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/2021/ticket2632_r14588_theta_sbcblk/doc/latex/NEMO/subfiles – NEMO

source: NEMO/branches/2021/ticket2632_r14588_theta_sbcblk/doc/latex/NEMO/subfiles/chap_ZDF.tex @ 15548

Last change on this file since 15548 was 15548, checked in by gsamson, 3 years ago

update branch to the head of the trunk (r15547); ticket #2632

File size: 100.1 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 (No information from an external wave model)}
281\label{subsubsec:ZDF_tke_wave} 
282
283Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to
284include the effect of surface wave breaking energetics.
285This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow.
286The \citet{mellor.blumberg_JPO04} modifications acts on surface length scale and TKE values and
287air-sea drag coefficient.
288The latter concerns the bulk formulae and is not discussed here.
289
290Following \citet{craig.banner_JPO94}, the boundary condition on surface TKE value is :
291\begin{equation}
292  \label{eq:ZDF_Esbc}
293  \bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o}
294\end{equation}
295where $\alpha_{CB}$ is the \citet{craig.banner_JPO94} constant of proportionality which depends on the ''wave age'',
296ranging from 57 for mature waves to 146 for younger waves \citep{mellor.blumberg_JPO04}.
297The boundary condition on the turbulent length scale follows the Charnock's relation:
298\begin{equation}
299  \label{eq:ZDF_Lsbc}
300  l_o = \kappa \beta \,\frac{|\tau|}{g\,\rho_o}
301\end{equation}
302where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant.
303\citet{mellor.blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by
304\citet{stacey_JPO99} citing observation evidence, and
305$\alpha_{CB} = 100$ the Craig and Banner's value.
306As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,
307with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter, setting \np[=67.83]{rn_ebb}{rn\_ebb} corresponds
308to $\alpha_{CB} = 100$.
309
310Further setting  \np[=.true.]{ln_mxl0}{ln\_mxl0},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale, with $\beta$ hard coded to the Stacey's value. Note that a minimal threshold of \np{rn_emin0}{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the surface $\bar{e}$ value.\\
311
312\subsubsection{Surface wave breaking parameterization (using information from an external wave model)}
313\label{subsubsec:ZDF_tke_waveco} 
314
315Surface boundary conditions for the turbulent kinetic energy, the mixing length scale and the dissipative length scale can be defined using wave fields provided from an external wave model (see \autoref{chap:SBC}, \autoref{sec:SBC_wave}).
316The injection of turbulent kinetic energy at the surface can be given by the dissipation of the wave field usually dominated by wave breaking. In coupled mode, the wave to ocean energy flux term ($\Phi_o$) from an external wave model can be provided and then converted into an ocean turbulence source by setting ln\_phioc=.true.
317
318The surface TKE can be defined by a Dirichlet boundary condition setting $nn\_bc\_surf=0$ in \nam{zdf}{tke} namelist:
319\begin{equation}
320  \bar{e}_o  = \frac{1}{2}\,\left( 15.8 \, \frac{\Phi_o}{\rho_o}\right) ^{2/3}
321\end{equation}
322
323Nevertheless, due to the definition of the computational grid, the TKE flux is not applied at the free surface but at the centre of the topmost grid cell ($z = z1$). To be more accurate, a Neumann boundary condition amounting to interpreter the half-grid cell at the top as a constant flux layer (consistent with the surface layer Monin–Obukhov theory) can be applied setting $nn\_bc\_surf=1$ in  \nam{zdf}{tke} namelist \citep{couvelard_2020}:
324
325\begin{equation}
326  \left(\frac{Km}{e_3}\,\partial_k e \right)_{z=z1} = \frac{\Phi_o}{\rho_o}
327\end{equation}
328
329
330The mixing length scale surface value $l_0$ can be estimated from the surface roughness length z0:
331\begin{equation}
332  l_o = \kappa \, \frac{ \left( C_k\,C_\epsilon \right) ^{1/4}}{C_k}\, z0
333\end{equation}
334where $z0$ is directly estimated from the significant wave height ($Hs$) provided by the external wave model as $z0=1.6Hs$. To use this option ln\_mxhsw as well as ln\_wave and ln\_sdw have to be set to .true.
335
336%% =================================================================================================
337\subsubsection{Langmuir cells}
338\label{subsubsec:ZDF_tke_langmuir}
339
340Langmuir circulations (LC) can be described as ordered large-scale vertical motions in
341the surface layer of the oceans.
342Although LC have nothing to do with convection, the circulation pattern is rather similar to
343so-called convective rolls in the atmospheric boundary layer.
344The detailed physics behind LC is described in, for example, \citet{craik.leibovich_JFM76}.
345The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and
346wind drift currents.
347
348Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by
349\citep{axell_JGR02} for a $k-\epsilon$ turbulent closure.
350The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in
351an extra source term of TKE, $P_{LC}$.
352The presence of $P_{LC}$ in \autoref{eq:ZDF_tke_e}, the TKE equation, is controlled by setting \np{ln_lc}{ln\_lc} to
353\forcode{.true.} in the \nam{zdf_tke}{zdf\_tke} namelist.
354
355By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}),
356$P_{LC}$ is assumed to be :
357\[
358P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}}
359\]
360where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth.
361
362For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at
363a finite depth $H_{LC}$ (which is often close to the mixed layer depth),
364and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).
365The resulting expression for $w_{LC}$ is :
366\[
367  w_{LC}  =
368  \begin{cases}
369    c_{LC} \,\|u_s^{LC}\| \,\sin(- \pi\,z / H_{LC} )    &      \text{if $-z \leq H_{LC}$}    \\
370    0                             &      \text{otherwise}
371  \end{cases}
372\]
373
374
375In the absence of information about the wave field, $w_{LC}$ is assumed to be proportional to
376the surface Stokes drift ($u_s^{LC}=u_{s0} $) empirically estimated by $ u_{s0} = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module
377\footnote{Following \citet{li.garrett_JMR93}, the surface Stoke drift velocity may be expressed as
378  $u_{s0} =  0.016 \,|U_{10m}|$.
379  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of
380  $1.5~10^{-3}$ give the expression used of $u_{s0}$ as a function of the module of surface stress
381}.
382
383In case of online coupling with an external wave model (see \autoref{chap:SBC} \autoref{sec:SBC_wave}), $w_{LC}$ is proportional to the component of the Stokes drift aligned with the wind \citep{couvelard_2020} and $ u_s^{LC}  = \max(u_{s0}.e_\tau,0)$ where $e_\tau$ is the unit vector in the wind stress direction and $u_{s0}$ is the surface Stokes drift provided by the external wave model.
384
385
386$c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data.
387The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimetres per second.
388The value of $c_{LC}$ is set through the \np{rn_lc}{rn\_lc} namelist parameter,
389having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}.
390
391The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations:
392$H_{LC}$ is the depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by
393converting its kinetic energy to potential energy, according to
394\[
395- \int_{-H_{LC}}^0 { N^2\;\;dz} = \frac{1}{2} \|u_s^{LC}\|^2
396\]
397
398%% =================================================================================================
399\subsubsection{Mixing just below the mixed layer}
400
401Vertical mixing parameterizations commonly used in ocean general circulation models tend to
402produce mixed-layer depths that are too shallow during summer months and windy conditions.
403This bias is particularly acute over the Southern Ocean.
404To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{rodgers.aumont.ea_B14}.
405The parameterization is an empirical one, \ie\ not derived from theoretical considerations,
406but rather is meant to account for observed processes that affect the density structure of
407the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme
408(\ie\ near-inertial oscillations and ocean swells and waves).
409
410When using this parameterization (\ie\ when \np[=1]{nn_etau}{nn\_etau}),
411the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations,
412swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition,
413plus a depth depend one given by:
414\begin{equation}
415  \label{eq:ZDF_Ehtau}
416  S = (1-f_i) \; f_r \; e_s \; e^{-z / h_\tau}
417\end{equation}
418where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that
419penetrates in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of
420the penetration, and $f_i$ is the ice concentration
421(no penetration if $f_i=1$, \ie\ if the ocean is entirely covered by sea-ice).
422The value of $f_r$, usually a few percents, is specified through \np{rn_efr}{rn\_efr} namelist parameter.
423The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np[=0]{nn_etau}{nn\_etau}) or
424a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes
425(\np[=1]{nn_etau}{nn\_etau}).
426
427Note that two other option exist, \np[=2, 3]{nn_etau}{nn\_etau}.
428They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer,
429or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean.
430Those two options are obsolescent features introduced for test purposes.
431They will be removed in the next release.
432
433% This should be explain better below what this rn_eice parameter is meant for:
434In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn_eice}{rn\_eice} namelist parameter.
435This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration
436is greater than 25\%.
437
438% from Burchard et al OM 2008 :
439% the most critical process not reproduced by statistical turbulence models is the activity of
440% internal waves and their interaction with turbulence. After the Reynolds decomposition,
441% internal waves are in principle included in the RANS equations, but later partially
442% excluded by the hydrostatic assumption and the model resolution.
443% Thus far, the representation of internal wave mixing in ocean models has been relatively crude
444% (\eg\ Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002).
445
446%% =================================================================================================
447\subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls})]{GLS: Generic Length Scale (\protect\np{ln_zdfgls}{ln\_zdfgls})}
448\label{subsec:ZDF_gls}
449
450\begin{listing}
451  \nlst{namzdf_gls}
452  \caption{\forcode{&namzdf_gls}}
453  \label{lst:namzdf_gls}
454\end{listing}
455
456The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations:
457one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale,
458$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}.
459This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,
460where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:ZDF_GLS} allows to recover a number of
461well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87},
462$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}).
463The GLS scheme is given by the following set of equations:
464\begin{equation}
465  \label{eq:ZDF_gls_e}
466  \frac{\partial \bar{e}}{\partial t} =
467  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
468      +\left( \frac{\partial v}{\partial k} \right)^2} \right]
469  -K_\rho \,N^2
470  +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right]
471  - \epsilon
472\end{equation}
473
474\[
475  % \label{eq:ZDF_gls_psi}
476  \begin{split}
477    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{
478      \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2
479          +\left( \frac{\partial v}{\partial k} \right)^2} \right]
480      - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\
481    &+\frac{1}{e_3\;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 }
482        \;\frac{\partial \psi}{\partial k}} \right]\;
483  \end{split}
484\]
485
486\[
487  % \label{eq:ZDF_gls_kz}
488  \begin{split}
489    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\
490    K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l
491  \end{split}
492\]
493
494\[
495  % \label{eq:ZDF_gls_eps}
496  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \;
497\]
498where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and
499$\epsilon$ the dissipation rate.
500The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of
501the choice of the turbulence model.
502Four different turbulent models are pre-defined (\autoref{tab:ZDF_GLS}).
503They are made available through the \np{nn_clo}{nn\_clo} namelist parameter.
504
505\begin{table}[htbp]
506  \centering
507  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c}
508  \begin{tabular}{ccccc}
509    &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\
510    % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\
511    \hline
512    \hline
513    \np{nn_clo}{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\
514    \hline
515    $( p , n , m )$         &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\
516    $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\
517    $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\
518    $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\
519    $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\
520    $C_3$              &      1.           &     1.              &      1.                &       1.           \\
521    $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\
522    \hline
523    \hline
524  \end{tabular}
525  \caption[Set of predefined GLS parameters or equivalently predefined turbulence models available]{
526    Set of predefined GLS parameters, or equivalently predefined turbulence models available with
527    \protect\np[=.true.]{ln_zdfgls}{ln\_zdfgls} and controlled by
528    the \protect\np{nn_clos}{nn\_clos} namelist variable in \protect\nam{zdf_gls}{zdf\_gls}.}
529  \label{tab:ZDF_GLS}
530\end{table}
531
532In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of
533the mixing length towards $\kappa z_b$ ($\kappa$ is the Von Karman constant and $z_b$ the rugosity length scale) value near physical boundaries
534(logarithmic boundary layer law).
535$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88},
536or by \citet{kantha.clayson_JGR94} or one of the two functions suggested by \citet{canuto.howard.ea_JPO01}
537(\np[=0, 3]{nn_stab_func}{nn\_stab\_func}, resp.).
538The value of $C_{0\mu}$ depends on the choice of the stability function.
539
540The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or
541Neumann condition through \np{nn_bc_surf}{nn\_bc\_surf} and \np{nn_bc_bot}{nn\_bc\_bot}, resp.
542As for TKE closure, the wave effect on the mixing is considered when
543\np[ > 0.]{rn_crban}{rn\_crban} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}.
544The \np{rn_crban}{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and
545\np{rn_charn}{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.
546
547The $\psi$ equation is known to fail in stably stratified flows, and for this reason
548almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.
549With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$.
550A value of $c_{lim} = 0.53$ is often used \citep{galperin.kantha.ea_JAS88}.
551\cite{umlauf.burchard_CSR05} show that the value of the clipping factor is of crucial importance for
552the entrainment depth predicted in stably stratified situations,
553and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes.
554The clipping is only activated if \np[=.true.]{ln_length_lim}{ln\_length\_lim},
555and the $c_{lim}$ is set to the \np{rn_clim_galp}{rn\_clim\_galp} value.
556
557The time and space discretization of the GLS equations follows the same energetic consideration as for
558the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{burchard_OM02}.
559Evaluation of the 4 GLS turbulent closure schemes can be found in \citet{warner.sherwood.ea_OM05} in ROMS model and
560 in \citet{reffray.bourdalle-badie.ea_GMD15} for the \NEMO\ model.
561
562% -------------------------------------------------------------------------------------------------------------
563%        OSM OSMOSIS BL Scheme
564% -------------------------------------------------------------------------------------------------------------
565\subsection[OSM: OSMOSIS boundary layer scheme (\forcode{ln_zdfosm = .true.})]
566{OSM: OSMOSIS boundary layer scheme (\protect\np{ln_zdfosm}{ln\_zdfosm})}
567\label{subsec:ZDF_osm}
568
569\begin{listing}
570  \nlst{namzdf_osm}
571  \caption{\forcode{&namzdf_osm}}
572  \label{lst:namzdf_osm}
573\end{listing}
574
575%--------------------------------------------------------------------------------------------------------------
576\paragraph{Namelist choices}
577Most of the namelist options refer to how to specify the Stokes
578surface drift and penetration depth. There are three options:
579\begin{description}
580  \item \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave} Default value in \texttt{namelist\_ref}. In this case the Stokes drift is
581      assumed to be parallel to the surface wind stress, with
582      magnitude consistent with a constant turbulent Langmuir number
583    $\mathrm{La}_t=$ \protect\np{rn_m_la} {rn\_m\_la} i.e.\
584    $u_{s0}=\tau/(\mathrm{La}_t^2\rho_0)$.  Default value of
585    \protect\np{rn_m_la}{rn\_m\_la} is 0.3. The Stokes penetration
586      depth $\delta = $ \protect\np{rn_osm_dstokes}{rn\_osm\_dstokes}; this has default value
587      of 5~m.
588
589  \item \protect\np[=1]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is
590      assumed to be parallel to the surface wind stress, with
591      magnitude as in the classical Pierson-Moskowitz wind-sea
592      spectrum.  Significant wave height and
593      wave-mean period taken from this spectrum are used to calculate the Stokes penetration
594      depth, following the approach set out in  \citet{breivik.janssen.ea_JPO14}.
595
596    \item \protect\np[=2]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is
597      taken from  ECMWF wave model output, though only the component parallel
598      to the wind stress is retained. Significant wave height and
599      wave-mean period from ECMWF wave model output are used to calculate the Stokes penetration
600      depth, again following \citet{breivik.janssen.ea_JPO14}.
601
602    \end{description}
603
604    Others refer to the treatment of diffusion and viscosity beneath
605    the surface boundary layer:
606\begin{description}
607   \item \protect\np{ln_kpprimix} {ln\_kpprimix}  Default is \np{.true.}. Switches on KPP-style Ri \#-dependent
608     mixing below the surface boundary layer. If this is set
609     \texttt{.true.}  the following variable settings are honoured:
610    \item \protect\np{rn_riinfty}{rn\_riinfty} Critical value of local Ri \# below which
611      shear instability increases vertical mixing from background value.
612    \item \protect\np{rn_difri} {rn\_difri} Maximum value of Ri \#-dependent mixing at $\mathrm{Ri}=0$.
613    \item \protect\np{ln_convmix}{ln\_convmix} If \texttt{.true.} then, where water column is unstable, specify
614       diffusivity equal to \protect\np{rn_dif_conv}{rn\_dif\_conv} (default value is 1 m~s$^{-2}$).
615 \end{description}
616 Diagnostic output is controlled by:
617  \begin{description}
618    \item \protect\np{ln_dia_osm}{ln\_dia\_osm} Default is \np{.false.}; allows XIOS output of OSMOSIS internal fields.
619  \end{description}
620Obsolete namelist parameters include:
621\begin{description}
622\item \protect\np{ln_use_osm_la}\np{ln\_use\_osm\_la} With \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave},
623  \protect\np{rn_osm_dstokes} {rn\_osm\_dstokes} is always used to specify the Stokes
624  penetration depth.
625\item \protect\np{nn_ave} {nn\_ave} Choice of averaging method for KPP-style Ri \#
626  mixing. Not taken account of.
627\item \protect\np{rn_osm_hbl0} {rn\_osm\_hbl0} Depth of initial boundary layer is now set
628  by a density criterion similar to that used in calculating \emph{hmlp} (output as \texttt{mldr10\_1}) in \mdl{zdfmxl}.
629\end{description}
630
631\subsubsection{Summary}
632Much of the time the turbulent motions in the ocean surface boundary
633layer (OSBL) are not given by
634classical shear turbulence. Instead they are in a regime known as
635`Langmuir turbulence',  dominated by an
636interaction between the currents and the Stokes drift of the surface waves \citep[e.g.][]{mcwilliams.sullivan.ea_JFM97}.
637This 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 $\mathrm{La}_{t}=(u_{\ast}/u_{s0}) > 0.4$.
638
639The OSMOSIS model is fundamentally based on results of Large Eddy
640Simulations (LES) of Langmuir turbulence and aims to fully describe
641this Langmuir regime. The description in this section is of necessity incomplete and further details are available in Grant. A (2019); in prep.
642
643The OSMOSIS turbulent closure scheme is a similarity-scale scheme in
644the same spirit as the K-profile
645parameterization (KPP) scheme of \citet{large.mcwilliams.ea_RG94}.
646A specified shape of diffusivity, scaled by the (OSBL) depth
647$h_{\mathrm{BL}}$ and a turbulent velocity scale, is imposed throughout the
648boundary layer
649$-h_{\mathrm{BL}}<z<\eta$. The turbulent closure model
650also includes fluxes of tracers and momentum that are``non-local'' (independent of the local property gradient).
651
652Rather than the OSBL
653depth being diagnosed in terms of a bulk Richardson number criterion,
654as in KPP, it is set by a prognostic equation that is informed by
655energy budget considerations reminiscent of the classical mixed layer
656models of \citet{kraus.turner_T67}.
657The model also includes an explicit parametrization of the structure
658of the pycnocline (the stratified region at the bottom of the OSBL).
659
660Presently, mixing below the OSBL is handled by the Richardson
661number-dependent mixing scheme used in \citet{large.mcwilliams.ea_RG94}.
662
663Convective parameterizations such as described in \autoref{sec:ZDF_conv}
664below should not be used with the OSMOSIS-OBL model: instabilities
665within the OSBL are part of the model, while instabilities below the
666ML are handled by the Ri \# dependent scheme.
667
668\subsubsection{Depth and velocity scales}
669The 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).
670\begin{figure}[!t]
671  \begin{center}
672    %\includegraphics[width=0.7\textwidth]{ZDF_OSM_structure_of_OSBL}
673    \caption{
674      \protect\label{fig:OSBL_structure}
675     The structure of the entraining  boundary layer. (a) Mean buoyancy profile. (b) Profile of the buoyancy flux.
676    }
677  \end{center}
678\end{figure}
679The 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.
680
681Consideration of the power input by wind acting on the Stokes drift suggests that the Langmuir turbulence has velocity scale:
682\begin{equation}
683  \label{eq:ZDF_w_La}
684  w_{*L}= \left(u_*^2 u_{s\,0}\right)^{1/3};
685\end{equation}
686but 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:
687\begin{equation}
688  \label{eq:ZDF_composite-nu}
689  \nu_{\ast}= \left\{ u_*^3 \left[1-\exp(-.5 \mathrm{La}_t^2)\right]+w_{*L}^3\right\}^{1/3}.
690\end{equation}
691For 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:
692\begin{equation}
693  \label{eq:ZDF_vel-scale-unstable}
694  \omega_* = \left(\nu_*^3 + 0.5 w_{*C}^3\right)^{1/3}.
695\end{equation}
696
697\subsubsection{The flux gradient model}
698The flux-gradient relationships used in the OSMOSIS scheme take the form:
699
700\begin{equation}
701  \label{eq:ZDF_flux-grad-gen}
702  \overline{w^\prime\chi^\prime}=-K\frac{\partial\overline{\chi}}{\partial z} + N_{\chi,s} +N_{\chi,b} +N_{\chi,t},
703\end{equation}
704
705where $\chi$ 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.
706
707In terms of the non-dimensionalized depth variables
708
709\begin{equation}
710  \label{eq:ZDF_sigma}
711  \sigma_{\mathrm{ml}}= -z/h_{\mathrm{ml}}; \;\sigma_{\mathrm{bl}}= -z/h_{\mathrm{bl}},
712\end{equation}
713
714in unstable conditions the eddy diffusivity ($K_d$) and eddy viscosity ($K_\nu$) profiles are parametrized as:
715
716\begin{align}
717  \label{eq:ZDF_diff-unstable}
718  K_d=&0.8\, \omega_*\, h_{\mathrm{ml}} \, \sigma_{\mathrm{ml}} \left(1-\beta_d \sigma_{\mathrm{ml}}\right)^{3/2}
719  \\
720  \label{eq:ZDF_visc-unstable}
721  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)
722\end{align}
723
724where $\beta_d$ 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
725
726\begin{equation}
727  \label{eq:ZDF_diff-wml-base}
728  K_{d,\mathrm{ml}}=K_{\nu,\mathrm{ml}}=\,0.16\,\omega_* \Delta h.
729\end{equation}
730
731For stable conditions the eddy diffusivity/viscosity profiles are given by:
732
733\begin{align}
734  \label{eq:ZDF_diff-stable}
735  K_d= & 0.75\,\, \nu_*\, h_{\mathrm{ml}}\,\,  \exp\left[-2.8
736       \left(h_{\mathrm{bl}}/L_L\right)^2\right]\sigma_{\mathrm{ml}}
737       \left(1-\sigma_{\mathrm{ml}}\right)^{3/2} \\
738  \label{eq:ZDF_visc-stable}
739  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).
740\end{align}
741
742The 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:
743\begin{equation}
744  \label{eq:ZDF_L_L}
745  L_L=-w_{*L}^3/\left<\overline{w^\prime b^\prime}\right>_L,
746\end{equation}
747with 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
748\begin{equation}
749  \label{eq:ZDF_stable-av-buoy-flux}
750  \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].
751\end{equation}
752
753In unstable conditions the eddy diffusivity and viscosity depend on stability through the velocity scale $\omega_*$, which depends on the two velocity scales $\nu_*$ and $w_{*C}$.
754
755Details 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).
756
757\subsubsection{Evolution of the boundary layer depth}
758
759The prognostic equation for the depth of the neutral/unstable boundary layer is given by \iffalse \citep{grant+etal18?}, \fi
760
761\begin{equation}
762  \label{eq:ZDF_dhdt-unstable}
763%\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}}
764   \frac{\partial h_\mathrm{bl}}{\partial t} = W_b - \frac{{\overline{w^\prime b^\prime}}_\mathrm{ent}}{\Delta B_\mathrm{bl}}
765\end{equation}
766where $h_\mathrm{bl}$ is the horizontally-varying depth of the OSBL,
767$\mathbf{U}_b$ and $W_b$ are the mean horizontal and vertical
768velocities at the base of the OSBL, ${\overline{w^\prime
769    b^\prime}}_\mathrm{ent}$ is the buoyancy flux due to entrainment
770and $\Delta B_\mathrm{bl}$ is the difference between the buoyancy
771averaged over the depth of the OSBL (i.e.\ including the ML and
772pycnocline) and the buoyancy just below the base of the OSBL. This
773equation for the case when the pycnocline has a finite thickness,
774based on the potential energy budget of the OSBL, is the leading term
775\iffalse \citep{grant+etal18?} \fi of a generalization of that used in mixed-layer
776models e.g.\ \citet{kraus.turner_T67}, in which the thickness of the pycnocline is taken to be zero.
777
778The entrainment flux for the combination of convective and Langmuir turbulence is given by
779\begin{equation}
780  \label{eq:ZDF_entrain-flux}
781  {\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}}}
782  + 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]
783\end{equation}
784where 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 $F(x)\equiv\tanh(x^{-1})^{0.69}$.
785
786For the stable boundary layer, the equation for the depth of the OSBL is:
787
788\begin{equation}
789  \label{eq:ZDF_dhdt-stable}
790\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.
791\end{equation}
792
793\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.
794
795
796%% =================================================================================================
797\subsection[ Discrete energy conservation for TKE and GLS schemes]{Discrete energy conservation for TKE and GLS schemes}
798\label{subsec:ZDF_tke_ene}
799
800\begin{figure}[!t]
801  \centering
802  \includegraphics[width=0.66\textwidth]{ZDF_TKE_time_scheme}
803  \caption[Subgrid kinetic energy integration in GLS and TKE schemes]{
804    Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and
805    its links to the momentum and tracer time integration.}
806  \label{fig:ZDF_TKE_time_scheme}
807\end{figure}
808
809The production of turbulence by vertical shear (the first term of the right hand side of
810\autoref{eq:ZDF_tke_e}) and  \autoref{eq:ZDF_gls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion
811(first line in \autoref{eq:MB_zdf}).
812To do so a special care has to be taken for both the time and space discretization of
813the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}.
814
815Let us first address the time stepping issue. \autoref{fig:ZDF_TKE_time_scheme} shows how
816the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with
817the one-level forward time stepping of the equation for $\bar{e}$.
818With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to
819the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and
820summing the result vertically:
821\begin{equation}
822  \label{eq:ZDF_energ1}
823  \begin{split}
824    \int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\
825    &= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}
826    - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz }
827  \end{split}
828\end{equation}
829Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$,
830known at time $t$ (\autoref{fig:ZDF_TKE_time_scheme}), as it is required when using the TKE scheme
831(see \autoref{sec:TD_forward_imp}).
832The first term of the right hand side of \autoref{eq:ZDF_energ1} represents the kinetic energy transfer at
833the surface (atmospheric forcing) and at the bottom (friction effect).
834The second term is always negative.
835It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$.
836\autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
837the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as
838${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$
839(and not by the more straightforward $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$).
840
841A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification
842(second term of the right hand side of \autoref{eq:ZDF_tke_e} and \autoref{eq:ZDF_gls_e}).
843This term must balance the input of potential energy resulting from vertical mixing.
844The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by
845multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically:
846\begin{equation}
847  \label{eq:ZDF_energ2}
848  \begin{split}
849    \int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\
850    &= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta}
851    - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\
852    &= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta}
853    + \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  }
854  \end{split}
855\end{equation}
856where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.
857The first term of the right hand side of \autoref{eq:ZDF_energ2} is always zero because
858there is no diffusive flux through the ocean surface and bottom).
859The second term is minus the destruction rate of  $\bar{e}$ due to stratification.
860Therefore \autoref{eq:ZDF_energ1} implies that, to be energetically consistent,
861the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:ZDF_tke_e} and  \autoref{eq:ZDF_gls_e}.
862
863Let us now address the space discretization issue.
864The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in
865the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:DOM_cell}).
866A space averaging is thus required to obtain the shear TKE production term.
867By redoing the \autoref{eq:ZDF_energ1} in the 3D case, it can be shown that the product of eddy coefficient by
868the shear at $t$ and $t-\rdt$ must be performed prior to the averaging.
869Furthermore, the time variation of $e_3$ has be taken into account.
870
871The above energetic considerations leads to the following final discrete form for the TKE equation:
872\begin{equation}
873  \label{eq:ZDF_tke_ene}
874  \begin{split}
875    \frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv
876    \Biggl\{ \Biggr.
877    &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} }
878        \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\
879    +&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} }
880        \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j}
881    \Biggr. \Biggr\}   \\
882    %
883    - &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\
884    %
885    +&\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]   \\
886    %
887    - &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt}
888  \end{split}
889\end{equation}
890where the last two terms in \autoref{eq:ZDF_tke_ene} (vertical diffusion and Kolmogorov dissipation)
891are time stepped using a backward scheme (see\autoref{sec:TD_forward_imp}).
892Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible.
893%The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as
894%they all appear in the right hand side of \autoref{eq:ZDF_tke_ene}.
895%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.
896
897%% =================================================================================================
898\section{Convection}
899\label{sec:ZDF_conv}
900
901Static instabilities (\ie\ light potential densities under heavy ones) may occur at particular ocean grid points.
902In nature, convective processes quickly re-establish the static stability of the water column.
903These processes have been removed from the model via the hydrostatic assumption so they must be parameterized.
904Three parameterisations are available to deal with convective processes:
905a non-penetrative convective adjustment or an enhanced vertical diffusion,
906or/and the use of a turbulent closure scheme.
907
908%% =================================================================================================
909\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc})]{Non-penetrative convective adjustment (\protect\np{ln_tranpc}{ln\_tranpc})}
910\label{subsec:ZDF_npc}
911
912\begin{figure}[!htb]
913  \centering
914  \includegraphics[width=0.66\textwidth]{ZDF_npc}
915  \caption[Unstable density profile treated by the non penetrative convective adjustment algorithm]{
916    Example of an unstable density profile treated by
917    the non penetrative convective adjustment algorithm.
918    $1^{st}$ step: the initial profile is checked from the surface to the bottom.
919    It is found to be unstable between levels 3 and 4.
920    They are mixed.
921    The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed.
922    The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed.
923    The $1^{st}$ step ends since the density profile is then stable below the level 3.
924    $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step:
925    levels 2 to 5 are mixed.
926    The new density profile is checked.
927    It is found stable: end of algorithm.}
928  \label{fig:ZDF_npc}
929\end{figure}
930
931Options are defined through the \nam{zdf}{zdf} namelist variables.
932The non-penetrative convective adjustment is used when \np[=.true.]{ln_zdfnpc}{ln\_zdfnpc}.
933It is applied at each \np{nn_npc}{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of
934the water column, but only until the density structure becomes neutrally stable
935(\ie\ until the mixed portion of the water column has \textit{exactly} the density of the water just below)
936\citep{madec.delecluse.ea_JPO91}.
937The associated algorithm is an iterative process used in the following way (\autoref{fig:ZDF_npc}):
938starting from the top of the ocean, the first instability is found.
939Assume in the following that the instability is located between levels $k$ and $k+1$.
940The temperature and salinity in the two levels are vertically mixed, conserving the heat and salt contents of
941the water column.
942The new density is then computed by a linear approximation.
943If the new density profile is still unstable between levels $k+1$ and $k+2$,
944levels $k$, $k+1$ and $k+2$ are then mixed.
945This process is repeated until stability is established below the level $k$
946(the mixing process can go down to the ocean bottom).
947The algorithm is repeated to check if the density profile between level $k-1$ and $k$ is unstable and/or
948if there is no deeper instability.
949
950This algorithm is significantly different from mixing statically unstable levels two by two.
951The latter procedure cannot converge with a finite number of iterations for some vertical profiles while
952the algorithm used in \NEMO\ converges for any profile in a number of iterations which is less than
953the number of vertical levels.
954This property is of paramount importance as pointed out by \citet{killworth_iprc89}:
955it avoids the existence of permanent and unrealistic static instabilities at the sea surface.
956This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in
957the north-western Mediterranean Sea \citep{madec.delecluse.ea_JPO91, madec.chartier.ea_DAO91, madec.crepon_iprc91}.
958
959The current implementation has been modified in order to deal with any non linear equation of seawater
960(L. Brodeau, personnal communication).
961Two main differences have been introduced compared to the original algorithm:
962$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency
963(not the difference in potential density);
964$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in
965the same way their temperature and salinity has been mixed.
966These two modifications allow the algorithm to perform properly and accurately with TEOS10 or EOS-80 without
967having to recompute the expansion coefficients at each mixing iteration.
968
969%% =================================================================================================
970\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd})]{Enhanced vertical diffusion (\protect\np{ln_zdfevd}{ln\_zdfevd})}
971\label{subsec:ZDF_evd}
972
973Options are defined through the  \nam{zdf}{zdf} namelist variables.
974The enhanced vertical diffusion parameterisation is used when \np[=.true.]{ln_zdfevd}{ln\_zdfevd}.
975In this case, the vertical eddy mixing coefficients are assigned very large values
976in regions where the stratification is unstable
977(\ie\ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}.
978This is done either on tracers only (\np[=0]{nn_evdm}{nn\_evdm}) or
979on both momentum and tracers (\np[=1]{nn_evdm}{nn\_evdm}).
980
981In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np[=1]{nn_evdm}{nn\_evdm},
982the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to
983the namelist parameter \np{rn_avevd}{rn\_avevd}.
984A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$.
985This parameterisation of convective processes is less time consuming than
986the convective adjustment algorithm presented above when mixing both tracers and
987momentum in the case of static instabilities.
988
989Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$.
990This removes a potential source of divergence of odd and even time step in
991a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:TD_mLF}).
992
993%% =================================================================================================
994\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}})}
995\label{subsec:ZDF_tcs}
996
997The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and
998\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,
999with statically unstable density profiles.
1000In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in
1001\autoref{eq:ZDF_tke_e} or \autoref{eq:ZDF_gls_e} becomes a source term, since $N^2$ is negative.
1002It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at
1003velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$).
1004These large values restore the static stability of the water column in a way similar to that of
1005the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}).
1006However, in the vicinity of the sea surface (first ocean layer), the eddy coefficients computed by
1007the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,
1008because the mixing length scale is bounded by the distance to the sea surface.
1009It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme,
1010\ie\ setting the \np{ln_zdfnpc}{ln\_zdfnpc} namelist parameter to true and
1011defining the turbulent closure (\np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} = \forcode{.true.}) all together.
1012
1013The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection,
1014%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp},
1015therefore \np[=.false.]{ln_zdfevd}{ln\_zdfevd} should be used with the OSMOSIS scheme.
1016% gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
1017
1018%% =================================================================================================
1019\section[Double diffusion mixing (\forcode{ln_zdfddm})]{Double diffusion mixing (\protect\np{ln_zdfddm}{ln\_zdfddm})}
1020\label{subsec:ZDF_ddm}
1021
1022%\nlst{namzdf_ddm}
1023
1024This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter
1025\np{ln_zdfddm}{ln\_zdfddm} in \nam{zdf}{zdf}.
1026Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa.
1027The former condition leads to salt fingering and the latter to diffusive convection.
1028Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean.
1029\citet{merryfield.holloway.ea_JPO99} include a parameterisation of such phenomena in a global ocean model and show that
1030it leads to relatively minor changes in circulation but exerts significant regional influences on
1031temperature and salinity.
1032
1033Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
1034\begin{align*}
1035  % \label{eq:ZDF_ddm_Kz}
1036  &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT} \\
1037  &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
1038\end{align*}
1039where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
1040and $o$ by processes other than double diffusion.
1041The rates of double-diffusive mixing depend on the buoyancy ratio
1042$R_\rho = \alpha \partial_z T / \beta \partial_z S$, where $\alpha$ and $\beta$ are coefficients of
1043thermal expansion and saline contraction (see \autoref{subsec:TRA_eos}).
1044To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt
1045(1981):
1046\begin{align}
1047  \label{eq:ZDF_ddm_f}
1048  A_f^{vS} &=
1049             \begin{cases}
1050               \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
1051               0                              &\text{otherwise}
1052             \end{cases}
1053  \\         \label{eq:ZDF_ddm_f_T}
1054  A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho
1055\end{align}
1056
1057\begin{figure}[!t]
1058  \centering
1059  \includegraphics[width=0.66\textwidth]{ZDF_ddm}
1060  \caption[Diapycnal diffusivities for temperature and salt in regions of salt fingering and
1061  diffusive convection]{
1062    From \citet{merryfield.holloway.ea_JPO99}:
1063    (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in
1064    regions of salt fingering.
1065    Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and
1066    thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$;
1067    (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in
1068    regions of diffusive convection.
1069    Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation.
1070    The latter is not implemented in \NEMO.}
1071  \label{fig:ZDF_ddm}
1072\end{figure}
1073
1074The factor 0.7 in \autoref{eq:ZDF_ddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of
1075buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}).
1076Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
1077
1078To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by
1079Federov (1988) is used:
1080\begin{align}
1081  % \label{eq:ZDF_ddm_d}
1082  A_d^{vT} &=
1083             \begin{cases}
1084               1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
1085               &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
1086               0                       &\text{otherwise}
1087             \end{cases}
1088                                       \nonumber \\
1089  \label{eq:ZDF_ddm_d_S}
1090  A_d^{vS} &=
1091             \begin{cases}
1092               A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right) &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
1093               A_d^{vT} \ 0.15 \ R_\rho               &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
1094               0                       &\text{otherwise}
1095             \end{cases}
1096\end{align}
1097
1098The dependencies of \autoref{eq:ZDF_ddm_f} to \autoref{eq:ZDF_ddm_d_S} on $R_\rho$ are illustrated in
1099\autoref{fig:ZDF_ddm}.
1100Implementing this requires computing $R_\rho$ at each grid point on every time step.
1101This is done in \mdl{eosbn2} at the same time as $N^2$ is computed.
1102This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive).
1103
1104%% =================================================================================================
1105\section[Bottom and top friction (\textit{zdfdrg.F90})]{Bottom and top friction (\protect\mdl{zdfdrg})}
1106\label{sec:ZDF_drg}
1107
1108\begin{listing}
1109  \nlst{namdrg}
1110  \caption{\forcode{&namdrg}}
1111  \label{lst:namdrg}
1112\end{listing}
1113
1114\begin{listing}
1115  \nlst{namdrg_top}
1116  \caption{\forcode{&namdrg_top}}
1117  \label{lst:namdrg_top}
1118\end{listing}
1119
1120\begin{listing}
1121  \nlst{namdrg_bot}
1122  \caption{\forcode{&namdrg_bot}}
1123  \label{lst:namdrg_bot}
1124\end{listing}
1125
1126Options to define the top and bottom friction are defined through the \nam{drg}{drg} namelist variables.
1127The bottom friction represents the friction generated by the bathymetry.
1128The top friction represents the friction generated by the ice shelf/ocean interface.
1129As the friction processes at the top and the bottom are treated in and identical way,
1130the description below considers mostly the bottom friction case, if not stated otherwise.
1131
1132Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as
1133a condition on the vertical diffusive flux.
1134For the bottom boundary layer, one has:
1135 \[
1136   % \label{eq:ZDF_bfr_flux}
1137   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U}
1138 \]
1139where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside
1140the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean).
1141How ${\cal F}_h^{\textbf U}$ influences the interior depends on the vertical resolution of the model near
1142the bottom relative to the Ekman layer depth.
1143For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m,
1144one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$
1145(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$).
1146With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
1147When the vertical mixing coefficient is this small, using a flux condition is equivalent to
1148entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or
1149bottom model layer.
1150To illustrate this, consider the equation for $u$ at $k$, the last ocean level:
1151\begin{equation}
1152  \label{eq:ZDF_drg_flux2}
1153  \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}}
1154\end{equation}
1155If the bottom layer thickness is 200~m, the Ekman transport will be distributed over that depth.
1156On the other hand, if the vertical resolution is high (1~m or less) and a turbulent closure model is used,
1157the turbulent Ekman layer will be represented explicitly by the model.
1158However, the logarithmic layer is never represented in current primitive equation model applications:
1159it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.
1160Two choices are available in \NEMO: a linear and a quadratic bottom friction.
1161Note that in both cases, the rotation between the interior velocity and the bottom friction is neglected in
1162the present release of \NEMO.
1163
1164In the code, the bottom friction is imposed by adding the trend due to the bottom friction to
1165 the general momentum trend in \mdl{dynzdf}.
1166For the time-split surface pressure gradient algorithm, the momentum trend due to
1167the barotropic component needs to be handled separately.
1168For this purpose it is convenient to compute and store coefficients which can be simply combined with
1169bottom velocities and geometric values to provide the momentum trend due to bottom friction.
1170 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where:
1171\begin{equation}
1172  \label{eq:ZDF_bfr_bdef}
1173  \frac{\partial {\textbf U_h}}{\partial t} =
1174  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b
1175\end{equation}
1176where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.
1177Note 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.
1178
1179%% =================================================================================================
1180\subsection[Linear top/bottom friction (\forcode{ln_lin})]{Linear top/bottom friction (\protect\np{ln_lin}{ln\_lin})}
1181\label{subsec:ZDF_drg_linear}
1182
1183The linear friction parameterisation (including the special case of a free-slip condition) assumes that
1184the friction is proportional to the interior velocity (\ie\ the velocity of the first/last model level):
1185\[
1186  % \label{eq:ZDF_bfr_linear}
1187  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
1188\]
1189where $r$ is a friction coefficient expressed in $m s^{-1}$.
1190This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,
1191and setting $r = H / \tau$, where $H$ is the ocean depth.
1192Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{weatherly_JMR84}.
1193A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models.
1194One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$
1195(\citet{gill_bk82}, Eq. 9.6.6).
1196For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$,
1197and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.
1198This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days.
1199It can be changed by specifying \np{rn_Uc0}{rn\_Uc0} (namelist parameter).
1200
1201 For the linear friction case the drag coefficient used in the general expression \autoref{eq:ZDF_bfr_bdef} is:
1202\[
1203  % \label{eq:ZDF_bfr_linbfr_b}
1204    c_b^T = - r
1205\]
1206When \np[=.true.]{ln_lin}{ln\_lin}, the value of $r$ used is \np{rn_Uc0}{rn\_Uc0}*\np{rn_Cd0}{rn\_Cd0}.
1207Setting \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.
1208
1209These values are assigned in \mdl{zdfdrg}.
1210Note that there is support for local enhancement of these values via an externally defined 2D mask array
1211(\np[=.true.]{ln_boost}{ln\_boost}) given in the \textit{bfr\_coef.nc} input NetCDF file.
1212The mask values should vary from 0 to 1.
1213Locations with a non-zero mask value will have the friction coefficient increased by
1214$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
1215
1216%% =================================================================================================
1217\subsection[Non-linear top/bottom friction (\forcode{ln_non_lin})]{Non-linear top/bottom friction (\protect\np{ln_non_lin}{ln\_non\_lin})}
1218\label{subsec:ZDF_drg_nonlinear}
1219
1220The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic:
1221\[
1222  % \label{eq:ZDF_drg_nonlinear}
1223  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
1224  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
1225\]
1226where $C_D$ is a drag coefficient, and $e_b $ a top/bottom turbulent kinetic energy due to tides,
1227internal waves breaking and other short time scale currents.
1228A typical value of the drag coefficient is $C_D = 10^{-3} $.
1229As an example, the CME experiment \citep{treguier_JGR92} uses $C_D = 10^{-3}$ and
1230$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and
1231$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.
1232The CME choices have been set as default values (\np{rn_Cd0}{rn\_Cd0} and \np{rn_ke0}{rn\_ke0} namelist parameters).
1233
1234As for the linear case, the friction is imposed in the code by adding the trend due to
1235the friction to the general momentum trend in \mdl{dynzdf}.
1236For the non-linear friction case the term computed in \mdl{zdfdrg} is:
1237\[
1238  % \label{eq:ZDF_drg_nonlinbfr}
1239    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2}
1240\]
1241
1242The coefficients that control the strength of the non-linear friction are initialised as namelist parameters:
1243$C_D$= \np{rn_Cd0}{rn\_Cd0}, and $e_b$ =\np{rn_bfeb2}{rn\_bfeb2}.
1244Note 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
1245(\np[=.true.]{ln_boost}{ln\_boost}).
1246This works in the same way as for the linear friction case with non-zero masked locations increased by
1247$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}.
1248
1249%% =================================================================================================
1250\subsection[Log-layer top/bottom friction (\forcode{ln_loglayer})]{Log-layer top/bottom friction (\protect\np{ln_loglayer}{ln\_loglayer})}
1251\label{subsec:ZDF_drg_loglayer}
1252
1253In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using
1254a "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.
1255If  \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):
1256\[
1257  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2
1258\]
1259
1260\noindent where $\kappa$ is the von-Karman constant and \np{rn_z0}{rn\_z0} is a roughness length provided via the namelist.
1261
1262The drag coefficient is bounded such that it is kept greater or equal to
1263the 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:
1264\np{rn_Cdmax}{rn\_Cdmax}, \ie
1265\[
1266  rn\_Cd0 \leq C_D \leq rn\_Cdmax
1267\]
1268
1269\noindent The log-layer enhancement can also be applied to the top boundary friction if
1270under ice-shelf cavities are activated (\np[=.true.]{ln_isfcav}{ln\_isfcav}).
1271%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}.
1272
1273%% =================================================================================================
1274\subsection[Explicit top/bottom friction (\forcode{ln_drgimp=.false.})]{Explicit top/bottom friction (\protect\np[=.false.]{ln_drgimp}{ln\_drgimp})}
1275\label{subsec:ZDF_drg_stability}
1276
1277Setting \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:
1278
1279At the top (below an ice shelf cavity):
1280\[
1281  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1282  = c_{t}^{\textbf{U}}\textbf{u}^{n-1}_{t}
1283\]
1284
1285At the bottom (above the sea floor):
1286\[
1287  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1288  = c_{b}^{\textbf{U}}\textbf{u}^{n-1}_{b}
1289\]
1290
1291Since 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.
1292For the purposes of stability analysis, an approximation to \autoref{eq:ZDF_drg_flux2} is:
1293\begin{equation}
1294  \label{eq:ZDF_Eqn_drgstab}
1295  \begin{split}
1296    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\
1297    &= -\frac{ru}{e_{3u}}\;2\rdt\\
1298  \end{split}
1299\end{equation}
1300\noindent where linear friction and a leapfrog timestep have been assumed.
1301To ensure that the friction cannot reverse the direction of flow it is necessary to have:
1302\[
1303  |\Delta u| < \;|u|
1304\]
1305\noindent which, using \autoref{eq:ZDF_Eqn_drgstab}, gives:
1306\[
1307  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\
1308\]
1309This same inequality can also be derived in the non-linear bottom friction case if
1310a velocity of 1 m.s$^{-1}$ is assumed.
1311Alternatively, this criterion can be rearranged to suggest a minimum bottom box thickness to ensure stability:
1312\[
1313  e_{3u} > 2\;r\;\rdt
1314\]
1315\noindent which it may be necessary to impose if partial steps are being used.
1316For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m.
1317For most applications, with physically sensible parameters these restrictions should not be of concern.
1318But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.
1319To ensure stability limits are imposed on the top/bottom friction coefficients both
1320during initialisation and at each time step.
1321Checks at initialisation are made in \mdl{zdfdrg} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case).
1322The number of breaches of the stability criterion are reported as well as
1323the minimum and maximum values that have been set.
1324The criterion is also checked at each time step, using the actual velocity, in \mdl{dynzdf}.
1325Values of the friction coefficient are reduced as necessary to ensure stability;
1326these changes are not reported.
1327
1328Limits on the top/bottom friction coefficient are not imposed if the user has elected to
1329handle the friction implicitly (see \autoref{subsec:ZDF_drg_imp}).
1330The number of potential breaches of the explicit stability criterion are still reported for information purposes.
1331
1332%% =================================================================================================
1333\subsection[Implicit top/bottom friction (\forcode{ln_drgimp=.true.})]{Implicit top/bottom friction (\protect\np[=.true.]{ln_drgimp}{ln\_drgimp})}
1334\label{subsec:ZDF_drg_imp}
1335
1336An optional implicit form of bottom friction has been implemented to improve model stability.
1337We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting.
1338This option can be invoked by setting \np{ln_drgimp}{ln\_drgimp} to \forcode{.true.} in the \nam{drg}{drg} namelist.
1339%This option requires \np{ln_zdfexp}{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf}{zdf} namelist.
1340
1341This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step:
1342
1343At the top (below an ice shelf cavity):
1344\[
1345  % \label{eq:ZDF_dynZDF__drg_top}
1346  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t}
1347  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t}
1348\]
1349
1350At the bottom (above the sea floor):
1351\[
1352  % \label{eq:ZDF_dynZDF__drg_bot}
1353  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b}
1354  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b}
1355\]
1356
1357where $t$ and $b$ refers to top and bottom layers respectively.
1358Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit.
1359
1360%% =================================================================================================
1361\subsection[Bottom friction with split-explicit free surface]{Bottom friction with split-explicit free surface}
1362\label{subsec:ZDF_drg_ts}
1363
1364With 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.
1365
1366The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO\ is as follows:
1367\begin{enumerate}
1368\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.
1369\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.
1370\end{enumerate}
1371
1372Note that other strategies are possible, like considering vertical diffusion step in advance, \ie\ prior barotropic integration.
1373
1374%% =================================================================================================
1375\section[Internal wave-driven mixing (\forcode{ln_zdfiwm})]{Internal wave-driven mixing (\protect\np{ln_zdfiwm}{ln\_zdfiwm})}
1376\label{subsec:ZDF_tmx_new}
1377
1378\begin{listing}
1379  \nlst{namzdf_iwm}
1380  \caption{\forcode{&namzdf_iwm}}
1381  \label{lst:namzdf_iwm}
1382\end{listing}
1383
1384The parameterization of mixing induced by breaking internal waves is a generalization of
1385the approach originally proposed by \citet{st-laurent.simmons.ea_GRL02}.
1386A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed,
1387and the resulting diffusivity is obtained as
1388\[
1389  % \label{eq:ZDF_Kwave}
1390  A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 }
1391\]
1392where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of
1393the energy available for mixing.
1394If the \np{ln_mevar}{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and
1395equal to 1/6 \citep{osborn_JPO80}.
1396In the opposite (recommended) case, $R_f$ is instead a function of
1397the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$,
1398with $\nu$ the molecular viscosity of seawater, following the model of \cite{bouffard.boegman_DAO13} and
1399the implementation of \cite{de-lavergne.madec.ea_JPO16}.
1400Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when
1401the mixing efficiency is constant.
1402
1403In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary
1404as a function of $Re_b$ by setting the \np{ln_tsdiff}{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice.
1405This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14},
1406is implemented as in \cite{de-lavergne.madec.ea_JPO16}.
1407
1408The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$,
1409is constructed from three static maps of column-integrated internal wave energy dissipation,
1410$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures:
1411
1412\begin{align*}
1413  F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\
1414  F_{pyc}(i,j,k) &\propto N^{n_p}\\
1415  F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} }
1416\end{align*}
1417In the above formula, $h_{ab}$ denotes the height above bottom,
1418$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by
1419\[
1420  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; ,
1421\]
1422The $n_p$ parameter (given by \np{nn_zpyc}{nn\_zpyc} in \nam{zdf_iwm}{zdf\_iwm} namelist)
1423controls the stratification-dependence of the pycnocline-intensified dissipation.
1424It can take values of $1$ (recommended) or $2$.
1425Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of
1426the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps.
1427$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and
1428$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of
1429the abyssal hill topography \citep{goff_JGR10} and the latitude.
1430% Jc: input files names ?
1431
1432%% =================================================================================================
1433\section[Surface wave-induced mixing (\forcode{ln_zdfswm})]{Surface wave-induced mixing (\protect\np{ln_zdfswm}{ln\_zdfswm})}
1434\label{subsec:ZDF_swm}
1435
1436Surface waves produce an enhanced mixing through wave-turbulence interaction.
1437In addition to breaking waves induced turbulence (\autoref{subsec:ZDF_tke}),
1438the influence of non-breaking waves can be accounted introducing
1439wave-induced viscosity and diffusivity as a function of the wave number spectrum.
1440Following \citet{qiao.yuan.ea_OD10}, a formulation of wave-induced mixing coefficient
1441is provided  as a function of wave amplitude, Stokes Drift and wave-number:
1442
1443\begin{equation}
1444  \label{eq:ZDF_Bv}
1445  B_{v} = \alpha {A} {U}_{st} {exp(3kz)}
1446\end{equation}
1447
1448Where $B_{v}$ is the wave-induced mixing coefficient, $A$ is the wave amplitude,
1449${U}_{st}$ is the Stokes Drift velocity, $k$ is the wave number and $\alpha$
1450is a constant which should be determined by observations or
1451numerical experiments and is set to be 1.
1452
1453The coefficient $B_{v}$ is then directly added to the vertical viscosity
1454and diffusivity coefficients.
1455
1456In order to account for this contribution set: \forcode{ln_zdfswm=.true.},
1457then wave interaction has to be activated through \forcode{ln_wave=.true.},
1458the Stokes Drift can be evaluated by setting \forcode{ln_sdw=.true.}
1459(see \autoref{subsec:SBC_wave_sdw})
1460and the needed wave fields (significant wave height and mean wave number) can be provided either in forcing or coupled mode
1461(for more information on wave parameters and settings see \autoref{sec:SBC_wave})
1462
1463%% =================================================================================================
1464\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp})]{Adaptive-implicit vertical advection(\protect\np{ln_zad_Aimp}{ln\_zad\_Aimp})}
1465\label{subsec:ZDF_aimp}
1466
1467The adaptive-implicit vertical advection option in NEMO is based on the work of
1468\citep{shchepetkin_OM15}.  In common with most ocean models, the timestep used with NEMO
1469needs to satisfy multiple criteria associated with different physical processes in order
1470to maintain numerical stability. \citep{shchepetkin_OM15} pointed out that the vertical
1471CFL criterion is commonly the most limiting. \citep{lemarie.debreu.ea_OM15} examined the
1472constraints for a range of time and space discretizations and provide the CFL stability
1473criteria for a range of advection schemes. The values for the Leap-Frog with Robert
1474asselin filter time-stepping (as used in NEMO) are reproduced in
1475\autoref{tab:ZDF_zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these
1476restrictions but at the cost of large dispersive errors and, possibly, large numerical
1477viscosity. The adaptive-implicit vertical advection option provides a targetted use of the
1478implicit scheme only when and where potential breaches of the vertical CFL condition
1479occur. In many practical applications these events may occur remote from the main area of
1480interest or due to short-lived conditions such that the extra numerical diffusion or
1481viscosity does not greatly affect the overall solution. With such applications, setting:
1482\forcode{ln_zad_Aimp=.true.} should allow much longer model timesteps to be used whilst
1483retaining the accuracy of the high order explicit schemes over most of the domain.
1484
1485\begin{table}[htbp]
1486  \centering
1487  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}}
1488  \begin{tabular}{r|ccc}
1489    \hline
1490    spatial discretization  & 2$^nd$ order centered & 3$^rd$ order upwind & 4$^th$ order compact \\
1491    advective CFL criterion &                 0.904 &              0.472  &                0.522 \\
1492    \hline
1493  \end{tabular}
1494  \caption[Advective CFL criteria for the leapfrog with Robert Asselin filter time-stepping]{
1495    The advective CFL criteria for a range of spatial discretizations for
1496    the leapfrog with Robert Asselin filter time-stepping
1497    ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}.}
1498  \label{tab:ZDF_zad_Aimp_CFLcrit}
1499\end{table}
1500
1501In particular, the advection scheme remains explicit everywhere except where and when
1502local vertical velocities exceed a threshold set just below the explicit stability limit.
1503Once the threshold is reached a tapered transition towards an implicit scheme is used by
1504partitioning the vertical velocity into a part that can be treated explicitly and any
1505excess that must be treated implicitly. The partitioning is achieved via a Courant-number
1506dependent weighting algorithm as described in \citep{shchepetkin_OM15}.
1507
1508The local cell Courant number ($Cu$) used for this partitioning is:
1509
1510\begin{equation}
1511  \label{eq:ZDF_Eqn_zad_Aimp_Courant}
1512  \begin{split}
1513    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 ]    \\
1514       &\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 ]
1515                     \big / e_{{1_t}ij}e_{{2_t}ij}            \\
1516       &\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 ]
1517                     \big / e_{{1_t}ij}e_{{2_t}ij} \bigg )    \\
1518  \end{split}
1519\end{equation}
1520
1521\noindent and the tapering algorithm follows \citep{shchepetkin_OM15} as:
1522
1523\begin{align}
1524  \label{eq:ZDF_Eqn_zad_Aimp_partition}
1525Cu_{min} &= 0.15 \nonumber \\
1526Cu_{max} &= 0.3  \nonumber \\
1527Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\
1528Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\
1529\cf &=
1530     \begin{cases}
1531        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\
1532        (Cu - Cu_{min})^2 / (Fcu +  (Cu - Cu_{min})^2)             &\text{else if $Cu < Cu_{cut}$} \\
1533        (Cu - Cu_{max}) / Cu                                       &\text{else}
1534     \end{cases}
1535\end{align}
1536
1537\begin{figure}[!t]
1538  \centering
1539  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_coeff}
1540  \caption[Partitioning coefficient used to partition vertical velocities into parts]{
1541    The value of the partitioning coefficient (\cf) used to partition vertical velocities into
1542    parts to be treated implicitly and explicitly for a range of typical Courant numbers
1543    (\forcode{ln_zad_Aimp=.true.}).}
1544  \label{fig:ZDF_zad_Aimp_coeff}
1545\end{figure}
1546
1547\noindent The partitioning coefficient is used to determine the part of the vertical
1548velocity that must be handled implicitly ($w_i$) and to subtract this from the total
1549vertical velocity ($w_n$) to leave that which can continue to be handled explicitly:
1550
1551\begin{align}
1552  \label{eq:ZDF_Eqn_zad_Aimp_partition2}
1553    w_{i_{ijk}} &= \cf_{ijk} w_{n_{ijk}}     \nonumber \\
1554    w_{n_{ijk}} &= (1-\cf_{ijk}) w_{n_{ijk}}
1555\end{align}
1556
1557\noindent Note that the coefficient is such that the treatment is never fully implicit;
1558the three cases from \autoref{eq:ZDF_Eqn_zad_Aimp_partition} can be considered as:
1559fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the
1560coefficient (\cf) varies as shown in \autoref{fig:ZDF_zad_Aimp_coeff}. Note with these values
1561the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly
1562implicit' is 0.45 which is just below the stability limited given in
1563\autoref{tab:ZDF_zad_Aimp_CFLcrit}  for a 3rd order scheme.
1564
1565The $w_i$ component is added to the implicit solvers for the vertical mixing in
1566\mdl{dynzdf} and \mdl{trazdf} in a similar way to \citep{shchepetkin_OM15}.  This is
1567sufficient for the flux-limited advection scheme (\forcode{ln_traadv_mus}) but further
1568intervention is required when using the flux-corrected scheme (\forcode{ln_traadv_fct}).
1569For these schemes the implicit upstream fluxes must be added to both the monotonic guess
1570and to the higher order solution when calculating the antidiffusive fluxes. The implicit
1571vertical fluxes are then removed since they are added by the implicit solver later on.
1572
1573The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be
1574used in a wide range of simulations. The following test simulation, however, does illustrate
1575the potential benefits and will hopefully encourage further testing and feedback from users:
1576
1577\begin{figure}[!t]
1578  \centering
1579  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_frames}
1580  \caption[OVERFLOW: time-series of temperature vertical cross-sections]{
1581    A time-series of temperature vertical cross-sections for the OVERFLOW test case.
1582    These results are for the default settings with \forcode{nn_rdt=10.0} and
1583    without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}).}
1584  \label{fig:ZDF_zad_Aimp_overflow_frames}
1585\end{figure}
1586
1587%% =================================================================================================
1588\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case}
1589
1590The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case}
1591provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case
1592by only a few extra physics choices namely:
1593
1594\begin{forlines}
1595     ln_dynldf_OFF = .false.
1596     ln_dynldf_lap = .true.
1597     ln_dynldf_hor = .true.
1598     ln_zdfnpc     = .true.
1599     ln_traadv_fct = .true.
1600        nn_fct_h   =  2
1601        nn_fct_v   =  2
1602\end{forlines}
1603
1604\noindent which were chosen to provide a slightly more stable and less noisy solution. The
1605result when using the default value of \forcode{nn_rdt=10.} without adaptive-implicit
1606vertical velocity is illustrated in \autoref{fig:ZDF_zad_Aimp_overflow_frames}. The mass of
1607cold water, initially sitting on the shelf, moves down the slope and forms a
1608bottom-trapped, dense plume. Even with these extra physics choices the model is close to
1609stability limits and attempts with \forcode{nn_rdt=30.} will fail after about 5.5 hours
1610with excessively high horizontal velocities. This time-scale corresponds with the time the
1611plume reaches the steepest part of the topography and, although detected as a horizontal
1612CFL breach, the instability originates from a breach of the vertical CFL limit. This is a good
1613candidate, therefore, for use of the adaptive-implicit vertical advection scheme.
1614
1615The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps
1616are shown in \autoref{fig:ZDF_zad_Aimp_overflow_all_rdt} (together with the equivalent
1617frames from the base run).  In this simple example the use of the adaptive-implicit
1618vertcal advection scheme has enabled a 12x increase in the model timestep without
1619significantly altering the solution (although at this extreme the plume is more diffuse
1620and has not travelled so far).  Notably, the solution with and without the scheme is
1621slightly different even with \forcode{nn_rdt=10.}; suggesting that the base run was
1622close enough to instability to trigger the scheme despite completing successfully.
1623To assist in diagnosing how active the scheme is, in both location and time, the 3D
1624implicit and explicit components of the vertical velocity are available via XIOS as
1625\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient
1626(\cf) is also available as \texttt{wi\_cff}. For a quick oversight of
1627the schemes activity the global maximum values of the absolute implicit component
1628of the vertical velocity and the partitioning coefficient are written to the netCDF
1629version of the run statistics file (\texttt{run.stat.nc}) if this is active (see
1630\autoref{sec:MISC_opt} for activation details).
1631
1632\autoref{fig:ZDF_zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for
1633the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is
1634active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the
1635test case is close to stability limits even with this value. At the larger timesteps, the
1636vertical velocity is treated mostly implicitly at some location throughout the run. The
1637oscillatory nature of this measure appears to be linked to the progress of the plume front
1638as each cusp is associated with the location of the maximum shifting to the adjacent cell.
1639This is illustrated in \autoref{fig:ZDF_zad_Aimp_maxCf_loc} where the i- and k- locations of the
1640maximum have been overlaid for the base run case.
1641
1642\medskip
1643\noindent Only limited tests have been performed in more realistic configurations. In the
1644ORCA2\_ICE\_PISCES reference configuration the scheme does activate and passes
1645restartability and reproducibility tests but it is unable to improve the model's stability
1646enough to allow an increase in the model time-step. A view of the time-series of maximum
1647partitioning coefficient (not shown here)  suggests that the default time-step of 5400s is
1648already pushing at stability limits, especially in the initial start-up phase. The
1649time-series does not, however, exhibit any of the 'cuspiness' found with the overflow
1650tests.
1651
1652\medskip
1653\noindent A short test with an eORCA1 configuration promises more since a test using a
1654time-step of 3600s remains stable with \forcode{ln_zad_Aimp=.true.} whereas the
1655time-step is limited to 2700s without.
1656
1657\begin{figure}[!t]
1658  \centering
1659  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_all_rdt}
1660  \caption[OVERFLOW: sample temperature vertical cross-sections from mid- and end-run]{
1661    Sample temperature vertical cross-sections from mid- and end-run using
1662    different values for \forcode{nn_rdt} and with or without adaptive implicit vertical advection.
1663    Without the adaptive implicit vertical advection
1664    only the run with the shortest timestep is able to run to completion.
1665    Note also that the colour-scale has been chosen to confirm that
1666    temperatures remain within the original range of 10$^o$ to 20$^o$.}
1667  \label{fig:ZDF_zad_Aimp_overflow_all_rdt}
1668\end{figure}
1669
1670\begin{figure}[!t]
1671  \centering
1672  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf}
1673  \caption[OVERFLOW: maximum partitioning coefficient during a series of test runs]{
1674    The maximum partitioning coefficient during a series of test runs with
1675    increasing model timestep length.
1676    At the larger timesteps,
1677    the vertical velocity is treated mostly implicitly at some location throughout the run.}
1678  \label{fig:ZDF_zad_Aimp_maxCf}
1679\end{figure}
1680
1681\begin{figure}[!t]
1682  \centering
1683  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf_loc}
1684  \caption[OVERFLOW: maximum partitioning coefficient for the case overlaid]{
1685    The maximum partitioning coefficient for the \forcode{nn_rdt=10.0} case overlaid with
1686    information on the gridcell i- and k-locations of the maximum value.}
1687  \label{fig:ZDF_zad_Aimp_maxCf_loc}
1688\end{figure}
1689
1690\subinc{\input{../../global/epilogue}}
1691
1692\end{document}
Note: See TracBrowser for help on using the repository browser.