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.
Changeset 11831 for NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc/latex/NEMO/subfiles/chap_ZDF.tex – NEMO

Ignore:
Timestamp:
2019-10-29T18:14:49+01:00 (4 years ago)
Author:
laurent
Message:

Update the branch to r11830 of the trunk!

Location:
NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc
Files:
5 edited

Legend:

Unmodified
Added
Removed
  • NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc

    • Property svn:ignore deleted
    • Property svn:externals set to
      ^/utils/badges badges
      ^/utils/logos logos
  • NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc/latex

    • Property svn:ignore deleted
  • NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc/latex/NEMO

    • Property svn:ignore deleted
    • Property svn:externals set to
      ^/utils/figures/NEMO figures
  • NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc/latex/NEMO/subfiles

    • Property svn:ignore
      •  

        old new  
        1 *.aux 
        2 *.bbl 
        3 *.blg 
        4 *.dvi 
        5 *.fdb* 
        6 *.fls 
        7 *.idx 
         1*.ind 
        82*.ilg 
        9 *.ind 
        10 *.log 
        11 *.maf 
        12 *.mtc* 
        13 *.out 
        14 *.pdf 
        15 *.toc 
        16 _minted-* 
  • NEMO/branches/2019/dev_r11085_ASINTER-05_Brodeau_Advanced_Bulk/doc/latex/NEMO/subfiles/chap_ZDF.tex

    r10442 r11831  
    11\documentclass[../main/NEMO_manual]{subfiles} 
    22 
     3%% Custom aliases 
     4\newcommand{\cf}{\ensuremath{C\kern-0.14em f}} 
     5 
    36\begin{document} 
    4 % ================================================================ 
    5 % Chapter  Vertical Ocean Physics (ZDF) 
    6 % ================================================================ 
     7 
    78\chapter{Vertical Ocean Physics (ZDF)} 
    89\label{chap:ZDF} 
    910 
    10 \minitoc 
    11  
    12 %gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN. 
    13  
    14 \newpage 
    15  
    16 % ================================================================ 
    17 % Vertical Mixing 
    18 % ================================================================ 
     11\thispagestyle{plain} 
     12 
     13\chaptertoc 
     14 
     15\paragraph{Changes record} ~\\ 
     16 
     17{\footnotesize 
     18  \begin{tabularx}{\textwidth}{l||X|X} 
     19    Release & Author(s) & Modifications \\ 
     20    \hline 
     21    {\em   4.0} & {\em ...} & {\em ...} \\ 
     22    {\em   3.6} & {\em ...} & {\em ...} \\ 
     23    {\em   3.4} & {\em ...} & {\em ...} \\ 
     24    {\em <=3.4} & {\em ...} & {\em ...} 
     25  \end{tabularx} 
     26} 
     27 
     28\clearpage 
     29 
     30\cmtgm{ Add here a small introduction to ZDF and naming of the different physics 
     31(similar to what have been written for TRA and DYN).} 
     32 
     33%% ================================================================================================= 
    1934\section{Vertical mixing} 
    20 \label{sec:ZDF_zdf} 
     35\label{sec:ZDF} 
    2136 
    2237The discrete form of the ocean subgrid scale physics has been presented in 
     
    2540At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}), 
    2641while at the bottom they are set to zero for heat and salt, 
    27 unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie \key{trabbl} defined, 
     42unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie\ \np{ln_trabbc}{ln\_trabbc} defined, 
    2843see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum 
    29 (see \autoref{sec:ZDF_bfr}). 
     44(see \autoref{sec:ZDF_drg}). 
    3045 
    3146In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and 
     
    3348respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}). 
    3449These coefficients can be assumed to be either constant, or a function of the local Richardson number, 
    35 or computed from a turbulent closure model (either TKE or GLS formulation). 
    36 The computation of these coefficients is initialized in the \mdl{zdfini} module and performed in 
    37 the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} modules. 
     50or computed from a turbulent closure model (either TKE or GLS or OSMOSIS formulation). 
     51The computation of these coefficients is initialized in the \mdl{zdfphy} module and performed in 
     52the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} or \mdl{zdfosm} modules. 
    3853The trends due to the vertical momentum and tracer diffusion, including the surface forcing, 
    39 are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.  
    40 These trends can be computed using either a forward time stepping scheme 
    41 (namelist parameter \np{ln\_zdfexp}\forcode{ = .true.}) or a backward time stepping scheme 
    42 (\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing coefficients, 
    43 and thus of the formulation used (see \autoref{chap:STP}). 
    44  
    45 % ------------------------------------------------------------------------------------------------------------- 
    46 %        Constant  
    47 % ------------------------------------------------------------------------------------------------------------- 
    48 \subsection{Constant (\protect\key{zdfcst})} 
     54are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively. 
     55%These trends can be computed using either a forward time stepping scheme 
     56%(namelist parameter \np[=.true.]{ln_zdfexp}{ln\_zdfexp}) or a backward time stepping scheme 
     57%(\np[=.false.]{ln_zdfexp}{ln\_zdfexp}) depending on the magnitude of the mixing coefficients, 
     58%and thus of the formulation used (see \autoref{chap:TD}). 
     59 
     60\begin{listing} 
     61  \nlst{namzdf} 
     62  \caption{\forcode{&namzdf}} 
     63  \label{lst:namzdf} 
     64\end{listing} 
     65 
     66%% ================================================================================================= 
     67\subsection[Constant (\forcode{ln_zdfcst})]{Constant (\protect\np{ln_zdfcst}{ln\_zdfcst})} 
    4968\label{subsec:ZDF_cst} 
    50 %--------------------------------------------namzdf--------------------------------------------------------- 
    51  
    52 \nlst{namzdf} 
    53 %-------------------------------------------------------------------------------------------------------------- 
    54  
    55 Options are defined through the \ngn{namzdf} namelist variables. 
    56 When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
     69 
     70Options are defined through the \nam{zdf}{zdf} namelist variables. 
     71When \np{ln_zdfcst}{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
    5772constant values over the whole ocean. 
    5873This is the crudest way to define the vertical ocean physics. 
    59 It is recommended that this option is only used in process studies, not in basin scale simulations. 
     74It is recommended to use this option only in process studies, not in basin scale simulations. 
    6075Typical values used in this case are: 
    6176\begin{align*} 
     
    6479\end{align*} 
    6580 
    66 These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.  
     81These values are set through the \np{rn_avm0}{rn\_avm0} and \np{rn_avt0}{rn\_avt0} namelist parameters. 
    6782In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity, 
    6883that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and 
    6984$\sim10^{-9}~m^2.s^{-1}$ for salinity. 
    7085 
    71 % ------------------------------------------------------------------------------------------------------------- 
    72 %        Richardson Number Dependent 
    73 % ------------------------------------------------------------------------------------------------------------- 
    74 \subsection{Richardson number dependent (\protect\key{zdfric})} 
     86%% ================================================================================================= 
     87\subsection[Richardson number dependent (\forcode{ln_zdfric})]{Richardson number dependent (\protect\np{ln_zdfric}{ln\_zdfric})} 
    7588\label{subsec:ZDF_ric} 
    7689 
    77 %--------------------------------------------namric--------------------------------------------------------- 
    78  
    79 \nlst{namzdf_ric} 
    80 %-------------------------------------------------------------------------------------------------------------- 
    81  
    82 When \key{zdfric} is defined, a local Richardson number dependent formulation for the vertical momentum and 
    83 tracer eddy coefficients is set through the \ngn{namzdf\_ric} namelist variables. 
    84 The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.  
     90\begin{listing} 
     91  \nlst{namzdf_ric} 
     92  \caption{\forcode{&namzdf_ric}} 
     93  \label{lst:namzdf_ric} 
     94\end{listing} 
     95 
     96When \np[=.true.]{ln_zdfric}{ln\_zdfric}, a local Richardson number dependent formulation for the vertical momentum and 
     97tracer eddy coefficients is set through the \nam{zdf_ric}{zdf\_ric} namelist variables. 
     98The vertical mixing coefficients are diagnosed from the large scale variables computed by the model. 
    8599\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures. 
    86100The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to 
    87101a dependency between the vertical eddy coefficients and the local Richardson number 
    88 (\ie the ratio of stratification to vertical shear). 
    89 Following \citet{Pacanowski_Philander_JPO81}, the following formulation has been implemented: 
    90 \[ 
    91   % \label{eq:zdfric} 
     102(\ie\ the ratio of stratification to vertical shear). 
     103Following \citet{pacanowski.philander_JPO81}, the following formulation has been implemented: 
     104\[ 
     105  % \label{eq:ZDF_ric} 
    92106  \left\{ 
    93107    \begin{aligned} 
     
    98112\] 
    99113where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number, 
    100 $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),  
     114$N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}), 
    101115$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the constant case 
    102116(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that 
    103117can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$. 
    104 The last three values can be modified by setting the \np{rn\_avmri}, \np{rn\_alp} and 
    105 \np{nn\_ric} namelist parameters, respectively. 
     118The last three values can be modified by setting the \np{rn_avmri}{rn\_avmri}, \np{rn_alp}{rn\_alp} and 
     119\np{nn_ric}{nn\_ric} namelist parameters, respectively. 
    106120 
    107121A simple mixing-layer model to transfer and dissipate the atmospheric forcings 
    108 (wind-stress and buoyancy fluxes) can be activated setting the \np{ln\_mldw}\forcode{ = .true.} in the namelist. 
     122(wind-stress and buoyancy fluxes) can be activated setting the \np[=.true.]{ln_mldw}{ln\_mldw} in the namelist. 
    109123 
    110124In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and 
     
    122136\] 
    123137is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$. 
    124 The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}. 
     138The final $h_{e}$ is further constrained by the adjustable bounds \np{rn_mldmin}{rn\_mldmin} and \np{rn_mldmax}{rn\_mldmax}. 
    125139Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to 
    126 the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{Lermusiaux2001}. 
    127  
    128 % ------------------------------------------------------------------------------------------------------------- 
    129 %        TKE Turbulent Closure Scheme  
    130 % ------------------------------------------------------------------------------------------------------------- 
    131 \subsection{TKE turbulent closure scheme (\protect\key{zdftke})} 
     140the empirical values \np{rn_wtmix}{rn\_wtmix} and \np{rn_wvmix}{rn\_wvmix} \citep{lermusiaux_JMS01}. 
     141 
     142%% ================================================================================================= 
     143\subsection[TKE turbulent closure scheme (\forcode{ln_zdftke})]{TKE turbulent closure scheme (\protect\np{ln_zdftke}{ln\_zdftke})} 
    132144\label{subsec:ZDF_tke} 
    133145 
    134 %--------------------------------------------namzdf_tke-------------------------------------------------- 
    135  
    136 \nlst{namzdf_tke} 
    137 %-------------------------------------------------------------------------------------------------------------- 
     146\begin{listing} 
     147  \nlst{namzdf_tke} 
     148  \caption{\forcode{&namzdf_tke}} 
     149  \label{lst:namzdf_tke} 
     150\end{listing} 
    138151 
    139152The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on 
    140153a prognostic equation for $\bar{e}$, the turbulent kinetic energy, 
    141154and a closure assumption for the turbulent length scales. 
    142 This turbulent closure model has been developed by \citet{Bougeault1989} in the atmospheric case, 
    143 adapted by \citet{Gaspar1990} for the oceanic case, and embedded in OPA, the ancestor of NEMO, 
    144 by \citet{Blanke1993} for equatorial Atlantic simulations. 
    145 Since then, significant modifications have been introduced by \citet{Madec1998} in both the implementation and 
     155This turbulent closure model has been developed by \citet{bougeault.lacarrere_MWR89} in the atmospheric case, 
     156adapted by \citet{gaspar.gregoris.ea_JGR90} for the oceanic case, and embedded in OPA, the ancestor of \NEMO, 
     157by \citet{blanke.delecluse_JPO93} for equatorial Atlantic simulations. 
     158Since then, significant modifications have been introduced by \citet{madec.delecluse.ea_NPM98} in both the implementation and 
    146159the formulation of the mixing length scale. 
    147160The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear, 
    148 its destruction through stratification, its vertical diffusion, and its dissipation of \citet{Kolmogorov1942} type: 
     161its destruction through stratification, its vertical diffusion, and its dissipation of \citet{kolmogorov_IANS42} type: 
    149162\begin{equation} 
    150   \label{eq:zdftke_e} 
     163  \label{eq:ZDF_tke_e} 
    151164  \frac{\partial \bar{e}}{\partial t} = 
    152165  \frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2 
     
    158171\end{equation} 
    159172\[ 
    160   % \label{eq:zdftke_kz} 
     173  % \label{eq:ZDF_tke_kz} 
    161174  \begin{split} 
    162175    K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }    \\ 
     
    164177  \end{split} 
    165178\] 
    166 where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),  
    167 $l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,  
     179where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}), 
     180$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales, 
    168181$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity and diffusivity coefficients. 
    169182The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with 
    170 vertical mixing at any depth \citep{Gaspar1990}.  
    171 They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}. 
    172 $P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function of the local Richardson number, $R_i$: 
     183vertical mixing at any depth \citep{gaspar.gregoris.ea_JGR90}. 
     184They are set through namelist parameters \np{nn_ediff}{nn\_ediff} and \np{nn_ediss}{nn\_ediss}. 
     185$P_{rt}$ can be set to unity or, following \citet{blanke.delecluse_JPO93}, be a function of the local Richardson number, $R_i$: 
    173186\begin{align*} 
    174   % \label{eq:prt} 
     187  % \label{eq:ZDF_prt} 
    175188  P_{rt} = 
    176189  \begin{cases} 
     
    180193  \end{cases} 
    181194\end{align*} 
    182 Options are defined through the  \ngn{namzdfy\_tke} namelist variables. 
    183 The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable. 
     195The choice of $P_{rt}$ is controlled by the \np{nn_pdl}{nn\_pdl} namelist variable. 
    184196 
    185197At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as 
    186 $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} namelist parameter. 
    187 The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}), however a much larger value can be used when 
     198$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter. 
     199The default value of $e_{bb}$ is 3.75. \citep{gaspar.gregoris.ea_JGR90}), however a much larger value can be used when 
    188200taking into account the surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}). 
    189201The bottom value of TKE is assumed to be equal to the value of the level just above. 
    190202The time integration of the $\bar{e}$ equation may formally lead to negative values because 
    191203the numerical scheme does not ensure its positivity. 
    192 To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} namelist parameter). 
    193 Following \citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
    194 This allows the subsequent formulations to match that of \citet{Gargett1984} for the diffusion in 
     204To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn_emin}{rn\_emin} namelist parameter). 
     205Following \citet{gaspar.gregoris.ea_JGR90}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
     206This allows the subsequent formulations to match that of \citet{gargett_JMR84} for the diffusion in 
    195207the thermocline and deep ocean :  $K_\rho = 10^{-3} / N$. 
    196208In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with 
    197209too weak vertical diffusion. 
    198 They must be specified at least larger than the molecular values, and are set through \np{rn\_avm0} and 
    199 \np{rn\_avt0} (namzdf namelist, see \autoref{subsec:ZDF_cst}). 
    200  
     210They must be specified at least larger than the molecular values, and are set through \np{rn_avm0}{rn\_avm0} and 
     211\np{rn_avt0}{rn\_avt0} (\nam{zdf}{zdf} namelist, see \autoref{subsec:ZDF_cst}). 
     212 
     213%% ================================================================================================= 
    201214\subsubsection{Turbulent length scale} 
    202215 
    203216For computational efficiency, the original formulation of the turbulent length scales proposed by 
    204 \citet{Gaspar1990} has been simplified. 
    205 Four formulations are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist parameter. 
    206 The first two are based on the following first order approximation \citep{Blanke1993}: 
     217\citet{gaspar.gregoris.ea_JGR90} has been simplified. 
     218Four formulations are proposed, the choice of which is controlled by the \np{nn_mxl}{nn\_mxl} namelist parameter. 
     219The first two are based on the following first order approximation \citep{blanke.delecluse_JPO93}: 
    207220\begin{equation} 
    208   \label{eq:tke_mxl0_1} 
     221  \label{eq:ZDF_tke_mxl0_1} 
    209222  l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N 
    210223\end{equation} 
    211224which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency. 
    212225The resulting length scale is bounded by the distance to the surface or to the bottom 
    213 (\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor (\np{nn\_mxl}\forcode{ = 1}). 
    214 \citet{Blanke1993} notice that this simplification has two major drawbacks: 
     226(\np[=0]{nn_mxl}{nn\_mxl}) or by the local vertical scale factor (\np[=1]{nn_mxl}{nn\_mxl}). 
     227\citet{blanke.delecluse_JPO93} notice that this simplification has two major drawbacks: 
    215228it makes no sense for locally unstable stratification and the computation no longer uses all 
    216229the information contained in the vertical density profile. 
    217 To overcome these drawbacks, \citet{Madec1998} introduces the \np{nn\_mxl}\forcode{ = 2..3} cases, 
     230To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np[=2, 3]{nn_mxl}{nn\_mxl} cases, 
    218231which add an extra assumption concerning the vertical gradient of the computed length scale. 
    219 So, the length scales are first evaluated as in \autoref{eq:tke_mxl0_1} and then bounded such that: 
     232So, the length scales are first evaluated as in \autoref{eq:ZDF_tke_mxl0_1} and then bounded such that: 
    220233\begin{equation} 
    221   \label{eq:tke_mxl_constraint} 
     234  \label{eq:ZDF_tke_mxl_constraint} 
    222235  \frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1 
    223236  \qquad \text{with }\  l =  l_k = l_\epsilon 
    224237\end{equation} 
    225 \autoref{eq:tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
     238\autoref{eq:ZDF_tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
    226239the variations of depth. 
    227 It provides a better approximation of the \citet{Gaspar1990} formulation while being much less  
     240It provides a better approximation of the \citet{gaspar.gregoris.ea_JGR90} formulation while being much less 
    228241time consuming. 
    229242In particular, it allows the length scale to be limited not only by the distance to the surface or 
    230243to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as 
    231 the thermocline (\autoref{fig:mixing_length}). 
    232 In order to impose the \autoref{eq:tke_mxl_constraint} constraint, we introduce two additional length scales: 
     244the thermocline (\autoref{fig:ZDF_mixing_length}). 
     245In order to impose the \autoref{eq:ZDF_tke_mxl_constraint} constraint, we introduce two additional length scales: 
    233246$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and 
    234247evaluate the dissipation and mixing length scales as 
    235248(and note that here we use numerical indexing): 
    236 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    237249\begin{figure}[!t] 
    238   \begin{center} 
    239     \includegraphics[width=1.00\textwidth]{Fig_mixing_length} 
    240     \caption{ 
    241       \protect\label{fig:mixing_length} 
    242       Illustration of the mixing length computation. 
    243     } 
    244   \end{center} 
     250  \centering 
     251  \includegraphics[width=0.66\textwidth]{ZDF_mixing_length} 
     252  \caption[Mixing length computation]{Illustration of the mixing length computation} 
     253  \label{fig:ZDF_mixing_length} 
    245254\end{figure} 
    246 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    247 \[ 
    248   % \label{eq:tke_mxl2} 
     255\[ 
     256  % \label{eq:ZDF_tke_mxl2} 
    249257  \begin{aligned} 
    250258    l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right) 
     
    254262  \end{aligned} 
    255263\] 
    256 where $l^{(k)}$ is computed using \autoref{eq:tke_mxl0_1}, \ie $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
    257  
    258 In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same value: 
    259 $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nn\_mxl}\forcode{ = 3} case, 
    260 the dissipation and mixing turbulent length scales are give as in \citet{Gaspar1990}: 
    261 \[ 
    262   % \label{eq:tke_mxl_gaspar} 
     264where $l^{(k)}$ is computed using \autoref{eq:ZDF_tke_mxl0_1}, \ie\ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
     265 
     266In the \np[=2]{nn_mxl}{nn\_mxl} case, the dissipation and mixing length scales take the same value: 
     267$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np[=3]{nn_mxl}{nn\_mxl} case, 
     268the dissipation and mixing turbulent length scales are give as in \citet{gaspar.gregoris.ea_JGR90}: 
     269\[ 
     270  % \label{eq:ZDF_tke_mxl_gaspar} 
    263271  \begin{aligned} 
    264272    & l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }   \\ 
     
    267275\] 
    268276 
    269 At the ocean surface, a non zero length scale is set through the  \np{rn\_mxl0} namelist parameter. 
     277At the ocean surface, a non zero length scale is set through the  \np{rn_mxl0}{rn\_mxl0} namelist parameter. 
    270278Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and 
    271279$z_o$ the roughness parameter of the surface. 
    272 Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}. 
     280Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn_mxl0}{rn\_mxl0}. 
    273281In the ocean interior a minimum length scale is set to recover the molecular viscosity when 
    274282$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ). 
    275283 
     284%% ================================================================================================= 
    276285\subsubsection{Surface wave breaking parameterization} 
    277 %-----------------------------------------------------------------------% 
    278  
    279 Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified to 
     286 
     287Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to 
    280288include the effect of surface wave breaking energetics. 
    281289This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow. 
    282 The \citet{Mellor_Blumberg_JPO04} modifications acts on surface length scale and TKE values and 
    283 air-sea drag coefficient.  
    284 The latter concerns the bulk formulea and is not discussed here.  
    285  
    286 Following \citet{Craig_Banner_JPO94}, the boundary condition on surface TKE value is : 
     290The \citet{mellor.blumberg_JPO04} modifications acts on surface length scale and TKE values and 
     291air-sea drag coefficient. 
     292The latter concerns the bulk formulae and is not discussed here. 
     293 
     294Following \citet{craig.banner_JPO94}, the boundary condition on surface TKE value is : 
    287295\begin{equation} 
    288296  \label{eq:ZDF_Esbc} 
    289297  \bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o} 
    290298\end{equation} 
    291 where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality which depends on the ''wave age'', 
    292 ranging from 57 for mature waves to 146 for younger waves \citep{Mellor_Blumberg_JPO04}.  
     299where $\alpha_{CB}$ is the \citet{craig.banner_JPO94} constant of proportionality which depends on the ''wave age'', 
     300ranging from 57 for mature waves to 146 for younger waves \citep{mellor.blumberg_JPO04}. 
    293301The boundary condition on the turbulent length scale follows the Charnock's relation: 
    294302\begin{equation} 
     
    297305\end{equation} 
    298306where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant. 
    299 \citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by 
    300 \citet{Stacey_JPO99} citing observation evidence, and 
     307\citet{mellor.blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by 
     308\citet{stacey_JPO99} citing observation evidence, and 
    301309$\alpha_{CB} = 100$ the Craig and Banner's value. 
    302310As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$, 
    303 with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}\forcode{ = 67.83} corresponds  
     311with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter, setting \np[=67.83]{rn_ebb}{rn\_ebb} corresponds 
    304312to $\alpha_{CB} = 100$. 
    305 Further setting  \np{ln\_mxl0} to true applies \autoref{eq:ZDF_Lsbc} as surface boundary condition on length scale, 
     313Further setting  \np[=.true.]{ln_mxl0}{ln\_mxl0},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale, 
    306314with $\beta$ hard coded to the Stacey's value. 
    307 Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on 
     315Note that a minimal threshold of \np{rn_emin0}{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the 
    308316surface $\bar{e}$ value. 
    309317 
     318%% ================================================================================================= 
    310319\subsubsection{Langmuir cells} 
    311 %--------------------------------------% 
    312320 
    313321Langmuir circulations (LC) can be described as ordered large-scale vertical motions in 
     
    315323Although LC have nothing to do with convection, the circulation pattern is rather similar to 
    316324so-called convective rolls in the atmospheric boundary layer. 
    317 The detailed physics behind LC is described in, for example, \citet{Craik_Leibovich_JFM76}. 
     325The detailed physics behind LC is described in, for example, \citet{craik.leibovich_JFM76}. 
    318326The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and 
    319 wind drift currents.  
     327wind drift currents. 
    320328 
    321329Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by 
    322 \citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure. 
     330\citep{axell_JGR02} for a $k-\epsilon$ turbulent closure. 
    323331The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in 
    324 an extra source terms of TKE, $P_{LC}$. 
    325 The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled by setting \np{ln\_lc} to 
    326 \forcode{.true.} in the namtke namelist. 
    327   
    328 By making an analogy with the characteristic convective velocity scale (\eg, \citet{D'Alessio_al_JPO98}), 
    329 $P_{LC}$ is assumed to be :  
     332an extra source term of TKE, $P_{LC}$. 
     333The presence of $P_{LC}$ in \autoref{eq:ZDF_tke_e}, the TKE equation, is controlled by setting \np{ln_lc}{ln\_lc} to 
     334\forcode{.true.} in the \nam{zdf_tke}{zdf\_tke} namelist. 
     335 
     336By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}), 
     337$P_{LC}$ is assumed to be : 
    330338\[ 
    331339P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}} 
    332340\] 
    333341where $w_{LC}(z)$ is the vertical velocity profile of LC, and $H_{LC}$ is the LC depth. 
    334 With no information about the wave field, $w_{LC}$ is assumed to be proportional to  
    335 the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module  
    336 \footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity may be expressed as 
     342With no information about the wave field, $w_{LC}$ is assumed to be proportional to 
     343the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module 
     344\footnote{Following \citet{li.garrett_JMR93}, the surface Stoke drift velocity may be expressed as 
    337345  $u_s =  0.016 \,|U_{10m}|$. 
    338346  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of 
     
    341349For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at 
    342350a finite depth $H_{LC}$ (which is often close to the mixed layer depth), 
    343 and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).  
     351and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures). 
    344352The resulting expression for $w_{LC}$ is : 
    345353\[ 
     
    350358  \end{cases} 
    351359\] 
    352 where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise to fit LES data. 
     360where $c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data. 
    353361The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second. 
    354 The value of $c_{LC}$ is set through the \np{rn\_lc} namelist parameter, 
    355 having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.  
     362The value of $c_{LC}$ is set through the \np{rn_lc}{rn\_lc} namelist parameter, 
     363having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}. 
    356364 
    357365The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations: 
    358 $H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by 
    359 converting its kinetic energy to potential energy, according to  
     366$H_{LC}$ is the depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by 
     367converting its kinetic energy to potential energy, according to 
    360368\[ 
    361369- \int_{-H_{LC}}^0 { N^2\;z  \;dz} = \frac{1}{2} u_s^2 
    362370\] 
    363371 
     372%% ================================================================================================= 
    364373\subsubsection{Mixing just below the mixed layer} 
    365 %--------------------------------------------------------------% 
    366374 
    367375Vertical mixing parameterizations commonly used in ocean general circulation models tend to 
    368376produce mixed-layer depths that are too shallow during summer months and windy conditions. 
    369377This bias is particularly acute over the Southern Ocean. 
    370 To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{Rodgers_2014}.  
    371 The parameterization is an empirical one, \ie not derived from theoretical considerations, 
    372 but rather is meant to account for observed processes that affect the density structure of  
    373 the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme  
    374 (\ie near-inertial oscillations and ocean swells and waves). 
    375  
    376 When using this parameterization (\ie when \np{nn\_etau}\forcode{ = 1}), 
     378To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{rodgers.aumont.ea_B14}. 
     379The parameterization is an empirical one, \ie\ not derived from theoretical considerations, 
     380but rather is meant to account for observed processes that affect the density structure of 
     381the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme 
     382(\ie\ near-inertial oscillations and ocean swells and waves). 
     383 
     384When using this parameterization (\ie\ when \np[=1]{nn_etau}{nn\_etau}), 
    377385the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations, 
    378386swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition, 
     
    383391\end{equation} 
    384392where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that 
    385 penetrate in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of 
     393penetrates in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of 
    386394the penetration, and $f_i$ is the ice concentration 
    387 (no penetration if $f_i=1$, that is if the ocean is entirely covered by sea-ice). 
    388 The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter. 
    389 The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0}) or 
     395(no penetration if $f_i=1$, \ie\ if the ocean is entirely covered by sea-ice). 
     396The value of $f_r$, usually a few percents, is specified through \np{rn_efr}{rn\_efr} namelist parameter. 
     397The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np[=0]{nn_etau}{nn\_etau}) or 
    390398a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes 
    391 (\np{nn\_etau}\forcode{ = 1}).  
    392  
    393 Note that two other option existe, \np{nn\_etau}\forcode{ = 2..3}. 
     399(\np[=1]{nn_etau}{nn\_etau}). 
     400 
     401Note that two other option exist, \np[=2, 3]{nn_etau}{nn\_etau}. 
    394402They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer, 
    395 or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrate the ocean.  
     403or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean. 
    396404Those two options are obsolescent features introduced for test purposes. 
    397 They will be removed in the next release.  
    398  
    399 % from Burchard et al OM 2008 :  
    400 % the most critical process not reproduced by statistical turbulence models is the activity of  
    401 % internal waves and their interaction with turbulence. After the Reynolds decomposition,  
    402 % internal waves are in principle included in the RANS equations, but later partially  
    403 % excluded by the hydrostatic assumption and the model resolution.  
    404 % Thus far, the representation of internal wave mixing in ocean models has been relatively crude  
    405 % (\eg Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002). 
     405They will be removed in the next release. 
     406 
     407% This should be explain better below what this rn_eice parameter is meant for: 
     408In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn_eice}{rn\_eice} namelist parameter. 
     409This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration 
     410is greater than 25\%. 
     411 
     412% from Burchard et al OM 2008 : 
     413% the most critical process not reproduced by statistical turbulence models is the activity of 
     414% internal waves and their interaction with turbulence. After the Reynolds decomposition, 
     415% internal waves are in principle included in the RANS equations, but later partially 
     416% excluded by the hydrostatic assumption and the model resolution. 
     417% Thus far, the representation of internal wave mixing in ocean models has been relatively crude 
     418% (\eg\ Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002). 
     419 
     420%% ================================================================================================= 
     421\subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls})]{GLS: Generic Length Scale (\protect\np{ln_zdfgls}{ln\_zdfgls})} 
     422\label{subsec:ZDF_gls} 
     423 
     424\begin{listing} 
     425  \nlst{namzdf_gls} 
     426  \caption{\forcode{&namzdf_gls}} 
     427  \label{lst:namzdf_gls} 
     428\end{listing} 
     429 
     430The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
     431one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale, 
     432$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}. 
     433This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$, 
     434where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:ZDF_GLS} allows to recover a number of 
     435well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87}, 
     436$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}). 
     437The GLS scheme is given by the following set of equations: 
     438\begin{equation} 
     439  \label{eq:ZDF_gls_e} 
     440  \frac{\partial \bar{e}}{\partial t} = 
     441  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     442      +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     443  -K_\rho \,N^2 
     444  +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right] 
     445  - \epsilon 
     446\end{equation} 
     447 
     448\[ 
     449  % \label{eq:ZDF_gls_psi} 
     450  \begin{split} 
     451    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
     452      \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     453          +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     454      - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\ 
     455    &+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 } 
     456        \;\frac{\partial \psi}{\partial k}} \right]\; 
     457  \end{split} 
     458\] 
     459 
     460\[ 
     461  % \label{eq:ZDF_gls_kz} 
     462  \begin{split} 
     463    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
     464    K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l 
     465  \end{split} 
     466\] 
     467 
     468\[ 
     469  % \label{eq:ZDF_gls_eps} 
     470  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
     471\] 
     472where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and 
     473$\epsilon$ the dissipation rate. 
     474The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
     475the choice of the turbulence model. 
     476Four different turbulent models are pre-defined (\autoref{tab:ZDF_GLS}). 
     477They are made available through the \np{nn_clo}{nn\_clo} namelist parameter. 
     478 
     479\begin{table}[htbp] 
     480  \centering 
     481  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
     482  \begin{tabular}{ccccc} 
     483    &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
     484    % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\ 
     485    \hline 
     486    \hline 
     487    \np{nn_clo}{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
     488    \hline 
     489    $( p , n , m )$         &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
     490    $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
     491    $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
     492    $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
     493    $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
     494    $C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
     495    $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
     496    \hline 
     497    \hline 
     498  \end{tabular} 
     499  \caption[Set of predefined GLS parameters or equivalently predefined turbulence models available]{ 
     500    Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
     501    \protect\np[=.true.]{ln_zdfgls}{ln\_zdfgls} and controlled by 
     502    the \protect\np{nn_clos}{nn\_clos} namelist variable in \protect\nam{zdf_gls}{zdf\_gls}.} 
     503  \label{tab:ZDF_GLS} 
     504\end{table} 
     505 
     506In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
     507the mixing length towards $\kappa z_b$ ($\kappa$ is the Von Karman constant and $z_b$ the rugosity length scale) value near physical boundaries 
     508(logarithmic boundary layer law). 
     509$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88}, 
     510or by \citet{kantha.clayson_JGR94} or one of the two functions suggested by \citet{canuto.howard.ea_JPO01} 
     511(\np[=0, 3]{nn_stab_func}{nn\_stab\_func}, resp.). 
     512The value of $C_{0\mu}$ depends on the choice of the stability function. 
     513 
     514The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
     515Neumann condition through \np{nn_bc_surf}{nn\_bc\_surf} and \np{nn_bc_bot}{nn\_bc\_bot}, resp. 
     516As for TKE closure, the wave effect on the mixing is considered when 
     517\np[ > 0.]{rn_crban}{rn\_crban} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}. 
     518The \np{rn_crban}{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
     519\np{rn_charn}{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}. 
     520 
     521The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
     522almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy. 
     523With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. 
     524A value of $c_{lim} = 0.53$ is often used \citep{galperin.kantha.ea_JAS88}. 
     525\cite{umlauf.burchard_CSR05} show that the value of the clipping factor is of crucial importance for 
     526the entrainment depth predicted in stably stratified situations, 
     527and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
     528The clipping is only activated if \np[=.true.]{ln_length_lim}{ln\_length\_lim}, 
     529and the $c_{lim}$ is set to the \np{rn_clim_galp}{rn\_clim\_galp} value. 
     530 
     531The time and space discretization of the GLS equations follows the same energetic consideration as for 
     532the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{burchard_OM02}. 
     533Evaluation of the 4 GLS turbulent closure schemes can be found in \citet{warner.sherwood.ea_OM05} in ROMS model and 
     534 in \citet{reffray.guillaume.ea_GMD15} for the \NEMO\ model. 
    406535 
    407536% ------------------------------------------------------------------------------------------------------------- 
    408 %        TKE discretization considerations 
     537%        OSM OSMOSIS BL Scheme 
    409538% ------------------------------------------------------------------------------------------------------------- 
    410 \subsection{TKE discretization considerations (\protect\key{zdftke})} 
    411 \label{subsec:ZDF_tke_ene} 
    412  
    413 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     539\subsection[OSM: OSMOSIS boundary layer scheme (\forcode{ln_zdfosm = .true.})] 
     540{OSM: OSMOSIS boundary layer scheme (\protect\np{ln_zdfosm}{ln\_zdfosm})} 
     541\label{subsec:ZDF_osm} 
     542 
     543\begin{listing} 
     544  \nlst{namzdf_osm} 
     545  \caption{\forcode{&namzdf_osm}} 
     546  \label{lst:namzdf_osm} 
     547\end{listing} 
     548 
     549%-------------------------------------------------------------------------------------------------------------- 
     550\paragraph{Namelist choices} 
     551Most of the namelist options refer to how to specify the Stokes 
     552surface drift and penetration depth. There are three options: 
     553\begin{description} 
     554  \item \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave} Default value in \texttt{namelist\_ref}. In this case the Stokes drift is 
     555      assumed to be parallel to the surface wind stress, with 
     556      magnitude consistent with a constant turbulent Langmuir number 
     557    $\mathrm{La}_t=$ \protect\np{rn_m_la} {rn\_m\_la} i.e.\ 
     558    $u_{s0}=\tau/(\mathrm{La}_t^2\rho_0)$.  Default value of 
     559    \protect\np{rn_m_la}{rn\_m\_la} is 0.3. The Stokes penetration 
     560      depth $\delta = $ \protect\np{rn_osm_dstokes}{rn\_osm\_dstokes}; this has default value 
     561      of 5~m. 
     562 
     563  \item \protect\np[=1]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is 
     564      assumed to be parallel to the surface wind stress, with 
     565      magnitude as in the classical Pierson-Moskowitz wind-sea 
     566      spectrum.  Significant wave height and 
     567      wave-mean period taken from this spectrum are used to calculate the Stokes penetration 
     568      depth, following the approach set out in  \citet{breivik.janssen.ea_JPO14}. 
     569 
     570    \item \protect\np[=2]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is 
     571      taken from  ECMWF wave model output, though only the component parallel 
     572      to the wind stress is retained. Significant wave height and 
     573      wave-mean period from ECMWF wave model output are used to calculate the Stokes penetration 
     574      depth, again following \citet{breivik.janssen.ea_JPO14}. 
     575 
     576    \end{description} 
     577 
     578    Others refer to the treatment of diffusion and viscosity beneath 
     579    the surface boundary layer: 
     580\begin{description} 
     581   \item \protect\np{ln_kpprimix} {ln\_kpprimix}  Default is \np{.true.}. Switches on KPP-style Ri \#-dependent 
     582     mixing below the surface boundary layer. If this is set 
     583     \texttt{.true.}  the following variable settings are honoured: 
     584    \item \protect\np{rn_riinfty}{rn\_riinfty} Critical value of local Ri \# below which 
     585      shear instability increases vertical mixing from background value. 
     586    \item \protect\np{rn_difri} {rn\_difri} Maximum value of Ri \#-dependent mixing at $\mathrm{Ri}=0$. 
     587    \item \protect\np{ln_convmix}{ln\_convmix} If \texttt{.true.} then, where water column is unstable, specify 
     588       diffusivity equal to \protect\np{rn_dif_conv}{rn\_dif\_conv} (default value is 1 m~s$^{-2}$). 
     589 \end{description} 
     590 Diagnostic output is controlled by: 
     591  \begin{description} 
     592    \item \protect\np{ln_dia_osm}{ln\_dia\_osm} Default is \np{.false.}; allows XIOS output of OSMOSIS internal fields. 
     593  \end{description} 
     594Obsolete namelist parameters include: 
     595\begin{description} 
     596   \item \protect\np{ln_use_osm_la}\np{ln\_use\_osm\_la} With \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave}, 
     597      \protect\np{rn_osm_dstokes} {rn\_osm\_dstokes} is always used to specify the Stokes 
     598      penetration depth. 
     599   \item \protect\np{nn_ave} {nn\_ave} Choice of averaging method for KPP-style Ri \# 
     600      mixing. Not taken account of. 
     601   \item \protect\np{rn_osm_hbl0} {rn\_osm\_hbl0} Depth of initial boundary layer is now set 
     602     by a density criterion similar to that used in calculating \emph{hmlp} (output as \texttt{mldr10\_1}) in \mdl{zdfmxl}. 
     603\end{description} 
     604 
     605\subsubsection{Summary} 
     606Much of the time the turbulent motions in the ocean surface boundary 
     607layer (OSBL) are not given by 
     608classical shear turbulence. Instead they are in a regime known as 
     609`Langmuir turbulence',  dominated by an 
     610interaction between the currents and the Stokes drift of the surface waves \citep[e.g.][]{mcwilliams.ea_JFM97}. 
     611This 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 $\mathrm{La}_{t}=(u_{\ast}/u_{s0}) > 0.4$. 
     612 
     613The OSMOSIS model is fundamentally based on results of Large Eddy 
     614Simulations (LES) of Langmuir turbulence and aims to fully describe 
     615this Langmuir regime. The description in this section is of necessity incomplete and further details are available in Grant. A (2019); in prep. 
     616 
     617The OSMOSIS turbulent closure scheme is a similarity-scale scheme in 
     618the same spirit as the K-profile 
     619parameterization (KPP) scheme of \citet{large.ea_RG97}. 
     620A specified shape of diffusivity, scaled by the (OSBL) depth 
     621$h_{\mathrm{BL}}$ and a turbulent velocity scale, is imposed throughout the 
     622boundary layer 
     623$-h_{\mathrm{BL}}<z<\eta$. The turbulent closure model 
     624also includes fluxes of tracers and momentum that are``non-local'' (independent of the local property gradient). 
     625 
     626Rather than the OSBL 
     627depth being diagnosed in terms of a bulk Richardson number criterion, 
     628as in KPP, it is set by a prognostic equation that is informed by 
     629energy budget considerations reminiscent of the classical mixed layer 
     630models of \citet{kraus.turner_tellus67}. 
     631The model also includes an explicit parametrization of the structure 
     632of the pycnocline (the stratified region at the bottom of the OSBL). 
     633 
     634Presently, mixing below the OSBL is handled by the Richardson 
     635number-dependent mixing scheme used in \citet{large.ea_RG97}. 
     636 
     637Convective parameterizations such as described in \ref{sec:ZDF_conv} 
     638below should not be used with the OSMOSIS-OBL model: instabilities 
     639within the OSBL are part of the model, while instabilities below the 
     640ML are handled by the Ri \# dependent scheme. 
     641 
     642\subsubsection{Depth and velocity scales} 
     643The 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 $\Delta h$; Fig.~\ref{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). 
    414644\begin{figure}[!t] 
    415645  \begin{center} 
    416     \includegraphics[width=1.00\textwidth]{Fig_ZDF_TKE_time_scheme} 
     646    %\includegraphics[width=0.7\textwidth]{ZDF_OSM_structure_of_OSBL} 
    417647    \caption{ 
    418       \protect\label{fig:TKE_time_scheme} 
    419       Illustration of the TKE time integration and its links to the momentum and tracer time integration. 
     648      \protect\label{fig: OSBL_structure} 
     649     The structure of the entraining  boundary layer. (a) Mean buoyancy profile. (b) Profile of the buoyancy flux. 
    420650    } 
    421   \end{center}   
     651  \end{center} 
    422652\end{figure} 
    423 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     653The 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. 
     654 
     655Consideration of the power input by wind acting on the Stokes drift suggests that the Langmuir turbulence has velocity scale: 
     656\begin{equation}\label{eq:w_La} 
     657w_{*L}= \left(u_*^2 u_{s\,0}\right)^{1/3}; 
     658\end{equation} 
     659but 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: 
     660\begin{equation}\label{eq:composite-nu} 
     661  \nu_{\ast}= \left\{ u_*^3 \left[1-\exp(-.5 \mathrm{La}_t^2)\right]+w_{*L}^3\right\}^{1/3}. 
     662\end{equation} 
     663For 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: 
     664\begin{equation}\label{eq:vel-scale-unstable} 
     665\omega_* = \left(\nu_*^3 + 0.5 w_{*C}^3\right)^{1/3}. 
     666\end{equation} 
     667 
     668\subsubsection{The flux gradient model} 
     669The flux-gradient relationships used in the OSMOSIS scheme take the form: 
     670% 
     671\begin{equation}\label{eq:flux-grad-gen} 
     672\overline{w^\prime\chi^\prime}=-K\frac{\partial\overline{\chi}}{\partial z} + N_{\chi,s} +N_{\chi,b} +N_{\chi,t}, 
     673\end{equation} 
     674% 
     675where $\chi$ 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. 
     676 
     677In terms of the non-dimensionalized depth variables 
     678% 
     679\begin{equation}\label{eq:sigma} 
     680\sigma_{\mathrm{ml}}= -z/h_{\mathrm{ml}}; \;\sigma_{\mathrm{bl}}= -z/h_{\mathrm{bl}}, 
     681\end{equation} 
     682% 
     683in unstable conditions the eddy diffusivity ($K_d$) and eddy viscosity ($K_\nu$) profiles are parametrized as: 
     684% 
     685\begin{align}\label{eq:diff-unstable} 
     686K_d=&0.8\, \omega_*\, h_{\mathrm{ml}} \, \sigma_{\mathrm{ml}} \left(1-\beta_d \sigma_{\mathrm{ml}}\right)^{3/2} 
     687\\\label{eq:visc-unstable} 
     688K_\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) 
     689\end{align} 
     690% 
     691where $\beta_d$ and $\beta_\nu$ are parameters that are determined by matching Eqs \ref{eq:diff-unstable} and \ref{eq:visc-unstable} to the eddy diffusivity and viscosity at the base of the well-mixed layer, given by 
     692% 
     693\begin{equation}\label{eq:diff-wml-base} 
     694K_{d,\mathrm{ml}}=K_{\nu,\mathrm{ml}}=\,0.16\,\omega_* \Delta h. 
     695\end{equation} 
     696% 
     697For stable conditions the eddy diffusivity/viscosity profiles are given by: 
     698% 
     699\begin{align}\label{diff-stable} 
     700K_d= & 0.75\,\, \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)^{3/2} \\\label{eq:visc-stable} 
     701K_\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). 
     702\end{align} 
     703% 
     704The 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: 
     705\begin{equation}\label{eq:L_L} 
     706  L_L=-w_{*L}^3/\left<\overline{w^\prime b^\prime}\right>_L, 
     707\end{equation} 
     708with 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 
     709\begin{equation} \label{eq:stable-av-buoy-flux} 
     710\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]. 
     711\end{equation} 
     712% 
     713In unstable conditions the eddy diffusivity and viscosity depend on stability through the velocity scale $\omega_*$, which depends on the two velocity scales $\nu_*$ and $w_{*C}$. 
     714 
     715Details of the non-gradient terms in \eqref{eq:flux-grad-gen} and of the fluxes within the pycnocline $-h_{\mathrm{bl}}<z<h_{\mathrm{ml}}$ can be found in Grant (2019). 
     716 
     717\subsubsection{Evolution of the boundary layer depth} 
     718 
     719The prognostic equation for the depth of the neutral/unstable boundary layer is given by \citep{grant+etal18}, 
     720 
     721\begin{equation} \label{eq:dhdt-unstable} 
     722%\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}} 
     723\frac{\partial h_\mathrm{bl}}{\partial t} = W_b - \frac{{\overline{w^\prime b^\prime}}_\mathrm{ent}}{\Delta B_\mathrm{bl}} 
     724\end{equation} 
     725where $h_\mathrm{bl}$ is the horizontally-varying depth of the OSBL, 
     726$\mathbf{U}_b$ and $W_b$ are the mean horizontal and vertical 
     727velocities at the base of the OSBL, ${\overline{w^\prime 
     728    b^\prime}}_\mathrm{ent}$ is the buoyancy flux due to entrainment 
     729and $\Delta B_\mathrm{bl}$ is the difference between the buoyancy 
     730averaged over the depth of the OSBL (i.e.\ including the ML and 
     731pycnocline) and the buoyancy just below the base of the OSBL. This 
     732equation for the case when the pycnocline has a finite thickness, 
     733based on the potential energy budget of the OSBL, is the leading term 
     734\citep{grant+etal18} of a generalization of that used in mixed-layer 
     735models e.g.\ \citet{kraus.turner_tellus67}, in which the thickness of the pycnocline is taken to be zero. 
     736 
     737The entrainment flux for the combination of convective and Langmuir turbulence is given by 
     738\begin{equation} \label{eq:entrain-flux} 
     739  {\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}}} 
     740  + 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] 
     741\end{equation} 
     742where 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 $F(x)\equiv\tanh(x^{-1})^{0.69}$. 
     743 
     744For the stable boundary layer, the equation for the depth of the OSBL is: 
     745 
     746\begin{equation}\label{eq:dhdt-stable} 
     747\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. 
     748\end{equation} 
     749 
     750Equation. \ref{eq: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 Eq. \ref{eq: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. 
     751 
     752 
     753%% ================================================================================================= 
     754\subsection[ Discrete energy conservation for TKE and GLS schemes]{Discrete energy conservation for TKE and GLS schemes} 
     755\label{subsec:ZDF_tke_ene} 
     756 
     757\begin{figure}[!t] 
     758  \centering 
     759  \includegraphics[width=0.66\textwidth]{ZDF_TKE_time_scheme} 
     760  \caption[Subgrid kinetic energy integration in GLS and TKE schemes]{ 
     761    Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and 
     762    its links to the momentum and tracer time integration.} 
     763  \label{fig:ZDF_TKE_time_scheme} 
     764\end{figure} 
    424765 
    425766The production of turbulence by vertical shear (the first term of the right hand side of 
    426 \autoref{eq:zdftke_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
    427 (first line in \autoref{eq:PE_zdf}). 
    428 To do so a special care have to be taken for both the time and space discretization of 
    429 the TKE equation \citep{Burchard_OM02,Marsaleix_al_OM08}. 
    430  
    431 Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows how 
     767\autoref{eq:ZDF_tke_e}) and  \autoref{eq:ZDF_gls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
     768(first line in \autoref{eq:MB_zdf}). 
     769To do so a special care has to be taken for both the time and space discretization of 
     770the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}. 
     771 
     772Let us first address the time stepping issue. \autoref{fig:ZDF_TKE_time_scheme} shows how 
    432773the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with 
    433 the one-level forward time stepping of TKE equation. 
     774the one-level forward time stepping of the equation for $\bar{e}$. 
    434775With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to 
    435776the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and 
    436 summing the result vertically:    
     777summing the result vertically: 
    437778\begin{equation} 
    438   \label{eq:energ1} 
     779  \label{eq:ZDF_energ1} 
    439780  \begin{split} 
    440781    \int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\ 
     
    444785\end{equation} 
    445786Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$, 
    446 known at time $t$ (\autoref{fig:TKE_time_scheme}), as it is required when using the TKE scheme 
    447 (see \autoref{sec:STP_forward_imp}). 
    448 The first term of the right hand side of \autoref{eq:energ1} represents the kinetic energy transfer at 
     787known at time $t$ (\autoref{fig:ZDF_TKE_time_scheme}), as it is required when using the TKE scheme 
     788(see \autoref{sec:TD_forward_imp}). 
     789The first term of the right hand side of \autoref{eq:ZDF_energ1} represents the kinetic energy transfer at 
    449790the surface (atmospheric forcing) and at the bottom (friction effect). 
    450791The second term is always negative. 
    451792It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$. 
    452 \autoref{eq:energ1} implies that, to be energetically consistent, 
     793\autoref{eq:ZDF_energ1} implies that, to be energetically consistent, 
    453794the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as 
    454795${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ 
     
    456797 
    457798A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification 
    458 (second term of the right hand side of \autoref{eq:zdftke_e}). 
     799(second term of the right hand side of \autoref{eq:ZDF_tke_e} and \autoref{eq:ZDF_gls_e}). 
    459800This term must balance the input of potential energy resulting from vertical mixing. 
    460 The rate of change of potential energy (in 1D for the demonstration) due vertical mixing is obtained by 
    461 multiplying vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
     801The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by 
     802multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
    462803\begin{equation} 
    463   \label{eq:energ2} 
     804  \label{eq:ZDF_energ2} 
    464805  \begin{split} 
    465806    \int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\ 
     
    470811  \end{split} 
    471812\end{equation} 
    472 where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.  
    473 The first term of the right hand side of \autoref{eq:energ2} is always zero because 
     813where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$. 
     814The first term of the right hand side of \autoref{eq:ZDF_energ2} is always zero because 
    474815there is no diffusive flux through the ocean surface and bottom). 
    475816The second term is minus the destruction rate of  $\bar{e}$ due to stratification. 
    476 Therefore \autoref{eq:energ1} implies that, to be energetically consistent, 
    477 the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e}, the TKE equation. 
     817Therefore \autoref{eq:ZDF_energ1} implies that, to be energetically consistent, 
     818the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:ZDF_tke_e} and  \autoref{eq:ZDF_gls_e}. 
    478819 
    479820Let us now address the space discretization issue. 
    480821The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in 
    481 the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:cell}). 
     822the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:DOM_cell}). 
    482823A space averaging is thus required to obtain the shear TKE production term. 
    483 By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
     824By redoing the \autoref{eq:ZDF_energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
    484825the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
    485 Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into account. 
     826Furthermore, the time variation of $e_3$ has be taken into account. 
    486827 
    487828The above energetic considerations leads to the following final discrete form for the TKE equation: 
    488829\begin{equation} 
    489   \label{eq:zdftke_ene} 
     830  \label{eq:ZDF_tke_ene} 
    490831  \begin{split} 
    491832    \frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv 
     
    504845  \end{split} 
    505846\end{equation} 
    506 where the last two terms in \autoref{eq:zdftke_ene} (vertical diffusion and Kolmogorov dissipation) 
    507 are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}). 
     847where the last two terms in \autoref{eq:ZDF_tke_ene} (vertical diffusion and Kolmogorov dissipation) 
     848are time stepped using a backward scheme (see\autoref{sec:TD_forward_imp}). 
    508849Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible. 
    509 The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as 
    510 they all appear in the right hand side of \autoref{eq:zdftke_ene}. 
    511 For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
    512  
    513 % ------------------------------------------------------------------------------------------------------------- 
    514 %        GLS Generic Length Scale Scheme  
    515 % ------------------------------------------------------------------------------------------------------------- 
    516 \subsection{GLS: Generic Length Scale (\protect\key{zdfgls})} 
    517 \label{subsec:ZDF_gls} 
    518  
    519 %--------------------------------------------namzdf_gls--------------------------------------------------------- 
    520  
    521 \nlst{namzdf_gls} 
    522 %-------------------------------------------------------------------------------------------------------------- 
    523  
    524 The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
    525 one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale, 
    526 $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}. 
    527 This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
    528 where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of 
    529 well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982}, $k$-$\epsilon$ \citep{Rodi_1987}, 
    530 $k$-$\omega$ \citep{Wilcox_1988} among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).  
    531 The GLS scheme is given by the following set of equations: 
    532 \begin{equation} 
    533   \label{eq:zdfgls_e} 
    534   \frac{\partial \bar{e}}{\partial t} = 
    535   \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
    536       +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
    537   -K_\rho \,N^2 
    538   +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right] 
    539   - \epsilon 
    540 \end{equation} 
    541  
    542 \[ 
    543   % \label{eq:zdfgls_psi} 
    544   \begin{split} 
    545     \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
    546       \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
    547           +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
    548       - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\ 
    549     &+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 } 
    550         \;\frac{\partial \psi}{\partial k}} \right]\; 
    551   \end{split} 
    552 \] 
    553  
    554 \[ 
    555   % \label{eq:zdfgls_kz} 
    556   \begin{split} 
    557     K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
    558     K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l 
    559   \end{split} 
    560 \] 
    561  
    562 \[ 
    563   % \label{eq:zdfgls_eps} 
    564   {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
    565 \] 
    566 where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and 
    567 $\epsilon$ the dissipation rate.  
    568 The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
    569 the choice of the turbulence model. 
    570 Four different turbulent models are pre-defined (Tab.\autoref{tab:GLS}). 
    571 They are made available through the \np{nn\_clo} namelist parameter.  
    572  
    573 %--------------------------------------------------TABLE-------------------------------------------------- 
    574 \begin{table}[htbp] 
    575   \begin{center} 
    576     % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
    577     \begin{tabular}{ccccc} 
    578       &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
    579       % & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\ 
    580       \hline 
    581       \hline 
    582       \np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
    583       \hline 
    584       $( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
    585       $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
    586       $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
    587       $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
    588       $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
    589       $C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
    590       $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
    591       \hline 
    592       \hline 
    593     \end{tabular} 
    594     \caption{ 
    595       \protect\label{tab:GLS} 
    596       Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
    597       \protect\key{zdfgls} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\ngn{namzdf\_gls}. 
    598     } 
    599   \end{center} 
    600 \end{table} 
    601 %-------------------------------------------------------------------------------------------------------------- 
    602  
    603 In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
    604 the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length) value near physical boundaries 
    605 (logarithmic boundary layer law). 
    606 $C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{Galperin_al_JAS88}, 
    607 or by \citet{Kantha_Clayson_1994} or one of the two functions suggested by \citet{Canuto_2001} 
    608 (\np{nn\_stab\_func}\forcode{ = 0..3}, resp.).  
    609 The value of $C_{0\mu}$ depends of the choice of the stability function. 
    610  
    611 The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
    612 Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp. 
    613 As for TKE closure, the wave effect on the mixing is considered when 
    614 \np{ln\_crban}\forcode{ = .true.} \citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. 
    615 The \np{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
    616 \np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.  
    617  
    618 The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
    619 almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy. 
    620 With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. 
    621 A value of $c_{lim} = 0.53$ is often used \citep{Galperin_al_JAS88}. 
    622 \cite{Umlauf_Burchard_CSR05} show that the value of the clipping factor is of crucial importance for 
    623 the entrainment depth predicted in stably stratified situations, 
    624 and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
    625 The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.}, 
    626 and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
    627  
    628 The time and space discretization of the GLS equations follows the same energetic consideration as for 
    629 the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{Burchard_OM02}. 
    630 Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}. 
    631  
    632 % ------------------------------------------------------------------------------------------------------------- 
    633 %        OSM OSMOSIS BL Scheme  
    634 % ------------------------------------------------------------------------------------------------------------- 
    635 \subsection{OSM: OSMOSIS boundary layer scheme (\protect\key{zdfosm})} 
    636 \label{subsec:ZDF_osm} 
    637  
    638 %--------------------------------------------namzdf_osm--------------------------------------------------------- 
    639  
    640 \nlst{namzdf_osm} 
    641 %-------------------------------------------------------------------------------------------------------------- 
    642  
    643 The OSMOSIS turbulent closure scheme is based on......   TBC 
    644  
    645 % ================================================================ 
    646 % Convection 
    647 % ================================================================ 
     850%The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as 
     851%they all appear in the right hand side of \autoref{eq:ZDF_tke_ene}. 
     852%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored. 
     853 
     854%% ================================================================================================= 
    648855\section{Convection} 
    649856\label{sec:ZDF_conv} 
    650857 
    651 %--------------------------------------------namzdf-------------------------------------------------------- 
    652  
    653 \nlst{namzdf} 
    654 %-------------------------------------------------------------------------------------------------------------- 
    655  
    656 Static instabilities (\ie light potential densities under heavy ones) may occur at particular ocean grid points. 
     858Static instabilities (\ie\ light potential densities under heavy ones) may occur at particular ocean grid points. 
    657859In nature, convective processes quickly re-establish the static stability of the water column. 
    658860These processes have been removed from the model via the hydrostatic assumption so they must be parameterized. 
     
    661863or/and the use of a turbulent closure scheme. 
    662864 
    663 % ------------------------------------------------------------------------------------------------------------- 
    664 %       Non-Penetrative Convective Adjustment  
    665 % ------------------------------------------------------------------------------------------------------------- 
    666 \subsection[Non-penetrative convective adjmt (\protect\np{ln\_tranpc}\forcode{ = .true.})] 
    667             {Non-penetrative convective adjustment (\protect\np{ln\_tranpc}\forcode{ = .true.})} 
     865%% ================================================================================================= 
     866\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc})]{Non-penetrative convective adjustment (\protect\np{ln_tranpc}{ln\_tranpc})} 
    668867\label{subsec:ZDF_npc} 
    669868 
    670 %--------------------------------------------namzdf-------------------------------------------------------- 
    671  
    672 \nlst{namzdf} 
    673 %-------------------------------------------------------------------------------------------------------------- 
    674  
    675 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    676869\begin{figure}[!htb] 
    677   \begin{center} 
    678     \includegraphics[width=0.90\textwidth]{Fig_npc} 
    679     \caption{ 
    680       \protect\label{fig:npc} 
    681       Example of an unstable density profile treated by the non penetrative convective adjustment algorithm. 
    682       $1^{st}$ step: the initial profile is checked from the surface to the bottom. 
    683       It is found to be unstable between levels 3 and 4. 
    684       They are mixed. 
    685       The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. 
    686       The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. 
    687       The $1^{st}$ step ends since the density profile is then stable below the level 3. 
    688       $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: 
    689       levels 2 to 5 are mixed. 
    690       The new density profile is checked. 
    691       It is found stable: end of algorithm. 
    692     } 
    693   \end{center} 
     870  \centering 
     871  \includegraphics[width=0.66\textwidth]{ZDF_npc} 
     872  \caption[Unstable density profile treated by the non penetrative convective adjustment algorithm]{ 
     873    Example of an unstable density profile treated by 
     874    the non penetrative convective adjustment algorithm. 
     875    $1^{st}$ step: the initial profile is checked from the surface to the bottom. 
     876    It is found to be unstable between levels 3 and 4. 
     877    They are mixed. 
     878    The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. 
     879    The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. 
     880    The $1^{st}$ step ends since the density profile is then stable below the level 3. 
     881    $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: 
     882    levels 2 to 5 are mixed. 
     883    The new density profile is checked. 
     884    It is found stable: end of algorithm.} 
     885  \label{fig:ZDF_npc} 
    694886\end{figure} 
    695 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    696  
    697 Options are defined through the \ngn{namzdf} namelist variables. 
    698 The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}\forcode{ = .true.}. 
    699 It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of 
     887 
     888Options are defined through the \nam{zdf}{zdf} namelist variables. 
     889The non-penetrative convective adjustment is used when \np[=.true.]{ln_zdfnpc}{ln\_zdfnpc}. 
     890It is applied at each \np{nn_npc}{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of 
    700891the water column, but only until the density structure becomes neutrally stable 
    701 (\ie until the mixed portion of the water column has \textit{exactly} the density of the water just below) 
    702 \citep{Madec_al_JPO91}. 
    703 The associated algorithm is an iterative process used in the following way (\autoref{fig:npc}): 
     892(\ie\ until the mixed portion of the water column has \textit{exactly} the density of the water just below) 
     893\citep{madec.delecluse.ea_JPO91}. 
     894The associated algorithm is an iterative process used in the following way (\autoref{fig:ZDF_npc}): 
    704895starting from the top of the ocean, the first instability is found. 
    705896Assume in the following that the instability is located between levels $k$ and $k+1$. 
     
    716907This algorithm is significantly different from mixing statically unstable levels two by two. 
    717908The latter procedure cannot converge with a finite number of iterations for some vertical profiles while 
    718 the algorithm used in \NEMO converges for any profile in a number of iterations which is less than 
     909the algorithm used in \NEMO\ converges for any profile in a number of iterations which is less than 
    719910the number of vertical levels. 
    720 This property is of paramount importance as pointed out by \citet{Killworth1989}: 
     911This property is of paramount importance as pointed out by \citet{killworth_iprc89}: 
    721912it avoids the existence of permanent and unrealistic static instabilities at the sea surface. 
    722913This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in 
    723 the north-western Mediterranean Sea \citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}. 
     914the north-western Mediterranean Sea \citep{madec.delecluse.ea_JPO91, madec.chartier.ea_DAO91, madec.crepon_iprc91}. 
    724915 
    725916The current implementation has been modified in order to deal with any non linear equation of seawater 
    726917(L. Brodeau, personnal communication). 
    727918Two main differences have been introduced compared to the original algorithm: 
    728 $(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency  
    729 (not the the difference in potential density);  
     919$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency 
     920(not the difference in potential density); 
    730921$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in 
    731922the same way their temperature and salinity has been mixed. 
     
    733924having to recompute the expansion coefficients at each mixing iteration. 
    734925 
    735 % ------------------------------------------------------------------------------------------------------------- 
    736 %       Enhanced Vertical Diffusion  
    737 % ------------------------------------------------------------------------------------------------------------- 
    738 \subsection{Enhanced vertical diffusion (\protect\np{ln\_zdfevd}\forcode{ = .true.})} 
     926%% ================================================================================================= 
     927\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd})]{Enhanced vertical diffusion (\protect\np{ln_zdfevd}{ln\_zdfevd})} 
    739928\label{subsec:ZDF_evd} 
    740929 
    741 %--------------------------------------------namzdf-------------------------------------------------------- 
    742  
    743 \nlst{namzdf} 
    744 %-------------------------------------------------------------------------------------------------------------- 
    745  
    746 Options are defined through the  \ngn{namzdf} namelist variables. 
    747 The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}. 
     930Options are defined through the  \nam{zdf}{zdf} namelist variables. 
     931The enhanced vertical diffusion parameterisation is used when \np[=.true.]{ln_zdfevd}{ln\_zdfevd}. 
    748932In this case, the vertical eddy mixing coefficients are assigned very large values 
    749 (a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable 
    750 (\ie when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar_PhD97, Lazar_al_JPO99}. 
    751 This is done either on tracers only (\np{nn\_evdm}\forcode{ = 0}) or 
    752 on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}). 
    753  
    754 In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np{nn\_evdm}\forcode{ = 1}, 
     933in regions where the stratification is unstable 
     934(\ie\ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}. 
     935This is done either on tracers only (\np[=0]{nn_evdm}{nn\_evdm}) or 
     936on both momentum and tracers (\np[=1]{nn_evdm}{nn\_evdm}). 
     937 
     938In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np[=1]{nn_evdm}{nn\_evdm}, 
    755939the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to 
    756 the namelist parameter \np{rn\_avevd}. 
     940the namelist parameter \np{rn_avevd}{rn\_avevd}. 
    757941A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. 
    758942This parameterisation of convective processes is less time consuming than 
    759943the convective adjustment algorithm presented above when mixing both tracers and 
    760944momentum in the case of static instabilities. 
    761 It requires the use of an implicit time stepping on vertical diffusion terms 
    762 (\ie np{ln\_zdfexp}\forcode{ = .false.}). 
    763945 
    764946Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$. 
    765947This removes a potential source of divergence of odd and even time step in 
    766 a leapfrog environment \citep{Leclair_PhD2010} (see \autoref{sec:STP_mLF}). 
    767  
    768 % ------------------------------------------------------------------------------------------------------------- 
    769 %       Turbulent Closure Scheme  
    770 % ------------------------------------------------------------------------------------------------------------- 
    771 \subsection[Turbulent closure scheme (\protect\key{zdf}\{tke,gls,osm\})]{Turbulent Closure Scheme (\protect\key{zdftke}, \protect\key{zdfgls} or \protect\key{zdfosm})} 
     948a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:TD_mLF}). 
     949 
     950%% ================================================================================================= 
     951\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}})} 
    772952\label{subsec:ZDF_tcs} 
    773953 
    774 The turbulent closure scheme presented in \autoref{subsec:ZDF_tke} and \autoref{subsec:ZDF_gls} 
    775 (\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically unstable density profiles. 
     954The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and 
     955\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, 
     956with statically unstable density profiles. 
    776957In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in 
    777 \autoref{eq:zdftke_e} or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative.  
    778 It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring $A_u^{vm} {and}\;A_v^{vm}$ 
    779 (up to $1\;m^2s^{-1}$). 
     958\autoref{eq:ZDF_tke_e} or \autoref{eq:ZDF_gls_e} becomes a source term, since $N^2$ is negative. 
     959It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at 
     960velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$). 
    780961These large values restore the static stability of the water column in a way similar to that of 
    781962the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}). 
     
    784965because the mixing length scale is bounded by the distance to the sea surface. 
    785966It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme, 
    786 \ie setting the \np{ln\_zdfnpc} namelist parameter to true and 
    787 defining the turbulent closure CPP key all together. 
    788  
    789 The KPP turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
    790 as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
    791 therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the KPP scheme. 
     967\ie\ setting the \np{ln_zdfnpc}{ln\_zdfnpc} namelist parameter to true and 
     968defining the turbulent closure (\np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} = \forcode{.true.}) all together. 
     969 
     970The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
     971%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
     972therefore \np[=.false.]{ln_zdfevd}{ln\_zdfevd} should be used with the OSMOSIS scheme. 
    792973% gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
    793974 
    794 % ================================================================ 
    795 % Double Diffusion Mixing 
    796 % ================================================================ 
    797 \section{Double diffusion mixing (\protect\key{zdfddm})} 
    798 \label{sec:ZDF_ddm} 
    799  
    800 %-------------------------------------------namzdf_ddm------------------------------------------------- 
    801 % 
     975%% ================================================================================================= 
     976\section[Double diffusion mixing (\forcode{ln_zdfddm})]{Double diffusion mixing (\protect\np{ln_zdfddm}{ln\_zdfddm})} 
     977\label{subsec:ZDF_ddm} 
     978 
    802979%\nlst{namzdf_ddm} 
    803 %-------------------------------------------------------------------------------------------------------------- 
    804  
    805 Options are defined through the  \ngn{namzdf\_ddm} namelist variables. 
     980 
     981This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter 
     982\np{ln_zdfddm}{ln\_zdfddm} in \nam{zdf}{zdf}. 
    806983Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa. 
    807984The former condition leads to salt fingering and the latter to diffusive convection. 
    808985Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean. 
    809 \citet{Merryfield1999} include a parameterisation of such phenomena in a global ocean model and show that  
     986\citet{merryfield.holloway.ea_JPO99} include a parameterisation of such phenomena in a global ocean model and show that 
    810987it leads to relatively minor changes in circulation but exerts significant regional influences on 
    811988temperature and salinity. 
    812 This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key. 
    813  
    814 Diapycnal mixing of S and T are described by diapycnal diffusion coefficients  
     989 
     990Diapycnal mixing of S and T are described by diapycnal diffusion coefficients 
    815991\begin{align*} 
    816   % \label{eq:zdfddm_Kz} 
     992  % \label{eq:ZDF_ddm_Kz} 
    817993  &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT} \\ 
    818994  &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS} 
     
    8261002(1981): 
    8271003\begin{align} 
    828   \label{eq:zdfddm_f} 
     1004  \label{eq:ZDF_ddm_f} 
    8291005  A_f^{vS} &= 
    8301006             \begin{cases} 
     
    8321008               0                              &\text{otherwise} 
    8331009             \end{cases} 
    834   \\         \label{eq:zdfddm_f_T} 
    835   A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho  
     1010  \\         \label{eq:ZDF_ddm_f_T} 
     1011  A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
    8361012\end{align} 
    8371013 
    838 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    8391014\begin{figure}[!t] 
    840   \begin{center} 
    841     \includegraphics[width=0.99\textwidth]{Fig_zdfddm} 
    842     \caption{ 
    843       \protect\label{fig:zdfddm} 
    844       From \citet{Merryfield1999} : 
    845       (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in regions of salt fingering. 
    846       Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$; 
    847       (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in regions of 
    848       diffusive convection. 
    849       Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation. 
    850       The latter is not implemented in \NEMO. 
    851     } 
    852   \end{center} 
     1015  \centering 
     1016  \includegraphics[width=0.66\textwidth]{ZDF_ddm} 
     1017  \caption[Diapycnal diffusivities for temperature and salt in regions of salt fingering and 
     1018  diffusive convection]{ 
     1019    From \citet{merryfield.holloway.ea_JPO99}: 
     1020    (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in 
     1021    regions of salt fingering. 
     1022    Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and 
     1023    thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$; 
     1024    (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in 
     1025    regions of diffusive convection. 
     1026    Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation. 
     1027    The latter is not implemented in \NEMO.} 
     1028  \label{fig:ZDF_ddm} 
    8531029\end{figure} 
    854 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    855  
    856 The factor 0.7 in \autoref{eq:zdfddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of 
    857 buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{McDougall_Taylor_JMR84}). 
    858 Following  \citet{Merryfield1999}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
     1030 
     1031The factor 0.7 in \autoref{eq:ZDF_ddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of 
     1032buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}). 
     1033Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
    8591034 
    8601035To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by 
    861 Federov (1988) is used:  
     1036Federov (1988) is used: 
    8621037\begin{align} 
    863   % \label{eq:zdfddm_d} 
     1038  % \label{eq:ZDF_ddm_d} 
    8641039  A_d^{vT} &= 
    8651040             \begin{cases} 
     
    8691044             \end{cases} 
    8701045                                       \nonumber \\ 
    871   \label{eq:zdfddm_d_S} 
     1046  \label{eq:ZDF_ddm_d_S} 
    8721047  A_d^{vS} &= 
    8731048             \begin{cases} 
     
    8781053\end{align} 
    8791054 
    880 The dependencies of \autoref{eq:zdfddm_f} to \autoref{eq:zdfddm_d_S} on $R_\rho$ are illustrated in 
    881 \autoref{fig:zdfddm}. 
     1055The dependencies of \autoref{eq:ZDF_ddm_f} to \autoref{eq:ZDF_ddm_d_S} on $R_\rho$ are illustrated in 
     1056\autoref{fig:ZDF_ddm}. 
    8821057Implementing this requires computing $R_\rho$ at each grid point on every time step. 
    8831058This is done in \mdl{eosbn2} at the same time as $N^2$ is computed. 
    8841059This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive). 
    8851060 
    886 % ================================================================ 
    887 % Bottom Friction 
    888 % ================================================================ 
    889 \section{Bottom and top friction (\protect\mdl{zdfbfr})} 
    890 \label{sec:ZDF_bfr} 
    891  
    892 %--------------------------------------------nambfr-------------------------------------------------------- 
    893 % 
    894 %\nlst{nambfr} 
    895 %-------------------------------------------------------------------------------------------------------------- 
    896  
    897 Options to define the top and bottom friction are defined through the \ngn{nambfr} namelist variables. 
     1061%% ================================================================================================= 
     1062\section[Bottom and top friction (\textit{zdfdrg.F90})]{Bottom and top friction (\protect\mdl{zdfdrg})} 
     1063\label{sec:ZDF_drg} 
     1064 
     1065\begin{listing} 
     1066  \nlst{namdrg} 
     1067  \caption{\forcode{&namdrg}} 
     1068  \label{lst:namdrg} 
     1069\end{listing} 
     1070\begin{listing} 
     1071  \nlst{namdrg_top} 
     1072  \caption{\forcode{&namdrg_top}} 
     1073  \label{lst:namdrg_top} 
     1074\end{listing} 
     1075\begin{listing} 
     1076  \nlst{namdrg_bot} 
     1077  \caption{\forcode{&namdrg_bot}} 
     1078  \label{lst:namdrg_bot} 
     1079\end{listing} 
     1080 
     1081Options to define the top and bottom friction are defined through the \nam{drg}{drg} namelist variables. 
    8981082The bottom friction represents the friction generated by the bathymetry. 
    8991083The top friction represents the friction generated by the ice shelf/ocean interface. 
    900 As the friction processes at the top and bottom are treated in similar way, 
    901 only the bottom friction is described in detail below. 
    902  
     1084As the friction processes at the top and the bottom are treated in and identical way, 
     1085the description below considers mostly the bottom friction case, if not stated otherwise. 
    9031086 
    9041087Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as 
    9051088a condition on the vertical diffusive flux. 
    9061089For the bottom boundary layer, one has: 
    907 \[ 
    908   % \label{eq:zdfbfr_flux} 
    909   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
    910 \] 
     1090 \[ 
     1091   % \label{eq:ZDF_bfr_flux} 
     1092   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
     1093 \] 
    9111094where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside 
    9121095the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean). 
     
    9161099one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$ 
    9171100(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$). 
    918 With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.  
     1101With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m. 
    9191102When the vertical mixing coefficient is this small, using a flux condition is equivalent to 
    9201103entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or 
     
    9221105To illustrate this, consider the equation for $u$ at $k$, the last ocean level: 
    9231106\begin{equation} 
    924   \label{eq:zdfbfr_flux2} 
     1107  \label{eq:ZDF_drg_flux2} 
    9251108  \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}} 
    9261109\end{equation} 
     
    9351118 
    9361119In the code, the bottom friction is imposed by adding the trend due to the bottom friction to 
    937 the general momentum trend in \mdl{dynbfr}. 
     1120 the general momentum trend in \mdl{dynzdf}. 
    9381121For the time-split surface pressure gradient algorithm, the momentum trend due to 
    9391122the barotropic component needs to be handled separately. 
    9401123For this purpose it is convenient to compute and store coefficients which can be simply combined with 
    9411124bottom velocities and geometric values to provide the momentum trend due to bottom friction. 
    942 These coefficients are computed in \mdl{zdfbfr} and generally take the form $c_b^{\textbf U}$ where: 
     1125 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where: 
    9431126\begin{equation} 
    944   \label{eq:zdfbfr_bdef} 
     1127  \label{eq:ZDF_bfr_bdef} 
    9451128  \frac{\partial {\textbf U_h}}{\partial t} = 
    9461129  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b 
    9471130\end{equation} 
    9481131where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity. 
    949  
    950 % ------------------------------------------------------------------------------------------------------------- 
    951 %       Linear Bottom Friction 
    952 % ------------------------------------------------------------------------------------------------------------- 
    953 \subsection{Linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 0..1})} 
    954 \label{subsec:ZDF_bfr_linear} 
    955  
    956 The linear bottom friction parameterisation (including the special case of a free-slip condition) assumes that 
    957 the bottom friction is proportional to the interior velocity (\ie the velocity of the last model level): 
    958 \[ 
    959   % \label{eq:zdfbfr_linear} 
     1132Note 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. 
     1133 
     1134%% ================================================================================================= 
     1135\subsection[Linear top/bottom friction (\forcode{ln_lin})]{Linear top/bottom friction (\protect\np{ln_lin}{ln\_lin})} 
     1136\label{subsec:ZDF_drg_linear} 
     1137 
     1138The linear friction parameterisation (including the special case of a free-slip condition) assumes that 
     1139the friction is proportional to the interior velocity (\ie\ the velocity of the first/last model level): 
     1140\[ 
     1141  % \label{eq:ZDF_bfr_linear} 
    9601142  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
    9611143\] 
    962 where $r$ is a friction coefficient expressed in ms$^{-1}$. 
    963 This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,  
     1144where $r$ is a friction coefficient expressed in $m s^{-1}$. 
     1145This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean, 
    9641146and setting $r = H / \tau$, where $H$ is the ocean depth. 
    965 Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}. 
     1147Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{weatherly_JMR84}. 
    9661148A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models. 
    9671149One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ 
    968 (\citet{Gill1982}, Eq. 9.6.6). 
     1150(\citet{gill_bk82}, Eq. 9.6.6). 
    9691151For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, 
    9701152and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$. 
    9711153This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days. 
    972 It can be changed by specifying \np{rn\_bfri1} (namelist parameter). 
    973  
    974 For the linear friction case the coefficients defined in the general expression \autoref{eq:zdfbfr_bdef} are:  
    975 \[ 
    976   % \label{eq:zdfbfr_linbfr_b} 
    977   \begin{split} 
    978     c_b^u &= - r\\ 
    979     c_b^v &= - r\\ 
    980   \end{split} 
    981 \] 
    982 When \np{nn\_botfr}\forcode{ = 1}, the value of $r$ used is \np{rn\_bfri1}. 
    983 Setting \np{nn\_botfr}\forcode{ = 0} is equivalent to setting $r=0$ and 
    984 leads to a free-slip bottom boundary condition. 
    985 These values are assigned in \mdl{zdfbfr}. 
    986 From v3.2 onwards there is support for local enhancement of these values via an externally defined 2D mask array 
    987 (\np{ln\_bfr2d}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file. 
     1154It can be changed by specifying \np{rn_Uc0}{rn\_Uc0} (namelist parameter). 
     1155 
     1156 For the linear friction case the drag coefficient used in the general expression \autoref{eq:ZDF_bfr_bdef} is: 
     1157\[ 
     1158  % \label{eq:ZDF_bfr_linbfr_b} 
     1159    c_b^T = - r 
     1160\] 
     1161When \np[=.true.]{ln_lin}{ln\_lin}, the value of $r$ used is \np{rn_Uc0}{rn\_Uc0}*\np{rn_Cd0}{rn\_Cd0}. 
     1162Setting \np[=.true.]{ln_OFF}{ln\_OFF} (and \forcode{ln_lin=.true.}) is equivalent to setting $r=0$ and leads to a free-slip boundary condition. 
     1163 
     1164These values are assigned in \mdl{zdfdrg}. 
     1165Note that there is support for local enhancement of these values via an externally defined 2D mask array 
     1166(\np[=.true.]{ln_boost}{ln\_boost}) given in the \ifile{bfr\_coef} input NetCDF file. 
    9881167The mask values should vary from 0 to 1. 
    9891168Locations with a non-zero mask value will have the friction coefficient increased by 
    990 $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri1}. 
    991  
    992 % ------------------------------------------------------------------------------------------------------------- 
    993 %       Non-Linear Bottom Friction 
    994 % ------------------------------------------------------------------------------------------------------------- 
    995 \subsection{Non-linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 2})} 
    996 \label{subsec:ZDF_bfr_nonlinear} 
    997  
    998 The non-linear bottom friction parameterisation assumes that the bottom friction is quadratic:  
    999 \[ 
    1000   % \label{eq:zdfbfr_nonlinear} 
     1169$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}. 
     1170 
     1171%% ================================================================================================= 
     1172\subsection[Non-linear top/bottom friction (\forcode{ln_non_lin})]{Non-linear top/bottom friction (\protect\np{ln_non_lin}{ln\_non\_lin})} 
     1173\label{subsec:ZDF_drg_nonlinear} 
     1174 
     1175The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic: 
     1176\[ 
     1177  % \label{eq:ZDF_drg_nonlinear} 
    10011178  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h 
    10021179  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b 
    10031180\] 
    1004 where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy due to tides, 
     1181where $C_D$ is a drag coefficient, and $e_b $ a top/bottom turbulent kinetic energy due to tides, 
    10051182internal waves breaking and other short time scale currents. 
    10061183A typical value of the drag coefficient is $C_D = 10^{-3} $. 
    1007 As an example, the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and 
    1008 $e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} uses $C_D = 1.4\;10^{-3}$ and 
     1184As an example, the CME experiment \citep{treguier_JGR92} uses $C_D = 10^{-3}$ and 
     1185$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and 
    10091186$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$. 
    1010 The CME choices have been set as default values (\np{rn\_bfri2} and \np{rn\_bfeb2} namelist parameters). 
    1011  
    1012 As for the linear case, the bottom friction is imposed in the code by adding the trend due to 
    1013 the bottom friction to the general momentum trend in \mdl{dynbfr}. 
    1014 For the non-linear friction case the terms computed in \mdl{zdfbfr} are: 
    1015 \[ 
    1016   % \label{eq:zdfbfr_nonlinbfr} 
    1017   \begin{split} 
    1018     c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\ 
    1019     c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\ 
    1020   \end{split} 
    1021 \] 
    1022  
    1023 The coefficients that control the strength of the non-linear bottom friction are initialised as namelist parameters: 
    1024 $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}. 
    1025 Note for applications which treat tides explicitly a low or even zero value of \np{rn\_bfeb2} is recommended. 
    1026 From v3.2 onwards a local enhancement of $C_D$ is possible via an externally defined 2D mask array 
    1027 (\np{ln\_bfr2d}\forcode{ = .true.}). 
    1028 This works in the same way as for the linear bottom friction case with non-zero masked locations increased by 
    1029 $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri2}. 
    1030  
    1031 % ------------------------------------------------------------------------------------------------------------- 
    1032 %       Bottom Friction Log-layer 
    1033 % ------------------------------------------------------------------------------------------------------------- 
    1034 \subsection[Log-layer btm frict enhncmnt (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})] 
    1035             {Log-layer bottom friction enhancement (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})} 
    1036 \label{subsec:ZDF_bfr_loglayer} 
    1037  
    1038 In the non-linear bottom friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
    1039 a "law of the wall" scaling. 
    1040 If  \np{ln\_loglayer} = .true., $C_D$ is no longer constant but is related to the thickness of 
    1041 the last wet layer in each column by: 
    1042 \[ 
    1043   C_D = \left ( {\kappa \over {\rm log}\left ( 0.5e_{3t}/rn\_bfrz0 \right ) } \right )^2 
    1044 \] 
    1045  
    1046 \noindent where $\kappa$ is the von-Karman constant and \np{rn\_bfrz0} is a roughness length provided via 
    1047 the namelist. 
    1048  
    1049 For stability, the drag coefficient is bounded such that it is kept greater or equal to 
    1050 the base \np{rn\_bfri2} value and it is not allowed to exceed the value of an additional namelist parameter: 
    1051 \np{rn\_bfri2\_max}, \ie 
    1052 \[ 
    1053   rn\_bfri2 \leq C_D \leq rn\_bfri2\_max 
    1054 \] 
    1055  
    1056 \noindent Note also that a log-layer enhancement can also be applied to the top boundary friction if 
    1057 under ice-shelf cavities are in use (\np{ln\_isfcav}\forcode{ = .true.}). 
    1058 In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}. 
    1059  
    1060 % ------------------------------------------------------------------------------------------------------------- 
    1061 %       Bottom Friction stability 
    1062 % ------------------------------------------------------------------------------------------------------------- 
    1063 \subsection{Bottom friction stability considerations} 
    1064 \label{subsec:ZDF_bfr_stability} 
    1065  
    1066 Some care needs to exercised over the choice of parameters to ensure that the implementation of 
    1067 bottom friction does not induce numerical instability. 
    1068 For the purposes of stability analysis, an approximation to \autoref{eq:zdfbfr_flux2} is: 
     1187The CME choices have been set as default values (\np{rn_Cd0}{rn\_Cd0} and \np{rn_ke0}{rn\_ke0} namelist parameters). 
     1188 
     1189As for the linear case, the friction is imposed in the code by adding the trend due to 
     1190the friction to the general momentum trend in \mdl{dynzdf}. 
     1191For the non-linear friction case the term computed in \mdl{zdfdrg} is: 
     1192\[ 
     1193  % \label{eq:ZDF_drg_nonlinbfr} 
     1194    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2} 
     1195\] 
     1196 
     1197The coefficients that control the strength of the non-linear friction are initialised as namelist parameters: 
     1198$C_D$= \np{rn_Cd0}{rn\_Cd0}, and $e_b$ =\np{rn_bfeb2}{rn\_bfeb2}. 
     1199Note 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 
     1200(\np[=.true.]{ln_boost}{ln\_boost}). 
     1201This works in the same way as for the linear friction case with non-zero masked locations increased by 
     1202$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}. 
     1203 
     1204%% ================================================================================================= 
     1205\subsection[Log-layer top/bottom friction (\forcode{ln_loglayer})]{Log-layer top/bottom friction (\protect\np{ln_loglayer}{ln\_loglayer})} 
     1206\label{subsec:ZDF_drg_loglayer} 
     1207 
     1208In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
     1209a "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. 
     1210If  \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): 
     1211\[ 
     1212  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2 
     1213\] 
     1214 
     1215\noindent where $\kappa$ is the von-Karman constant and \np{rn_z0}{rn\_z0} is a roughness length provided via the namelist. 
     1216 
     1217The drag coefficient is bounded such that it is kept greater or equal to 
     1218the 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: 
     1219\np{rn_Cdmax}{rn\_Cdmax}, \ie 
     1220\[ 
     1221  rn\_Cd0 \leq C_D \leq rn\_Cdmax 
     1222\] 
     1223 
     1224\noindent The log-layer enhancement can also be applied to the top boundary friction if 
     1225under ice-shelf cavities are activated (\np[=.true.]{ln_isfcav}{ln\_isfcav}). 
     1226%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}. 
     1227 
     1228%% ================================================================================================= 
     1229\subsection[Explicit top/bottom friction (\forcode{ln_drgimp=.false.})]{Explicit top/bottom friction (\protect\np[=.false.]{ln_drgimp}{ln\_drgimp})} 
     1230\label{subsec:ZDF_drg_stability} 
     1231 
     1232Setting \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: 
     1233 
     1234At the top (below an ice shelf cavity): 
     1235\[ 
     1236  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t} 
     1237  = c_{t}^{\textbf{U}}\textbf{u}^{n-1}_{t} 
     1238\] 
     1239 
     1240At the bottom (above the sea floor): 
     1241\[ 
     1242  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b} 
     1243  = c_{b}^{\textbf{U}}\textbf{u}^{n-1}_{b} 
     1244\] 
     1245 
     1246Since 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. 
     1247For the purposes of stability analysis, an approximation to \autoref{eq:ZDF_drg_flux2} is: 
    10691248\begin{equation} 
    1070   \label{eq:Eqn_bfrstab} 
     1249  \label{eq:ZDF_Eqn_drgstab} 
    10711250  \begin{split} 
    10721251    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\ 
     
    10741253  \end{split} 
    10751254\end{equation} 
    1076 \noindent where linear bottom friction and a leapfrog timestep have been assumed. 
    1077 To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have: 
    1078 \[ 
    1079   |\Delta u| < \;|u|  
    1080 \] 
    1081 \noindent which, using \autoref{eq:Eqn_bfrstab}, gives: 
     1255\noindent where linear friction and a leapfrog timestep have been assumed. 
     1256To ensure that the friction cannot reverse the direction of flow it is necessary to have: 
     1257\[ 
     1258  |\Delta u| < \;|u| 
     1259\] 
     1260\noindent which, using \autoref{eq:ZDF_Eqn_drgstab}, gives: 
    10821261\[ 
    10831262  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\ 
     
    10921271For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m. 
    10931272For most applications, with physically sensible parameters these restrictions should not be of concern. 
    1094 But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.  
    1095 To ensure stability limits are imposed on the bottom friction coefficients both 
     1273But caution may be necessary if attempts are made to locally enhance the bottom friction parameters. 
     1274To ensure stability limits are imposed on the top/bottom friction coefficients both 
    10961275during initialisation and at each time step. 
    1097 Checks at initialisation are made in \mdl{zdfbfr} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
     1276Checks at initialisation are made in \mdl{zdfdrg} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
    10981277The number of breaches of the stability criterion are reported as well as 
    10991278the minimum and maximum values that have been set. 
    1100 The criterion is also checked at each time step, using the actual velocity, in \mdl{dynbfr}. 
    1101 Values of the bottom friction coefficient are reduced as necessary to ensure stability; 
     1279The criterion is also checked at each time step, using the actual velocity, in \mdl{dynzdf}. 
     1280Values of the friction coefficient are reduced as necessary to ensure stability; 
    11021281these changes are not reported. 
    11031282 
    1104 Limits on the bottom friction coefficient are not imposed if the user has elected to 
    1105 handle the bottom friction implicitly (see \autoref{subsec:ZDF_bfr_imp}). 
     1283Limits on the top/bottom friction coefficient are not imposed if the user has elected to 
     1284handle the friction implicitly (see \autoref{subsec:ZDF_drg_imp}). 
    11061285The number of potential breaches of the explicit stability criterion are still reported for information purposes. 
    11071286 
    1108 % ------------------------------------------------------------------------------------------------------------- 
    1109 %       Implicit Bottom Friction 
    1110 % ------------------------------------------------------------------------------------------------------------- 
    1111 \subsection{Implicit bottom friction (\protect\np{ln\_bfrimp}\forcode{ = .true.})} 
    1112 \label{subsec:ZDF_bfr_imp} 
     1287%% ================================================================================================= 
     1288\subsection[Implicit top/bottom friction (\forcode{ln_drgimp=.true.})]{Implicit top/bottom friction (\protect\np[=.true.]{ln_drgimp}{ln\_drgimp})} 
     1289\label{subsec:ZDF_drg_imp} 
    11131290 
    11141291An optional implicit form of bottom friction has been implemented to improve model stability. 
    1115 We recommend this option for shelf sea and coastal ocean applications, especially for split-explicit time splitting. 
    1116 This option can be invoked by setting \np{ln\_bfrimp} to \forcode{.true.} in the \textit{nambfr} namelist. 
    1117 This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \textit{namzdf} namelist.  
    1118  
    1119 This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, 
    1120 the bottom boundary condition is implemented implicitly. 
    1121  
    1122 \[ 
    1123   % \label{eq:dynzdf_bfr} 
    1124   \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{mbk} 
    1125   = \binom{c_{b}^{u}u^{n+1}_{mbk}}{c_{b}^{v}v^{n+1}_{mbk}} 
    1126 \] 
    1127  
    1128 where $mbk$ is the layer number of the bottom wet layer. 
    1129 Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so, it is implicit. 
    1130  
    1131 If split-explicit time splitting is used, care must be taken to avoid the double counting of the bottom friction in 
    1132 the 2-D barotropic momentum equations. 
    1133 As NEMO only updates the barotropic pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, 
    1134 we need to remove the bottom friction induced by these two terms which has been included in the 3-D momentum trend  
    1135 and update it with the latest value. 
    1136 On the other hand, the bottom friction contributed by the other terms 
    1137 (\eg the advection term, viscosity term) has been included in the 3-D momentum equations and 
    1138 should not be added in the 2-D barotropic mode. 
    1139  
    1140 The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the following: 
    1141  
    1142 \[ 
    1143   % \label{eq:dynspg_ts_bfr1} 
    1144   \frac{\textbf{U}_{med}-\textbf{U}^{m-1}}{2\Delta t}=-g\nabla\eta-f\textbf{k}\times\textbf{U}^{m}+c_{b} 
    1145   \left(\textbf{U}_{med}-\textbf{U}^{m-1}\right) 
    1146 \] 
    1147 \[ 
    1148   \frac{\textbf{U}^{m+1}-\textbf{U}_{med}}{2\Delta t}=\textbf{T}+ 
    1149   \left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{U}^{'}\right)- 
    1150   2\Delta t_{bc}c_{b}\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{u}_{b}\right) 
    1151 \] 
    1152  
    1153 where $\textbf{T}$ is the vertical integrated 3-D momentum trend. 
    1154 We assume the leap-frog time-stepping is used here. 
    1155 $\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step. 
    1156 $c_{b}$ is the friction coefficient. 
    1157 $\eta$ is the sea surface level calculated in the barotropic loops while $\eta^{'}$ is the sea surface level used in 
    1158 the 3-D baroclinic mode. 
    1159 $\textbf{u}_{b}$ is the bottom layer horizontal velocity. 
    1160  
    1161 % ------------------------------------------------------------------------------------------------------------- 
    1162 %       Bottom Friction with split-explicit time splitting 
    1163 % ------------------------------------------------------------------------------------------------------------- 
    1164 \subsection[Bottom friction w/ split-explicit time splitting (\protect\np{ln\_bfrimp})] 
    1165             {Bottom friction with split-explicit time splitting (\protect\np{ln\_bfrimp})} 
    1166 \label{subsec:ZDF_bfr_ts} 
    1167  
    1168 When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, 
    1169 the bottom velocity at the before time step is used. 
    1170 This velocity includes both the baroclinic and barotropic components which is appropriate when 
    1171 using either the explicit or filtered surface pressure gradient algorithms 
    1172 (\key{dynspg\_exp} or \key{dynspg\_flt}). 
    1173 Extra attention is required, however, when using split-explicit time stepping (\key{dynspg\_ts}). 
    1174 In this case the free surface equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, 
    1175 while the three dimensional prognostic variables are solved with the longer time step of \np{rn\_rdt} seconds. 
    1176 The trend in the barotropic momentum due to bottom friction appropriate to this method is that given by 
    1177 the selected parameterisation (\ie linear or non-linear bottom friction) computed with 
    1178 the evolving velocities at each barotropic timestep.  
    1179  
    1180 In the case of non-linear bottom friction, we have elected to partially linearise the problem by 
    1181 keeping the coefficients fixed throughout the barotropic time-stepping to those computed in 
    1182 \mdl{zdfbfr} using the now timestep. 
    1183 This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to: 
    1184  
     1292We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting. 
     1293This option can be invoked by setting \np{ln_drgimp}{ln\_drgimp} to \forcode{.true.} in the \nam{drg}{drg} namelist. 
     1294%This option requires \np{ln_zdfexp}{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf}{zdf} namelist. 
     1295 
     1296This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step: 
     1297 
     1298At the top (below an ice shelf cavity): 
     1299\[ 
     1300  % \label{eq:ZDF_dynZDF__drg_top} 
     1301  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t} 
     1302  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t} 
     1303\] 
     1304 
     1305At the bottom (above the sea floor): 
     1306\[ 
     1307  % \label{eq:ZDF_dynZDF__drg_bot} 
     1308  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b} 
     1309  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b} 
     1310\] 
     1311 
     1312where $t$ and $b$ refers to top and bottom layers respectively. 
     1313Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit. 
     1314 
     1315%% ================================================================================================= 
     1316\subsection[Bottom friction with split-explicit free surface]{Bottom friction with split-explicit free surface} 
     1317\label{subsec:ZDF_drg_ts} 
     1318 
     1319With 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. 
     1320 
     1321The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO\ is as follows: 
    11851322\begin{enumerate} 
    1186 \item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before barotropic velocity to 
    1187   the bottom friction component of the vertically integrated momentum trend. 
    1188   Note the same stability check that is carried out on the bottom friction coefficient in \mdl{dynbfr} has to 
    1189   be applied here to ensure that the trend removed matches that which was added in \mdl{dynbfr}. 
    1190 \item At each barotropic step, compute the contribution of the current barotropic velocity to 
    1191   the trend due to bottom friction. 
    1192   Add this contribution to the vertically integrated momentum trend. 
    1193   This contribution is handled implicitly which eliminates the need to impose a stability criteria on 
    1194   the values of the bottom friction coefficient within the barotropic loop.  
     1323\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. 
     1324\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. 
    11951325\end{enumerate} 
    11961326 
    1197 Note that the use of an implicit formulation within the barotropic loop for the bottom friction trend means that 
    1198 any limiting of the bottom friction coefficient in \mdl{dynbfr} does not adversely affect the solution when 
    1199 using split-explicit time splitting. 
    1200 This is because the major contribution to bottom friction is likely to come from the barotropic component which 
    1201 uses the unrestricted value of the coefficient. 
    1202 However, if the limiting is thought to be having a major effect 
    1203 (a more likely prospect in coastal and shelf seas applications) then 
    1204 the fully implicit form of the bottom friction should be used (see \autoref{subsec:ZDF_bfr_imp}) 
    1205 which can be selected by setting \np{ln\_bfrimp} $=$ \forcode{.true.}. 
    1206  
    1207 Otherwise, the implicit formulation takes the form: 
    1208 \[ 
    1209   % \label{eq:zdfbfr_implicitts} 
    1210   \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ] 
    1211 \] 
    1212 where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),  
    1213 $c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and 
    1214 $RHS$ represents all the components to the vertically integrated momentum trend except for 
    1215 that due to bottom friction. 
    1216  
    1217 % ================================================================ 
    1218 % Tidal Mixing 
    1219 % ================================================================ 
    1220 \section{Tidal mixing (\protect\key{zdftmx})} 
    1221 \label{sec:ZDF_tmx} 
    1222  
    1223 %--------------------------------------------namzdf_tmx-------------------------------------------------- 
    1224 % 
    1225 %\nlst{namzdf_tmx} 
    1226 %-------------------------------------------------------------------------------------------------------------- 
    1227  
    1228  
    1229 % ------------------------------------------------------------------------------------------------------------- 
    1230 %        Bottom intensified tidal mixing  
    1231 % ------------------------------------------------------------------------------------------------------------- 
    1232 \subsection{Bottom intensified tidal mixing} 
    1233 \label{subsec:ZDF_tmx_bottom} 
    1234  
    1235 Options are defined through the  \ngn{namzdf\_tmx} namelist variables. 
    1236 The parameterization of tidal mixing follows the general formulation for the vertical eddy diffusivity proposed by 
    1237 \citet{St_Laurent_al_GRL02} and first introduced in an OGCM by \citep{Simmons_al_OM04}.  
    1238 In this formulation an additional vertical diffusivity resulting from internal tide breaking, 
    1239 $A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, 
    1240 the energy transfer from barotropic tides to baroclinic tides: 
    1241 \begin{equation} 
    1242   \label{eq:Ktides} 
    1243   A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 } 
    1244 \end{equation} 
    1245 where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}), 
    1246 $\rho$ the density, $q$ the tidal dissipation efficiency, and $F(z)$ the vertical structure function.  
    1247  
    1248 The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) and 
    1249 is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).  
    1250 The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter) 
    1251 represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally, 
    1252 with the remaining $1-q$ radiating away as low mode internal waves and 
    1253 contributing to the background internal wave field. 
    1254 A value of $q=1/3$ is typically used \citet{St_Laurent_al_GRL02}. 
    1255 The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical. 
    1256 It is implemented as a simple exponential decaying upward away from the bottom, 
    1257 with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, 
    1258 with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},  
    1259 \[ 
    1260   % \label{eq:Fz} 
    1261   F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) } 
    1262 \] 
    1263 and is normalized so that vertical integral over the water column is unity.  
    1264  
    1265 The associated vertical viscosity is calculated from the vertical diffusivity assuming a Prandtl number of 1, 
    1266 \ie $A^{vm}_{tides}=A^{vT}_{tides}$. 
    1267 In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity is capped at $300\,cm^2/s$ and 
    1268 impose a lower limit on $N^2$ of \np{rn\_n2min} usually set to $10^{-8} s^{-2}$. 
    1269 These bounds are usually rarely encountered. 
    1270  
    1271 The internal wave energy map, $E(x, y)$ in \autoref{eq:Ktides}, is derived from a barotropic model of 
    1272 the tides utilizing a parameterization of the conversion of barotropic tidal energy into internal waves. 
    1273 The essential goal of the parameterization is to represent the momentum exchange between the barotropic tides and 
    1274 the unrepresented internal waves induced by the tidal flow over rough topography in a stratified ocean. 
    1275 In the current version of \NEMO, the map is built from the output of 
    1276 the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}. 
    1277 This model provides the dissipation associated with internal wave energy for the M2 and K1 tides component 
    1278 (\autoref{fig:ZDF_M2_K1_tmx}). 
    1279 The S2 dissipation is simply approximated as being $1/4$ of the M2 one. 
    1280 The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$. 
    1281 Its global mean value is $1.1$ TW, 
    1282 in agreement with independent estimates \citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.  
    1283  
    1284 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    1285 \begin{figure}[!t] 
    1286   \begin{center} 
    1287     \includegraphics[width=0.90\textwidth]{Fig_ZDF_M2_K1_tmx} 
    1288     \caption{ 
    1289       \protect\label{fig:ZDF_M2_K1_tmx} 
    1290       (a) M2 and (b) K1 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). 
    1291     } 
    1292   \end{center} 
    1293 \end{figure} 
    1294 %>>>>>>>>>>>>>>>>>>>>>>>>>>>>  
    1295   
    1296 % ------------------------------------------------------------------------------------------------------------- 
    1297 %        Indonesian area specific treatment  
    1298 % ------------------------------------------------------------------------------------------------------------- 
    1299 \subsection{Indonesian area specific treatment (\protect\np{ln\_zdftmx\_itf})} 
    1300 \label{subsec:ZDF_tmx_itf} 
    1301  
    1302 When the Indonesian Through Flow (ITF) area is included in the model domain, 
    1303 a specific treatment of tidal induced mixing in this area can be used. 
    1304 It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide an input NetCDF file, 
    1305 \ifile{mask\_itf}, which contains a mask array defining the ITF area where the specific treatment is applied. 
    1306  
    1307 When \np{ln\_tmx\_itf}\forcode{ = .true.}, the two key parameters $q$ and $F(z)$ are adjusted following 
    1308 the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}: 
    1309  
    1310 First, the Indonesian archipelago is a complex geographic region with a series of 
    1311 large, deep, semi-enclosed basins connected via numerous narrow straits. 
    1312 Once generated, internal tides remain confined within this semi-enclosed area and hardly radiate away. 
    1313 Therefore all the internal tides energy is consumed within this area. 
    1314 So it is assumed that $q = 1$, \ie all the energy generated is available for mixing. 
    1315 Note that for test purposed, the ITF tidal dissipation efficiency is a namelist parameter (\np{rn\_tfe\_itf}). 
    1316 A value of $1$ or close to is this recommended for this parameter. 
    1317  
    1318 Second, the vertical structure function, $F(z)$, is no more associated with a bottom intensification of the mixing, 
    1319 but with a maximum of energy available within the thermocline. 
    1320 \citet{Koch-Larrouy_al_GRL07} have suggested that the vertical distribution of 
    1321 the energy dissipation proportional to $N^2$ below the core of the thermocline and to $N$ above.  
    1322 The resulting $F(z)$ is: 
    1323 \[ 
    1324   % \label{eq:Fz_itf} 
    1325   F(i,j,k) \sim     \left\{ 
    1326     \begin{aligned} 
    1327       \frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\ 
    1328       \frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$} 
    1329     \end{aligned} 
    1330   \right. 
    1331 \] 
    1332  
    1333 Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,  
    1334 which agrees with the independent estimates inferred from observations. 
    1335 Introduced in a regional OGCM, the parameterization improves the water mass characteristics in 
    1336 the different Indonesian seas, suggesting that the horizontal and vertical distributions of 
    1337 the mixing are adequately prescribed \citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}. 
    1338 Note also that such a parameterisation has a significant impact on the behaviour of 
    1339 global coupled GCMs \citep{Koch-Larrouy_al_CD10}. 
    1340  
    1341 % ================================================================ 
    1342 % Internal wave-driven mixing 
    1343 % ================================================================ 
    1344 \section{Internal wave-driven mixing (\protect\key{zdftmx\_new})} 
    1345 \label{sec:ZDF_tmx_new} 
    1346  
    1347 %--------------------------------------------namzdf_tmx_new------------------------------------------ 
    1348 % 
    1349 %\nlst{namzdf_tmx_new} 
    1350 %-------------------------------------------------------------------------------------------------------------- 
     1327Note that other strategies are possible, like considering vertical diffusion step in advance, \ie\ prior barotropic integration. 
     1328 
     1329%% ================================================================================================= 
     1330\section[Internal wave-driven mixing (\forcode{ln_zdfiwm})]{Internal wave-driven mixing (\protect\np{ln_zdfiwm}{ln\_zdfiwm})} 
     1331\label{subsec:ZDF_tmx_new} 
     1332 
     1333\begin{listing} 
     1334  \nlst{namzdf_iwm} 
     1335  \caption{\forcode{&namzdf_iwm}} 
     1336  \label{lst:namzdf_iwm} 
     1337\end{listing} 
    13511338 
    13521339The parameterization of mixing induced by breaking internal waves is a generalization of 
    1353 the approach originally proposed by \citet{St_Laurent_al_GRL02}. 
     1340the approach originally proposed by \citet{st-laurent.simmons.ea_GRL02}. 
    13541341A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed, 
    1355 and the resulting diffusivity is obtained as  
    1356 \[ 
    1357   % \label{eq:Kwave} 
     1342and the resulting diffusivity is obtained as 
     1343\[ 
     1344  % \label{eq:ZDF_Kwave} 
    13581345  A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 } 
    13591346\] 
    13601347where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of 
    13611348the energy available for mixing. 
    1362 If the \np{ln\_mevar} namelist parameter is set to false, the mixing efficiency is taken as constant and 
    1363 equal to 1/6 \citep{Osborn_JPO80}. 
     1349If the \np{ln_mevar}{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and 
     1350equal to 1/6 \citep{osborn_JPO80}. 
    13641351In the opposite (recommended) case, $R_f$ is instead a function of 
    13651352the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$, 
    1366 with $\nu$ the molecular viscosity of seawater, following the model of \cite{Bouffard_Boegman_DAO2013} and 
    1367 the implementation of \cite{de_lavergne_JPO2016_efficiency}. 
     1353with $\nu$ the molecular viscosity of seawater, following the model of \cite{bouffard.boegman_DAO13} and 
     1354the implementation of \cite{de-lavergne.madec.ea_JPO16}. 
    13681355Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when 
    13691356the mixing efficiency is constant. 
    13701357 
    1371 In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary  
    1372 as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to true, a recommended choice.  
    1373 This parameterization of differential mixing, due to \cite{Jackson_Rehmann_JPO2014}, 
    1374 is implemented as in \cite{de_lavergne_JPO2016_efficiency}. 
     1358In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary 
     1359as a function of $Re_b$ by setting the \np{ln_tsdiff}{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice. 
     1360This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14}, 
     1361is implemented as in \cite{de-lavergne.madec.ea_JPO16}. 
    13751362 
    13761363The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$, 
    13771364is constructed from three static maps of column-integrated internal wave energy dissipation, 
    1378 $E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures 
    1379 (de Lavergne et al., in prep): 
     1365$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures: 
     1366 
    13801367\begin{align*} 
    13811368  F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\ 
    1382   F_{pyc}(i,j,k) &\propto N^{n\_p}\\ 
     1369  F_{pyc}(i,j,k) &\propto N^{n_p}\\ 
    13831370  F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} } 
    1384 \end{align*}  
     1371\end{align*} 
    13851372In the above formula, $h_{ab}$ denotes the height above bottom, 
    13861373$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by 
     
    13881375  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; , 
    13891376\] 
    1390 The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_tmx\_new} namelist) 
     1377The $n_p$ parameter (given by \np{nn_zpyc}{nn\_zpyc} in \nam{zdf_iwm}{zdf\_iwm} namelist) 
    13911378controls the stratification-dependence of the pycnocline-intensified dissipation. 
    1392 It can take values of 1 (recommended) or 2. 
     1379It can take values of $1$ (recommended) or $2$. 
    13931380Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of 
    13941381the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps. 
    13951382$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and 
    13961383$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of 
    1397 the abyssal hill topography \citep{Goff_JGR2010} and the latitude. 
    1398  
    1399 % ================================================================ 
    1400  
    1401 \biblio 
    1402  
    1403 \pindex 
     1384the abyssal hill topography \citep{goff_JGR10} and the latitude. 
     1385% Jc: input files names ? 
     1386 
     1387%% ================================================================================================= 
     1388\section[Surface wave-induced mixing (\forcode{ln_zdfswm})]{Surface wave-induced mixing (\protect\np{ln_zdfswm}{ln\_zdfswm})} 
     1389\label{subsec:ZDF_swm} 
     1390 
     1391Surface waves produce an enhanced mixing through wave-turbulence interaction. 
     1392In addition to breaking waves induced turbulence (\autoref{subsec:ZDF_tke}), 
     1393the influence of non-breaking waves can be accounted introducing 
     1394wave-induced viscosity and diffusivity as a function of the wave number spectrum. 
     1395Following \citet{qiao.yuan.ea_OD10}, a formulation of wave-induced mixing coefficient 
     1396is provided  as a function of wave amplitude, Stokes Drift and wave-number: 
     1397 
     1398\begin{equation} 
     1399  \label{eq:ZDF_Bv} 
     1400  B_{v} = \alpha {A} {U}_{st} {exp(3kz)} 
     1401\end{equation} 
     1402 
     1403Where $B_{v}$ is the wave-induced mixing coefficient, $A$ is the wave amplitude, 
     1404${U}_{st}$ is the Stokes Drift velocity, $k$ is the wave number and $\alpha$ 
     1405is a constant which should be determined by observations or 
     1406numerical experiments and is set to be 1. 
     1407 
     1408The coefficient $B_{v}$ is then directly added to the vertical viscosity 
     1409and diffusivity coefficients. 
     1410 
     1411In order to account for this contribution set: \forcode{ln_zdfswm=.true.}, 
     1412then wave interaction has to be activated through \forcode{ln_wave=.true.}, 
     1413the Stokes Drift can be evaluated by setting \forcode{ln_sdw=.true.} 
     1414(see \autoref{subsec:SBC_wave_sdw}) 
     1415and the needed wave fields can be provided either in forcing or coupled mode 
     1416(for more information on wave parameters and settings see \autoref{sec:SBC_wave}) 
     1417 
     1418%% ================================================================================================= 
     1419\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp})]{Adaptive-implicit vertical advection(\protect\np{ln_zad_Aimp}{ln\_zad\_Aimp})} 
     1420\label{subsec:ZDF_aimp} 
     1421 
     1422The adaptive-implicit vertical advection option in NEMO is based on the work of 
     1423\citep{shchepetkin_OM15}.  In common with most ocean models, the timestep used with NEMO 
     1424needs to satisfy multiple criteria associated with different physical processes in order 
     1425to maintain numerical stability. \citep{shchepetkin_OM15} pointed out that the vertical 
     1426CFL criterion is commonly the most limiting. \citep{lemarie.debreu.ea_OM15} examined the 
     1427constraints for a range of time and space discretizations and provide the CFL stability 
     1428criteria for a range of advection schemes. The values for the Leap-Frog with Robert 
     1429asselin filter time-stepping (as used in NEMO) are reproduced in 
     1430\autoref{tab:ZDF_zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these 
     1431restrictions but at the cost of large dispersive errors and, possibly, large numerical 
     1432viscosity. The adaptive-implicit vertical advection option provides a targetted use of the 
     1433implicit scheme only when and where potential breaches of the vertical CFL condition 
     1434occur. In many practical applications these events may occur remote from the main area of 
     1435interest or due to short-lived conditions such that the extra numerical diffusion or 
     1436viscosity does not greatly affect the overall solution. With such applications, setting: 
     1437\forcode{ln_zad_Aimp=.true.} should allow much longer model timesteps to be used whilst 
     1438retaining the accuracy of the high order explicit schemes over most of the domain. 
     1439 
     1440\begin{table}[htbp] 
     1441  \centering 
     1442  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}} 
     1443  \begin{tabular}{r|ccc} 
     1444    \hline 
     1445    spatial discretization  & 2$^nd$ order centered & 3$^rd$ order upwind & 4$^th$ order compact \\ 
     1446    advective CFL criterion &                 0.904 &              0.472  &                0.522 \\ 
     1447    \hline 
     1448  \end{tabular} 
     1449  \caption[Advective CFL criteria for the leapfrog with Robert Asselin filter time-stepping]{ 
     1450    The advective CFL criteria for a range of spatial discretizations for 
     1451    the leapfrog with Robert Asselin filter time-stepping 
     1452    ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}.} 
     1453  \label{tab:ZDF_zad_Aimp_CFLcrit} 
     1454\end{table} 
     1455 
     1456In particular, the advection scheme remains explicit everywhere except where and when 
     1457local vertical velocities exceed a threshold set just below the explicit stability limit. 
     1458Once the threshold is reached a tapered transition towards an implicit scheme is used by 
     1459partitioning the vertical velocity into a part that can be treated explicitly and any 
     1460excess that must be treated implicitly. The partitioning is achieved via a Courant-number 
     1461dependent weighting algorithm as described in \citep{shchepetkin_OM15}. 
     1462 
     1463The local cell Courant number ($Cu$) used for this partitioning is: 
     1464 
     1465\begin{equation} 
     1466  \label{eq:ZDF_Eqn_zad_Aimp_Courant} 
     1467  \begin{split} 
     1468    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 ]    \\ 
     1469       &\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 ] 
     1470                     \big / e_{{1_t}ij}e_{{2_t}ij}            \\ 
     1471       &\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 ] 
     1472                     \big / e_{{1_t}ij}e_{{2_t}ij} \bigg )    \\ 
     1473  \end{split} 
     1474\end{equation} 
     1475 
     1476\noindent and the tapering algorithm follows \citep{shchepetkin_OM15} as: 
     1477 
     1478\begin{align} 
     1479  \label{eq:ZDF_Eqn_zad_Aimp_partition} 
     1480Cu_{min} &= 0.15 \nonumber \\ 
     1481Cu_{max} &= 0.3  \nonumber \\ 
     1482Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\ 
     1483Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\ 
     1484\cf &= 
     1485     \begin{cases} 
     1486        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\ 
     1487        (Cu - Cu_{min})^2 / (Fcu +  (Cu - Cu_{min})^2)             &\text{else if $Cu < Cu_{cut}$} \\ 
     1488        (Cu - Cu_{max}) / Cu                                       &\text{else} 
     1489     \end{cases} 
     1490\end{align} 
     1491 
     1492\begin{figure}[!t] 
     1493  \centering 
     1494  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_coeff} 
     1495  \caption[Partitioning coefficient used to partition vertical velocities into parts]{ 
     1496    The value of the partitioning coefficient (\cf) used to partition vertical velocities into 
     1497    parts to be treated implicitly and explicitly for a range of typical Courant numbers 
     1498    (\forcode{ln_zad_Aimp=.true.}).} 
     1499  \label{fig:ZDF_zad_Aimp_coeff} 
     1500\end{figure} 
     1501 
     1502\noindent The partitioning coefficient is used to determine the part of the vertical 
     1503velocity that must be handled implicitly ($w_i$) and to subtract this from the total 
     1504vertical velocity ($w_n$) to leave that which can continue to be handled explicitly: 
     1505 
     1506\begin{align} 
     1507  \label{eq:ZDF_Eqn_zad_Aimp_partition2} 
     1508    w_{i_{ijk}} &= \cf_{ijk} w_{n_{ijk}}     \nonumber \\ 
     1509    w_{n_{ijk}} &= (1-\cf_{ijk}) w_{n_{ijk}} 
     1510\end{align} 
     1511 
     1512\noindent Note that the coefficient is such that the treatment is never fully implicit; 
     1513the three cases from \autoref{eq:ZDF_Eqn_zad_Aimp_partition} can be considered as: 
     1514fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the 
     1515coefficient (\cf) varies as shown in \autoref{fig:ZDF_zad_Aimp_coeff}. Note with these values 
     1516the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly 
     1517implicit' is 0.45 which is just below the stability limited given in 
     1518\autoref{tab:ZDF_zad_Aimp_CFLcrit}  for a 3rd order scheme. 
     1519 
     1520The $w_i$ component is added to the implicit solvers for the vertical mixing in 
     1521\mdl{dynzdf} and \mdl{trazdf} in a similar way to \citep{shchepetkin_OM15}.  This is 
     1522sufficient for the flux-limited advection scheme (\forcode{ln_traadv_mus}) but further 
     1523intervention is required when using the flux-corrected scheme (\forcode{ln_traadv_fct}). 
     1524For these schemes the implicit upstream fluxes must be added to both the monotonic guess 
     1525and to the higher order solution when calculating the antidiffusive fluxes. The implicit 
     1526vertical fluxes are then removed since they are added by the implicit solver later on. 
     1527 
     1528The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be 
     1529used in a wide range of simulations. The following test simulation, however, does illustrate 
     1530the potential benefits and will hopefully encourage further testing and feedback from users: 
     1531 
     1532\begin{figure}[!t] 
     1533  \centering 
     1534  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_frames} 
     1535  \caption[OVERFLOW: time-series of temperature vertical cross-sections]{ 
     1536    A time-series of temperature vertical cross-sections for the OVERFLOW test case. 
     1537    These results are for the default settings with \forcode{nn_rdt=10.0} and 
     1538    without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}).} 
     1539  \label{fig:ZDF_zad_Aimp_overflow_frames} 
     1540\end{figure} 
     1541 
     1542%% ================================================================================================= 
     1543\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case} 
     1544 
     1545The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case} 
     1546provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case 
     1547by only a few extra physics choices namely: 
     1548 
     1549\begin{verbatim} 
     1550     ln_dynldf_OFF = .false. 
     1551     ln_dynldf_lap = .true. 
     1552     ln_dynldf_hor = .true. 
     1553     ln_zdfnpc     = .true. 
     1554     ln_traadv_fct = .true. 
     1555        nn_fct_h   =  2 
     1556        nn_fct_v   =  2 
     1557\end{verbatim} 
     1558 
     1559\noindent which were chosen to provide a slightly more stable and less noisy solution. The 
     1560result when using the default value of \forcode{nn_rdt=10.} without adaptive-implicit 
     1561vertical velocity is illustrated in \autoref{fig:ZDF_zad_Aimp_overflow_frames}. The mass of 
     1562cold water, initially sitting on the shelf, moves down the slope and forms a 
     1563bottom-trapped, dense plume. Even with these extra physics choices the model is close to 
     1564stability limits and attempts with \forcode{nn_rdt=30.} will fail after about 5.5 hours 
     1565with excessively high horizontal velocities. This time-scale corresponds with the time the 
     1566plume reaches the steepest part of the topography and, although detected as a horizontal 
     1567CFL breach, the instability originates from a breach of the vertical CFL limit. This is a good 
     1568candidate, therefore, for use of the adaptive-implicit vertical advection scheme. 
     1569 
     1570The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps 
     1571are shown in \autoref{fig:ZDF_zad_Aimp_overflow_all_rdt} (together with the equivalent 
     1572frames from the base run).  In this simple example the use of the adaptive-implicit 
     1573vertcal advection scheme has enabled a 12x increase in the model timestep without 
     1574significantly altering the solution (although at this extreme the plume is more diffuse 
     1575and has not travelled so far).  Notably, the solution with and without the scheme is 
     1576slightly different even with \forcode{nn_rdt=10.}; suggesting that the base run was 
     1577close enough to instability to trigger the scheme despite completing successfully. 
     1578To assist in diagnosing how active the scheme is, in both location and time, the 3D 
     1579implicit and explicit components of the vertical velocity are available via XIOS as 
     1580\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient 
     1581(\cf) is also available as \texttt{wi\_cff}. For a quick oversight of 
     1582the schemes activity the global maximum values of the absolute implicit component 
     1583of the vertical velocity and the partitioning coefficient are written to the netCDF 
     1584version of the run statistics file (\texttt{run.stat.nc}) if this is active (see 
     1585\autoref{sec:MISC_opt} for activation details). 
     1586 
     1587\autoref{fig:ZDF_zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for 
     1588the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is 
     1589active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the 
     1590test case is close to stability limits even with this value. At the larger timesteps, the 
     1591vertical velocity is treated mostly implicitly at some location throughout the run. The 
     1592oscillatory nature of this measure appears to be linked to the progress of the plume front 
     1593as each cusp is associated with the location of the maximum shifting to the adjacent cell. 
     1594This is illustrated in \autoref{fig:ZDF_zad_Aimp_maxCf_loc} where the i- and k- locations of the 
     1595maximum have been overlaid for the base run case. 
     1596 
     1597\medskip 
     1598\noindent Only limited tests have been performed in more realistic configurations. In the 
     1599ORCA2\_ICE\_PISCES reference configuration the scheme does activate and passes 
     1600restartability and reproducibility tests but it is unable to improve the model's stability 
     1601enough to allow an increase in the model time-step. A view of the time-series of maximum 
     1602partitioning coefficient (not shown here)  suggests that the default time-step of 5400s is 
     1603already pushing at stability limits, especially in the initial start-up phase. The 
     1604time-series does not, however, exhibit any of the 'cuspiness' found with the overflow 
     1605tests. 
     1606 
     1607\medskip 
     1608\noindent A short test with an eORCA1 configuration promises more since a test using a 
     1609time-step of 3600s remains stable with \forcode{ln_zad_Aimp=.true.} whereas the 
     1610time-step is limited to 2700s without. 
     1611 
     1612\begin{figure}[!t] 
     1613  \centering 
     1614  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_overflow_all_rdt} 
     1615  \caption[OVERFLOW: sample temperature vertical cross-sections from mid- and end-run]{ 
     1616    Sample temperature vertical cross-sections from mid- and end-run using 
     1617    different values for \forcode{nn_rdt} and with or without adaptive implicit vertical advection. 
     1618    Without the adaptive implicit vertical advection 
     1619    only the run with the shortest timestep is able to run to completion. 
     1620    Note also that the colour-scale has been chosen to confirm that 
     1621    temperatures remain within the original range of 10$^o$ to 20$^o$.} 
     1622  \label{fig:ZDF_zad_Aimp_overflow_all_rdt} 
     1623\end{figure} 
     1624 
     1625\begin{figure}[!t] 
     1626  \centering 
     1627  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf} 
     1628  \caption[OVERFLOW: maximum partitioning coefficient during a series of test runs]{ 
     1629    The maximum partitioning coefficient during a series of test runs with 
     1630    increasing model timestep length. 
     1631    At the larger timesteps, 
     1632    the vertical velocity is treated mostly implicitly at some location throughout the run.} 
     1633  \label{fig:ZDF_zad_Aimp_maxCf} 
     1634\end{figure} 
     1635 
     1636\begin{figure}[!t] 
     1637  \centering 
     1638  \includegraphics[width=0.66\textwidth]{ZDF_zad_Aimp_maxCf_loc} 
     1639  \caption[OVERFLOW: maximum partitioning coefficient for the case overlaid]{ 
     1640    The maximum partitioning coefficient for the \forcode{nn_rdt=10.0} case overlaid with 
     1641    information on the gridcell i- and k-locations of the maximum value.} 
     1642  \label{fig:ZDF_zad_Aimp_maxCf_loc} 
     1643\end{figure} 
     1644 
     1645\subinc{\input{../../global/epilogue}} 
    14041646 
    14051647\end{document} 
Note: See TracChangeset for help on using the changeset viewer.