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 11799 for NEMO/branches/2019/dev_r11470_HPC_12_mpi3/doc/latex/NEMO/subfiles/chap_ZDF.tex – NEMO

Ignore:
Timestamp:
2019-10-25T16:27:34+02:00 (4 years ago)
Author:
mocavero
Message:

Update the branch to v4.0.1 of the trunk

Location:
NEMO/branches/2019/dev_r11470_HPC_12_mpi3/doc
Files:
5 edited

Legend:

Unmodified
Added
Removed
  • NEMO/branches/2019/dev_r11470_HPC_12_mpi3/doc

    • Property svn:externals set to
      ^/utils/badges badges
      ^/utils/logos logos
  • NEMO/branches/2019/dev_r11470_HPC_12_mpi3/doc/latex

    • Property svn:ignore deleted
  • NEMO/branches/2019/dev_r11470_HPC_12_mpi3/doc/latex/NEMO

    • Property svn:externals set to
      ^/utils/figures/NEMO figures
  • NEMO/branches/2019/dev_r11470_HPC_12_mpi3/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_r11470_HPC_12_mpi3/doc/latex/NEMO/subfiles/chap_ZDF.tex

    r11435 r11799  
    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 
     11\thispagestyle{plain} 
     12 
    1013\chaptertoc 
    1114 
     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 
    1230%gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN. 
    1331 
    14 \newpage 
    15  
    16 % ================================================================ 
    17 % Vertical Mixing 
    18 % ================================================================ 
     32%% ================================================================================================= 
    1933\section{Vertical mixing} 
    20 \label{sec:ZDF_zdf} 
     34\label{sec:ZDF} 
    2135 
    2236The discrete form of the ocean subgrid scale physics has been presented in 
     
    2539At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}), 
    2640while 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\ \np{ln\_trabbc} defined, 
     41unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie\ \np{ln_trabbc}{ln\_trabbc} defined, 
    2842see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum 
    2943(see \autoref{sec:ZDF_drg}). 
     
    3953are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively. 
    4054%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 %--------------------------------------------namzdf-------------------------------------------------------- 
    46  
    47 \nlst{namzdf} 
    48 %-------------------------------------------------------------------------------------------------------------- 
    49  
    50 % ------------------------------------------------------------------------------------------------------------- 
    51 %        Constant 
    52 % ------------------------------------------------------------------------------------------------------------- 
    53 \subsection[Constant (\forcode{ln_zdfcst = .true.})] 
    54 {Constant (\protect\np{ln\_zdfcst}\forcode{ = .true.})} 
     55%(namelist parameter \np[=.true.]{ln_zdfexp}{ln\_zdfexp}) or a backward time stepping scheme 
     56%(\np[=.false.]{ln_zdfexp}{ln\_zdfexp}) depending on the magnitude of the mixing coefficients, 
     57%and thus of the formulation used (see \autoref{chap:TD}). 
     58 
     59\begin{listing} 
     60  \nlst{namzdf} 
     61  \caption{\forcode{&namzdf}} 
     62  \label{lst:namzdf} 
     63\end{listing} 
     64 
     65%% ================================================================================================= 
     66\subsection[Constant (\forcode{ln_zdfcst})]{Constant (\protect\np{ln_zdfcst}{ln\_zdfcst})} 
    5567\label{subsec:ZDF_cst} 
    5668 
    57 Options are defined through the \nam{zdf} namelist variables. 
    58 When \np{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
     69Options are defined through the \nam{zdf}{zdf} namelist variables. 
     70When \np{ln_zdfcst}{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
    5971constant values over the whole ocean. 
    6072This is the crudest way to define the vertical ocean physics. 
     
    6678\end{align*} 
    6779 
    68 These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters. 
     80These values are set through the \np{rn_avm0}{rn\_avm0} and \np{rn_avt0}{rn\_avt0} namelist parameters. 
    6981In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity, 
    7082that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and 
    7183$\sim10^{-9}~m^2.s^{-1}$ for salinity. 
    7284 
    73 % ------------------------------------------------------------------------------------------------------------- 
    74 %        Richardson Number Dependent 
    75 % ------------------------------------------------------------------------------------------------------------- 
    76 \subsection[Richardson number dependent (\forcode{ln_zdfric = .true.})] 
    77 {Richardson number dependent (\protect\np{ln\_zdfric}\forcode{ = .true.})} 
     85%% ================================================================================================= 
     86\subsection[Richardson number dependent (\forcode{ln_zdfric})]{Richardson number dependent (\protect\np{ln_zdfric}{ln\_zdfric})} 
    7887\label{subsec:ZDF_ric} 
    7988 
    80 %--------------------------------------------namric--------------------------------------------------------- 
    81  
    82 \nlst{namzdf_ric} 
    83 %-------------------------------------------------------------------------------------------------------------- 
    84  
    85 When \np{ln\_zdfric}\forcode{ = .true.}, a local Richardson number dependent formulation for the vertical momentum and 
    86 tracer eddy coefficients is set through the \nam{zdf\_ric} namelist variables. 
     89\begin{listing} 
     90  \nlst{namzdf_ric} 
     91  \caption{\forcode{&namzdf_ric}} 
     92  \label{lst:namzdf_ric} 
     93\end{listing} 
     94 
     95When \np[=.true.]{ln_zdfric}{ln\_zdfric}, a local Richardson number dependent formulation for the vertical momentum and 
     96tracer eddy coefficients is set through the \nam{zdf_ric}{zdf\_ric} namelist variables. 
    8797The vertical mixing coefficients are diagnosed from the large scale variables computed by the model. 
    8898\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures. 
     
    92102Following \citet{pacanowski.philander_JPO81}, the following formulation has been implemented: 
    93103\[ 
    94   % \label{eq:zdfric} 
     104  % \label{eq:ZDF_ric} 
    95105  \left\{ 
    96106    \begin{aligned} 
     
    105115(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that 
    106116can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$. 
    107 The last three values can be modified by setting the \np{rn\_avmri}, \np{rn\_alp} and 
    108 \np{nn\_ric} namelist parameters, respectively. 
     117The last three values can be modified by setting the \np{rn_avmri}{rn\_avmri}, \np{rn_alp}{rn\_alp} and 
     118\np{nn_ric}{nn\_ric} namelist parameters, respectively. 
    109119 
    110120A simple mixing-layer model to transfer and dissipate the atmospheric forcings 
    111 (wind-stress and buoyancy fluxes) can be activated setting the \np{ln\_mldw}\forcode{ = .true.} in the namelist. 
     121(wind-stress and buoyancy fluxes) can be activated setting the \np[=.true.]{ln_mldw}{ln\_mldw} in the namelist. 
    112122 
    113123In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and 
     
    125135\] 
    126136is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$. 
    127 The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}. 
     137The final $h_{e}$ is further constrained by the adjustable bounds \np{rn_mldmin}{rn\_mldmin} and \np{rn_mldmax}{rn\_mldmax}. 
    128138Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to 
    129 the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{lermusiaux_JMS01}. 
    130  
    131 % ------------------------------------------------------------------------------------------------------------- 
    132 %        TKE Turbulent Closure Scheme 
    133 % ------------------------------------------------------------------------------------------------------------- 
    134 \subsection[TKE turbulent closure scheme (\forcode{ln_zdftke = .true.})] 
    135 {TKE turbulent closure scheme (\protect\np{ln\_zdftke}\forcode{ = .true.})} 
     139the empirical values \np{rn_wtmix}{rn\_wtmix} and \np{rn_wvmix}{rn\_wvmix} \citep{lermusiaux_JMS01}. 
     140 
     141%% ================================================================================================= 
     142\subsection[TKE turbulent closure scheme (\forcode{ln_zdftke})]{TKE turbulent closure scheme (\protect\np{ln_zdftke}{ln\_zdftke})} 
    136143\label{subsec:ZDF_tke} 
    137 %--------------------------------------------namzdf_tke-------------------------------------------------- 
    138  
    139 \nlst{namzdf_tke} 
    140 %-------------------------------------------------------------------------------------------------------------- 
     144 
     145\begin{listing} 
     146  \nlst{namzdf_tke} 
     147  \caption{\forcode{&namzdf_tke}} 
     148  \label{lst:namzdf_tke} 
     149\end{listing} 
    141150 
    142151The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on 
     
    151160its destruction through stratification, its vertical diffusion, and its dissipation of \citet{kolmogorov_IANS42} type: 
    152161\begin{equation} 
    153   \label{eq:zdftke_e} 
     162  \label{eq:ZDF_tke_e} 
    154163  \frac{\partial \bar{e}}{\partial t} = 
    155164  \frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2 
     
    161170\end{equation} 
    162171\[ 
    163   % \label{eq:zdftke_kz} 
     172  % \label{eq:ZDF_tke_kz} 
    164173  \begin{split} 
    165174    K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }    \\ 
     
    172181The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with 
    173182vertical mixing at any depth \citep{gaspar.gregoris.ea_JGR90}. 
    174 They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}. 
     183They are set through namelist parameters \np{nn_ediff}{nn\_ediff} and \np{nn_ediss}{nn\_ediss}. 
    175184$P_{rt}$ can be set to unity or, following \citet{blanke.delecluse_JPO93}, be a function of the local Richardson number, $R_i$: 
    176185\begin{align*} 
    177   % \label{eq:prt} 
     186  % \label{eq:ZDF_prt} 
    178187  P_{rt} = 
    179188  \begin{cases} 
     
    183192  \end{cases} 
    184193\end{align*} 
    185 The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable. 
     194The choice of $P_{rt}$ is controlled by the \np{nn_pdl}{nn\_pdl} namelist variable. 
    186195 
    187196At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as 
    188 $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} namelist parameter. 
     197$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter. 
    189198The default value of $e_{bb}$ is 3.75. \citep{gaspar.gregoris.ea_JGR90}), however a much larger value can be used when 
    190199taking into account the surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}). 
     
    192201The time integration of the $\bar{e}$ equation may formally lead to negative values because 
    193202the numerical scheme does not ensure its positivity. 
    194 To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} namelist parameter). 
     203To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn_emin}{rn\_emin} namelist parameter). 
    195204Following \citet{gaspar.gregoris.ea_JGR90}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
    196205This allows the subsequent formulations to match that of \citet{gargett_JMR84} for the diffusion in 
     
    198207In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with 
    199208too weak vertical diffusion. 
    200 They must be specified at least larger than the molecular values, and are set through \np{rn\_avm0} and 
    201 \np{rn\_avt0} (\nam{zdf} namelist, see \autoref{subsec:ZDF_cst}). 
    202  
     209They must be specified at least larger than the molecular values, and are set through \np{rn_avm0}{rn\_avm0} and 
     210\np{rn_avt0}{rn\_avt0} (\nam{zdf}{zdf} namelist, see \autoref{subsec:ZDF_cst}). 
     211 
     212%% ================================================================================================= 
    203213\subsubsection{Turbulent length scale} 
    204214 
    205215For computational efficiency, the original formulation of the turbulent length scales proposed by 
    206216\citet{gaspar.gregoris.ea_JGR90} has been simplified. 
    207 Four formulations are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist parameter. 
     217Four formulations are proposed, the choice of which is controlled by the \np{nn_mxl}{nn\_mxl} namelist parameter. 
    208218The first two are based on the following first order approximation \citep{blanke.delecluse_JPO93}: 
    209219\begin{equation} 
    210   \label{eq:tke_mxl0_1} 
     220  \label{eq:ZDF_tke_mxl0_1} 
    211221  l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N 
    212222\end{equation} 
    213223which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency. 
    214224The resulting length scale is bounded by the distance to the surface or to the bottom 
    215 (\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor (\np{nn\_mxl}\forcode{ = 1}). 
     225(\np[=0]{nn_mxl}{nn\_mxl}) or by the local vertical scale factor (\np[=1]{nn_mxl}{nn\_mxl}). 
    216226\citet{blanke.delecluse_JPO93} notice that this simplification has two major drawbacks: 
    217227it makes no sense for locally unstable stratification and the computation no longer uses all 
    218228the information contained in the vertical density profile. 
    219 To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np{nn\_mxl}\forcode{ = 2, 3} cases, 
     229To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np[=2, 3]{nn_mxl}{nn\_mxl} cases, 
    220230which add an extra assumption concerning the vertical gradient of the computed length scale. 
    221 So, the length scales are first evaluated as in \autoref{eq:tke_mxl0_1} and then bounded such that: 
     231So, the length scales are first evaluated as in \autoref{eq:ZDF_tke_mxl0_1} and then bounded such that: 
    222232\begin{equation} 
    223   \label{eq:tke_mxl_constraint} 
     233  \label{eq:ZDF_tke_mxl_constraint} 
    224234  \frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1 
    225235  \qquad \text{with }\  l =  l_k = l_\epsilon 
    226236\end{equation} 
    227 \autoref{eq:tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
     237\autoref{eq:ZDF_tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
    228238the variations of depth. 
    229239It provides a better approximation of the \citet{gaspar.gregoris.ea_JGR90} formulation while being much less 
     
    231241In particular, it allows the length scale to be limited not only by the distance to the surface or 
    232242to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as 
    233 the thermocline (\autoref{fig:mixing_length}). 
    234 In order to impose the \autoref{eq:tke_mxl_constraint} constraint, we introduce two additional length scales: 
     243the thermocline (\autoref{fig:ZDF_mixing_length}). 
     244In order to impose the \autoref{eq:ZDF_tke_mxl_constraint} constraint, we introduce two additional length scales: 
    235245$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and 
    236246evaluate the dissipation and mixing length scales as 
    237247(and note that here we use numerical indexing): 
    238 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    239248\begin{figure}[!t] 
    240   \begin{center} 
    241     \includegraphics[width=\textwidth]{Fig_mixing_length} 
    242     \caption{ 
    243       \protect\label{fig:mixing_length} 
    244       Illustration of the mixing length computation. 
    245     } 
    246   \end{center} 
     249  \centering 
     250  \includegraphics[width=0.66\textwidth]{Fig_mixing_length} 
     251  \caption[Mixing length computation]{Illustration of the mixing length computation} 
     252  \label{fig:ZDF_mixing_length} 
    247253\end{figure} 
    248 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    249 \[ 
    250   % \label{eq:tke_mxl2} 
     254\[ 
     255  % \label{eq:ZDF_tke_mxl2} 
    251256  \begin{aligned} 
    252257    l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right) 
     
    256261  \end{aligned} 
    257262\] 
    258 where $l^{(k)}$ is computed using \autoref{eq:tke_mxl0_1}, \ie\ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
    259  
    260 In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same value: 
    261 $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nn\_mxl}\forcode{ = 3} case, 
     263where $l^{(k)}$ is computed using \autoref{eq:ZDF_tke_mxl0_1}, \ie\ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
     264 
     265In the \np[=2]{nn_mxl}{nn\_mxl} case, the dissipation and mixing length scales take the same value: 
     266$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np[=3]{nn_mxl}{nn\_mxl} case, 
    262267the dissipation and mixing turbulent length scales are give as in \citet{gaspar.gregoris.ea_JGR90}: 
    263268\[ 
    264   % \label{eq:tke_mxl_gaspar} 
     269  % \label{eq:ZDF_tke_mxl_gaspar} 
    265270  \begin{aligned} 
    266271    & l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }   \\ 
     
    269274\] 
    270275 
    271 At the ocean surface, a non zero length scale is set through the  \np{rn\_mxl0} namelist parameter. 
     276At the ocean surface, a non zero length scale is set through the  \np{rn_mxl0}{rn\_mxl0} namelist parameter. 
    272277Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and 
    273278$z_o$ the roughness parameter of the surface. 
    274 Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}. 
     279Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn_mxl0}{rn\_mxl0}. 
    275280In the ocean interior a minimum length scale is set to recover the molecular viscosity when 
    276281$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ). 
    277282 
     283%% ================================================================================================= 
    278284\subsubsection{Surface wave breaking parameterization} 
    279 %-----------------------------------------------------------------------% 
    280285 
    281286Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to 
     
    303308$\alpha_{CB} = 100$ the Craig and Banner's value. 
    304309As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$, 
    305 with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}\forcode{ = 67.83} corresponds 
     310with $e_{bb}$ the \np{rn_ebb}{rn\_ebb} namelist parameter, setting \np[=67.83]{rn_ebb}{rn\_ebb} corresponds 
    306311to $\alpha_{CB} = 100$. 
    307 Further setting  \np{ln\_mxl0}\forcode{ =.true.},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale, 
     312Further setting  \np[=.true.]{ln_mxl0}{ln\_mxl0},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale, 
    308313with $\beta$ hard coded to the Stacey's value. 
    309 Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the 
     314Note that a minimal threshold of \np{rn_emin0}{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the 
    310315surface $\bar{e}$ value. 
    311316 
     317%% ================================================================================================= 
    312318\subsubsection{Langmuir cells} 
    313 %--------------------------------------% 
    314319 
    315320Langmuir circulations (LC) can be described as ordered large-scale vertical motions in 
     
    325330The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in 
    326331an extra source term of TKE, $P_{LC}$. 
    327 The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled by setting \np{ln\_lc} to 
    328 \forcode{.true.} in the \nam{zdf\_tke} namelist. 
     332The presence of $P_{LC}$ in \autoref{eq:ZDF_tke_e}, the TKE equation, is controlled by setting \np{ln_lc}{ln\_lc} to 
     333\forcode{.true.} in the \nam{zdf_tke}{zdf\_tke} namelist. 
    329334 
    330335By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}), 
     
    354359where $c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data. 
    355360The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second. 
    356 The value of $c_{LC}$ is set through the \np{rn\_lc} namelist parameter, 
     361The value of $c_{LC}$ is set through the \np{rn_lc}{rn\_lc} namelist parameter, 
    357362having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}. 
    358363 
     
    364369\] 
    365370 
     371%% ================================================================================================= 
    366372\subsubsection{Mixing just below the mixed layer} 
    367 %--------------------------------------------------------------% 
    368373 
    369374Vertical mixing parameterizations commonly used in ocean general circulation models tend to 
     
    376381(\ie\ near-inertial oscillations and ocean swells and waves). 
    377382 
    378 When using this parameterization (\ie\ when \np{nn\_etau}\forcode{ = 1}), 
     383When using this parameterization (\ie\ when \np[=1]{nn_etau}{nn\_etau}), 
    379384the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations, 
    380385swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition, 
     
    388393the penetration, and $f_i$ is the ice concentration 
    389394(no penetration if $f_i=1$, \ie\ if the ocean is entirely covered by sea-ice). 
    390 The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter. 
    391 The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0}) or 
     395The value of $f_r$, usually a few percents, is specified through \np{rn_efr}{rn\_efr} namelist parameter. 
     396The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np[=0]{nn_etau}{nn\_etau}) or 
    392397a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes 
    393 (\np{nn\_etau}\forcode{ = 1}). 
    394  
    395 Note that two other option exist, \np{nn\_etau}\forcode{ = 2, 3}. 
     398(\np[=1]{nn_etau}{nn\_etau}). 
     399 
     400Note that two other option exist, \np[=2, 3]{nn_etau}{nn\_etau}. 
    396401They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer, 
    397402or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean. 
     
    400405 
    401406% This should be explain better below what this rn_eice parameter is meant for: 
    402 In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn\_eice} namelist parameter. 
     407In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn_eice}{rn\_eice} namelist parameter. 
    403408This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration 
    404409is greater than 25\%. 
     
    412417% (\eg\ Mellor, 1989; Large et al., 1994; Meier, 2001; Axell, 2002; St. Laurent and Garrett, 2002). 
    413418 
    414 % ------------------------------------------------------------------------------------------------------------- 
    415 %        GLS Generic Length Scale Scheme 
    416 % ------------------------------------------------------------------------------------------------------------- 
    417 \subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls = .true.})] 
    418 {GLS: Generic Length Scale (\protect\np{ln\_zdfgls}\forcode{ = .true.})} 
     419%% ================================================================================================= 
     420\subsection[GLS: Generic Length Scale (\forcode{ln_zdfgls})]{GLS: Generic Length Scale (\protect\np{ln_zdfgls}{ln\_zdfgls})} 
    419421\label{subsec:ZDF_gls} 
    420422 
    421 %--------------------------------------------namzdf_gls--------------------------------------------------------- 
    422  
    423 \nlst{namzdf_gls} 
    424 %-------------------------------------------------------------------------------------------------------------- 
     423\begin{listing} 
     424  \nlst{namzdf_gls} 
     425  \caption{\forcode{&namzdf_gls}} 
     426  \label{lst:namzdf_gls} 
     427\end{listing} 
    425428 
    426429The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
     
    428431$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}. 
    429432This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$, 
    430 where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of 
     433where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:ZDF_GLS} allows to recover a number of 
    431434well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87}, 
    432435$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}). 
    433436The GLS scheme is given by the following set of equations: 
    434437\begin{equation} 
    435   \label{eq:zdfgls_e} 
     438  \label{eq:ZDF_gls_e} 
    436439  \frac{\partial \bar{e}}{\partial t} = 
    437440  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     
    443446 
    444447\[ 
    445   % \label{eq:zdfgls_psi} 
     448  % \label{eq:ZDF_gls_psi} 
    446449  \begin{split} 
    447450    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
     
    455458 
    456459\[ 
    457   % \label{eq:zdfgls_kz} 
     460  % \label{eq:ZDF_gls_kz} 
    458461  \begin{split} 
    459462    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
     
    463466 
    464467\[ 
    465   % \label{eq:zdfgls_eps} 
     468  % \label{eq:ZDF_gls_eps} 
    466469  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
    467470\] 
     
    470473The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
    471474the choice of the turbulence model. 
    472 Four different turbulent models are pre-defined (\autoref{tab:GLS}). 
    473 They are made available through the \np{nn\_clo} namelist parameter. 
    474  
    475 %--------------------------------------------------TABLE-------------------------------------------------- 
     475Four different turbulent models are pre-defined (\autoref{tab:ZDF_GLS}). 
     476They are made available through the \np{nn_clo}{nn\_clo} namelist parameter. 
     477 
    476478\begin{table}[htbp] 
    477   \begin{center} 
    478     % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
    479     \begin{tabular}{ccccc} 
    480       &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
    481       % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\ 
    482       \hline 
    483       \hline 
    484       \np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
    485       \hline 
    486       $( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
    487       $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
    488       $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
    489       $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
    490       $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
    491       $C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
    492       $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
    493       \hline 
    494       \hline 
    495     \end{tabular} 
    496     \caption{ 
    497       \protect\label{tab:GLS} 
    498       Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
    499       \protect\np{ln\_zdfgls}\forcode{ = .true.} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\nam{zdf\_gls}. 
    500     } 
    501   \end{center} 
     479  \centering 
     480  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
     481  \begin{tabular}{ccccc} 
     482    &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
     483    % & \citep{mellor.yamada_RG82} &  \citep{rodi_JGR87}       & \citep{wilcox_AJ88} &                 \\ 
     484    \hline 
     485    \hline 
     486    \np{nn_clo}{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
     487    \hline 
     488    $( p , n , m )$         &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
     489    $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
     490    $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
     491    $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
     492    $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
     493    $C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
     494    $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
     495    \hline 
     496    \hline 
     497  \end{tabular} 
     498  \caption[Set of predefined GLS parameters or equivalently predefined turbulence models available]{ 
     499    Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
     500    \protect\np[=.true.]{ln_zdfgls}{ln\_zdfgls} and controlled by 
     501    the \protect\np{nn_clos}{nn\_clos} namelist variable in \protect\nam{zdf_gls}{zdf\_gls}.} 
     502  \label{tab:ZDF_GLS} 
    502503\end{table} 
    503 %-------------------------------------------------------------------------------------------------------------- 
    504504 
    505505In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
     
    508508$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88}, 
    509509or by \citet{kantha.clayson_JGR94} or one of the two functions suggested by \citet{canuto.howard.ea_JPO01} 
    510 (\np{nn\_stab\_func}\forcode{ = 0, 3}, resp.). 
     510(\np[=0, 3]{nn_stab_func}{nn\_stab\_func}, resp.). 
    511511The value of $C_{0\mu}$ depends on the choice of the stability function. 
    512512 
    513513The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
    514 Neumann condition through \np{nn\_bc\_surf} and \np{nn\_bc\_bot}, resp. 
     514Neumann condition through \np{nn_bc_surf}{nn\_bc\_surf} and \np{nn_bc_bot}{nn\_bc\_bot}, resp. 
    515515As for TKE closure, the wave effect on the mixing is considered when 
    516 \np{rn\_crban}\forcode{ > 0.} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}. 
    517 The \np{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
    518 \np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}. 
     516\np[ > 0.]{rn_crban}{rn\_crban} \citep{craig.banner_JPO94, mellor.blumberg_JPO04}. 
     517The \np{rn_crban}{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
     518\np{rn_charn}{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}. 
    519519 
    520520The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
     
    525525the entrainment depth predicted in stably stratified situations, 
    526526and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
    527 The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.}, 
    528 and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
     527The clipping is only activated if \np[=.true.]{ln_length_lim}{ln\_length\_lim}, 
     528and the $c_{lim}$ is set to the \np{rn_clim_galp}{rn\_clim\_galp} value. 
    529529 
    530530The time and space discretization of the GLS equations follows the same energetic consideration as for 
     
    533533 in \citet{reffray.guillaume.ea_GMD15} for the \NEMO\ model. 
    534534 
    535  
    536535% ------------------------------------------------------------------------------------------------------------- 
    537 %        OSM OSMOSIS BL Scheme 
     536%        OSM OSMOSIS BL Scheme  
    538537% ------------------------------------------------------------------------------------------------------------- 
    539 \subsection[OSM: OSMosis boundary layer scheme (\forcode{ln_zdfosm = .true.})] 
    540 {OSM: OSMosis boundary layer scheme (\protect\np{ln\_zdfosm}\forcode{ = .true.})} 
     538\subsection[OSM: OSMOSIS boundary layer scheme (\forcode{ln_zdfosm = .true.})] 
     539{OSM: OSMOSIS boundary layer scheme (\protect\np{ln_zdfosm}{ln\_zdfosm})} 
    541540\label{subsec:ZDF_osm} 
    542 %--------------------------------------------namzdf_osm--------------------------------------------------------- 
    543  
    544 \nlst{namzdf_osm} 
     541 
     542\begin{listing} 
     543  \nlst{namzdf_osm} 
     544  \caption{\forcode{&namzdf_osm}} 
     545  \label{lst:namzdf_osm} 
     546\end{listing} 
     547 
    545548%-------------------------------------------------------------------------------------------------------------- 
    546  
    547 The OSMOSIS turbulent closure scheme is based on......   TBC 
    548  
    549 % ------------------------------------------------------------------------------------------------------------- 
    550 %        TKE and GLS discretization considerations 
    551 % ------------------------------------------------------------------------------------------------------------- 
    552 \subsection[ Discrete energy conservation for TKE and GLS schemes] 
    553 {Discrete energy conservation for TKE and GLS schemes} 
    554 \label{subsec:ZDF_tke_ene} 
    555  
    556 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     549\paragraph{Namelist choices} 
     550Most of the namelist options refer to how to specify the Stokes 
     551surface drift and penetration depth. There are three options: 
     552\begin{description} 
     553  \item \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave} Default value in \texttt{namelist\_ref}. In this case the Stokes drift is 
     554      assumed to be parallel to the surface wind stress, with 
     555      magnitude consistent with a constant turbulent Langmuir number 
     556    $\mathrm{La}_t=$ \protect\np{rn_m_la} {rn\_m\_la} i.e.\ 
     557    $u_{s0}=\tau/(\mathrm{La}_t^2\rho_0)$.  Default value of 
     558    \protect\np{rn_m_la}{rn\_m\_la} is 0.3. The Stokes penetration 
     559      depth $\delta = $ \protect\np{rn_osm_dstokes}{rn\_osm\_dstokes}; this has default value 
     560      of 5~m. 
     561  
     562  \item \protect\np[=1]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is 
     563      assumed to be parallel to the surface wind stress, with 
     564      magnitude as in the classical Pierson-Moskowitz wind-sea 
     565      spectrum.  Significant wave height and 
     566      wave-mean period taken from this spectrum are used to calculate the Stokes penetration 
     567      depth, following the approach set out in  \citet{breivik.janssen.ea_JPO14}. 
     568  
     569    \item \protect\np[=2]{nn_osm_wave}{nn\_osm\_wave} In this case the Stokes drift is 
     570      taken from  ECMWF wave model output, though only the component parallel 
     571      to the wind stress is retained. Significant wave height and 
     572      wave-mean period from ECMWF wave model output are used to calculate the Stokes penetration 
     573      depth, again following \citet{breivik.janssen.ea_JPO14}. 
     574 
     575    \end{description} 
     576 
     577    Others refer to the treatment of diffusion and viscosity beneath 
     578    the surface boundary layer: 
     579\begin{description} 
     580   \item \protect\np{ln_kpprimix} {ln\_kpprimix}  Default is \np{.true.}. Switches on KPP-style Ri \#-dependent 
     581     mixing below the surface boundary layer. If this is set 
     582     \texttt{.true.}  the following variable settings are honoured: 
     583    \item \protect\np{rn_riinfty}{rn\_riinfty} Critical value of local Ri \# below which 
     584      shear instability increases vertical mixing from background value. 
     585    \item \protect\np{rn_difri} {rn\_difri} Maximum value of Ri \#-dependent mixing at $\mathrm{Ri}=0$. 
     586    \item \protect\np{ln_convmix}{ln\_convmix} If \texttt{.true.} then, where water column is unstable, specify 
     587       diffusivity equal to \protect\np{rn_dif_conv}{rn\_dif\_conv} (default value is 1 m~s$^{-2}$).  
     588 \end{description} 
     589 Diagnostic output is controlled by: 
     590  \begin{description} 
     591    \item \protect\np{ln_dia_osm}{ln\_dia\_osm} Default is \np{.false.}; allows XIOS output of OSMOSIS internal fields. 
     592  \end{description} 
     593Obsolete namelist parameters include: 
     594\begin{description} 
     595   \item \protect\np{ln_use_osm_la}\np{ln\_use\_osm\_la} With \protect\np[=0]{nn_osm_wave}{nn\_osm\_wave}, 
     596      \protect\np{rn_osm_dstokes} {rn\_osm\_dstokes} is always used to specify the Stokes 
     597      penetration depth. 
     598   \item \protect\np{nn_ave} {nn\_ave} Choice of averaging method for KPP-style Ri \# 
     599      mixing. Not taken account of. 
     600   \item \protect\np{rn_osm_hbl0} {rn\_osm\_hbl0} Depth of initial boundary layer is now set 
     601     by a density criterion similar to that used in calculating \emph{hmlp} (output as \texttt{mldr10\_1}) in \mdl{zdfmxl}. 
     602\end{description} 
     603 
     604\subsubsection{Summary} 
     605Much of the time the turbulent motions in the ocean surface boundary 
     606layer (OSBL) are not given by 
     607classical shear turbulence. Instead they are in a regime known as 
     608`Langmuir turbulence',  dominated by an 
     609interaction between the currents and the Stokes drift of the surface waves \citep[e.g.][]{mcwilliams.ea_JFM97}. 
     610This 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 $\mathrm{La}_{t}=(u_{\ast}/u_{s0}) > 0.4$. 
     611 
     612The OSMOSIS model is fundamentally based on results of Large Eddy 
     613Simulations (LES) of Langmuir turbulence and aims to fully describe 
     614this Langmuir regime. The description in this section is of necessity incomplete and further details are available in Grant. A (2019); in prep. 
     615 
     616The OSMOSIS turbulent closure scheme is a similarity-scale scheme in 
     617the same spirit as the K-profile 
     618parameterization (KPP) scheme of \citet{large.ea_RG97}. 
     619A specified shape of diffusivity, scaled by the (OSBL) depth 
     620$h_{\mathrm{BL}}$ and a turbulent velocity scale, is imposed throughout the  
     621boundary layer 
     622$-h_{\mathrm{BL}}<z<\eta$. The turbulent closure model 
     623also includes fluxes of tracers and momentum that are``non-local'' (independent of the local property gradient). 
     624 
     625Rather than the OSBL 
     626depth being diagnosed in terms of a bulk Richardson number criterion, 
     627as in KPP, it is set by a prognostic equation that is informed by 
     628energy budget considerations reminiscent of the classical mixed layer 
     629models of \citet{kraus.turner_tellus67}.  
     630The model also includes an explicit parametrization of the structure 
     631of the pycnocline (the stratified region at the bottom of the OSBL). 
     632 
     633Presently, mixing below the OSBL is handled by the Richardson 
     634number-dependent mixing scheme used in \citet{large.ea_RG97}. 
     635 
     636Convective parameterizations such as described in \ref{sec:ZDF_conv} 
     637below should not be used with the OSMOSIS-OBL model: instabilities 
     638within the OSBL are part of the model, while instabilities below the 
     639ML are handled by the Ri \# dependent scheme. 
     640 
     641\subsubsection{Depth and velocity scales} 
     642The 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). 
    557643\begin{figure}[!t] 
    558644  \begin{center} 
    559     \includegraphics[width=\textwidth]{Fig_ZDF_TKE_time_scheme} 
     645    \includegraphics[width=0.7\textwidth]{Fig_ZDF_OSM_structure_of_OSBL} 
    560646    \caption{ 
    561       \protect\label{fig:TKE_time_scheme} 
    562       Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and its links to the momentum and tracer time integration. 
     647      \protect\label{fig: OSBL_structure} 
     648     The structure of the entraining  boundary layer. (a) Mean buoyancy profile. (b) Profile of the buoyancy flux. 
    563649    } 
    564650  \end{center} 
    565651\end{figure} 
    566 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     652The 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. 
     653 
     654Consideration of the power input by wind acting on the Stokes drift suggests that the Langmuir turbulence has velocity scale: 
     655\begin{equation}\label{eq:w_La} 
     656w_{*L}= \left(u_*^2 u_{s\,0}\right)^{1/3}; 
     657\end{equation}  
     658but 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: 
     659\begin{equation}\label{eq:composite-nu} 
     660  \nu_{\ast}= \left\{ u_*^3 \left[1-\exp(-.5 \mathrm{La}_t^2)\right]+w_{*L}^3\right\}^{1/3}. 
     661\end{equation} 
     662For 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: 
     663\begin{equation}\label{eq:vel-scale-unstable} 
     664\omega_* = \left(\nu_*^3 + 0.5 w_{*C}^3\right)^{1/3}. 
     665\end{equation} 
     666 
     667\subsubsection{The flux gradient model} 
     668The flux-gradient relationships used in the OSMOSIS scheme take the form: 
     669% 
     670\begin{equation}\label{eq:flux-grad-gen} 
     671\overline{w^\prime\chi^\prime}=-K\frac{\partial\overline{\chi}}{\partial z} + N_{\chi,s} +N_{\chi,b} +N_{\chi,t}, 
     672\end{equation} 
     673% 
     674where $\chi$ is a general variable and $N_{\chi,s}, N_{\chi,b} \mathrm{and} N_{\chi,t}$  are the non-gradient terms, and represent the effects of the different terms in the turbulent flux-budget on the transport of $\chi$. $N_{\chi,s}$ represents the effects that the Stokes shear has on the transport of $\chi$, $N_{\chi,b}$  the effect of buoyancy, and $N_{\chi,t}$ the effect of the turbulent transport.  The same general form for the flux-gradient relationship is used to parametrize the transports of momentum, heat and salinity. 
     675 
     676In terms of the non-dimensionalized depth variables 
     677% 
     678\begin{equation}\label{eq:sigma} 
     679\sigma_{\mathrm{ml}}= -z/h_{\mathrm{ml}}; \;\sigma_{\mathrm{bl}}= -z/h_{\mathrm{bl}}, 
     680\end{equation} 
     681% 
     682in unstable conditions the eddy diffusivity ($K_d$) and eddy viscosity ($K_\nu$) profiles are parametrized as: 
     683% 
     684\begin{align}\label{eq:diff-unstable} 
     685K_d=&0.8\, \omega_*\, h_{\mathrm{ml}} \, \sigma_{\mathrm{ml}} \left(1-\beta_d \sigma_{\mathrm{ml}}\right)^{3/2} 
     686\\\label{eq:visc-unstable} 
     687K_\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) 
     688\end{align} 
     689% 
     690where $\beta_d$ 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 
     691% 
     692\begin{equation}\label{eq:diff-wml-base}  
     693K_{d,\mathrm{ml}}=K_{\nu,\mathrm{ml}}=\,0.16\,\omega_* \Delta h. 
     694\end{equation} 
     695% 
     696For stable conditions the eddy diffusivity/viscosity profiles are given by: 
     697% 
     698\begin{align}\label{diff-stable} 
     699K_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} 
     700K_\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). 
     701\end{align} 
     702% 
     703The 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: 
     704\begin{equation}\label{eq:L_L}  
     705  L_L=-w_{*L}^3/\left<\overline{w^\prime b^\prime}\right>_L, 
     706\end{equation} 
     707with 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 
     708\begin{equation} \label{eq:stable-av-buoy-flux} 
     709\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]. 
     710\end{equation} 
     711% 
     712In unstable conditions the eddy diffusivity and viscosity depend on stability through the velocity scale $\omega_*$, which depends on the two velocity scales $\nu_*$ and $w_{*C}$. 
     713 
     714Details 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). 
     715 
     716\subsubsection{Evolution of the boundary layer depth} 
     717 
     718The prognostic equation for the depth of the neutral/unstable boundary layer is given by \citep{grant+etal18}, 
     719 
     720\begin{equation} \label{eq:dhdt-unstable} 
     721%\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}} 
     722\frac{\partial h_\mathrm{bl}}{\partial t} = W_b - \frac{{\overline{w^\prime b^\prime}}_\mathrm{ent}}{\Delta B_\mathrm{bl}} 
     723\end{equation} 
     724where $h_\mathrm{bl}$ is the horizontally-varying depth of the OSBL, 
     725$\mathbf{U}_b$ and $W_b$ are the mean horizontal and vertical 
     726velocities at the base of the OSBL, ${\overline{w^\prime 
     727    b^\prime}}_\mathrm{ent}$ is the buoyancy flux due to entrainment 
     728and $\Delta B_\mathrm{bl}$ is the difference between the buoyancy 
     729averaged over the depth of the OSBL (i.e.\ including the ML and 
     730pycnocline) and the buoyancy just below the base of the OSBL. This 
     731equation for the case when the pycnocline has a finite thickness, 
     732based on the potential energy budget of the OSBL, is the leading term 
     733\citep{grant+etal18} of a generalization of that used in mixed-layer 
     734models e.g.\ \citet{kraus.turner_tellus67}, in which the thickness of the pycnocline is taken to be zero. 
     735 
     736The entrainment flux for the combination of convective and Langmuir turbulence is given by 
     737\begin{equation} \label{eq:entrain-flux} 
     738  {\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}}} 
     739  + 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] 
     740\end{equation} 
     741where 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 $F(x)\equiv\tanh(x^{-1})^{0.69}$. 
     742 
     743For the stable boundary layer, the equation for the depth of the OSBL is: 
     744 
     745\begin{equation}\label{eq:dhdt-stable} 
     746\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. 
     747\end{equation}  
     748 
     749Equation. \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. 
     750 
     751 
     752%% ================================================================================================= 
     753\subsection[ Discrete energy conservation for TKE and GLS schemes]{Discrete energy conservation for TKE and GLS schemes} 
     754\label{subsec:ZDF_tke_ene} 
     755 
     756\begin{figure}[!t] 
     757  \centering 
     758  \includegraphics[width=0.66\textwidth]{Fig_ZDF_TKE_time_scheme} 
     759  \caption[Subgrid kinetic energy integration in GLS and TKE schemes]{ 
     760    Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and 
     761    its links to the momentum and tracer time integration.} 
     762  \label{fig:ZDF_TKE_time_scheme} 
     763\end{figure} 
    567764 
    568765The production of turbulence by vertical shear (the first term of the right hand side of 
    569 \autoref{eq:zdftke_e}) and  \autoref{eq:zdfgls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
    570 (first line in \autoref{eq:PE_zdf}). 
     766\autoref{eq:ZDF_tke_e}) and  \autoref{eq:ZDF_gls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
     767(first line in \autoref{eq:MB_zdf}). 
    571768To do so a special care has to be taken for both the time and space discretization of 
    572769the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}. 
    573770 
    574 Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows how 
     771Let us first address the time stepping issue. \autoref{fig:ZDF_TKE_time_scheme} shows how 
    575772the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with 
    576773the one-level forward time stepping of the equation for $\bar{e}$. 
     
    579776summing the result vertically: 
    580777\begin{equation} 
    581   \label{eq:energ1} 
     778  \label{eq:ZDF_energ1} 
    582779  \begin{split} 
    583780    \int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\ 
     
    587784\end{equation} 
    588785Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$, 
    589 known at time $t$ (\autoref{fig:TKE_time_scheme}), as it is required when using the TKE scheme 
    590 (see \autoref{sec:STP_forward_imp}). 
    591 The first term of the right hand side of \autoref{eq:energ1} represents the kinetic energy transfer at 
     786known at time $t$ (\autoref{fig:ZDF_TKE_time_scheme}), as it is required when using the TKE scheme 
     787(see \autoref{sec:TD_forward_imp}). 
     788The first term of the right hand side of \autoref{eq:ZDF_energ1} represents the kinetic energy transfer at 
    592789the surface (atmospheric forcing) and at the bottom (friction effect). 
    593790The second term is always negative. 
    594791It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$. 
    595 \autoref{eq:energ1} implies that, to be energetically consistent, 
     792\autoref{eq:ZDF_energ1} implies that, to be energetically consistent, 
    596793the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as 
    597794${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ 
     
    599796 
    600797A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification 
    601 (second term of the right hand side of \autoref{eq:zdftke_e} and \autoref{eq:zdfgls_e}). 
     798(second term of the right hand side of \autoref{eq:ZDF_tke_e} and \autoref{eq:ZDF_gls_e}). 
    602799This term must balance the input of potential energy resulting from vertical mixing. 
    603800The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by 
    604801multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
    605802\begin{equation} 
    606   \label{eq:energ2} 
     803  \label{eq:ZDF_energ2} 
    607804  \begin{split} 
    608805    \int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\ 
     
    614811\end{equation} 
    615812where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$. 
    616 The first term of the right hand side of \autoref{eq:energ2} is always zero because 
     813The first term of the right hand side of \autoref{eq:ZDF_energ2} is always zero because 
    617814there is no diffusive flux through the ocean surface and bottom). 
    618815The second term is minus the destruction rate of  $\bar{e}$ due to stratification. 
    619 Therefore \autoref{eq:energ1} implies that, to be energetically consistent, 
    620 the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e} and  \autoref{eq:zdfgls_e}. 
     816Therefore \autoref{eq:ZDF_energ1} implies that, to be energetically consistent, 
     817the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:ZDF_tke_e} and  \autoref{eq:ZDF_gls_e}. 
    621818 
    622819Let us now address the space discretization issue. 
    623820The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in 
    624 the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:cell}). 
     821the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:DOM_cell}). 
    625822A space averaging is thus required to obtain the shear TKE production term. 
    626 By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
     823By redoing the \autoref{eq:ZDF_energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
    627824the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
    628825Furthermore, the time variation of $e_3$ has be taken into account. 
     
    630827The above energetic considerations leads to the following final discrete form for the TKE equation: 
    631828\begin{equation} 
    632   \label{eq:zdftke_ene} 
     829  \label{eq:ZDF_tke_ene} 
    633830  \begin{split} 
    634831    \frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv 
     
    647844  \end{split} 
    648845\end{equation} 
    649 where the last two terms in \autoref{eq:zdftke_ene} (vertical diffusion and Kolmogorov dissipation) 
    650 are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}). 
     846where the last two terms in \autoref{eq:ZDF_tke_ene} (vertical diffusion and Kolmogorov dissipation) 
     847are time stepped using a backward scheme (see\autoref{sec:TD_forward_imp}). 
    651848Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible. 
    652849%The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as 
    653 %they all appear in the right hand side of \autoref{eq:zdftke_ene}. 
     850%they all appear in the right hand side of \autoref{eq:ZDF_tke_ene}. 
    654851%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored. 
    655852 
    656 % ================================================================ 
    657 % Convection 
    658 % ================================================================ 
     853%% ================================================================================================= 
    659854\section{Convection} 
    660855\label{sec:ZDF_conv} 
     
    667862or/and the use of a turbulent closure scheme. 
    668863 
    669 % ------------------------------------------------------------------------------------------------------------- 
    670 %       Non-Penetrative Convective Adjustment 
    671 % ------------------------------------------------------------------------------------------------------------- 
    672 \subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc = .true.})] 
    673 {Non-penetrative convective adjustment (\protect\np{ln\_tranpc}\forcode{ = .true.})} 
     864%% ================================================================================================= 
     865\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc})]{Non-penetrative convective adjustment (\protect\np{ln_tranpc}{ln\_tranpc})} 
    674866\label{subsec:ZDF_npc} 
    675867 
    676 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    677868\begin{figure}[!htb] 
    678   \begin{center} 
    679     \includegraphics[width=\textwidth]{Fig_npc} 
    680     \caption{ 
    681       \protect\label{fig:npc} 
    682       Example of an unstable density profile treated by the non penetrative convective adjustment algorithm. 
    683       $1^{st}$ step: the initial profile is checked from the surface to the bottom. 
    684       It is found to be unstable between levels 3 and 4. 
    685       They are mixed. 
    686       The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. 
    687       The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. 
    688       The $1^{st}$ step ends since the density profile is then stable below the level 3. 
    689       $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: 
    690       levels 2 to 5 are mixed. 
    691       The new density profile is checked. 
    692       It is found stable: end of algorithm. 
    693     } 
    694   \end{center} 
     869  \centering 
     870  \includegraphics[width=0.66\textwidth]{Fig_npc} 
     871  \caption[Unstable density profile treated by the non penetrative convective adjustment algorithm]{ 
     872    Example of an unstable density profile treated by 
     873    the non penetrative convective adjustment algorithm. 
     874    $1^{st}$ step: the initial profile is checked from the surface to the bottom. 
     875    It is found to be unstable between levels 3 and 4. 
     876    They are mixed. 
     877    The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. 
     878    The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. 
     879    The $1^{st}$ step ends since the density profile is then stable below the level 3. 
     880    $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: 
     881    levels 2 to 5 are mixed. 
     882    The new density profile is checked. 
     883    It is found stable: end of algorithm.} 
     884  \label{fig:ZDF_npc} 
    695885\end{figure} 
    696 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    697  
    698 Options are defined through the \nam{zdf} namelist variables. 
    699 The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}\forcode{ = .true.}. 
    700 It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of 
     886 
     887Options are defined through the \nam{zdf}{zdf} namelist variables. 
     888The non-penetrative convective adjustment is used when \np[=.true.]{ln_zdfnpc}{ln\_zdfnpc}. 
     889It is applied at each \np{nn_npc}{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of 
    701890the water column, but only until the density structure becomes neutrally stable 
    702891(\ie\ until the mixed portion of the water column has \textit{exactly} the density of the water just below) 
    703892\citep{madec.delecluse.ea_JPO91}. 
    704 The associated algorithm is an iterative process used in the following way (\autoref{fig:npc}): 
     893The associated algorithm is an iterative process used in the following way (\autoref{fig:ZDF_npc}): 
    705894starting from the top of the ocean, the first instability is found. 
    706895Assume in the following that the instability is located between levels $k$ and $k+1$. 
     
    734923having to recompute the expansion coefficients at each mixing iteration. 
    735924 
    736 % ------------------------------------------------------------------------------------------------------------- 
    737 %       Enhanced Vertical Diffusion 
    738 % ------------------------------------------------------------------------------------------------------------- 
    739 \subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd = .true.})] 
    740 {Enhanced vertical diffusion (\protect\np{ln\_zdfevd}\forcode{ = .true.})} 
     925%% ================================================================================================= 
     926\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd})]{Enhanced vertical diffusion (\protect\np{ln_zdfevd}{ln\_zdfevd})} 
    741927\label{subsec:ZDF_evd} 
    742928 
    743 Options are defined through the  \nam{zdf} namelist variables. 
    744 The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}. 
     929Options are defined through the  \nam{zdf}{zdf} namelist variables. 
     930The enhanced vertical diffusion parameterisation is used when \np[=.true.]{ln_zdfevd}{ln\_zdfevd}. 
    745931In this case, the vertical eddy mixing coefficients are assigned very large values 
    746932in regions where the stratification is unstable 
    747933(\ie\ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}. 
    748 This is done either on tracers only (\np{nn\_evdm}\forcode{ = 0}) or 
    749 on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}). 
    750  
    751 In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np{nn\_evdm}\forcode{ = 1}, 
     934This is done either on tracers only (\np[=0]{nn_evdm}{nn\_evdm}) or 
     935on both momentum and tracers (\np[=1]{nn_evdm}{nn\_evdm}). 
     936 
     937In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np[=1]{nn_evdm}{nn\_evdm}, 
    752938the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to 
    753 the namelist parameter \np{rn\_avevd}. 
     939the namelist parameter \np{rn_avevd}{rn\_avevd}. 
    754940A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. 
    755941This parameterisation of convective processes is less time consuming than 
     
    759945Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$. 
    760946This removes a potential source of divergence of odd and even time step in 
    761 a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:STP_mLF}). 
    762  
    763 % ------------------------------------------------------------------------------------------------------------- 
    764 %       Turbulent Closure Scheme 
    765 % ------------------------------------------------------------------------------------------------------------- 
    766 \subsection{Handling convection with turbulent closure schemes (\forcode{ln_zdf\{tke,gls,osm\} = .true.})} 
     947a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:TD_mLF}). 
     948 
     949%% ================================================================================================= 
     950\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}})} 
    767951\label{subsec:ZDF_tcs} 
    768952 
    769  
    770953The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and 
    771 \autoref{subsec:ZDF_osm} (\ie\ \np{ln\_zdftke} or \np{ln\_zdfgls} or \np{ln\_zdfosm} defined) deal, in theory, 
     954\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, 
    772955with statically unstable density profiles. 
    773956In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in 
    774 \autoref{eq:zdftke_e} or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative. 
     957\autoref{eq:ZDF_tke_e} or \autoref{eq:ZDF_gls_e} becomes a source term, since $N^2$ is negative. 
    775958It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at 
    776959velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$). 
     
    781964because the mixing length scale is bounded by the distance to the sea surface. 
    782965It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme, 
    783 \ie\ setting the \np{ln\_zdfnpc} namelist parameter to true and 
    784 defining the turbulent closure (\np{ln\_zdftke} or \np{ln\_zdfgls} = \forcode{.true.}) all together. 
     966\ie\ setting the \np{ln_zdfnpc}{ln\_zdfnpc} namelist parameter to true and 
     967defining the turbulent closure (\np{ln_zdftke}{ln\_zdftke} or \np{ln_zdfgls}{ln\_zdfgls} = \forcode{.true.}) all together. 
    785968 
    786969The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
    787970%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
    788 therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the OSMOSIS scheme. 
     971therefore \np[=.false.]{ln_zdfevd}{ln\_zdfevd} should be used with the OSMOSIS scheme. 
    789972% gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
    790973 
    791 % ================================================================ 
    792 % Double Diffusion Mixing 
    793 % ================================================================ 
    794 \section[Double diffusion mixing (\forcode{ln_zdfddm = .true.})] 
    795 {Double diffusion mixing (\protect\np{ln\_zdfddm}\forcode{ = .true.})} 
     974%% ================================================================================================= 
     975\section[Double diffusion mixing (\forcode{ln_zdfddm})]{Double diffusion mixing (\protect\np{ln_zdfddm}{ln\_zdfddm})} 
    796976\label{subsec:ZDF_ddm} 
    797977 
    798  
    799 %-------------------------------------------namzdf_ddm------------------------------------------------- 
    800 % 
    801978%\nlst{namzdf_ddm} 
    802 %-------------------------------------------------------------------------------------------------------------- 
    803979 
    804980This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter 
    805 \np{ln\_zdfddm} in \nam{zdf}. 
     981\np{ln_zdfddm}{ln\_zdfddm} in \nam{zdf}{zdf}. 
    806982Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa. 
    807983The former condition leads to salt fingering and the latter to diffusive convection. 
     
    811987temperature and salinity. 
    812988 
    813  
    814989Diapycnal mixing of S and T are described by diapycnal diffusion coefficients 
    815990\begin{align*} 
    816   % \label{eq:zdfddm_Kz} 
     991  % \label{eq:ZDF_ddm_Kz} 
    817992  &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT} \\ 
    818993  &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS} 
     
    8261001(1981): 
    8271002\begin{align} 
    828   \label{eq:zdfddm_f} 
     1003  \label{eq:ZDF_ddm_f} 
    8291004  A_f^{vS} &= 
    8301005             \begin{cases} 
     
    8321007               0                              &\text{otherwise} 
    8331008             \end{cases} 
    834   \\         \label{eq:zdfddm_f_T} 
     1009  \\         \label{eq:ZDF_ddm_f_T} 
    8351010  A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
    8361011\end{align} 
    8371012 
    838 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    8391013\begin{figure}[!t] 
    840   \begin{center} 
    841     \includegraphics[width=\textwidth]{Fig_zdfddm} 
    842     \caption{ 
    843       \protect\label{fig:zdfddm} 
    844       From \citet{merryfield.holloway.ea_JPO99} : 
    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} 
     1014  \centering 
     1015  \includegraphics[width=0.66\textwidth]{Fig_zdfddm} 
     1016  \caption[Diapycnal diffusivities for temperature and salt in regions of salt fingering and 
     1017  diffusive convection]{ 
     1018    From \citet{merryfield.holloway.ea_JPO99}: 
     1019    (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in 
     1020    regions of salt fingering. 
     1021    Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and 
     1022    thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$; 
     1023    (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in 
     1024    regions of diffusive convection. 
     1025    Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation. 
     1026    The latter is not implemented in \NEMO.} 
     1027  \label{fig:ZDF_ddm} 
    8531028\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 
     1029 
     1030The factor 0.7 in \autoref{eq:ZDF_ddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of 
    8571031buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}). 
    8581032Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
     
    8611035Federov (1988) is used: 
    8621036\begin{align} 
    863   % \label{eq:zdfddm_d} 
     1037  % \label{eq:ZDF_ddm_d} 
    8641038  A_d^{vT} &= 
    8651039             \begin{cases} 
     
    8691043             \end{cases} 
    8701044                                       \nonumber \\ 
    871   \label{eq:zdfddm_d_S} 
     1045  \label{eq:ZDF_ddm_d_S} 
    8721046  A_d^{vS} &= 
    8731047             \begin{cases} 
     
    8781052\end{align} 
    8791053 
    880 The dependencies of \autoref{eq:zdfddm_f} to \autoref{eq:zdfddm_d_S} on $R_\rho$ are illustrated in 
    881 \autoref{fig:zdfddm}. 
     1054The dependencies of \autoref{eq:ZDF_ddm_f} to \autoref{eq:ZDF_ddm_d_S} on $R_\rho$ are illustrated in 
     1055\autoref{fig:ZDF_ddm}. 
    8821056Implementing this requires computing $R_\rho$ at each grid point on every time step. 
    8831057This is done in \mdl{eosbn2} at the same time as $N^2$ is computed. 
    8841058This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive). 
    8851059 
    886 % ================================================================ 
    887 % Bottom Friction 
    888 % ================================================================ 
    889  \section[Bottom and top friction (\textit{zdfdrg.F90})] 
    890  {Bottom and top friction (\protect\mdl{zdfdrg})} 
    891  \label{sec:ZDF_drg} 
    892  
    893 %--------------------------------------------nambfr-------------------------------------------------------- 
    894 % 
    895 \nlst{namdrg} 
    896 \nlst{namdrg_top} 
    897 \nlst{namdrg_bot} 
    898  
    899 %-------------------------------------------------------------------------------------------------------------- 
    900  
    901 Options to define the top and bottom friction are defined through the \nam{drg} namelist variables. 
     1060%% ================================================================================================= 
     1061\section[Bottom and top friction (\textit{zdfdrg.F90})]{Bottom and top friction (\protect\mdl{zdfdrg})} 
     1062\label{sec:ZDF_drg} 
     1063 
     1064\begin{listing} 
     1065  \nlst{namdrg} 
     1066  \caption{\forcode{&namdrg}} 
     1067  \label{lst:namdrg} 
     1068\end{listing} 
     1069\begin{listing} 
     1070  \nlst{namdrg_top} 
     1071  \caption{\forcode{&namdrg_top}} 
     1072  \label{lst:namdrg_top} 
     1073\end{listing} 
     1074\begin{listing} 
     1075  \nlst{namdrg_bot} 
     1076  \caption{\forcode{&namdrg_bot}} 
     1077  \label{lst:namdrg_bot} 
     1078\end{listing} 
     1079 
     1080Options to define the top and bottom friction are defined through the \nam{drg}{drg} namelist variables. 
    9021081The bottom friction represents the friction generated by the bathymetry. 
    9031082The top friction represents the friction generated by the ice shelf/ocean interface. 
     
    9051084the description below considers mostly the bottom friction case, if not stated otherwise. 
    9061085 
    907  
    9081086Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as 
    9091087a condition on the vertical diffusive flux. 
    9101088For the bottom boundary layer, one has: 
    9111089 \[ 
    912    % \label{eq:zdfbfr_flux} 
     1090   % \label{eq:ZDF_bfr_flux} 
    9131091   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
    9141092 \] 
     
    9261104To illustrate this, consider the equation for $u$ at $k$, the last ocean level: 
    9271105\begin{equation} 
    928   \label{eq:zdfdrg_flux2} 
     1106  \label{eq:ZDF_drg_flux2} 
    9291107  \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}} 
    9301108\end{equation} 
     
    9461124 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where: 
    9471125\begin{equation} 
    948   \label{eq:zdfbfr_bdef} 
     1126  \label{eq:ZDF_bfr_bdef} 
    9491127  \frac{\partial {\textbf U_h}}{\partial t} = 
    9501128  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b 
     
    9531131Note 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. 
    9541132 
    955 % ------------------------------------------------------------------------------------------------------------- 
    956 %       Linear Bottom Friction 
    957 % ------------------------------------------------------------------------------------------------------------- 
    958  \subsection[Linear top/bottom friction (\forcode{ln_lin = .true.})] 
    959  {Linear top/bottom friction (\protect\np{ln\_lin}\forcode{ = .true.)}} 
    960  \label{subsec:ZDF_drg_linear} 
     1133%% ================================================================================================= 
     1134\subsection[Linear top/bottom friction (\forcode{ln_lin})]{Linear top/bottom friction (\protect\np{ln_lin}{ln\_lin})} 
     1135\label{subsec:ZDF_drg_linear} 
    9611136 
    9621137The linear friction parameterisation (including the special case of a free-slip condition) assumes that 
    9631138the friction is proportional to the interior velocity (\ie\ the velocity of the first/last model level): 
    9641139\[ 
    965   % \label{eq:zdfbfr_linear} 
     1140  % \label{eq:ZDF_bfr_linear} 
    9661141  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
    9671142\] 
     
    9761151and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$. 
    9771152This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days. 
    978 It can be changed by specifying \np{rn\_Uc0} (namelist parameter). 
    979  
    980  For the linear friction case the drag coefficient used in the general expression \autoref{eq:zdfbfr_bdef} is: 
    981 \[ 
    982   % \label{eq:zdfbfr_linbfr_b} 
     1153It can be changed by specifying \np{rn_Uc0}{rn\_Uc0} (namelist parameter). 
     1154 
     1155 For the linear friction case the drag coefficient used in the general expression \autoref{eq:ZDF_bfr_bdef} is: 
     1156\[ 
     1157  % \label{eq:ZDF_bfr_linbfr_b} 
    9831158    c_b^T = - r 
    9841159\] 
    985 When \np{ln\_lin} \forcode{= .true.}, the value of $r$ used is \np{rn\_Uc0}*\np{rn\_Cd0}. 
    986 Setting \np{ln\_OFF} \forcode{= .true.} (and \forcode{ln_lin = .true.}) is equivalent to setting $r=0$ and leads to a free-slip boundary condition. 
     1160When \np[=.true.]{ln_lin}{ln\_lin}, the value of $r$ used is \np{rn_Uc0}{rn\_Uc0}*\np{rn_Cd0}{rn\_Cd0}. 
     1161Setting \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. 
    9871162 
    9881163These values are assigned in \mdl{zdfdrg}. 
    9891164Note that there is support for local enhancement of these values via an externally defined 2D mask array 
    990 (\np{ln\_boost}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file. 
     1165(\np[=.true.]{ln_boost}{ln\_boost}) given in the \ifile{bfr\_coef} input NetCDF file. 
    9911166The mask values should vary from 0 to 1. 
    9921167Locations with a non-zero mask value will have the friction coefficient increased by 
    993 $mask\_value$ * \np{rn\_boost} * \np{rn\_Cd0}. 
    994  
    995 % ------------------------------------------------------------------------------------------------------------- 
    996 %       Non-Linear Bottom Friction 
    997 % ------------------------------------------------------------------------------------------------------------- 
    998  \subsection[Non-linear top/bottom friction (\forcode{ln_non_lin = .true.})] 
    999  {Non-linear top/bottom friction (\protect\np{ln\_non\_lin}\forcode{ = .true.})} 
    1000  \label{subsec:ZDF_drg_nonlinear} 
     1168$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}. 
     1169 
     1170%% ================================================================================================= 
     1171\subsection[Non-linear top/bottom friction (\forcode{ln_non_lin})]{Non-linear top/bottom friction (\protect\np{ln_non_lin}{ln\_non\_lin})} 
     1172\label{subsec:ZDF_drg_nonlinear} 
    10011173 
    10021174The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic: 
    10031175\[ 
    1004   % \label{eq:zdfdrg_nonlinear} 
     1176  % \label{eq:ZDF_drg_nonlinear} 
    10051177  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h 
    10061178  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b 
     
    10121184$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and 
    10131185$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$. 
    1014 The CME choices have been set as default values (\np{rn\_Cd0} and \np{rn\_ke0} namelist parameters). 
     1186The CME choices have been set as default values (\np{rn_Cd0}{rn\_Cd0} and \np{rn_ke0}{rn\_ke0} namelist parameters). 
    10151187 
    10161188As for the linear case, the friction is imposed in the code by adding the trend due to 
     
    10181190For the non-linear friction case the term computed in \mdl{zdfdrg} is: 
    10191191\[ 
    1020   % \label{eq:zdfdrg_nonlinbfr} 
     1192  % \label{eq:ZDF_drg_nonlinbfr} 
    10211193    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2} 
    10221194\] 
    10231195 
    10241196The coefficients that control the strength of the non-linear friction are initialised as namelist parameters: 
    1025 $C_D$= \np{rn\_Cd0}, and $e_b$ =\np{rn\_bfeb2}. 
    1026 Note that for applications which consider tides explicitly, a low or even zero value of \np{rn\_bfeb2} is recommended. A local enhancement of $C_D$ is again possible via an externally defined 2D mask array 
    1027 (\np{ln\_boost}\forcode{ = .true.}). 
     1197$C_D$= \np{rn_Cd0}{rn\_Cd0}, and $e_b$ =\np{rn_bfeb2}{rn\_bfeb2}. 
     1198Note 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 
     1199(\np[=.true.]{ln_boost}{ln\_boost}). 
    10281200This works in the same way as for the linear friction case with non-zero masked locations increased by 
    1029 $mask\_value$ * \np{rn\_boost} * \np{rn\_Cd0}. 
    1030  
    1031 % ------------------------------------------------------------------------------------------------------------- 
    1032 %       Bottom Friction Log-layer 
    1033 % ------------------------------------------------------------------------------------------------------------- 
    1034  \subsection[Log-layer top/bottom friction (\forcode{ln_loglayer = .true.})] 
    1035  {Log-layer top/bottom friction (\protect\np{ln\_loglayer}\forcode{ = .true.})} 
    1036  \label{subsec:ZDF_drg_loglayer} 
     1201$mask\_value$ * \np{rn_boost}{rn\_boost} * \np{rn_Cd0}{rn\_Cd0}. 
     1202 
     1203%% ================================================================================================= 
     1204\subsection[Log-layer top/bottom friction (\forcode{ln_loglayer})]{Log-layer top/bottom friction (\protect\np{ln_loglayer}{ln\_loglayer})} 
     1205\label{subsec:ZDF_drg_loglayer} 
    10371206 
    10381207In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
    10391208a "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. 
    1040 If  \np{ln\_loglayer} \forcode{= .true.}, $C_D$ is no longer constant but is related to the distance to the wall (or equivalently to the half of the top/bottom layer thickness): 
     1209If  \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): 
    10411210\[ 
    10421211  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2 
    10431212\] 
    10441213 
    1045 \noindent where $\kappa$ is the von-Karman constant and \np{rn\_z0} is a roughness length provided via the namelist. 
     1214\noindent where $\kappa$ is the von-Karman constant and \np{rn_z0}{rn\_z0} is a roughness length provided via the namelist. 
    10461215 
    10471216The drag coefficient is bounded such that it is kept greater or equal to 
    1048 the base \np{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: 
    1049 \np{rn\_Cdmax}, \ie 
     1217the 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: 
     1218\np{rn_Cdmax}{rn\_Cdmax}, \ie 
    10501219\[ 
    10511220  rn\_Cd0 \leq C_D \leq rn\_Cdmax 
     
    10531222 
    10541223\noindent The log-layer enhancement can also be applied to the top boundary friction if 
    1055 under ice-shelf cavities are activated (\np{ln\_isfcav}\forcode{ = .true.}). 
    1056 %In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}. 
    1057  
    1058 % ------------------------------------------------------------------------------------------------------------- 
    1059 %       Explicit bottom Friction 
    1060 % ------------------------------------------------------------------------------------------------------------- 
    1061  \subsection{Explicit top/bottom friction (\forcode{ln_drgimp = .false.})} 
    1062  \label{subsec:ZDF_drg_stability} 
    1063  
    1064 Setting \np{ln\_drgimp} \forcode{= .false.} 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: 
     1224under ice-shelf cavities are activated (\np[=.true.]{ln_isfcav}{ln\_isfcav}). 
     1225%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}. 
     1226 
     1227%% ================================================================================================= 
     1228\subsection[Explicit top/bottom friction (\forcode{ln_drgimp=.false.})]{Explicit top/bottom friction (\protect\np[=.false.]{ln_drgimp}{ln\_drgimp})} 
     1229\label{subsec:ZDF_drg_stability} 
     1230 
     1231Setting \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: 
    10651232 
    10661233At the top (below an ice shelf cavity): 
     
    10771244 
    10781245Since 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. 
    1079 For the purposes of stability analysis, an approximation to \autoref{eq:zdfdrg_flux2} is: 
     1246For the purposes of stability analysis, an approximation to \autoref{eq:ZDF_drg_flux2} is: 
    10801247\begin{equation} 
    1081   \label{eq:Eqn_drgstab} 
     1248  \label{eq:ZDF_Eqn_drgstab} 
    10821249  \begin{split} 
    10831250    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\ 
     
    10901257  |\Delta u| < \;|u| 
    10911258\] 
    1092 \noindent which, using \autoref{eq:Eqn_drgstab}, gives: 
     1259\noindent which, using \autoref{eq:ZDF_Eqn_drgstab}, gives: 
    10931260\[ 
    10941261  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\ 
     
    11171284The number of potential breaches of the explicit stability criterion are still reported for information purposes. 
    11181285 
    1119 % ------------------------------------------------------------------------------------------------------------- 
    1120 %       Implicit Bottom Friction 
    1121 % ------------------------------------------------------------------------------------------------------------- 
    1122  \subsection[Implicit top/bottom friction (\forcode{ln_drgimp = .true.})] 
    1123  {Implicit top/bottom friction (\protect\np{ln\_drgimp}\forcode{ = .true.})} 
    1124  \label{subsec:ZDF_drg_imp} 
     1286%% ================================================================================================= 
     1287\subsection[Implicit top/bottom friction (\forcode{ln_drgimp=.true.})]{Implicit top/bottom friction (\protect\np[=.true.]{ln_drgimp}{ln\_drgimp})} 
     1288\label{subsec:ZDF_drg_imp} 
    11251289 
    11261290An optional implicit form of bottom friction has been implemented to improve model stability. 
    11271291We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting. 
    1128 This option can be invoked by setting \np{ln\_drgimp} to \forcode{.true.} in the \nam{drg} namelist. 
    1129 %This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf} namelist. 
     1292This option can be invoked by setting \np{ln_drgimp}{ln\_drgimp} to \forcode{.true.} in the \nam{drg}{drg} namelist. 
     1293%This option requires \np{ln_zdfexp}{ln\_zdfexp} to be \forcode{.false.} in the \nam{zdf}{zdf} namelist. 
    11301294 
    11311295This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step: 
     
    11331297At the top (below an ice shelf cavity): 
    11341298\[ 
    1135   % \label{eq:dynzdf_drg_top} 
     1299  % \label{eq:ZDF_dynZDF__drg_top} 
    11361300  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t} 
    11371301  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t} 
     
    11401304At the bottom (above the sea floor): 
    11411305\[ 
    1142   % \label{eq:dynzdf_drg_bot} 
     1306  % \label{eq:ZDF_dynZDF__drg_bot} 
    11431307  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b} 
    11441308  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b} 
     
    11481312Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit. 
    11491313 
    1150 % ------------------------------------------------------------------------------------------------------------- 
    1151 %       Bottom Friction with split-explicit free surface 
    1152 % ------------------------------------------------------------------------------------------------------------- 
    1153  \subsection[Bottom friction with split-explicit free surface] 
    1154  {Bottom friction with split-explicit free surface} 
    1155  \label{subsec:ZDF_drg_ts} 
    1156  
    1157 With 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{ln\_drgimp}\forcode{= .false.} 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{ln\_drgimp}\forcode{= .true.},  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. 
     1314%% ================================================================================================= 
     1315\subsection[Bottom friction with split-explicit free surface]{Bottom friction with split-explicit free surface} 
     1316\label{subsec:ZDF_drg_ts} 
     1317 
     1318With 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. 
    11581319 
    11591320The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO\ is as follows: 
     
    11651326Note that other strategies are possible, like considering vertical diffusion step in advance, \ie\ prior barotropic integration. 
    11661327 
    1167  
    1168 % ================================================================ 
    1169 % Internal wave-driven mixing 
    1170 % ================================================================ 
    1171 \section[Internal wave-driven mixing (\forcode{ln_zdfiwm = .true.})] 
    1172 {Internal wave-driven mixing (\protect\np{ln\_zdfiwm}\forcode{ = .true.})} 
     1328%% ================================================================================================= 
     1329\section[Internal wave-driven mixing (\forcode{ln_zdfiwm})]{Internal wave-driven mixing (\protect\np{ln_zdfiwm}{ln\_zdfiwm})} 
    11731330\label{subsec:ZDF_tmx_new} 
    11741331 
    1175 %--------------------------------------------namzdf_iwm------------------------------------------ 
    1176 % 
    1177 \nlst{namzdf_iwm} 
    1178 %-------------------------------------------------------------------------------------------------------------- 
     1332\begin{listing} 
     1333  \nlst{namzdf_iwm} 
     1334  \caption{\forcode{&namzdf_iwm}} 
     1335  \label{lst:namzdf_iwm} 
     1336\end{listing} 
    11791337 
    11801338The parameterization of mixing induced by breaking internal waves is a generalization of 
     
    11831341and the resulting diffusivity is obtained as 
    11841342\[ 
    1185   % \label{eq:Kwave} 
     1343  % \label{eq:ZDF_Kwave} 
    11861344  A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 } 
    11871345\] 
    11881346where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of 
    11891347the energy available for mixing. 
    1190 If the \np{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and 
     1348If the \np{ln_mevar}{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and 
    11911349equal to 1/6 \citep{osborn_JPO80}. 
    11921350In the opposite (recommended) case, $R_f$ is instead a function of 
     
    11981356 
    11991357In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary 
    1200 as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice. 
     1358as a function of $Re_b$ by setting the \np{ln_tsdiff}{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice. 
    12011359This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14}, 
    12021360is implemented as in \cite{de-lavergne.madec.ea_JPO16}. 
     
    12161374  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; , 
    12171375\] 
    1218 The $n_p$ parameter (given by \np{nn\_zpyc} in \nam{zdf\_iwm} namelist) 
     1376The $n_p$ parameter (given by \np{nn_zpyc}{nn\_zpyc} in \nam{zdf_iwm}{zdf\_iwm} namelist) 
    12191377controls the stratification-dependence of the pycnocline-intensified dissipation. 
    12201378It can take values of $1$ (recommended) or $2$. 
     
    12241382$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of 
    12251383the abyssal hill topography \citep{goff_JGR10} and the latitude. 
    1226 % 
    12271384% Jc: input files names ? 
    12281385 
    1229 % ================================================================ 
    1230 % surface wave-induced mixing 
    1231 % ================================================================ 
    1232 \section[Surface wave-induced mixing (\forcode{ln_zdfswm = .true.})] 
    1233 {Surface wave-induced mixing (\protect\np{ln\_zdfswm}\forcode{ = .true.})} 
     1386%% ================================================================================================= 
     1387\section[Surface wave-induced mixing (\forcode{ln_zdfswm})]{Surface wave-induced mixing (\protect\np{ln_zdfswm}{ln\_zdfswm})} 
    12341388\label{subsec:ZDF_swm} 
    12351389 
     
    12421396 
    12431397\begin{equation} 
    1244   \label{eq:Bv} 
     1398  \label{eq:ZDF_Bv} 
    12451399  B_{v} = \alpha {A} {U}_{st} {exp(3kz)} 
    12461400\end{equation} 
     
    12541408and diffusivity coefficients. 
    12551409 
    1256 In order to account for this contribution set: \forcode{ln_zdfswm = .true.}, 
    1257 then wave interaction has to be activated through \forcode{ln_wave = .true.}, 
    1258 the Stokes Drift can be evaluated by setting \forcode{ln_sdw = .true.} 
     1410In order to account for this contribution set: \forcode{ln_zdfswm=.true.}, 
     1411then wave interaction has to be activated through \forcode{ln_wave=.true.}, 
     1412the Stokes Drift can be evaluated by setting \forcode{ln_sdw=.true.} 
    12591413(see \autoref{subsec:SBC_wave_sdw}) 
    12601414and the needed wave fields can be provided either in forcing or coupled mode 
    12611415(for more information on wave parameters and settings see \autoref{sec:SBC_wave}) 
    12621416 
    1263 % ================================================================ 
    1264 % Adaptive-implicit vertical advection 
    1265 % ================================================================ 
    1266 \section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp = .true.})] 
    1267 {Adaptive-implicit vertical advection(\protect\np{ln\_zad\_Aimp}\forcode{ = .true.})} 
     1417%% ================================================================================================= 
     1418\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp})]{Adaptive-implicit vertical advection(\protect\np{ln_zad_Aimp}{ln\_zad\_Aimp})} 
    12681419\label{subsec:ZDF_aimp} 
    12691420 
     
    12761427criteria for a range of advection schemes. The values for the Leap-Frog with Robert 
    12771428asselin filter time-stepping (as used in NEMO) are reproduced in 
    1278 \autoref{tab:zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these 
     1429\autoref{tab:ZDF_zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these 
    12791430restrictions but at the cost of large dispersive errors and, possibly, large numerical 
    12801431viscosity. The adaptive-implicit vertical advection option provides a targetted use of the 
     
    12831434interest or due to short-lived conditions such that the extra numerical diffusion or 
    12841435viscosity does not greatly affect the overall solution. With such applications, setting: 
    1285 \forcode{ln_zad_Aimp = .true.} should allow much longer model timesteps to be used whilst 
     1436\forcode{ln_zad_Aimp=.true.} should allow much longer model timesteps to be used whilst 
    12861437retaining the accuracy of the high order explicit schemes over most of the domain. 
    12871438 
    12881439\begin{table}[htbp] 
    1289   \begin{center} 
    1290     % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}} 
    1291     \begin{tabular}{r|ccc} 
    1292       \hline 
    1293       spatial discretization &   2nd order centered   & 3rd order upwind & 4th order compact  \\ 
    1294       advective CFL criterion     & 0.904 &   0.472  &   0.522    \\ 
    1295       \hline 
    1296     \end{tabular} 
    1297     \caption{ 
    1298       \protect\label{tab:zad_Aimp_CFLcrit} 
    1299       The advective CFL criteria for a range of spatial discretizations for the Leap-Frog with Robert Asselin filter time-stepping 
    1300       ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}. 
    1301     } 
    1302   \end{center} 
     1440  \centering 
     1441  % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}} 
     1442  \begin{tabular}{r|ccc} 
     1443    \hline 
     1444    spatial discretization  & 2$^nd$ order centered & 3$^rd$ order upwind & 4$^th$ order compact \\ 
     1445    advective CFL criterion &                 0.904 &              0.472  &                0.522 \\ 
     1446    \hline 
     1447  \end{tabular} 
     1448  \caption[Advective CFL criteria for the leapfrog with Robert Asselin filter time-stepping]{ 
     1449    The advective CFL criteria for a range of spatial discretizations for 
     1450    the leapfrog with Robert Asselin filter time-stepping 
     1451    ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}.} 
     1452  \label{tab:ZDF_zad_Aimp_CFLcrit} 
    13031453\end{table} 
    13041454 
     
    13131463 
    13141464\begin{equation} 
    1315   \label{eq:Eqn_zad_Aimp_Courant} 
     1465  \label{eq:ZDF_Eqn_zad_Aimp_Courant} 
    13161466  \begin{split} 
    13171467    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 ]    \\ 
     
    13261476 
    13271477\begin{align} 
    1328   \label{eq:Eqn_zad_Aimp_partition} 
     1478  \label{eq:ZDF_Eqn_zad_Aimp_partition} 
    13291479Cu_{min} &= 0.15 \nonumber \\ 
    13301480Cu_{max} &= 0.3  \nonumber \\ 
    13311481Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\ 
    13321482Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\ 
    1333 C\kern-0.14em f &= 
     1483\cf &= 
    13341484     \begin{cases} 
    13351485        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\ 
     
    13401490 
    13411491\begin{figure}[!t] 
    1342   \begin{center} 
    1343     \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_coeff} 
    1344     \caption{ 
    1345       \protect\label{fig:zad_Aimp_coeff} 
    1346       The value of the partitioning coefficient ($C\kern-0.14em f$) used to partition vertical velocities into parts to 
    1347       be treated implicitly and explicitly for a range of typical Courant numbers (\forcode{ln_zad_Aimp=.true.}) 
    1348     } 
    1349   \end{center} 
     1492  \centering 
     1493  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_coeff} 
     1494  \caption[Partitioning coefficient used to partition vertical velocities into parts]{ 
     1495    The value of the partitioning coefficient (\cf) used to partition vertical velocities into 
     1496    parts to be treated implicitly and explicitly for a range of typical Courant numbers 
     1497    (\forcode{ln_zad_Aimp=.true.}).} 
     1498  \label{fig:ZDF_zad_Aimp_coeff} 
    13501499\end{figure} 
    13511500 
     
    13551504 
    13561505\begin{align} 
    1357   \label{eq:Eqn_zad_Aimp_partition2} 
    1358     w_{i_{ijk}} &= C\kern-0.14em f_{ijk} w_{n_{ijk}}     \nonumber \\ 
    1359     w_{n_{ijk}} &= (1-C\kern-0.14em f_{ijk}) w_{n_{ijk}}            
     1506  \label{eq:ZDF_Eqn_zad_Aimp_partition2} 
     1507    w_{i_{ijk}} &= \cf_{ijk} w_{n_{ijk}}     \nonumber \\ 
     1508    w_{n_{ijk}} &= (1-\cf_{ijk}) w_{n_{ijk}} 
    13601509\end{align} 
    13611510 
    13621511\noindent Note that the coefficient is such that the treatment is never fully implicit; 
    1363 the three cases from \autoref{eq:Eqn_zad_Aimp_partition} can be considered as: 
     1512the three cases from \autoref{eq:ZDF_Eqn_zad_Aimp_partition} can be considered as: 
    13641513fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the 
    1365 coefficient ($C\kern-0.14em f$) varies as shown in \autoref{fig:zad_Aimp_coeff}. Note with these values 
     1514coefficient (\cf) varies as shown in \autoref{fig:ZDF_zad_Aimp_coeff}. Note with these values 
    13661515the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly 
    13671516implicit' is 0.45 which is just below the stability limited given in 
    1368 \autoref{tab:zad_Aimp_CFLcrit}  for a 3rd order scheme. 
     1517\autoref{tab:ZDF_zad_Aimp_CFLcrit}  for a 3rd order scheme. 
    13691518 
    13701519The $w_i$ component is added to the implicit solvers for the vertical mixing in 
     
    13761525vertical fluxes are then removed since they are added by the implicit solver later on. 
    13771526 
    1378 The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be  
     1527The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be 
    13791528used in a wide range of simulations. The following test simulation, however, does illustrate 
    13801529the potential benefits and will hopefully encourage further testing and feedback from users: 
    13811530 
    13821531\begin{figure}[!t] 
    1383   \begin{center} 
    1384     \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_overflow_frames} 
    1385     \caption{ 
    1386       \protect\label{fig:zad_Aimp_overflow_frames} 
    1387       A time-series of temperature vertical cross-sections for the OVERFLOW test case. These results are for the default 
    1388       settings with \forcode{nn_rdt=10.0} and without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}). 
    1389     } 
    1390   \end{center} 
     1532  \centering 
     1533  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_overflow_frames} 
     1534  \caption[OVERFLOW: time-series of temperature vertical cross-sections]{ 
     1535    A time-series of temperature vertical cross-sections for the OVERFLOW test case. 
     1536    These results are for the default settings with \forcode{nn_rdt=10.0} and 
     1537    without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}).} 
     1538  \label{fig:ZDF_zad_Aimp_overflow_frames} 
    13911539\end{figure} 
    13921540 
     1541%% ================================================================================================= 
    13931542\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case} 
     1543 
    13941544The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case} 
    13951545provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case 
     
    14071557 
    14081558\noindent which were chosen to provide a slightly more stable and less noisy solution. The 
    1409 result when using the default value of \forcode{nn_rdt = 10.} without adaptive-implicit 
    1410 vertical velocity is illustrated in \autoref{fig:zad_Aimp_overflow_frames}. The mass of 
     1559result when using the default value of \forcode{nn_rdt=10.} without adaptive-implicit 
     1560vertical velocity is illustrated in \autoref{fig:ZDF_zad_Aimp_overflow_frames}. The mass of 
    14111561cold water, initially sitting on the shelf, moves down the slope and forms a 
    14121562bottom-trapped, dense plume. Even with these extra physics choices the model is close to 
    1413 stability limits and attempts with \forcode{nn_rdt = 30.} will fail after about 5.5 hours 
     1563stability limits and attempts with \forcode{nn_rdt=30.} will fail after about 5.5 hours 
    14141564with excessively high horizontal velocities. This time-scale corresponds with the time the 
    14151565plume reaches the steepest part of the topography and, although detected as a horizontal 
     
    14181568 
    14191569The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps 
    1420 are shown in \autoref{fig:zad_Aimp_overflow_all_rdt} (together with the equivalent 
     1570are shown in \autoref{fig:ZDF_zad_Aimp_overflow_all_rdt} (together with the equivalent 
    14211571frames from the base run).  In this simple example the use of the adaptive-implicit 
    14221572vertcal advection scheme has enabled a 12x increase in the model timestep without 
    14231573significantly altering the solution (although at this extreme the plume is more diffuse 
    14241574and has not travelled so far).  Notably, the solution with and without the scheme is 
    1425 slightly different even with \forcode{nn_rdt = 10.}; suggesting that the base run was 
     1575slightly different even with \forcode{nn_rdt=10.}; suggesting that the base run was 
    14261576close enough to instability to trigger the scheme despite completing successfully. 
    14271577To assist in diagnosing how active the scheme is, in both location and time, the 3D 
    14281578implicit and explicit components of the vertical velocity are available via XIOS as 
    14291579\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient 
    1430 ($C\kern-0.14em f$) is also available as \texttt{wi\_cff}. For a quick oversight of 
     1580(\cf) is also available as \texttt{wi\_cff}. For a quick oversight of 
    14311581the schemes activity the global maximum values of the absolute implicit component 
    14321582of the vertical velocity and the partitioning coefficient are written to the netCDF 
     
    14341584\autoref{sec:MISC_opt} for activation details). 
    14351585 
    1436 \autoref{fig:zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for 
     1586\autoref{fig:ZDF_zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for 
    14371587the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is 
    14381588active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the 
     
    14411591oscillatory nature of this measure appears to be linked to the progress of the plume front 
    14421592as each cusp is associated with the location of the maximum shifting to the adjacent cell. 
    1443 This is illustrated in \autoref{fig:zad_Aimp_maxCf_loc} where the i- and k- locations of the 
     1593This is illustrated in \autoref{fig:ZDF_zad_Aimp_maxCf_loc} where the i- and k- locations of the 
    14441594maximum have been overlaid for the base run case. 
    14451595 
     
    14601610 
    14611611\begin{figure}[!t] 
    1462   \begin{center} 
    1463     \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_overflow_all_rdt} 
    1464     \caption{ 
    1465       \protect\label{fig:zad_Aimp_overflow_all_rdt} 
    1466       Sample temperature vertical cross-sections from mid- and end-run using different values for \forcode{nn_rdt}  
    1467       and with or without adaptive implicit vertical advection. Without the adaptive implicit vertical advection only 
    1468       the run with the shortest timestep is able to run to completion. Note also that the colour-scale has been 
    1469       chosen to confirm that temperatures remain within the original range of 10$^o$ to 20$^o$. 
    1470     } 
    1471   \end{center} 
     1612  \centering 
     1613  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_overflow_all_rdt} 
     1614  \caption[OVERFLOW: sample temperature vertical cross-sections from mid- and end-run]{ 
     1615    Sample temperature vertical cross-sections from mid- and end-run using 
     1616    different values for \forcode{nn_rdt} and with or without adaptive implicit vertical advection. 
     1617    Without the adaptive implicit vertical advection 
     1618    only the run with the shortest timestep is able to run to completion. 
     1619    Note also that the colour-scale has been chosen to confirm that 
     1620    temperatures remain within the original range of 10$^o$ to 20$^o$.} 
     1621  \label{fig:ZDF_zad_Aimp_overflow_all_rdt} 
    14721622\end{figure} 
    14731623 
    14741624\begin{figure}[!t] 
    1475   \begin{center} 
    1476     \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_maxCf} 
    1477     \caption{ 
    1478       \protect\label{fig:zad_Aimp_maxCf} 
    1479       The maximum partitioning coefficient during a series of test runs with increasing model timestep length. 
    1480       At the larger timesteps, the vertical velocity is treated mostly implicitly at some location throughout  
    1481       the run. 
    1482     } 
    1483   \end{center} 
     1625  \centering 
     1626  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_maxCf} 
     1627  \caption[OVERFLOW: maximum partitioning coefficient during a series of test runs]{ 
     1628    The maximum partitioning coefficient during a series of test runs with 
     1629    increasing model timestep length. 
     1630    At the larger timesteps, 
     1631    the vertical velocity is treated mostly implicitly at some location throughout the run.} 
     1632  \label{fig:ZDF_zad_Aimp_maxCf} 
    14841633\end{figure} 
    14851634 
    14861635\begin{figure}[!t] 
    1487   \begin{center} 
    1488     \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_maxCf_loc} 
    1489     \caption{ 
    1490       \protect\label{fig:zad_Aimp_maxCf_loc} 
    1491       The maximum partitioning coefficient for the  \forcode{nn_rdt=10.0s} case overlaid with  information on the gridcell i- and k- 
    1492       locations of the maximum value.  
    1493     } 
    1494   \end{center} 
     1636  \centering 
     1637  \includegraphics[width=0.66\textwidth]{Fig_ZDF_zad_Aimp_maxCf_loc} 
     1638  \caption[OVERFLOW: maximum partitioning coefficient for the case overlaid]{ 
     1639    The maximum partitioning coefficient for the \forcode{nn_rdt=10.0} case overlaid with 
     1640    information on the gridcell i- and k-locations of the maximum value.} 
     1641  \label{fig:ZDF_zad_Aimp_maxCf_loc} 
    14951642\end{figure} 
    14961643 
    1497 % ================================================================ 
    1498  
    1499 \biblio 
    1500  
    1501 \pindex 
     1644\onlyinsubfile{\input{../../global/epilogue}} 
    15021645 
    15031646\end{document} 
Note: See TracChangeset for help on using the changeset viewer.