1180(\np[=.true.]{ln_boost}{ln\_boost}) given in the \textit{bfr\_coef.nc} input NetCDF file.
1181The mask values should vary from 0 to 1.
1182Locations with a non-zero mask value will have the friction coefficient increased by
1183$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
1184
1185%% =================================================================================================
1186\subsection[Non-linear top/bottom friction (\forcode{ln_non_lin})]{Non-linear top/bottom friction (\protect\np{ln_non_lin}{ln\_non\_lin})}
1187\label{subsec:ZDF_drg_nonlinear}
1188
1189The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic:
1190\[
1191  % \label{eq:ZDF_drg_nonlinear}
1192  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
1193  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
1194\]
1195where $C_D$ is a drag coefficient, and $e_b $ a top/bottom turbulent kinetic energy due to tides,
1196internal waves breaking and other short time scale currents.
1197A typical value of the drag coefficient is $C_D = 10^{-3} $.
1198As an example, the CME experiment \citep{treguier_JGR92} uses $C_D = 10^{-3}$ and
1199$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and
1200$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
1201The CME choices have been set as default values (\np{rn_Cd0}{rn\_Cd0} and \np{rn_ke0}{rn\_ke0} namelist parameters).
1202
1203As for the linear case, the friction is imposed in the code by adding the trend due to
1204the friction to the general momentum trend in \mdl{dynzdf}.
1205For the non-linear friction case the term computed in \mdl{zdfdrg} is:
1206\[
1207  % \label{eq:ZDF_drg_nonlinbfr}
1208    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2}
1209\]
1210
1211The coefficients that control the strength of the non-linear friction are initialised as namelist parameters:
1212$C_D$= \np{rn_Cd0}{rn\_Cd0}, and $e_b$ =\np{rn_bfeb2}{rn\_bfeb2}.
1213Note 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
1214(\np[=.true.]{ln_boost}{ln\_boost}).
1215This works in the same way as for the linear friction case with non-zero masked locations increased by
1216$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
1217
1218%% =================================================================================================
1219\subsection[Log-layer top/bottom friction (\forcode{ln_loglayer})]{Log-layer top/bottom friction (\protect\np{ln_loglayer}{ln\_loglayer})}
1220\label{subsec:ZDF_drg_loglayer}
1221
1222In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using
1223a "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.
1224If  \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):
1225\[
1226  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2
1227\]
1228
1229\noindent where $\kappa$ is the von-Karman constant and \np{rn_z0}{rn\_z0} is a roughness length provided via the namelist.
1230
1231The drag coefficient is bounded such that it is kept greater or equal to
1232the 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:
1233\np{rn_Cdmax}{rn\_Cdmax}, \ie
1234\[
1235  rn\_Cd0 \leq C_D \leq rn\_Cdmax
1236\]
1237
1238\noindent The log-layer enhancement can also be applied to the top boundary friction if
1239under ice-shelf cavities are activated (\np[=.true.]{ln_isfcav}{ln\_isfcav}).
1240%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}.
1241
1242%% =================================================================================================
1243\subsection[Explicit top/bottom friction (\forcode{ln_drgimp=.false.})]{Explicit top/bottom friction (\protect\np[=.false.]{ln_drgimp}{ln\_drgimp})}
1244\label{subsec:ZDF_drg_stability}
1245
1246Setting \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:
1247
1248At the top (below an ice shelf cavity):
1249\[
1250  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1251  = c_{t}^{\textbf{U}}\textbf{u}^{n-1}_{t}
1252\]
1253
1254At the bottom (above the sea floor):
1255\[
1256  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1257  = c_{b}^{\textbf{U}}\textbf{u}^{n-1}_{b}
1258\]
1259
1260Since 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.
1261For the purposes of stability analysis, an approximation to \autoref{eq:ZDF_drg_flux2} is:
1262\begin{equation}
1263  \label{eq:ZDF_Eqn_drgstab}
1264  \begin{split}
1265    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
1266    &= -\frac{ru}{e_{3u}}\;2\rdt\\
1267  \end{split}
1268\end{equation}
1269\noindent where linear friction and a leapfrog timestep have been assumed.
1270To ensure that the friction cannot reverse the direction of flow it is necessary to have:
1271\[
1272  |\Delta u| < \;|u|
1273\]
1274\noindent which, using \autoref{eq:ZDF_Eqn_drgstab}, gives:
1275\[
1276  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
1277\]
1278This same inequality can also be derived in the non-linear bottom friction case if
1279a velocity of 1 m.s$^{-1}$ is assumed.
1280Alternatively, this criterion can be rearranged to suggest a minimum bottom box thickness to ensure stability:
1281\[
1282  e_{3u} > 2\;r\;\rdt
1283\]
1284\noindent which it may be necessary to impose if partial steps are being used.
1285For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m.
1286For most applications, with physically sensible parameters these restrictions should not be of concern.
1287But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.
1288To ensure stability limits are imposed on the top/bottom friction coefficients both
1289during initialisation and at each time step.
1290Checks at initialisation are made in \mdl{zdfdrg} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1291The number of breaches of the stability criterion are reported as well as
1292the minimum and maximum values that have been set.
1293The criterion is also checked at each time step, using the actual velocity, in \mdl{dynzdf}.
1294Values of the friction coefficient are reduced as necessary to ensure stability;
1295these changes are not reported.
1296
1297Limits on the top/bottom friction coefficient are not imposed if the user has elected to
1298handle the friction implicitly (see \autoref{subsec:ZDF_drg_imp}).
1299The number of potential breaches of the explicit stability criterion are still reported for information purposes.
1300
1301%% =================================================================================================
1302\subsection[Implicit top/bottom friction (\forcode{ln_drgimp=.true.})]{Implicit top/bottom friction (\protect\np[=.true.]{ln_drgimp}{ln\_drgimp})}
1303\label{subsec:ZDF_drg_imp}
1304
1305An optional implicit form of bottom friction has been implemented to improve model stability.
1306We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting.
1307This option can be invoked by setting \np{ln_drgimp}{ln\_drgimp} to \forcode{.true.} in the \nam{drg}{drg} namelist.
1308%This option requires \np{ln_zdfexp}{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf}{zdf} namelist.
1309
1310This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step:
1311
1312At the top (below an ice shelf cavity):
1313\[
1314  % \label{eq:ZDF_dynZDF__drg_top}
1315  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1316  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t}
1317\]
1318
1319At the bottom (above the sea floor):
1320\[
1321  % \label{eq:ZDF_dynZDF__drg_bot}
1322  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1323  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b}
1324\]
1325
1326where $t$ and $b$ refers to top and bottom layers respectively.
1327Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit.
1328
1329%% =================================================================================================
1330\subsection[Bottom friction with split-explicit free surface]{Bottom friction with split-explicit free surface}
1331\label{subsec:ZDF_drg_ts}
1332
1333With 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.
1334
1335The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO\ is as follows:
1336\begin{enumerate}
1337\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.
1338\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.
1339\end{enumerate}
1340
1341Note that other strategies are possible, like considering vertical diffusion step in advance, \ie\ prior barotropic integration.
1342
1343%% =================================================================================================
1344\section[Internal wave-driven mixing (\forcode{ln_zdfiwm})]{Internal wave-driven mixing (\protect\np{ln_zdfiwm}{ln\_zdfiwm})}
1345\label{subsec:ZDF_tmx_new}
1346
1347\begin{listing}
1348  \nlst{namzdf_iwm}
1349  \caption{\forcode{&namzdf_iwm}}
1350  \label{lst:namzdf_iwm}
1351\end{listing}
1352
1353The parameterization of mixing induced by breaking internal waves is a generalization of
1354the approach originally proposed by \citet{st-laurent.simmons.ea_GRL02}.
1355A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed,
1356and the resulting diffusivity is obtained as
1357\[
1358  % \label{eq:ZDF_Kwave}
1359  A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 }
1360\]
1361where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of
1362the energy available for mixing.
1363If the \np{ln_mevar}{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and
1364equal to 1/6 \citep{osborn_JPO80}.
1365In the opposite (recommended) case, $R_f$ is instead a function of
1366the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$,
1367with $\nu$ the molecular viscosity of seawater, following the model of \cite{bouffard.boegman_DAO13} and
1368the implementation of \cite{de-lavergne.madec.ea_JPO16}.
1369Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when
1370the mixing efficiency is constant.
1371
1372In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary
1373as a function of $Re_b$ by setting the \np{ln_tsdiff}{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice.
1374This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14},
1375is implemented as in \cite{de-lavergne.madec.ea_JPO16}.
1376
1377The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$,
1378is constructed from three static maps of column-integrated internal wave energy dissipation,
1379$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures:
1380
1381\begin{align*}
1382  F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\
1383  F_{pyc}(i,j,k) &\propto N^{n_p}\\
1384  F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} }
1385\end{align*}
1386In the above formula, $h_{ab}$ denotes the height above bottom,
1387$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by
1388\[
1389  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; ,
1390\]
1391The $n_p$ parameter (given by \np{nn_zpyc}{nn\_zpyc} in \nam{zdf_iwm}{zdf\_iwm} namelist)
1392controls the stratification-dependence of the pycnocline-intensified dissipation.
1393It can take values of $1$ (recommended) or $2$.
1394Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of
1395the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps.
1396$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and
1397$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of
1398the abyssal hill topography \citep{goff_JGR10} and the latitude.
1399% Jc: input files names ?
1400
1401%% =================================================================================================
1402\section[Surface wave-induced mixing (\forcode{ln_zdfswm})]{Surface wave-induced mixing (\protect\np{ln_zdfswm}{ln\_zdfswm})}
1403\label{subsec:ZDF_swm}
1404
1405Surface waves produce an enhanced mixing through wave-turbulence interaction.
1406In addition to breaking waves induced turbulence (\autoref{subsec:ZDF_tke}),
1407the influence of non-breaking waves can be accounted introducing
1408wave-induced viscosity and diffusivity as a function of the wave number spectrum.
1409Following \citet{qiao.yuan.ea_OD10}, a formulation of wave-induced mixing coefficient
1410is provided  as a function of wave amplitude, Stokes Drift and wave-number:
1411
1412\begin{equation}
1413  \label{eq:ZDF_Bv}
1414  B_{v} = \alpha {A} {U}_{st} {exp(3kz)}
1415\end{equation}
1416
1417Where $B_{v}$ is the wave-induced mixing coefficient, $A$ is the wave amplitude,
1418${U}_{st}$ is the Stokes Drift velocity, $k$ is the wave number and $\alpha$
1419is a constant which should be determined by observations or
1420numerical experiments and is set to be 1.
1421
1422The coefficient $B_{v}$ is then directly added to the vertical viscosity
1423and diffusivity coefficients.
1424
1425In order to account for this contribution set: \forcode{ln_zdfswm=.true.},
1426then wave interaction has to be activated through \forcode{ln_wave=.true.},
1427the Stokes Drift can be evaluated by setting \forcode{ln_sdw=.true.}
1428(see \autoref{subsec:SBC_wave_sdw})
1429and the needed wave fields can be provided either in forcing or coupled mode
1430(for more information on wave parameters and settings see \autoref{sec:SBC_wave})
1431
1432%% =================================================================================================
1433\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp})]{Adaptive-implicit vertical advection(\protect\np{ln_zad_Aimp}{ln\_zad\_Aimp})}
1434\label{subsec:ZDF_aimp}
1435
1436The adaptive-implicit vertical advection option in NEMO is based on the work of
1437\citep{shchepetkin_OM15}.  In common with most ocean models, the timestep used with NEMO
1438needs to satisfy multiple criteria associated with different physical processes in order
1439to maintain numerical stability. \citep{shchepetkin_OM15} pointed out that the vertical
1440CFL criterion is commonly the most limiting. \citep{lemarie.debreu.ea_OM15} examined the
1441constraints for a range of time and space discretizations and provide the CFL stability
1442criteria for a range of advection schemes. The values for the Leap-Frog with Robert
1443asselin filter time-stepping (as used in NEMO) are reproduced in
1444\autoref{tab:ZDF_zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these
1445restrictions but at the cost of large dispersive errors and, possibly, large numerical
1446viscosity. The adaptive-implicit vertical advection option provides a targetted use of the
1447implicit scheme only when and where potential breaches of the vertical CFL condition
1448occur. In many practical applications these events may occur remote from the main area of
1449interest or due to short-lived conditions such that the extra numerical diffusion or
1450viscosity does not greatly affect the overall solution. With such applications, setting:
1451\forcode{ln_zad_Aimp=.true.} should allow much longer model timesteps to be used whilst
1452retaining the accuracy of the high order explicit schemes over most of the domain.
1453
1454\begin{table}[htbp]
1455  \centering
1456  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}}
1457  \begin{tabular}{r|ccc}
1458    \hline
1459    spatial discretization  & 2$^nd$ order centered & 3$^rd$ order upwind & 4$^th$ order compact \\
1460    advective CFL criterion &                 0.904 &              0.472  &                0.522 \\
1461    \hline
1462  \end{tabular}
1463  \caption[Advective CFL criteria for the leapfrog with Robert Asselin filter time-stepping]{
1464    The advective CFL criteria for a range of spatial discretizations for
1465    the leapfrog with Robert Asselin filter time-stepping
1466    ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}.}
1467  \label{tab:ZDF_zad_Aimp_CFLcrit}
1468\end{table}
1469
1470In particular, the advection scheme remains explicit everywhere except where and when
1471local vertical velocities exceed a threshold set just below the explicit stability limit.
1472Once the threshold is reached a tapered transition towards an implicit scheme is used by
1473partitioning the vertical velocity into a part that can be treated explicitly and any
1474excess that must be treated implicitly. The partitioning is achieved via a Courant-number
1475dependent weighting algorithm as described in \citep{shchepetkin_OM15}.
1476
1477The local cell Courant number ($Cu$) used for this partitioning is:
1478
1479\begin{equation}
1480  \label{eq:ZDF_Eqn_zad_Aimp_Courant}
1481  \begin{split}
1482    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 ]    \\
1483       &\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 ]
1484                     \big / e_{{1_t}ij}e_{{2_t}ij}            \\
1485       &\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 ]
1486                     \big / e_{{1_t}ij}e_{{2_t}ij} \bigg )    \\
1487  \end{split}
1488\end{equation}
1489
1490\noindent and the tapering algorithm follows \citep{shchepetkin_OM15} as:
1491
1492\begin{align}
1493  \label{eq:ZDF_Eqn_zad_Aimp_partition}
1494Cu_{min} &= 0.15 \nonumber \\
1495Cu_{max} &= 0.3  \nonumber \\
1496Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\
1497Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\
1498\cf &=
1499     \begin{cases}
1500        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\
1501        (Cu - Cu_{min})^2 / (Fcu +  (Cu - Cu_{min})^2)             &\text{else if $Cu < Cu_{cut}$} \\
1502        (Cu - Cu_{max}) / Cu                                       &\text{else}
1503     \end{cases}
1504\end{align}
1505
1506\begin{figure}[!t]
1507  \centering
1508  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_coeff}
1509  \caption[Partitioning coefficient used to partition vertical velocities into parts]{
1510    The value of the partitioning coefficient (\cf) used to partition vertical velocities into
1511    parts to be treated implicitly and explicitly for a range of typical Courant numbers
1512    (\forcode{ln_zad_Aimp=.true.}).}
1513  \label{fig:ZDF_zad_Aimp_coeff}
1514\end{figure}
1515
1516\noindent The partitioning coefficient is used to determine the part of the vertical
1517velocity that must be handled implicitly ($w_i$) and to subtract this from the total
1518vertical velocity ($w_n$) to leave that which can continue to be handled explicitly:
1519
1520\begin{align}
1521  \label{eq:ZDF_Eqn_zad_Aimp_partition2}
1522    w_{i_{ijk}} &= \cf_{ijk} w_{n_{ijk}}     \nonumber \\
1523    w_{n_{ijk}} &= (1-\cf_{ijk}) w_{n_{ijk}}
1524\end{align}
1525
1526\noindent Note that the coefficient is such that the treatment is never fully implicit;
1527the three cases from \autoref{eq:ZDF_Eqn_zad_Aimp_partition} can be considered as:
1528fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the
1529coefficient (\cf) varies as shown in \autoref{fig:ZDF_zad_Aimp_coeff}. Note with these values
1530the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly
1531implicit' is 0.45 which is just below the stability limited given in
1532\autoref{tab:ZDF_zad_Aimp_CFLcrit}  for a 3rd order scheme.
1533
1534The $w_i$ component is added to the implicit solvers for the vertical mixing in
1535\mdl{dynzdf} and \mdl{trazdf} in a similar way to \citep{shchepetkin_OM15}.  This is
1536sufficient for the flux-limited advection scheme (\forcode{ln_traadv_mus}) but further
1537intervention is required when using the flux-corrected scheme (\forcode{ln_traadv_fct}).
1538For these schemes the implicit upstream fluxes must be added to both the monotonic guess
1539and to the higher order solution when calculating the antidiffusive fluxes. The implicit
1540vertical fluxes are then removed since they are added by the implicit solver later on.
1541
1542The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be
1543used in a wide range of simulations. The following test simulation, however, does illustrate
1544the potential benefits and will hopefully encourage further testing and feedback from users:
1545
1546\begin{figure}[!t]
1547  \centering
1548  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_frames}
1549  \caption[OVERFLOW: time-series of temperature vertical cross-sections]{
1550    A time-series of temperature vertical cross-sections for the OVERFLOW test case.
1551    These results are for the default settings with \forcode{nn_rdt=10.0} and
1552    without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}).}
1553  \label{fig:ZDF_zad_Aimp_overflow_frames}
1554\end{figure}
1555
1556%% =================================================================================================
1557\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case}
1558
1559The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case}
1560provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case
1561by only a few extra physics choices namely:
1562
1563\begin{forlines}
1564     ln_dynldf_OFF = .false.
1565     ln_dynldf_lap = .true.
1566     ln_dynldf_hor = .true.
1567     ln_zdfnpc     = .true.
1568     ln_traadv_fct = .true.
1569        nn_fct_h   =  2
1570        nn_fct_v   =  2
1571\end{forlines}
1572
1573\noindent which were chosen to provide a slightly more stable and less noisy solution. The
1574result when using the default value of \forcode{nn_rdt=10.} without adaptive-implicit
1575vertical velocity is illustrated in \autoref{fig:ZDF_zad_Aimp_overflow_frames}. The mass of
1576cold water, initially sitting on the shelf, moves down the slope and forms a
1577bottom-trapped, dense plume. Even with these extra physics choices the model is close to
1578stability limits and attempts with \forcode{nn_rdt=30.} will fail after about 5.5 hours
1579with excessively high horizontal velocities. This time-scale corresponds with the time the
1580plume reaches the steepest part of the topography and, although detected as a horizontal
1581CFL breach, the instability originates from a breach of the vertical CFL limit. This is a good
1582candidate, therefore, for use of the adaptive-implicit vertical advection scheme.
1583
1584The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps
1585are shown in \autoref{fig:ZDF_zad_Aimp_overflow_all_rdt} (together with the equivalent
1586frames from the base run).  In this simple example the use of the adaptive-implicit
1587vertcal advection scheme has enabled a 12x increase in the model timestep without
1588significantly altering the solution (although at this extreme the plume is more diffuse
1589and has not travelled so far).  Notably, the solution with and without the scheme is
1590slightly different even with \forcode{nn_rdt=10.}; suggesting that the base run was
1591close enough to instability to trigger the scheme despite completing successfully.
1592To assist in diagnosing how active the scheme is, in both location and time, the 3D
1593implicit and explicit components of the vertical velocity are available via XIOS as
1594\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient
1595(\cf) is also available as \texttt{wi\_cff}. For a quick oversight of
1596the schemes activity the global maximum values of the absolute implicit component
1597of the vertical velocity and the partitioning coefficient are written to the netCDF
1598version of the run statistics file (\texttt{run.stat.nc}) if this is active (see
1599\autoref{sec:MISC_opt} for activation details).
1600
1601\autoref{fig:ZDF_zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for
1602the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is
1603active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the
1604test case is close to stability limits even with this value. At the larger timesteps, the
1605vertical velocity is treated mostly implicitly at some location throughout the run. The
1606oscillatory nature of this measure appears to be linked to the progress of the plume front
1607as each cusp is associated with the location of the maximum shifting to the adjacent cell.
1608This is illustrated in \autoref{fig:ZDF_zad_Aimp_maxCf_loc} where the i- and k- locations of the
1609maximum have been overlaid for the base run case.
1610
1611\medskip
1612\noindent Only limited tests have been performed in more realistic configurations. In the
1613ORCA2\_ICE\_PISCES reference configuration the scheme does activate and passes
1614restartability and reproducibility tests but it is unable to improve the model's stability
1615enough to allow an increase in the model time-step. A view of the time-series of maximum
1616partitioning coefficient (not shown here)  suggests that the default time-step of 5400s is
1617already pushing at stability limits, especially in the initial start-up phase. The
1618time-series does not, however, exhibit any of the 'cuspiness' found with the overflow
1619tests.
1620
1621\medskip
1622\noindent A short test with an eORCA1 configuration promises more since a test using a
1623time-step of 3600s remains stable with \forcode{ln_zad_Aimp=.true.} whereas the
1624time-step is limited to 2700s without.
1625
1626\begin{figure}[!t]
1627  \centering
1628  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_all_rdt}
1629  \caption[OVERFLOW: sample temperature vertical cross-sections from mid- and end-run]{
1630    Sample temperature vertical cross-sections from mid- and end-run using
1631    different values for \forcode{nn_rdt} and with or without adaptive implicit vertical advection.
1632    Without the adaptive implicit vertical advection
1633    only the run with the shortest timestep is able to run to completion.
1634    Note also that the colour-scale has been chosen to confirm that
1635    temperatures remain within the original range of 10$^o$ to 20$^o$.}
1636  \label{fig:ZDF_zad_Aimp_overflow_all_rdt}
1637\end{figure}
1638
1639\begin{figure}[!t]
1640  \centering
1641  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf}
1642  \caption[OVERFLOW: maximum partitioning coefficient during a series of test runs]{
1643    The maximum partitioning coefficient during a series of test runs with
1644    increasing model timestep length.
1645    At the larger timesteps,
1646    the vertical velocity is treated mostly implicitly at some location throughout the run.}
1647  \label{fig:ZDF_zad_Aimp_maxCf}
1648\end{figure}
1649
1650\begin{figure}[!t]
1651  \centering
1652  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf_loc}
1653  \caption[OVERFLOW: maximum partitioning coefficient for the case overlaid]{
1654    The maximum partitioning coefficient for the \forcode{nn_rdt=10.0} case overlaid with
1655    information on the gridcell i- and k-locations of the maximum value.}
1656  \label{fig:ZDF_zad_Aimp_maxCf_loc}
1657\end{figure}
1658
1659\subinc{\input{../../global/epilogue}}
1660
1661\end{document}
Note: See TracBrowser for help on using the repository browser.