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 11422 for NEMO/branches/2019/fix_vvl_ticket1791/doc/latex/NEMO/subfiles/chap_ZDF.tex – NEMO

Ignore:
Timestamp:
2019-08-08T15:40:47+02:00 (5 years ago)
Author:
jchanut
Message:

#1791, merge with trunk

Location:
NEMO/branches/2019/fix_vvl_ticket1791/doc
Files:
4 edited

Legend:

Unmodified
Added
Removed
  • NEMO/branches/2019/fix_vvl_ticket1791/doc

    • Property svn:ignore deleted
  • NEMO/branches/2019/fix_vvl_ticket1791/doc/latex

    • Property svn:ignore
      •  

        old new  
        1 *.aux 
        2 *.bbl 
        3 *.blg 
        4 *.dvi 
        5 *.fdb* 
        6 *.fls 
        7 *.idx 
        8 *.ilg 
        9 *.ind 
        10 *.log 
        11 *.maf 
        12 *.mtc* 
        13 *.out 
        14 *.pdf 
        15 *.toc 
        16 _minted-* 
         1figures 
  • NEMO/branches/2019/fix_vvl_ticket1791/doc/latex/NEMO

    • Property svn:ignore deleted
  • NEMO/branches/2019/fix_vvl_ticket1791/doc/latex/NEMO/subfiles/chap_ZDF.tex

    r10442 r11422  
    2525At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}), 
    2626while at the bottom they are set to zero for heat and salt, 
    27 unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie \key{trabbl} defined, 
     27unless a geothermal flux forcing is prescribed as a bottom boundary condition (\ie \np{ln\_trabbc} defined, 
    2828see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum 
    29 (see \autoref{sec:ZDF_bfr}). 
     29(see \autoref{sec:ZDF_drg}).  
    3030 
    3131In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and 
     
    3333respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}). 
    3434These coefficients can be assumed to be either constant, or a function of the local Richardson number, 
    35 or computed from a turbulent closure model (either TKE or GLS formulation). 
    36 The computation of these coefficients is initialized in the \mdl{zdfini} module and performed in 
    37 the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} modules. 
     35or computed from a turbulent closure model (either TKE or GLS or OSMOSIS formulation). 
     36The computation of these coefficients is initialized in the \mdl{zdfphy} module and performed in 
     37the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} or \mdl{zdfosm} modules. 
    3838The trends due to the vertical momentum and tracer diffusion, including the surface forcing, 
    3939are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.  
    40 These trends can be computed using either a forward time stepping scheme 
    41 (namelist parameter \np{ln\_zdfexp}\forcode{ = .true.}) or a backward time stepping scheme 
    42 (\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing coefficients, 
    43 and thus of the formulation used (see \autoref{chap:STP}). 
    44  
    45 % ------------------------------------------------------------------------------------------------------------- 
    46 %        Constant  
    47 % ------------------------------------------------------------------------------------------------------------- 
    48 \subsection{Constant (\protect\key{zdfcst})} 
    49 \label{subsec:ZDF_cst} 
    50 %--------------------------------------------namzdf--------------------------------------------------------- 
     40%These trends can be computed using either a forward time stepping scheme 
     41%(namelist parameter \np{ln\_zdfexp}\forcode{ = .true.}) or a backward time stepping scheme 
     42%(\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing coefficients, 
     43%and thus of the formulation used (see \autoref{chap:STP}). 
     44 
     45%--------------------------------------------namzdf-------------------------------------------------------- 
    5146 
    5247\nlst{namzdf} 
    5348%-------------------------------------------------------------------------------------------------------------- 
    5449 
     50% ------------------------------------------------------------------------------------------------------------- 
     51%        Constant  
     52% ------------------------------------------------------------------------------------------------------------- 
     53\subsection[Constant (\forcode{ln_zdfcst = .true.})] 
     54{Constant (\protect\np{ln\_zdfcst}\forcode{ = .true.})} 
     55\label{subsec:ZDF_cst} 
     56 
    5557Options are defined through the \ngn{namzdf} namelist variables. 
    56 When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
     58When \np{ln\_zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
    5759constant values over the whole ocean. 
    5860This is the crudest way to define the vertical ocean physics. 
    59 It is recommended that this option is only used in process studies, not in basin scale simulations. 
     61It is recommended to use this option only in process studies, not in basin scale simulations. 
    6062Typical values used in this case are: 
    6163\begin{align*} 
     
    7274%        Richardson Number Dependent 
    7375% ------------------------------------------------------------------------------------------------------------- 
    74 \subsection{Richardson number dependent (\protect\key{zdfric})} 
     76\subsection[Richardson number dependent (\forcode{ln_zdfric = .true.})] 
     77{Richardson number dependent (\protect\np{ln\_zdfric}\forcode{ = .true.})} 
    7578\label{subsec:ZDF_ric} 
    7679 
     
    8083%-------------------------------------------------------------------------------------------------------------- 
    8184 
    82 When \key{zdfric} is defined, a local Richardson number dependent formulation for the vertical momentum and 
     85When \np{ln\_zdfric}\forcode{ = .true.}, a local Richardson number dependent formulation for the vertical momentum and 
    8386tracer eddy coefficients is set through the \ngn{namzdf\_ric} namelist variables. 
    8487The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.  
     
    8790a dependency between the vertical eddy coefficients and the local Richardson number 
    8891(\ie the ratio of stratification to vertical shear). 
    89 Following \citet{Pacanowski_Philander_JPO81}, the following formulation has been implemented: 
     92Following \citet{pacanowski.philander_JPO81}, the following formulation has been implemented: 
    9093\[ 
    9194  % \label{eq:zdfric} 
     
    124127The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}. 
    125128Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to 
    126 the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{Lermusiaux2001}. 
     129the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{lermusiaux_JMS01}. 
    127130 
    128131% ------------------------------------------------------------------------------------------------------------- 
    129132%        TKE Turbulent Closure Scheme  
    130133% ------------------------------------------------------------------------------------------------------------- 
    131 \subsection{TKE turbulent closure scheme (\protect\key{zdftke})} 
     134\subsection[TKE turbulent closure scheme (\forcode{ln_zdftke = .true.})] 
     135{TKE turbulent closure scheme (\protect\np{ln\_zdftke}\forcode{ = .true.})} 
    132136\label{subsec:ZDF_tke} 
    133  
    134137%--------------------------------------------namzdf_tke-------------------------------------------------- 
    135138 
     
    140143a prognostic equation for $\bar{e}$, the turbulent kinetic energy, 
    141144and a closure assumption for the turbulent length scales. 
    142 This turbulent closure model has been developed by \citet{Bougeault1989} in the atmospheric case, 
    143 adapted by \citet{Gaspar1990} for the oceanic case, and embedded in OPA, the ancestor of NEMO, 
    144 by \citet{Blanke1993} for equatorial Atlantic simulations. 
    145 Since then, significant modifications have been introduced by \citet{Madec1998} in both the implementation and 
     145This turbulent closure model has been developed by \citet{bougeault.lacarrere_MWR89} in the atmospheric case, 
     146adapted by \citet{gaspar.gregoris.ea_JGR90} for the oceanic case, and embedded in OPA, the ancestor of NEMO, 
     147by \citet{blanke.delecluse_JPO93} for equatorial Atlantic simulations. 
     148Since then, significant modifications have been introduced by \citet{madec.delecluse.ea_NPM98} in both the implementation and 
    146149the formulation of the mixing length scale. 
    147150The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear, 
    148 its destruction through stratification, its vertical diffusion, and its dissipation of \citet{Kolmogorov1942} type: 
     151its destruction through stratification, its vertical diffusion, and its dissipation of \citet{kolmogorov_IANS42} type: 
    149152\begin{equation} 
    150153  \label{eq:zdftke_e} 
     
    168171$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity and diffusivity coefficients. 
    169172The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with 
    170 vertical mixing at any depth \citep{Gaspar1990}.  
     173vertical mixing at any depth \citep{gaspar.gregoris.ea_JGR90}.  
    171174They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}. 
    172 $P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function of the local Richardson number, $R_i$: 
     175$P_{rt}$ can be set to unity or, following \citet{blanke.delecluse_JPO93}, be a function of the local Richardson number, $R_i$: 
    173176\begin{align*} 
    174177  % \label{eq:prt} 
     
    180183  \end{cases} 
    181184\end{align*} 
    182 Options are defined through the  \ngn{namzdfy\_tke} namelist variables. 
    183185The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable. 
    184186 
    185187At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as 
    186188$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} namelist parameter. 
    187 The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}), however a much larger value can be used when 
     189The default value of $e_{bb}$ is 3.75. \citep{gaspar.gregoris.ea_JGR90}), however a much larger value can be used when 
    188190taking into account the surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}). 
    189191The bottom value of TKE is assumed to be equal to the value of the level just above. 
     
    191193the numerical scheme does not ensure its positivity. 
    192194To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} namelist parameter). 
    193 Following \citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
    194 This allows the subsequent formulations to match that of \citet{Gargett1984} for the diffusion in 
     195Following \citet{gaspar.gregoris.ea_JGR90}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
     196This allows the subsequent formulations to match that of \citet{gargett_JMR84} for the diffusion in 
    195197the thermocline and deep ocean :  $K_\rho = 10^{-3} / N$. 
    196198In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with 
    197199too weak vertical diffusion. 
    198200They must be specified at least larger than the molecular values, and are set through \np{rn\_avm0} and 
    199 \np{rn\_avt0} (namzdf namelist, see \autoref{subsec:ZDF_cst}). 
     201\np{rn\_avt0} (\ngn{namzdf} namelist, see \autoref{subsec:ZDF_cst}). 
    200202 
    201203\subsubsection{Turbulent length scale} 
    202204 
    203205For computational efficiency, the original formulation of the turbulent length scales proposed by 
    204 \citet{Gaspar1990} has been simplified. 
     206\citet{gaspar.gregoris.ea_JGR90} has been simplified. 
    205207Four formulations are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist parameter. 
    206 The first two are based on the following first order approximation \citep{Blanke1993}: 
     208The first two are based on the following first order approximation \citep{blanke.delecluse_JPO93}: 
    207209\begin{equation} 
    208210  \label{eq:tke_mxl0_1} 
     
    212214The resulting length scale is bounded by the distance to the surface or to the bottom 
    213215(\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor (\np{nn\_mxl}\forcode{ = 1}). 
    214 \citet{Blanke1993} notice that this simplification has two major drawbacks: 
     216\citet{blanke.delecluse_JPO93} notice that this simplification has two major drawbacks: 
    215217it makes no sense for locally unstable stratification and the computation no longer uses all 
    216218the information contained in the vertical density profile. 
    217 To overcome these drawbacks, \citet{Madec1998} introduces the \np{nn\_mxl}\forcode{ = 2..3} cases, 
     219To overcome these drawbacks, \citet{madec.delecluse.ea_NPM98} introduces the \np{nn\_mxl}\forcode{ = 2, 3} cases, 
    218220which add an extra assumption concerning the vertical gradient of the computed length scale. 
    219221So, the length scales are first evaluated as in \autoref{eq:tke_mxl0_1} and then bounded such that: 
     
    225227\autoref{eq:tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
    226228the variations of depth. 
    227 It provides a better approximation of the \citet{Gaspar1990} formulation while being much less  
     229It provides a better approximation of the \citet{gaspar.gregoris.ea_JGR90} formulation while being much less  
    228230time consuming. 
    229231In particular, it allows the length scale to be limited not only by the distance to the surface or 
     
    237239\begin{figure}[!t] 
    238240  \begin{center} 
    239     \includegraphics[width=1.00\textwidth]{Fig_mixing_length} 
     241    \includegraphics[width=\textwidth]{Fig_mixing_length} 
    240242    \caption{ 
    241243      \protect\label{fig:mixing_length} 
     
    258260In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same value: 
    259261$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nn\_mxl}\forcode{ = 3} case, 
    260 the dissipation and mixing turbulent length scales are give as in \citet{Gaspar1990}: 
     262the dissipation and mixing turbulent length scales are give as in \citet{gaspar.gregoris.ea_JGR90}: 
    261263\[ 
    262264  % \label{eq:tke_mxl_gaspar} 
     
    270272Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and 
    271273$z_o$ the roughness parameter of the surface. 
    272 Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}. 
     274Assuming $z_o=0.1$~m \citep{craig.banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}. 
    273275In the ocean interior a minimum length scale is set to recover the molecular viscosity when 
    274276$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ). 
     
    277279%-----------------------------------------------------------------------% 
    278280 
    279 Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified to 
     281Following \citet{mellor.blumberg_JPO04}, the TKE turbulence closure model has been modified to 
    280282include the effect of surface wave breaking energetics. 
    281283This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow. 
    282 The \citet{Mellor_Blumberg_JPO04} modifications acts on surface length scale and TKE values and 
     284The \citet{mellor.blumberg_JPO04} modifications acts on surface length scale and TKE values and 
    283285air-sea drag coefficient.  
    284 The latter concerns the bulk formulea and is not discussed here.  
    285  
    286 Following \citet{Craig_Banner_JPO94}, the boundary condition on surface TKE value is : 
     286The latter concerns the bulk formulae and is not discussed here.  
     287 
     288Following \citet{craig.banner_JPO94}, the boundary condition on surface TKE value is : 
    287289\begin{equation} 
    288290  \label{eq:ZDF_Esbc} 
    289291  \bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o} 
    290292\end{equation} 
    291 where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality which depends on the ''wave age'', 
    292 ranging from 57 for mature waves to 146 for younger waves \citep{Mellor_Blumberg_JPO04}.  
     293where $\alpha_{CB}$ is the \citet{craig.banner_JPO94} constant of proportionality which depends on the ''wave age'', 
     294ranging from 57 for mature waves to 146 for younger waves \citep{mellor.blumberg_JPO04}.  
    293295The boundary condition on the turbulent length scale follows the Charnock's relation: 
    294296\begin{equation} 
     
    297299\end{equation} 
    298300where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant. 
    299 \citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by 
    300 \citet{Stacey_JPO99} citing observation evidence, and 
     301\citet{mellor.blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by 
     302\citet{stacey_JPO99} citing observation evidence, and 
    301303$\alpha_{CB} = 100$ the Craig and Banner's value. 
    302304As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$, 
    303305with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}\forcode{ = 67.83} corresponds  
    304306to $\alpha_{CB} = 100$. 
    305 Further setting  \np{ln\_mxl0} to true applies \autoref{eq:ZDF_Lsbc} as surface boundary condition on length scale, 
     307Further setting  \np{ln\_mxl0=.true.},  applies \autoref{eq:ZDF_Lsbc} as the surface boundary condition on the length scale, 
    306308with $\beta$ hard coded to the Stacey's value. 
    307 Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on 
     309Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on the  
    308310surface $\bar{e}$ value. 
    309311 
     
    315317Although LC have nothing to do with convection, the circulation pattern is rather similar to 
    316318so-called convective rolls in the atmospheric boundary layer. 
    317 The detailed physics behind LC is described in, for example, \citet{Craik_Leibovich_JFM76}. 
     319The detailed physics behind LC is described in, for example, \citet{craik.leibovich_JFM76}. 
    318320The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and 
    319321wind drift currents.  
    320322 
    321323Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by 
    322 \citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure. 
     324\citep{axell_JGR02} for a $k-\epsilon$ turbulent closure. 
    323325The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in 
    324 an extra source terms of TKE, $P_{LC}$. 
     326an extra source term of TKE, $P_{LC}$. 
    325327The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled by setting \np{ln\_lc} to 
    326 \forcode{.true.} in the namtke namelist. 
     328\forcode{.true.} in the \ngn{namzdf\_tke} namelist. 
    327329  
    328 By making an analogy with the characteristic convective velocity scale (\eg, \citet{D'Alessio_al_JPO98}), 
     330By making an analogy with the characteristic convective velocity scale (\eg, \citet{dalessio.abdella.ea_JPO98}), 
    329331$P_{LC}$ is assumed to be :  
    330332\[ 
     
    334336With no information about the wave field, $w_{LC}$ is assumed to be proportional to  
    335337the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module  
    336 \footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity may be expressed as 
     338\footnote{Following \citet{li.garrett_JMR93}, the surface Stoke drift velocity may be expressed as 
    337339  $u_s =  0.016 \,|U_{10m}|$. 
    338340  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of 
     
    350352  \end{cases} 
    351353\] 
    352 where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise to fit LES data. 
     354where $c_{LC} = 0.15$ has been chosen by \citep{axell_JGR02} as a good compromise to fit LES data. 
    353355The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second. 
    354356The value of $c_{LC}$ is set through the \np{rn\_lc} namelist parameter, 
    355 having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.  
     357having in mind that it should stay between 0.15 and 0.54 \citep{axell_JGR02}.  
    356358 
    357359The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations: 
    358 $H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by 
     360$H_{LC}$ is the depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by 
    359361converting its kinetic energy to potential energy, according to  
    360362\[ 
     
    368370produce mixed-layer depths that are too shallow during summer months and windy conditions. 
    369371This bias is particularly acute over the Southern Ocean. 
    370 To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{Rodgers_2014}.  
     372To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{rodgers.aumont.ea_B14}.  
    371373The parameterization is an empirical one, \ie not derived from theoretical considerations, 
    372374but rather is meant to account for observed processes that affect the density structure of  
     
    383385\end{equation} 
    384386where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that 
    385 penetrate in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of 
     387penetrates in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of 
    386388the penetration, and $f_i$ is the ice concentration 
    387 (no penetration if $f_i=1$, that is if the ocean is entirely covered by sea-ice). 
     389(no penetration if $f_i=1$, \ie if the ocean is entirely covered by sea-ice). 
    388390The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter. 
    389391The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0}) or 
     
    391393(\np{nn\_etau}\forcode{ = 1}).  
    392394 
    393 Note that two other option existe, \np{nn\_etau}\forcode{ = 2..3}. 
     395Note that two other option exist, \np{nn\_etau}\forcode{ = 2, 3}. 
    394396They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer, 
    395 or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrate the ocean.  
     397or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrates the ocean.  
    396398Those two options are obsolescent features introduced for test purposes. 
    397399They will be removed in the next release.  
     400 
     401% This should be explain better below what this rn_eice parameter is meant for: 
     402In presence of Sea Ice, the value of this mixing can be modulated by the \np{rn\_eice} namelist parameter. 
     403This parameter varies from \forcode{0} for no effect to \forcode{4} to suppress the TKE input into the ocean when Sea Ice concentration 
     404is greater than 25\%.  
    398405 
    399406% from Burchard et al OM 2008 :  
     
    406413 
    407414% ------------------------------------------------------------------------------------------------------------- 
    408 %        TKE discretization considerations 
    409 % ------------------------------------------------------------------------------------------------------------- 
    410 \subsection{TKE discretization considerations (\protect\key{zdftke})} 
     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\label{subsec:ZDF_gls} 
     420 
     421%--------------------------------------------namzdf_gls--------------------------------------------------------- 
     422 
     423\nlst{namzdf_gls} 
     424%-------------------------------------------------------------------------------------------------------------- 
     425 
     426The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
     427one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale, 
     428$\psi$ \citep{umlauf.burchard_JMR03, umlauf.burchard_CSR05}. 
     429This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
     430where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of 
     431well-known turbulent closures ($k$-$kl$ \citep{mellor.yamada_RG82}, $k$-$\epsilon$ \citep{rodi_JGR87}, 
     432$k$-$\omega$ \citep{wilcox_AJ88} among others \citep{umlauf.burchard_JMR03,kantha.carniel_JMR03}).  
     433The GLS scheme is given by the following set of equations: 
     434\begin{equation} 
     435  \label{eq:zdfgls_e} 
     436  \frac{\partial \bar{e}}{\partial t} = 
     437  \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     438      +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     439  -K_\rho \,N^2 
     440  +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right] 
     441  - \epsilon 
     442\end{equation} 
     443 
     444\[ 
     445  % \label{eq:zdfgls_psi} 
     446  \begin{split} 
     447    \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
     448      \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     449          +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     450      - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\ 
     451    &+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 } 
     452        \;\frac{\partial \psi}{\partial k}} \right]\; 
     453  \end{split} 
     454\] 
     455 
     456\[ 
     457  % \label{eq:zdfgls_kz} 
     458  \begin{split} 
     459    K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
     460    K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l 
     461  \end{split} 
     462\] 
     463 
     464\[ 
     465  % \label{eq:zdfgls_eps} 
     466  {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
     467\] 
     468where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and 
     469$\epsilon$ the dissipation rate.  
     470The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
     471the choice of the turbulence model. 
     472Four different turbulent models are pre-defined (\autoref{tab:GLS}). 
     473They are made available through the \np{nn\_clo} namelist parameter.  
     474 
     475%--------------------------------------------------TABLE-------------------------------------------------- 
     476\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\ngn{namzdf\_gls}. 
     500    } 
     501  \end{center} 
     502\end{table} 
     503%-------------------------------------------------------------------------------------------------------------- 
     504 
     505In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
     506the mixing length towards $\kappa z_b$ ($\kappa$ is the Von Karman constant and $z_b$ the rugosity length scale) value near physical boundaries 
     507(logarithmic boundary layer law). 
     508$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{galperin.kantha.ea_JAS88}, 
     509or 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.).   
     511The value of $C_{0\mu}$ depends on the choice of the stability function. 
     512 
     513The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
     514Neumann condition through \np{nn\_bc\_surf} and \np{nn\_bc\_bot}, resp. 
     515As 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}. 
     517The \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}.  
     519 
     520The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
     521almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy. 
     522With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. 
     523A value of $c_{lim} = 0.53$ is often used \citep{galperin.kantha.ea_JAS88}. 
     524\cite{umlauf.burchard_CSR05} show that the value of the clipping factor is of crucial importance for 
     525the entrainment depth predicted in stably stratified situations, 
     526and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
     527The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.}, 
     528and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
     529 
     530The time and space discretization of the GLS equations follows the same energetic consideration as for 
     531the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{burchard_OM02}. 
     532Evaluation of the 4 GLS turbulent closure schemes can be found in \citet{warner.sherwood.ea_OM05} in ROMS model and 
     533 in \citet{reffray.guillaume.ea_GMD15} for the \NEMO model. 
     534 
     535 
     536% ------------------------------------------------------------------------------------------------------------- 
     537%        OSM OSMOSIS BL Scheme  
     538% ------------------------------------------------------------------------------------------------------------- 
     539\subsection[OSM: OSMosis boundary layer scheme (\forcode{ln_zdfosm = .true.})] 
     540{OSM: OSMosis boundary layer scheme (\protect\np{ln\_zdfosm}\forcode{ = .true.})} 
     541\label{subsec:ZDF_osm} 
     542%--------------------------------------------namzdf_osm--------------------------------------------------------- 
     543 
     544\nlst{namzdf_osm} 
     545%-------------------------------------------------------------------------------------------------------------- 
     546 
     547The 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} 
    411554\label{subsec:ZDF_tke_ene} 
    412555 
     
    414557\begin{figure}[!t] 
    415558  \begin{center} 
    416     \includegraphics[width=1.00\textwidth]{Fig_ZDF_TKE_time_scheme} 
     559    \includegraphics[width=\textwidth]{Fig_ZDF_TKE_time_scheme} 
    417560    \caption{ 
    418561      \protect\label{fig:TKE_time_scheme} 
    419       Illustration of the TKE time integration and its links to the momentum and tracer time integration. 
     562      Illustration of the subgrid kinetic energy integration in GLS and TKE schemes and its links to the momentum and tracer time integration. 
    420563    } 
    421564  \end{center}   
     
    424567 
    425568The production of turbulence by vertical shear (the first term of the right hand side of 
    426 \autoref{eq:zdftke_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
     569\autoref{eq:zdftke_e}) and  \autoref{eq:zdfgls_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
    427570(first line in \autoref{eq:PE_zdf}). 
    428 To do so a special care have to be taken for both the time and space discretization of 
    429 the TKE equation \citep{Burchard_OM02,Marsaleix_al_OM08}. 
     571To do so a special care has to be taken for both the time and space discretization of 
     572the kinetic energy equation \citep{burchard_OM02,marsaleix.auclair.ea_OM08}. 
    430573 
    431574Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows how 
    432575the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with 
    433 the one-level forward time stepping of TKE equation. 
     576the one-level forward time stepping of the equation for $\bar{e}$. 
    434577With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to 
    435578the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and 
     
    456599 
    457600A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification 
    458 (second term of the right hand side of \autoref{eq:zdftke_e}). 
     601(second term of the right hand side of \autoref{eq:zdftke_e} and \autoref{eq:zdfgls_e}). 
    459602This term must balance the input of potential energy resulting from vertical mixing. 
    460 The rate of change of potential energy (in 1D for the demonstration) due vertical mixing is obtained by 
    461 multiplying vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
     603The rate of change of potential energy (in 1D for the demonstration) due to vertical mixing is obtained by 
     604multiplying the vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
    462605\begin{equation} 
    463606  \label{eq:energ2} 
     
    475618The second term is minus the destruction rate of  $\bar{e}$ due to stratification. 
    476619Therefore \autoref{eq:energ1} implies that, to be energetically consistent, 
    477 the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e}, the TKE equation. 
     620the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e} and  \autoref{eq:zdfgls_e}. 
    478621 
    479622Let us now address the space discretization issue. 
     
    483626By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
    484627the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
    485 Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into account. 
     628Furthermore, the time variation of $e_3$ has be taken into account. 
    486629 
    487630The above energetic considerations leads to the following final discrete form for the TKE equation: 
     
    507650are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}). 
    508651Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible. 
    509 The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as 
    510 they all appear in the right hand side of \autoref{eq:zdftke_ene}. 
    511 For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
    512  
    513 % ------------------------------------------------------------------------------------------------------------- 
    514 %        GLS Generic Length Scale Scheme  
    515 % ------------------------------------------------------------------------------------------------------------- 
    516 \subsection{GLS: Generic Length Scale (\protect\key{zdfgls})} 
    517 \label{subsec:ZDF_gls} 
    518  
    519 %--------------------------------------------namzdf_gls--------------------------------------------------------- 
    520  
    521 \nlst{namzdf_gls} 
    522 %-------------------------------------------------------------------------------------------------------------- 
    523  
    524 The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
    525 one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale, 
    526 $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}. 
    527 This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
    528 where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of 
    529 well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982}, $k$-$\epsilon$ \citep{Rodi_1987}, 
    530 $k$-$\omega$ \citep{Wilcox_1988} among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).  
    531 The GLS scheme is given by the following set of equations: 
    532 \begin{equation} 
    533   \label{eq:zdfgls_e} 
    534   \frac{\partial \bar{e}}{\partial t} = 
    535   \frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
    536       +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
    537   -K_\rho \,N^2 
    538   +\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right] 
    539   - \epsilon 
    540 \end{equation} 
    541  
    542 \[ 
    543   % \label{eq:zdfgls_psi} 
    544   \begin{split} 
    545     \frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
    546       \frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
    547           +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
    548       - C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\ 
    549     &+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 } 
    550         \;\frac{\partial \psi}{\partial k}} \right]\; 
    551   \end{split} 
    552 \] 
    553  
    554 \[ 
    555   % \label{eq:zdfgls_kz} 
    556   \begin{split} 
    557     K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
    558     K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l 
    559   \end{split} 
    560 \] 
    561  
    562 \[ 
    563   % \label{eq:zdfgls_eps} 
    564   {\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
    565 \] 
    566 where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and 
    567 $\epsilon$ the dissipation rate.  
    568 The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
    569 the choice of the turbulence model. 
    570 Four different turbulent models are pre-defined (Tab.\autoref{tab:GLS}). 
    571 They are made available through the \np{nn\_clo} namelist parameter.  
    572  
    573 %--------------------------------------------------TABLE-------------------------------------------------- 
    574 \begin{table}[htbp] 
    575   \begin{center} 
    576     % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
    577     \begin{tabular}{ccccc} 
    578       &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\ 
    579       % & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\ 
    580       \hline 
    581       \hline 
    582       \np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\ 
    583       \hline 
    584       $( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
    585       $\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
    586       $\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
    587       $C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
    588       $C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
    589       $C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
    590       $F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
    591       \hline 
    592       \hline 
    593     \end{tabular} 
    594     \caption{ 
    595       \protect\label{tab:GLS} 
    596       Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
    597       \protect\key{zdfgls} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\ngn{namzdf\_gls}. 
    598     } 
    599   \end{center} 
    600 \end{table} 
    601 %-------------------------------------------------------------------------------------------------------------- 
    602  
    603 In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
    604 the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length) value near physical boundaries 
    605 (logarithmic boundary layer law). 
    606 $C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{Galperin_al_JAS88}, 
    607 or by \citet{Kantha_Clayson_1994} or one of the two functions suggested by \citet{Canuto_2001} 
    608 (\np{nn\_stab\_func}\forcode{ = 0..3}, resp.).  
    609 The value of $C_{0\mu}$ depends of the choice of the stability function. 
    610  
    611 The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
    612 Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp. 
    613 As for TKE closure, the wave effect on the mixing is considered when 
    614 \np{ln\_crban}\forcode{ = .true.} \citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. 
    615 The \np{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
    616 \np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.  
    617  
    618 The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
    619 almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy. 
    620 With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. 
    621 A value of $c_{lim} = 0.53$ is often used \citep{Galperin_al_JAS88}. 
    622 \cite{Umlauf_Burchard_CSR05} show that the value of the clipping factor is of crucial importance for 
    623 the entrainment depth predicted in stably stratified situations, 
    624 and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
    625 The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.}, 
    626 and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
    627  
    628 The time and space discretization of the GLS equations follows the same energetic consideration as for 
    629 the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{Burchard_OM02}. 
    630 Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}. 
    631  
    632 % ------------------------------------------------------------------------------------------------------------- 
    633 %        OSM OSMOSIS BL Scheme  
    634 % ------------------------------------------------------------------------------------------------------------- 
    635 \subsection{OSM: OSMOSIS boundary layer scheme (\protect\key{zdfosm})} 
    636 \label{subsec:ZDF_osm} 
    637  
    638 %--------------------------------------------namzdf_osm--------------------------------------------------------- 
    639  
    640 \nlst{namzdf_osm} 
    641 %-------------------------------------------------------------------------------------------------------------- 
    642  
    643 The OSMOSIS turbulent closure scheme is based on......   TBC 
     652%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}. 
     654%For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
    644655 
    645656% ================================================================ 
     
    648659\section{Convection} 
    649660\label{sec:ZDF_conv} 
    650  
    651 %--------------------------------------------namzdf-------------------------------------------------------- 
    652  
    653 \nlst{namzdf} 
    654 %-------------------------------------------------------------------------------------------------------------- 
    655661 
    656662Static instabilities (\ie light potential densities under heavy ones) may occur at particular ocean grid points. 
     
    664670%       Non-Penetrative Convective Adjustment  
    665671% ------------------------------------------------------------------------------------------------------------- 
    666 \subsection[Non-penetrative convective adjmt (\protect\np{ln\_tranpc}\forcode{ = .true.})] 
    667             {Non-penetrative convective adjustment (\protect\np{ln\_tranpc}\forcode{ = .true.})} 
     672\subsection[Non-penetrative convective adjustment (\forcode{ln_tranpc = .true.})] 
     673{Non-penetrative convective adjustment (\protect\np{ln\_tranpc}\forcode{ = .true.})} 
    668674\label{subsec:ZDF_npc} 
    669  
    670 %--------------------------------------------namzdf-------------------------------------------------------- 
    671  
    672 \nlst{namzdf} 
    673 %-------------------------------------------------------------------------------------------------------------- 
    674675 
    675676%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    676677\begin{figure}[!htb] 
    677678  \begin{center} 
    678     \includegraphics[width=0.90\textwidth]{Fig_npc} 
     679    \includegraphics[width=\textwidth]{Fig_npc} 
    679680    \caption{ 
    680681      \protect\label{fig:npc} 
     
    700701the water column, but only until the density structure becomes neutrally stable 
    701702(\ie until the mixed portion of the water column has \textit{exactly} the density of the water just below) 
    702 \citep{Madec_al_JPO91}. 
     703\citep{madec.delecluse.ea_JPO91}. 
    703704The associated algorithm is an iterative process used in the following way (\autoref{fig:npc}): 
    704705starting from the top of the ocean, the first instability is found. 
     
    718719the algorithm used in \NEMO converges for any profile in a number of iterations which is less than 
    719720the number of vertical levels. 
    720 This property is of paramount importance as pointed out by \citet{Killworth1989}: 
     721This property is of paramount importance as pointed out by \citet{killworth_iprc89}: 
    721722it avoids the existence of permanent and unrealistic static instabilities at the sea surface. 
    722723This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in 
    723 the north-western Mediterranean Sea \citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}. 
     724the north-western Mediterranean Sea \citep{madec.delecluse.ea_JPO91, madec.chartier.ea_DAO91, madec.crepon_iprc91}. 
    724725 
    725726The current implementation has been modified in order to deal with any non linear equation of seawater 
     
    727728Two main differences have been introduced compared to the original algorithm: 
    728729$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency  
    729 (not the the difference in potential density);  
     730(not the difference in potential density);  
    730731$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in 
    731732the same way their temperature and salinity has been mixed. 
     
    736737%       Enhanced Vertical Diffusion  
    737738% ------------------------------------------------------------------------------------------------------------- 
    738 \subsection{Enhanced vertical diffusion (\protect\np{ln\_zdfevd}\forcode{ = .true.})} 
     739\subsection[Enhanced vertical diffusion (\forcode{ln_zdfevd = .true.})] 
     740{Enhanced vertical diffusion (\protect\np{ln\_zdfevd}\forcode{ = .true.})} 
    739741\label{subsec:ZDF_evd} 
    740  
    741 %--------------------------------------------namzdf-------------------------------------------------------- 
    742  
    743 \nlst{namzdf} 
    744 %-------------------------------------------------------------------------------------------------------------- 
    745742 
    746743Options are defined through the  \ngn{namzdf} namelist variables. 
    747744The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}. 
    748 In this case, the vertical eddy mixing coefficients are assigned very large values 
    749 (a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable 
    750 (\ie when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar_PhD97, Lazar_al_JPO99}. 
     745In this case, the vertical eddy mixing coefficients are assigned very large values  
     746in regions where the stratification is unstable 
     747(\ie when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{lazar_phd97, lazar.madec.ea_JPO99}. 
    751748This is done either on tracers only (\np{nn\_evdm}\forcode{ = 0}) or 
    752749on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}). 
     
    759756the convective adjustment algorithm presented above when mixing both tracers and 
    760757momentum in the case of static instabilities. 
    761 It requires the use of an implicit time stepping on vertical diffusion terms 
    762 (\ie np{ln\_zdfexp}\forcode{ = .false.}). 
    763758 
    764759Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$. 
    765760This removes a potential source of divergence of odd and even time step in 
    766 a leapfrog environment \citep{Leclair_PhD2010} (see \autoref{sec:STP_mLF}). 
     761a leapfrog environment \citep{leclair_phd10} (see \autoref{sec:STP_mLF}). 
    767762 
    768763% ------------------------------------------------------------------------------------------------------------- 
    769764%       Turbulent Closure Scheme  
    770765% ------------------------------------------------------------------------------------------------------------- 
    771 \subsection[Turbulent closure scheme (\protect\key{zdf}\{tke,gls,osm\})]{Turbulent Closure Scheme (\protect\key{zdftke}, \protect\key{zdfgls} or \protect\key{zdfosm})} 
     766\subsection[Handling convection with turbulent closure schemes (\forcode{ln_zdf/tke/gls/osm = .true.})] 
     767{Handling convection with turbulent closure schemes (\protect\np{ln\_zdf/tke/gls/osm}\forcode{ = .true.})} 
    772768\label{subsec:ZDF_tcs} 
    773769 
    774 The turbulent closure scheme presented in \autoref{subsec:ZDF_tke} and \autoref{subsec:ZDF_gls} 
    775 (\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically unstable density profiles. 
     770 
     771The turbulent closure schemes presented in \autoref{subsec:ZDF_tke}, \autoref{subsec:ZDF_gls} and 
     772\autoref{subsec:ZDF_osm} (\ie \np{ln\_zdftke} or \np{ln\_zdfgls} or \np{ln\_zdfosm} defined) deal, in theory,  
     773with statically unstable density profiles. 
    776774In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in 
    777775\autoref{eq:zdftke_e} or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative.  
    778 It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring $A_u^{vm} {and}\;A_v^{vm}$ 
    779 (up to $1\;m^2s^{-1}$). 
     776It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also of the four neighboring values at  
     777velocity points $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1}$). 
    780778These large values restore the static stability of the water column in a way similar to that of 
    781779the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}). 
     
    785783It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme, 
    786784\ie setting the \np{ln\_zdfnpc} namelist parameter to true and 
    787 defining the turbulent closure CPP key all together. 
    788  
    789 The KPP turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
    790 as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
    791 therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the KPP scheme. 
     785defining the turbulent closure (\np{ln\_zdftke} or \np{ln\_zdfgls} = \forcode{.true.}) all together. 
     786 
     787The OSMOSIS turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
     788%as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
     789therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the OSMOSIS scheme. 
    792790% gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
    793791 
     
    795793% Double Diffusion Mixing 
    796794% ================================================================ 
    797 \section{Double diffusion mixing (\protect\key{zdfddm})} 
    798 \label{sec:ZDF_ddm} 
     795\section[Double diffusion mixing (\forcode{ln_zdfddm = .true.})] 
     796{Double diffusion mixing (\protect\np{ln\_zdfddm}\forcode{ = .true.})} 
     797\label{subsec:ZDF_ddm} 
     798 
    799799 
    800800%-------------------------------------------namzdf_ddm------------------------------------------------- 
     
    803803%-------------------------------------------------------------------------------------------------------------- 
    804804 
    805 Options are defined through the  \ngn{namzdf\_ddm} namelist variables. 
     805This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the namelist parameter 
     806\np{ln\_zdfddm} in \ngn{namzdf}. 
    806807Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa. 
    807808The former condition leads to salt fingering and the latter to diffusive convection. 
    808809Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean. 
    809 \citet{Merryfield1999} include a parameterisation of such phenomena in a global ocean model and show that  
     810\citet{merryfield.holloway.ea_JPO99} include a parameterisation of such phenomena in a global ocean model and show that  
    810811it leads to relatively minor changes in circulation but exerts significant regional influences on 
    811812temperature and salinity. 
    812 This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key. 
     813 
    813814 
    814815Diapycnal mixing of S and T are described by diapycnal diffusion coefficients  
     
    839840\begin{figure}[!t] 
    840841  \begin{center} 
    841     \includegraphics[width=0.99\textwidth]{Fig_zdfddm} 
     842    \includegraphics[width=\textwidth]{Fig_zdfddm} 
    842843    \caption{ 
    843844      \protect\label{fig:zdfddm} 
    844       From \citet{Merryfield1999} : 
     845      From \citet{merryfield.holloway.ea_JPO99} : 
    845846      (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in regions of salt fingering. 
    846847      Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$; 
     
    855856 
    856857The factor 0.7 in \autoref{eq:zdfddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of 
    857 buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{McDougall_Taylor_JMR84}). 
    858 Following  \citet{Merryfield1999}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
     858buoyancy flux of heat to buoyancy flux of salt (\eg, \citet{mcdougall.taylor_JMR84}). 
     859Following  \citet{merryfield.holloway.ea_JPO99}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
    859860 
    860861To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by 
     
    887888% Bottom Friction 
    888889% ================================================================ 
    889 \section{Bottom and top friction (\protect\mdl{zdfbfr})} 
    890 \label{sec:ZDF_bfr} 
     890 \section[Bottom and top friction (\textit{zdfdrg.F90})] 
     891 {Bottom and top friction (\protect\mdl{zdfdrg})} 
     892 \label{sec:ZDF_drg} 
    891893 
    892894%--------------------------------------------nambfr-------------------------------------------------------- 
    893895% 
    894 %\nlst{nambfr} 
     896\nlst{namdrg} 
     897\nlst{namdrg_top} 
     898\nlst{namdrg_bot} 
     899 
    895900%-------------------------------------------------------------------------------------------------------------- 
    896901 
    897 Options to define the top and bottom friction are defined through the \ngn{nambfr} namelist variables. 
     902Options to define the top and bottom friction are defined through the \ngn{namdrg} namelist variables. 
    898903The bottom friction represents the friction generated by the bathymetry. 
    899904The top friction represents the friction generated by the ice shelf/ocean interface. 
    900 As the friction processes at the top and bottom are treated in similar way, 
    901 only the bottom friction is described in detail below. 
     905As the friction processes at the top and the bottom are treated in and identical way,  
     906the description below considers mostly the bottom friction case, if not stated otherwise. 
    902907 
    903908 
     
    905910a condition on the vertical diffusive flux. 
    906911For the bottom boundary layer, one has: 
    907 \[ 
    908   % \label{eq:zdfbfr_flux} 
    909   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
    910 \] 
     912 \[ 
     913   % \label{eq:zdfbfr_flux} 
     914   A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
     915 \] 
    911916where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside 
    912917the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean). 
     
    922927To illustrate this, consider the equation for $u$ at $k$, the last ocean level: 
    923928\begin{equation} 
    924   \label{eq:zdfbfr_flux2} 
     929  \label{eq:zdfdrg_flux2} 
    925930  \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}} 
    926931\end{equation} 
     
    935940 
    936941In the code, the bottom friction is imposed by adding the trend due to the bottom friction to 
    937 the general momentum trend in \mdl{dynbfr}. 
     942 the general momentum trend in \mdl{dynzdf}. 
    938943For the time-split surface pressure gradient algorithm, the momentum trend due to 
    939944the barotropic component needs to be handled separately. 
    940945For this purpose it is convenient to compute and store coefficients which can be simply combined with 
    941946bottom velocities and geometric values to provide the momentum trend due to bottom friction. 
    942 These coefficients are computed in \mdl{zdfbfr} and generally take the form $c_b^{\textbf U}$ where: 
     947 These coefficients are computed in \mdl{zdfdrg} and generally take the form $c_b^{\textbf U}$ where: 
    943948\begin{equation} 
    944949  \label{eq:zdfbfr_bdef} 
     
    946951  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b 
    947952\end{equation} 
    948 where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity. 
     953where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity.  
     954Note 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. 
    949955 
    950956% ------------------------------------------------------------------------------------------------------------- 
    951957%       Linear Bottom Friction 
    952958% ------------------------------------------------------------------------------------------------------------- 
    953 \subsection{Linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 0..1})} 
    954 \label{subsec:ZDF_bfr_linear} 
    955  
    956 The linear bottom friction parameterisation (including the special case of a free-slip condition) assumes that 
    957 the bottom friction is proportional to the interior velocity (\ie the velocity of the last model level): 
     959 \subsection[Linear top/bottom friction (\forcode{ln_lin = .true.})] 
     960 {Linear top/bottom friction (\protect\np{ln\_lin}\forcode{ = .true.)}} 
     961 \label{subsec:ZDF_drg_linear} 
     962 
     963The linear friction parameterisation (including the special case of a free-slip condition) assumes that 
     964the friction is proportional to the interior velocity (\ie the velocity of the first/last model level): 
    958965\[ 
    959966  % \label{eq:zdfbfr_linear} 
    960967  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
    961968\] 
    962 where $r$ is a friction coefficient expressed in ms$^{-1}$. 
     969where $r$ is a friction coefficient expressed in $m s^{-1}$. 
    963970This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,  
    964971and setting $r = H / \tau$, where $H$ is the ocean depth. 
    965 Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}. 
     972Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{weatherly_JMR84}. 
    966973A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models. 
    967974One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ 
    968 (\citet{Gill1982}, Eq. 9.6.6). 
     975(\citet{gill_bk82}, Eq. 9.6.6). 
    969976For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, 
    970977and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$. 
    971978This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days. 
    972 It can be changed by specifying \np{rn\_bfri1} (namelist parameter). 
    973  
    974 For the linear friction case the coefficients defined in the general expression \autoref{eq:zdfbfr_bdef} are:  
     979It can be changed by specifying \np{rn\_Uc0} (namelist parameter). 
     980 
     981 For the linear friction case the drag coefficient used in the general expression \autoref{eq:zdfbfr_bdef} is:  
    975982\[ 
    976983  % \label{eq:zdfbfr_linbfr_b} 
    977   \begin{split} 
    978     c_b^u &= - r\\ 
    979     c_b^v &= - r\\ 
    980   \end{split} 
    981 \] 
    982 When \np{nn\_botfr}\forcode{ = 1}, the value of $r$ used is \np{rn\_bfri1}. 
    983 Setting \np{nn\_botfr}\forcode{ = 0} is equivalent to setting $r=0$ and 
    984 leads to a free-slip bottom boundary condition. 
    985 These values are assigned in \mdl{zdfbfr}. 
    986 From v3.2 onwards there is support for local enhancement of these values via an externally defined 2D mask array 
    987 (\np{ln\_bfr2d}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file. 
     984    c_b^T = - r 
     985\] 
     986When \np{ln\_lin} \forcode{= .true.}, the value of $r$ used is \np{rn\_Uc0}*\np{rn\_Cd0}. 
     987Setting \np{ln\_OFF} \forcode{= .true.} (and \forcode{ln_lin = .true.}) is equivalent to setting $r=0$ and leads to a free-slip boundary condition. 
     988 
     989These values are assigned in \mdl{zdfdrg}. 
     990Note that there is support for local enhancement of these values via an externally defined 2D mask array 
     991(\np{ln\_boost}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file. 
    988992The mask values should vary from 0 to 1. 
    989993Locations with a non-zero mask value will have the friction coefficient increased by 
    990 $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri1}. 
     994$mask\_value$ * \np{rn\_boost} * \np{rn\_Cd0}. 
    991995 
    992996% ------------------------------------------------------------------------------------------------------------- 
    993997%       Non-Linear Bottom Friction 
    994998% ------------------------------------------------------------------------------------------------------------- 
    995 \subsection{Non-linear bottom friction (\protect\np{nn\_botfr}\forcode{ = 2})} 
    996 \label{subsec:ZDF_bfr_nonlinear} 
    997  
    998 The non-linear bottom friction parameterisation assumes that the bottom friction is quadratic:  
    999 \[ 
    1000   % \label{eq:zdfbfr_nonlinear} 
     999 \subsection[Non-linear top/bottom friction (\forcode{ln_non_lin = .true.})] 
     1000 {Non-linear top/bottom friction (\protect\np{ln\_non\_lin}\forcode{ = .true.})} 
     1001 \label{subsec:ZDF_drg_nonlinear} 
     1002 
     1003The non-linear bottom friction parameterisation assumes that the top/bottom friction is quadratic:  
     1004\[ 
     1005  % \label{eq:zdfdrg_nonlinear} 
    10011006  {\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h 
    10021007  }{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b 
    10031008\] 
    1004 where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy due to tides, 
     1009where $C_D$ is a drag coefficient, and $e_b $ a top/bottom turbulent kinetic energy due to tides, 
    10051010internal waves breaking and other short time scale currents. 
    10061011A typical value of the drag coefficient is $C_D = 10^{-3} $. 
    1007 As an example, the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and 
    1008 $e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} uses $C_D = 1.4\;10^{-3}$ and 
     1012As an example, the CME experiment \citep{treguier_JGR92} uses $C_D = 10^{-3}$ and 
     1013$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{killworth_JPO92} uses $C_D = 1.4\;10^{-3}$ and 
    10091014$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$. 
    1010 The CME choices have been set as default values (\np{rn\_bfri2} and \np{rn\_bfeb2} namelist parameters). 
    1011  
    1012 As for the linear case, the bottom friction is imposed in the code by adding the trend due to 
    1013 the bottom friction to the general momentum trend in \mdl{dynbfr}. 
    1014 For the non-linear friction case the terms computed in \mdl{zdfbfr} are: 
    1015 \[ 
    1016   % \label{eq:zdfbfr_nonlinbfr} 
    1017   \begin{split} 
    1018     c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\ 
    1019     c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\ 
    1020   \end{split} 
    1021 \] 
    1022  
    1023 The coefficients that control the strength of the non-linear bottom friction are initialised as namelist parameters: 
    1024 $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}. 
    1025 Note for applications which treat tides explicitly a low or even zero value of \np{rn\_bfeb2} is recommended. 
    1026 From v3.2 onwards a local enhancement of $C_D$ is possible via an externally defined 2D mask array 
    1027 (\np{ln\_bfr2d}\forcode{ = .true.}). 
    1028 This works in the same way as for the linear bottom friction case with non-zero masked locations increased by 
    1029 $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri2}. 
     1015The CME choices have been set as default values (\np{rn\_Cd0} and \np{rn\_ke0} namelist parameters). 
     1016 
     1017As for the linear case, the friction is imposed in the code by adding the trend due to 
     1018the friction to the general momentum trend in \mdl{dynzdf}. 
     1019For the non-linear friction case the term computed in \mdl{zdfdrg} is: 
     1020\[ 
     1021  % \label{eq:zdfdrg_nonlinbfr} 
     1022    c_b^T = - \; C_D\;\left[ \left(\bar{u_b}^{i}\right)^2 + \left(\bar{v_b}^{j}\right)^2 + e_b \right]^{1/2} 
     1023\] 
     1024 
     1025The coefficients that control the strength of the non-linear friction are initialised as namelist parameters: 
     1026$C_D$= \np{rn\_Cd0}, and $e_b$ =\np{rn\_bfeb2}. 
     1027Note 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 
     1028(\np{ln\_boost}\forcode{ = .true.}). 
     1029This works in the same way as for the linear friction case with non-zero masked locations increased by 
     1030$mask\_value$ * \np{rn\_boost} * \np{rn\_Cd0}. 
    10301031 
    10311032% ------------------------------------------------------------------------------------------------------------- 
    10321033%       Bottom Friction Log-layer 
    10331034% ------------------------------------------------------------------------------------------------------------- 
    1034 \subsection[Log-layer btm frict enhncmnt (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})] 
    1035             {Log-layer bottom friction enhancement (\protect\np{nn\_botfr}\forcode{ = 2}, \protect\np{ln\_loglayer}\forcode{ = .true.})} 
    1036 \label{subsec:ZDF_bfr_loglayer} 
    1037  
    1038 In the non-linear bottom friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
    1039 a "law of the wall" scaling. 
    1040 If  \np{ln\_loglayer} = .true., $C_D$ is no longer constant but is related to the thickness of 
    1041 the last wet layer in each column by: 
    1042 \[ 
    1043   C_D = \left ( {\kappa \over {\rm log}\left ( 0.5e_{3t}/rn\_bfrz0 \right ) } \right )^2 
    1044 \] 
    1045  
    1046 \noindent where $\kappa$ is the von-Karman constant and \np{rn\_bfrz0} is a roughness length provided via 
    1047 the namelist. 
    1048  
    1049 For stability, the drag coefficient is bounded such that it is kept greater or equal to 
    1050 the base \np{rn\_bfri2} value and it is not allowed to exceed the value of an additional namelist parameter: 
    1051 \np{rn\_bfri2\_max}, \ie 
    1052 \[ 
    1053   rn\_bfri2 \leq C_D \leq rn\_bfri2\_max 
    1054 \] 
    1055  
    1056 \noindent Note also that a log-layer enhancement can also be applied to the top boundary friction if 
    1057 under ice-shelf cavities are in use (\np{ln\_isfcav}\forcode{ = .true.}). 
    1058 In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}. 
    1059  
    1060 % ------------------------------------------------------------------------------------------------------------- 
    1061 %       Bottom Friction stability 
    1062 % ------------------------------------------------------------------------------------------------------------- 
    1063 \subsection{Bottom friction stability considerations} 
    1064 \label{subsec:ZDF_bfr_stability} 
    1065  
    1066 Some care needs to exercised over the choice of parameters to ensure that the implementation of 
    1067 bottom friction does not induce numerical instability. 
    1068 For the purposes of stability analysis, an approximation to \autoref{eq:zdfbfr_flux2} is: 
     1035 \subsection[Log-layer top/bottom friction (\forcode{ln_loglayer = .true.})] 
     1036 {Log-layer top/bottom friction (\protect\np{ln\_loglayer}\forcode{ = .true.})} 
     1037 \label{subsec:ZDF_drg_loglayer} 
     1038 
     1039In the non-linear friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
     1040a "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. 
     1041If  \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): 
     1042\[ 
     1043  C_D = \left ( {\kappa \over {\mathrm log}\left ( 0.5 \; e_{3b} / rn\_{z0} \right ) } \right )^2 
     1044\] 
     1045 
     1046\noindent where $\kappa$ is the von-Karman constant and \np{rn\_z0} is a roughness length provided via the namelist. 
     1047 
     1048The drag coefficient is bounded such that it is kept greater or equal to 
     1049the 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: 
     1050\np{rn\_Cdmax}, \ie 
     1051\[ 
     1052  rn\_Cd0 \leq C_D \leq rn\_Cdmax 
     1053\] 
     1054 
     1055\noindent The log-layer enhancement can also be applied to the top boundary friction if 
     1056under ice-shelf cavities are activated (\np{ln\_isfcav}\forcode{ = .true.}). 
     1057%In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}. 
     1058 
     1059% ------------------------------------------------------------------------------------------------------------- 
     1060%       Explicit bottom Friction 
     1061% ------------------------------------------------------------------------------------------------------------- 
     1062 \subsection{Explicit top/bottom friction (\forcode{ln_drgimp = .false.})} 
     1063 \label{subsec:ZDF_drg_stability} 
     1064 
     1065Setting \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: 
     1066 
     1067At the top (below an ice shelf cavity): 
     1068\[ 
     1069  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t} 
     1070  = c_{t}^{\textbf{U}}\textbf{u}^{n-1}_{t} 
     1071\] 
     1072 
     1073At the bottom (above the sea floor): 
     1074\[ 
     1075  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b} 
     1076  = c_{b}^{\textbf{U}}\textbf{u}^{n-1}_{b} 
     1077\] 
     1078 
     1079Since 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. 
     1080For the purposes of stability analysis, an approximation to \autoref{eq:zdfdrg_flux2} is: 
    10691081\begin{equation} 
    1070   \label{eq:Eqn_bfrstab} 
     1082  \label{eq:Eqn_drgstab} 
    10711083  \begin{split} 
    10721084    \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\ 
     
    10741086  \end{split} 
    10751087\end{equation} 
    1076 \noindent where linear bottom friction and a leapfrog timestep have been assumed. 
    1077 To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have: 
     1088\noindent where linear friction and a leapfrog timestep have been assumed. 
     1089To ensure that the friction cannot reverse the direction of flow it is necessary to have: 
    10781090\[ 
    10791091  |\Delta u| < \;|u|  
    10801092\] 
    1081 \noindent which, using \autoref{eq:Eqn_bfrstab}, gives: 
     1093\noindent which, using \autoref{eq:Eqn_drgstab}, gives: 
    10821094\[ 
    10831095  r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\ 
     
    10931105For most applications, with physically sensible parameters these restrictions should not be of concern. 
    10941106But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.  
    1095 To ensure stability limits are imposed on the bottom friction coefficients both 
     1107To ensure stability limits are imposed on the top/bottom friction coefficients both 
    10961108during initialisation and at each time step. 
    1097 Checks at initialisation are made in \mdl{zdfbfr} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
     1109Checks at initialisation are made in \mdl{zdfdrg} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
    10981110The number of breaches of the stability criterion are reported as well as 
    10991111the minimum and maximum values that have been set. 
    1100 The criterion is also checked at each time step, using the actual velocity, in \mdl{dynbfr}. 
    1101 Values of the bottom friction coefficient are reduced as necessary to ensure stability; 
     1112The criterion is also checked at each time step, using the actual velocity, in \mdl{dynzdf}. 
     1113Values of the friction coefficient are reduced as necessary to ensure stability; 
    11021114these changes are not reported. 
    11031115 
    1104 Limits on the bottom friction coefficient are not imposed if the user has elected to 
    1105 handle the bottom friction implicitly (see \autoref{subsec:ZDF_bfr_imp}). 
     1116Limits on the top/bottom friction coefficient are not imposed if the user has elected to 
     1117handle the friction implicitly (see \autoref{subsec:ZDF_drg_imp}). 
    11061118The number of potential breaches of the explicit stability criterion are still reported for information purposes. 
    11071119 
     
    11091121%       Implicit Bottom Friction 
    11101122% ------------------------------------------------------------------------------------------------------------- 
    1111 \subsection{Implicit bottom friction (\protect\np{ln\_bfrimp}\forcode{ = .true.})} 
    1112 \label{subsec:ZDF_bfr_imp} 
     1123 \subsection[Implicit top/bottom friction (\forcode{ln_drgimp = .true.})] 
     1124 {Implicit top/bottom friction (\protect\np{ln\_drgimp}\forcode{ = .true.})} 
     1125 \label{subsec:ZDF_drg_imp} 
    11131126 
    11141127An optional implicit form of bottom friction has been implemented to improve model stability. 
    1115 We recommend this option for shelf sea and coastal ocean applications, especially for split-explicit time splitting. 
    1116 This option can be invoked by setting \np{ln\_bfrimp} to \forcode{.true.} in the \textit{nambfr} namelist. 
    1117 This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \textit{namzdf} namelist.  
    1118  
    1119 This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, 
    1120 the bottom boundary condition is implemented implicitly. 
    1121  
    1122 \[ 
    1123   % \label{eq:dynzdf_bfr} 
    1124   \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{mbk} 
    1125   = \binom{c_{b}^{u}u^{n+1}_{mbk}}{c_{b}^{v}v^{n+1}_{mbk}} 
    1126 \] 
    1127  
    1128 where $mbk$ is the layer number of the bottom wet layer. 
    1129 Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so, it is implicit. 
    1130  
    1131 If split-explicit time splitting is used, care must be taken to avoid the double counting of the bottom friction in 
    1132 the 2-D barotropic momentum equations. 
    1133 As NEMO only updates the barotropic pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, 
    1134 we need to remove the bottom friction induced by these two terms which has been included in the 3-D momentum trend  
    1135 and update it with the latest value. 
    1136 On the other hand, the bottom friction contributed by the other terms 
    1137 (\eg the advection term, viscosity term) has been included in the 3-D momentum equations and 
    1138 should not be added in the 2-D barotropic mode. 
    1139  
    1140 The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the following: 
    1141  
    1142 \[ 
    1143   % \label{eq:dynspg_ts_bfr1} 
    1144   \frac{\textbf{U}_{med}-\textbf{U}^{m-1}}{2\Delta t}=-g\nabla\eta-f\textbf{k}\times\textbf{U}^{m}+c_{b} 
    1145   \left(\textbf{U}_{med}-\textbf{U}^{m-1}\right) 
    1146 \] 
    1147 \[ 
    1148   \frac{\textbf{U}^{m+1}-\textbf{U}_{med}}{2\Delta t}=\textbf{T}+ 
    1149   \left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{U}^{'}\right)- 
    1150   2\Delta t_{bc}c_{b}\left(g\nabla\eta^{'}+f\textbf{k}\times\textbf{u}_{b}\right) 
    1151 \] 
    1152  
    1153 where $\textbf{T}$ is the vertical integrated 3-D momentum trend. 
    1154 We assume the leap-frog time-stepping is used here. 
    1155 $\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step. 
    1156 $c_{b}$ is the friction coefficient. 
    1157 $\eta$ is the sea surface level calculated in the barotropic loops while $\eta^{'}$ is the sea surface level used in 
    1158 the 3-D baroclinic mode. 
    1159 $\textbf{u}_{b}$ is the bottom layer horizontal velocity. 
    1160  
    1161 % ------------------------------------------------------------------------------------------------------------- 
    1162 %       Bottom Friction with split-explicit time splitting 
    1163 % ------------------------------------------------------------------------------------------------------------- 
    1164 \subsection[Bottom friction w/ split-explicit time splitting (\protect\np{ln\_bfrimp})] 
    1165             {Bottom friction with split-explicit time splitting (\protect\np{ln\_bfrimp})} 
    1166 \label{subsec:ZDF_bfr_ts} 
    1167  
    1168 When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, 
    1169 the bottom velocity at the before time step is used. 
    1170 This velocity includes both the baroclinic and barotropic components which is appropriate when 
    1171 using either the explicit or filtered surface pressure gradient algorithms 
    1172 (\key{dynspg\_exp} or \key{dynspg\_flt}). 
    1173 Extra attention is required, however, when using split-explicit time stepping (\key{dynspg\_ts}). 
    1174 In this case the free surface equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, 
    1175 while the three dimensional prognostic variables are solved with the longer time step of \np{rn\_rdt} seconds. 
    1176 The trend in the barotropic momentum due to bottom friction appropriate to this method is that given by 
    1177 the selected parameterisation (\ie linear or non-linear bottom friction) computed with 
    1178 the evolving velocities at each barotropic timestep.  
    1179  
    1180 In the case of non-linear bottom friction, we have elected to partially linearise the problem by 
    1181 keeping the coefficients fixed throughout the barotropic time-stepping to those computed in 
    1182 \mdl{zdfbfr} using the now timestep. 
    1183 This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to: 
    1184  
     1128We recommend this option for shelf sea and coastal ocean applications. %, especially for split-explicit time splitting. 
     1129This option can be invoked by setting \np{ln\_drgimp} to \forcode{.true.} in the \textit{namdrg} namelist. 
     1130%This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \textit{namzdf} namelist.  
     1131 
     1132This implementation is performed in \mdl{dynzdf} where the following boundary conditions are set while solving the fully implicit diffusion step:  
     1133 
     1134At the top (below an ice shelf cavity): 
     1135\[ 
     1136  % \label{eq:dynzdf_drg_top} 
     1137  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{t} 
     1138  = c_{t}^{\textbf{U}}\textbf{u}^{n+1}_{t} 
     1139\] 
     1140 
     1141At the bottom (above the sea floor): 
     1142\[ 
     1143  % \label{eq:dynzdf_drg_bot} 
     1144  \left.{\left( {\frac{A^{vm} }{e_3 }\ \frac{\partial \textbf{U}_h}{\partial k}} \right)} \right|_{b} 
     1145  = c_{b}^{\textbf{U}}\textbf{u}^{n+1}_{b} 
     1146\] 
     1147 
     1148where $t$ and $b$ refers to top and bottom layers respectively.  
     1149Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so it is implicit. 
     1150 
     1151% ------------------------------------------------------------------------------------------------------------- 
     1152%       Bottom Friction with split-explicit free surface 
     1153% ------------------------------------------------------------------------------------------------------------- 
     1154 \subsection[Bottom friction with split-explicit free surface] 
     1155 {Bottom friction with split-explicit free surface} 
     1156 \label{subsec:ZDF_drg_ts} 
     1157 
     1158With 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.  
     1159 
     1160The strategy to handle top/bottom stresses with split-explicit free surface in \NEMO is as follows: 
    11851161\begin{enumerate} 
    1186 \item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before barotropic velocity to 
    1187   the bottom friction component of the vertically integrated momentum trend. 
    1188   Note the same stability check that is carried out on the bottom friction coefficient in \mdl{dynbfr} has to 
    1189   be applied here to ensure that the trend removed matches that which was added in \mdl{dynbfr}. 
    1190 \item At each barotropic step, compute the contribution of the current barotropic velocity to 
    1191   the trend due to bottom friction. 
    1192   Add this contribution to the vertically integrated momentum trend. 
    1193   This contribution is handled implicitly which eliminates the need to impose a stability criteria on 
    1194   the values of the bottom friction coefficient within the barotropic loop.  
    1195 \end{enumerate} 
    1196  
    1197 Note that the use of an implicit formulation within the barotropic loop for the bottom friction trend means that 
    1198 any limiting of the bottom friction coefficient in \mdl{dynbfr} does not adversely affect the solution when 
    1199 using split-explicit time splitting. 
    1200 This is because the major contribution to bottom friction is likely to come from the barotropic component which 
    1201 uses the unrestricted value of the coefficient. 
    1202 However, if the limiting is thought to be having a major effect 
    1203 (a more likely prospect in coastal and shelf seas applications) then 
    1204 the fully implicit form of the bottom friction should be used (see \autoref{subsec:ZDF_bfr_imp}) 
    1205 which can be selected by setting \np{ln\_bfrimp} $=$ \forcode{.true.}. 
    1206  
    1207 Otherwise, the implicit formulation takes the form: 
    1208 \[ 
    1209   % \label{eq:zdfbfr_implicitts} 
    1210   \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ] 
    1211 \] 
    1212 where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),  
    1213 $c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and 
    1214 $RHS$ represents all the components to the vertically integrated momentum trend except for 
    1215 that due to bottom friction. 
    1216  
    1217 % ================================================================ 
    1218 % Tidal Mixing 
    1219 % ================================================================ 
    1220 \section{Tidal mixing (\protect\key{zdftmx})} 
    1221 \label{sec:ZDF_tmx} 
    1222  
    1223 %--------------------------------------------namzdf_tmx-------------------------------------------------- 
     1162\item To extend the stability of the barotropic sub-stepping, bottom stresses are refreshed at each sub-iteration. The baroclinic part of the flow entering the stresses is frozen at the initial time of the barotropic iteration. In case of non-linear friction, the drag coefficient is also constant. 
     1163\item In case of an implicit drag, specific computations are performed in \mdl{dynzdf} which renders the overall scheme mixed explicit/implicit: the barotropic components of 3d velocities are removed before seeking for the implicit vertical diffusion result. Top/bottom stresses due to the barotropic components are explicitly accounted for thanks to the updated values of barotropic velocities. Then the implicit solution of 3d velocities is obtained. Lastly, the residual barotropic component is replaced by the time split estimate. 
     1164\end{enumerate}  
     1165 
     1166Note that other strategies are possible, like considering vertical diffusion step in advance, \ie prior barotropic integration.   
     1167 
     1168 
     1169% ================================================================ 
     1170% Internal wave-driven mixing 
     1171% ================================================================ 
     1172\section[Internal wave-driven mixing (\forcode{ln_zdfiwm = .true.})] 
     1173{Internal wave-driven mixing (\protect\np{ln\_zdfiwm}\forcode{ = .true.})} 
     1174\label{subsec:ZDF_tmx_new} 
     1175 
     1176%--------------------------------------------namzdf_iwm------------------------------------------ 
    12241177% 
    1225 %\nlst{namzdf_tmx} 
     1178\nlst{namzdf_iwm} 
    12261179%-------------------------------------------------------------------------------------------------------------- 
    12271180 
    1228  
    1229 % ------------------------------------------------------------------------------------------------------------- 
    1230 %        Bottom intensified tidal mixing  
    1231 % ------------------------------------------------------------------------------------------------------------- 
    1232 \subsection{Bottom intensified tidal mixing} 
    1233 \label{subsec:ZDF_tmx_bottom} 
    1234  
    1235 Options are defined through the  \ngn{namzdf\_tmx} namelist variables. 
    1236 The parameterization of tidal mixing follows the general formulation for the vertical eddy diffusivity proposed by 
    1237 \citet{St_Laurent_al_GRL02} and first introduced in an OGCM by \citep{Simmons_al_OM04}.  
    1238 In this formulation an additional vertical diffusivity resulting from internal tide breaking, 
    1239 $A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, 
    1240 the energy transfer from barotropic tides to baroclinic tides: 
    1241 \begin{equation} 
    1242   \label{eq:Ktides} 
    1243   A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 } 
    1244 \end{equation} 
    1245 where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}), 
    1246 $\rho$ the density, $q$ the tidal dissipation efficiency, and $F(z)$ the vertical structure function.  
    1247  
    1248 The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) and 
    1249 is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).  
    1250 The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter) 
    1251 represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally, 
    1252 with the remaining $1-q$ radiating away as low mode internal waves and 
    1253 contributing to the background internal wave field. 
    1254 A value of $q=1/3$ is typically used \citet{St_Laurent_al_GRL02}. 
    1255 The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical. 
    1256 It is implemented as a simple exponential decaying upward away from the bottom, 
    1257 with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, 
    1258 with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},  
    1259 \[ 
    1260   % \label{eq:Fz} 
    1261   F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) } 
    1262 \] 
    1263 and is normalized so that vertical integral over the water column is unity.  
    1264  
    1265 The associated vertical viscosity is calculated from the vertical diffusivity assuming a Prandtl number of 1, 
    1266 \ie $A^{vm}_{tides}=A^{vT}_{tides}$. 
    1267 In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity is capped at $300\,cm^2/s$ and 
    1268 impose a lower limit on $N^2$ of \np{rn\_n2min} usually set to $10^{-8} s^{-2}$. 
    1269 These bounds are usually rarely encountered. 
    1270  
    1271 The internal wave energy map, $E(x, y)$ in \autoref{eq:Ktides}, is derived from a barotropic model of 
    1272 the tides utilizing a parameterization of the conversion of barotropic tidal energy into internal waves. 
    1273 The essential goal of the parameterization is to represent the momentum exchange between the barotropic tides and 
    1274 the unrepresented internal waves induced by the tidal flow over rough topography in a stratified ocean. 
    1275 In the current version of \NEMO, the map is built from the output of 
    1276 the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}. 
    1277 This model provides the dissipation associated with internal wave energy for the M2 and K1 tides component 
    1278 (\autoref{fig:ZDF_M2_K1_tmx}). 
    1279 The S2 dissipation is simply approximated as being $1/4$ of the M2 one. 
    1280 The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$. 
    1281 Its global mean value is $1.1$ TW, 
    1282 in agreement with independent estimates \citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.  
    1283  
    1284 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    1285 \begin{figure}[!t] 
    1286   \begin{center} 
    1287     \includegraphics[width=0.90\textwidth]{Fig_ZDF_M2_K1_tmx} 
    1288     \caption{ 
    1289       \protect\label{fig:ZDF_M2_K1_tmx} 
    1290       (a) M2 and (b) K1 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). 
    1291     } 
    1292   \end{center} 
    1293 \end{figure} 
    1294 %>>>>>>>>>>>>>>>>>>>>>>>>>>>>  
    1295   
    1296 % ------------------------------------------------------------------------------------------------------------- 
    1297 %        Indonesian area specific treatment  
    1298 % ------------------------------------------------------------------------------------------------------------- 
    1299 \subsection{Indonesian area specific treatment (\protect\np{ln\_zdftmx\_itf})} 
    1300 \label{subsec:ZDF_tmx_itf} 
    1301  
    1302 When the Indonesian Through Flow (ITF) area is included in the model domain, 
    1303 a specific treatment of tidal induced mixing in this area can be used. 
    1304 It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide an input NetCDF file, 
    1305 \ifile{mask\_itf}, which contains a mask array defining the ITF area where the specific treatment is applied. 
    1306  
    1307 When \np{ln\_tmx\_itf}\forcode{ = .true.}, the two key parameters $q$ and $F(z)$ are adjusted following 
    1308 the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}: 
    1309  
    1310 First, the Indonesian archipelago is a complex geographic region with a series of 
    1311 large, deep, semi-enclosed basins connected via numerous narrow straits. 
    1312 Once generated, internal tides remain confined within this semi-enclosed area and hardly radiate away. 
    1313 Therefore all the internal tides energy is consumed within this area. 
    1314 So it is assumed that $q = 1$, \ie all the energy generated is available for mixing. 
    1315 Note that for test purposed, the ITF tidal dissipation efficiency is a namelist parameter (\np{rn\_tfe\_itf}). 
    1316 A value of $1$ or close to is this recommended for this parameter. 
    1317  
    1318 Second, the vertical structure function, $F(z)$, is no more associated with a bottom intensification of the mixing, 
    1319 but with a maximum of energy available within the thermocline. 
    1320 \citet{Koch-Larrouy_al_GRL07} have suggested that the vertical distribution of 
    1321 the energy dissipation proportional to $N^2$ below the core of the thermocline and to $N$ above.  
    1322 The resulting $F(z)$ is: 
    1323 \[ 
    1324   % \label{eq:Fz_itf} 
    1325   F(i,j,k) \sim     \left\{ 
    1326     \begin{aligned} 
    1327       \frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\ 
    1328       \frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$} 
    1329     \end{aligned} 
    1330   \right. 
    1331 \] 
    1332  
    1333 Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,  
    1334 which agrees with the independent estimates inferred from observations. 
    1335 Introduced in a regional OGCM, the parameterization improves the water mass characteristics in 
    1336 the different Indonesian seas, suggesting that the horizontal and vertical distributions of 
    1337 the mixing are adequately prescribed \citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}. 
    1338 Note also that such a parameterisation has a significant impact on the behaviour of 
    1339 global coupled GCMs \citep{Koch-Larrouy_al_CD10}. 
    1340  
    1341 % ================================================================ 
    1342 % Internal wave-driven mixing 
    1343 % ================================================================ 
    1344 \section{Internal wave-driven mixing (\protect\key{zdftmx\_new})} 
    1345 \label{sec:ZDF_tmx_new} 
    1346  
    1347 %--------------------------------------------namzdf_tmx_new------------------------------------------ 
    1348 % 
    1349 %\nlst{namzdf_tmx_new} 
    1350 %-------------------------------------------------------------------------------------------------------------- 
    1351  
    13521181The parameterization of mixing induced by breaking internal waves is a generalization of 
    1353 the approach originally proposed by \citet{St_Laurent_al_GRL02}. 
     1182the approach originally proposed by \citet{st-laurent.simmons.ea_GRL02}. 
    13541183A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed, 
    13551184and the resulting diffusivity is obtained as  
     
    13601189where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of 
    13611190the energy available for mixing. 
    1362 If the \np{ln\_mevar} namelist parameter is set to false, the mixing efficiency is taken as constant and 
    1363 equal to 1/6 \citep{Osborn_JPO80}. 
     1191If the \np{ln\_mevar} namelist parameter is set to \forcode{.false.}, the mixing efficiency is taken as constant and 
     1192equal to 1/6 \citep{osborn_JPO80}. 
    13641193In the opposite (recommended) case, $R_f$ is instead a function of 
    13651194the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$, 
    1366 with $\nu$ the molecular viscosity of seawater, following the model of \cite{Bouffard_Boegman_DAO2013} and 
    1367 the implementation of \cite{de_lavergne_JPO2016_efficiency}. 
     1195with $\nu$ the molecular viscosity of seawater, following the model of \cite{bouffard.boegman_DAO13} and 
     1196the implementation of \cite{de-lavergne.madec.ea_JPO16}. 
    13681197Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when 
    13691198the mixing efficiency is constant. 
    13701199 
    13711200In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary  
    1372 as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to true, a recommended choice.  
    1373 This parameterization of differential mixing, due to \cite{Jackson_Rehmann_JPO2014}, 
    1374 is implemented as in \cite{de_lavergne_JPO2016_efficiency}. 
     1201as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to \forcode{.true.}, a recommended choice.  
     1202This parameterization of differential mixing, due to \cite{jackson.rehmann_JPO14}, 
     1203is implemented as in \cite{de-lavergne.madec.ea_JPO16}. 
    13751204 
    13761205The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$, 
    13771206is constructed from three static maps of column-integrated internal wave energy dissipation, 
    1378 $E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures 
    1379 (de Lavergne et al., in prep): 
     1207$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures: 
     1208 
    13801209\begin{align*} 
    13811210  F_{cri}(i,j,k) &\propto e^{-h_{ab} / h_{cri} }\\ 
    1382   F_{pyc}(i,j,k) &\propto N^{n\_p}\\ 
     1211  F_{pyc}(i,j,k) &\propto N^{n_p}\\ 
    13831212  F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} } 
    13841213\end{align*}  
     
    13881217  h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; , 
    13891218\] 
    1390 The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_tmx\_new} namelist) 
     1219The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_iwm} namelist) 
    13911220controls the stratification-dependence of the pycnocline-intensified dissipation. 
    1392 It can take values of 1 (recommended) or 2. 
     1221It can take values of $1$ (recommended) or $2$. 
    13931222Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of 
    13941223the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps. 
    13951224$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and 
    13961225$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of 
    1397 the abyssal hill topography \citep{Goff_JGR2010} and the latitude. 
     1226the abyssal hill topography \citep{goff_JGR10} and the latitude. 
     1227% 
     1228% Jc: input files names ? 
     1229 
     1230% ================================================================ 
     1231% surface wave-induced mixing  
     1232% ================================================================ 
     1233\section[Surface wave-induced mixing (\forcode{ln_zdfswm = .true.})] 
     1234{Surface wave-induced mixing (\protect\np{ln\_zdfswm}\forcode{ = .true.})} 
     1235\label{subsec:ZDF_swm} 
     1236 
     1237Surface waves produce an enhanced mixing through wave-turbulence interaction. 
     1238In addition to breaking waves induced turbulence (\autoref{subsec:ZDF_tke}), 
     1239the influence of non-breaking waves can be accounted introducing  
     1240wave-induced viscosity and diffusivity as a function of the wave number spectrum. 
     1241Following \citet{qiao.yuan.ea_OD10}, a formulation of wave-induced mixing coefficient 
     1242is provided  as a function of wave amplitude, Stokes Drift and wave-number: 
     1243 
     1244\begin{equation} 
     1245  \label{eq:Bv} 
     1246  B_{v} = \alpha {A} {U}_{st} {exp(3kz)} 
     1247\end{equation} 
     1248 
     1249Where $B_{v}$ is the wave-induced mixing coefficient, $A$ is the wave amplitude,  
     1250${U}_{st}$ is the Stokes Drift velocity, $k$ is the wave number and $\alpha$  
     1251is a constant which should be determined by observations or  
     1252numerical experiments and is set to be 1. 
     1253 
     1254The coefficient $B_{v}$ is then directly added to the vertical viscosity  
     1255and diffusivity coefficients. 
     1256 
     1257In order to account for this contribution set: \forcode{ln_zdfswm = .true.}, 
     1258then wave interaction has to be activated through \forcode{ln_wave = .true.}, 
     1259the Stokes Drift can be evaluated by setting \forcode{ln_sdw = .true.}  
     1260(see \autoref{subsec:SBC_wave_sdw}) 
     1261and the needed wave fields can be provided either in forcing or coupled mode 
     1262(for more information on wave parameters and settings see \autoref{sec:SBC_wave}) 
     1263 
     1264% ================================================================ 
     1265% Adaptive-implicit vertical advection 
     1266% ================================================================ 
     1267\section[Adaptive-implicit vertical advection (\forcode{ln_zad_Aimp = .true.})] 
     1268{Adaptive-implicit vertical advection(\protect\np{ln\_zad\_Aimp}\forcode{ = .true.})} 
     1269\label{subsec:ZDF_aimp} 
     1270 
     1271The adaptive-implicit vertical advection option in NEMO is based on the work of 
     1272\citep{shchepetkin_OM15}.  In common with most ocean models, the timestep used with NEMO 
     1273needs to satisfy multiple criteria associated with different physical processes in order 
     1274to maintain numerical stability. \citep{shchepetkin_OM15} pointed out that the vertical 
     1275CFL criterion is commonly the most limiting. \citep{lemarie.debreu.ea_OM15} examined the 
     1276constraints for a range of time and space discretizations and provide the CFL stability 
     1277criteria for a range of advection schemes. The values for the Leap-Frog with Robert 
     1278asselin filter time-stepping (as used in NEMO) are reproduced in 
     1279\autoref{tab:zad_Aimp_CFLcrit}. Treating the vertical advection implicitly can avoid these 
     1280restrictions but at the cost of large dispersive errors and, possibly, large numerical 
     1281viscosity. The adaptive-implicit vertical advection option provides a targetted use of the 
     1282implicit scheme only when and where potential breaches of the vertical CFL condition 
     1283occur. In many practical applications these events may occur remote from the main area of 
     1284interest or due to short-lived conditions such that the extra numerical diffusion or 
     1285viscosity does not greatly affect the overall solution. With such applications, setting: 
     1286\forcode{ln_zad_Aimp = .true.} should allow much longer model timesteps to be used whilst 
     1287retaining the accuracy of the high order explicit schemes over most of the domain. 
     1288 
     1289\begin{table}[htbp] 
     1290  \begin{center} 
     1291    % \begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}} 
     1292    \begin{tabular}{r|ccc} 
     1293      \hline 
     1294      spatial discretization &   2nd order centered   & 3rd order upwind & 4th order compact  \\ 
     1295      advective CFL criterion     & 0.904 &   0.472  &   0.522    \\ 
     1296      \hline 
     1297    \end{tabular} 
     1298    \caption{ 
     1299      \protect\label{tab:zad_Aimp_CFLcrit} 
     1300      The advective CFL criteria for a range of spatial discretizations for the Leap-Frog with Robert Asselin filter time-stepping 
     1301      ($\nu=0.1$) as given in \citep{lemarie.debreu.ea_OM15}. 
     1302    } 
     1303  \end{center} 
     1304\end{table} 
     1305 
     1306In particular, the advection scheme remains explicit everywhere except where and when 
     1307local vertical velocities exceed a threshold set just below the explicit stability limit. 
     1308Once the threshold is reached a tapered transition towards an implicit scheme is used by 
     1309partitioning the vertical velocity into a part that can be treated explicitly and any 
     1310excess that must be treated implicitly. The partitioning is achieved via a Courant-number 
     1311dependent weighting algorithm as described in \citep{shchepetkin_OM15}. 
     1312 
     1313The local cell Courant number ($Cu$) used for this partitioning is: 
     1314 
     1315\begin{equation} 
     1316  \label{eq:Eqn_zad_Aimp_Courant} 
     1317  \begin{split} 
     1318    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 ]    \\ 
     1319       &\phantom{=} +\big [ \texttt{Max}(e_{{2_u}ij}e^n_{{3_{u}}ijk}u^n_{ijk},0.0) - \texttt{Min}(e_{{2_u}i-1j}e^n_{{3_{u}}i-1jk}u^n_{i-1jk},0.0) \big ] 
     1320                     \big / e_{{1_t}ij}e_{{2_t}ij}            \\ 
     1321       &\phantom{=} +\big [ \texttt{Max}(e_{{1_v}ij}e^n_{{3_{v}}ijk}v^n_{ijk},0.0) - \texttt{Min}(e_{{1_v}ij-1}e^n_{{3_{v}}ij-1k}v^n_{ij-1k},0.0) \big ] 
     1322                     \big / e_{{1_t}ij}e_{{2_t}ij} \bigg )    \\ 
     1323  \end{split} 
     1324\end{equation} 
     1325 
     1326\noindent and the tapering algorithm follows \citep{shchepetkin_OM15} as: 
     1327 
     1328\begin{align} 
     1329  \label{eq:Eqn_zad_Aimp_partition} 
     1330Cu_{min} &= 0.15 \nonumber \\ 
     1331Cu_{max} &= 0.3  \nonumber \\ 
     1332Cu_{cut} &= 2Cu_{max} - Cu_{min} \nonumber \\ 
     1333Fcu    &= 4Cu_{max}*(Cu_{max}-Cu_{min}) \nonumber \\ 
     1334C\kern-0.14em f &= 
     1335     \begin{cases} 
     1336        0.0                                                        &\text{if $Cu \leq Cu_{min}$} \\ 
     1337        (Cu - Cu_{min})^2 / (Fcu +  (Cu - Cu_{min})^2)             &\text{else if $Cu < Cu_{cut}$} \\ 
     1338        (Cu - Cu_{max}) / Cu                                       &\text{else} 
     1339     \end{cases} 
     1340\end{align} 
     1341 
     1342\begin{figure}[!t] 
     1343  \begin{center} 
     1344    \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_coeff} 
     1345    \caption{ 
     1346      \protect\label{fig:zad_Aimp_coeff} 
     1347      The value of the partitioning coefficient ($C\kern-0.14em f$) used to partition vertical velocities into parts to 
     1348      be treated implicitly and explicitly for a range of typical Courant numbers (\forcode{ln_zad_Aimp=.true.}) 
     1349    } 
     1350  \end{center} 
     1351\end{figure} 
     1352 
     1353\noindent The partitioning coefficient is used to determine the part of the vertical 
     1354velocity that must be handled implicitly ($w_i$) and to subtract this from the total 
     1355vertical velocity ($w_n$) to leave that which can continue to be handled explicitly: 
     1356 
     1357\begin{align} 
     1358  \label{eq:Eqn_zad_Aimp_partition2} 
     1359    w_{i_{ijk}} &= C\kern-0.14em f_{ijk} w_{n_{ijk}}     \nonumber \\ 
     1360    w_{n_{ijk}} &= (1-C\kern-0.14em f_{ijk}) w_{n_{ijk}}            
     1361\end{align} 
     1362 
     1363\noindent Note that the coefficient is such that the treatment is never fully implicit; 
     1364the three cases from \autoref{eq:Eqn_zad_Aimp_partition} can be considered as: 
     1365fully-explicit; mixed explicit/implicit and mostly-implicit.  With the settings shown the 
     1366coefficient ($C\kern-0.14em f$) varies as shown in \autoref{fig:zad_Aimp_coeff}. Note with these values 
     1367the $Cu_{cut}$ boundary between the mixed implicit-explicit treatment and 'mostly 
     1368implicit' is 0.45 which is just below the stability limited given in 
     1369\autoref{tab:zad_Aimp_CFLcrit}  for a 3rd order scheme. 
     1370 
     1371The $w_i$ component is added to the implicit solvers for the vertical mixing in 
     1372\mdl{dynzdf} and \mdl{trazdf} in a similar way to \citep{shchepetkin_OM15}.  This is 
     1373sufficient for the flux-limited advection scheme (\forcode{ln_traadv_mus}) but further 
     1374intervention is required when using the flux-corrected scheme (\forcode{ln_traadv_fct}). 
     1375For these schemes the implicit upstream fluxes must be added to both the monotonic guess 
     1376and to the higher order solution when calculating the antidiffusive fluxes. The implicit 
     1377vertical fluxes are then removed since they are added by the implicit solver later on. 
     1378 
     1379The adaptive-implicit vertical advection option is new to NEMO at v4.0 and has yet to be  
     1380used in a wide range of simulations. The following test simulation, however, does illustrate 
     1381the potential benefits and will hopefully encourage further testing and feedback from users: 
     1382 
     1383\begin{figure}[!t] 
     1384  \begin{center} 
     1385    \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_overflow_frames} 
     1386    \caption{ 
     1387      \protect\label{fig:zad_Aimp_overflow_frames} 
     1388      A time-series of temperature vertical cross-sections for the OVERFLOW test case. These results are for the default 
     1389      settings with \forcode{nn_rdt=10.0} and without adaptive implicit vertical advection (\forcode{ln_zad_Aimp=.false.}). 
     1390    } 
     1391  \end{center} 
     1392\end{figure} 
     1393 
     1394\subsection{Adaptive-implicit vertical advection in the OVERFLOW test-case} 
     1395The \href{https://forge.ipsl.jussieu.fr/nemo/chrome/site/doc/NEMO/guide/html/test\_cases.html\#overflow}{OVERFLOW test case} 
     1396provides a simple illustration of the adaptive-implicit advection in action. The example here differs from the basic test case 
     1397by only a few extra physics choices namely: 
     1398 
     1399\begin{verbatim} 
     1400     ln_dynldf_OFF = .false. 
     1401     ln_dynldf_lap = .true. 
     1402     ln_dynldf_hor = .true. 
     1403     ln_zdfnpc     = .true. 
     1404     ln_traadv_fct = .true. 
     1405        nn_fct_h   =  2 
     1406        nn_fct_v   =  2 
     1407\end{verbatim} 
     1408 
     1409\noindent which were chosen to provide a slightly more stable and less noisy solution. The 
     1410result when using the default value of \forcode{nn_rdt = 10.} without adaptive-implicit 
     1411vertical velocity is illustrated in \autoref{fig:zad_Aimp_overflow_frames}. The mass of 
     1412cold water, initially sitting on the shelf, moves down the slope and forms a 
     1413bottom-trapped, dense plume. Even with these extra physics choices the model is close to 
     1414stability limits and attempts with \forcode{nn_rdt = 30.} will fail after about 5.5 hours 
     1415with excessively high horizontal velocities. This time-scale corresponds with the time the 
     1416plume reaches the steepest part of the topography and, although detected as a horizontal 
     1417CFL breach, the instability originates from a breach of the vertical CFL limit. This is a good 
     1418candidate, therefore, for use of the adaptive-implicit vertical advection scheme. 
     1419 
     1420The results with \forcode{ln_zad_Aimp=.true.} and a variety of model timesteps 
     1421are shown in \autoref{fig:zad_Aimp_overflow_all_rdt} (together with the equivalent 
     1422frames from the base run).  In this simple example the use of the adaptive-implicit 
     1423vertcal advection scheme has enabled a 12x increase in the model timestep without 
     1424significantly altering the solution (although at this extreme the plume is more diffuse 
     1425and has not travelled so far).  Notably, the solution with and without the scheme is 
     1426slightly different even with \forcode{nn_rdt = 10.}; suggesting that the base run was 
     1427close enough to instability to trigger the scheme despite completing successfully. 
     1428To assist in diagnosing how active the scheme is, in both location and time, the 3D 
     1429implicit and explicit components of the vertical velocity are available via XIOS as 
     1430\texttt{wimp} and \texttt{wexp} respectively.  Likewise, the partitioning coefficient 
     1431($C\kern-0.14em f$) is also available as \texttt{wi\_cff}. For a quick oversight of 
     1432the schemes activity the global maximum values of the absolute implicit component 
     1433of the vertical velocity and the partitioning coefficient are written to the netCDF 
     1434version of the run statistics file (\texttt{run.stat.nc}) if this is active (see 
     1435\autoref{sec:MISC_opt} for activation details). 
     1436 
     1437\autoref{fig:zad_Aimp_maxCf} shows examples of the maximum partitioning coefficient for 
     1438the various overflow tests.  Note that the adaptive-implicit vertical advection scheme is 
     1439active even in the base run with \forcode{nn_rdt=10.0s} adding to the evidence that the 
     1440test case is close to stability limits even with this value. At the larger timesteps, the 
     1441vertical velocity is treated mostly implicitly at some location throughout the run. The 
     1442oscillatory nature of this measure appears to be linked to the progress of the plume front 
     1443as each cusp is associated with the location of the maximum shifting to the adjacent cell. 
     1444This is illustrated in \autoref{fig:zad_Aimp_maxCf_loc} where the i- and k- locations of the 
     1445maximum have been overlaid for the base run case. 
     1446 
     1447\medskip 
     1448\noindent Only limited tests have been performed in more realistic configurations. In the 
     1449ORCA2\_ICE\_PISCES reference configuration the scheme does activate and passes 
     1450restartability and reproducibility tests but it is unable to improve the model's stability 
     1451enough to allow an increase in the model time-step. A view of the time-series of maximum 
     1452partitioning coefficient (not shown here)  suggests that the default time-step of 5400s is 
     1453already pushing at stability limits, especially in the initial start-up phase. The 
     1454time-series does not, however, exhibit any of the 'cuspiness' found with the overflow 
     1455tests. 
     1456 
     1457\medskip 
     1458\noindent A short test with an eORCA1 configuration promises more since a test using a 
     1459time-step of 3600s remains stable with \forcode{ln_zad_Aimp=.true.} whereas the 
     1460time-step is limited to 2700s without. 
     1461 
     1462\begin{figure}[!t] 
     1463  \begin{center} 
     1464    \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_overflow_all_rdt} 
     1465    \caption{ 
     1466      \protect\label{fig:zad_Aimp_overflow_all_rdt} 
     1467      Sample temperature vertical cross-sections from mid- and end-run using different values for \forcode{nn_rdt}  
     1468      and with or without adaptive implicit vertical advection. Without the adaptive implicit vertical advection only 
     1469      the run with the shortest timestep is able to run to completion. Note also that the colour-scale has been 
     1470      chosen to confirm that temperatures remain within the original range of 10$^o$ to 20$^o$. 
     1471    } 
     1472  \end{center} 
     1473\end{figure} 
     1474 
     1475\begin{figure}[!t] 
     1476  \begin{center} 
     1477    \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_maxCf} 
     1478    \caption{ 
     1479      \protect\label{fig:zad_Aimp_maxCf} 
     1480      The maximum partitioning coefficient during a series of test runs with increasing model timestep length. 
     1481      At the larger timesteps, the vertical velocity is treated mostly implicitly at some location throughout  
     1482      the run. 
     1483    } 
     1484  \end{center} 
     1485\end{figure} 
     1486 
     1487\begin{figure}[!t] 
     1488  \begin{center} 
     1489    \includegraphics[width=\textwidth]{Fig_ZDF_zad_Aimp_maxCf_loc} 
     1490    \caption{ 
     1491      \protect\label{fig:zad_Aimp_maxCf_loc} 
     1492      The maximum partitioning coefficient for the  \forcode{nn_rdt=10.0s} case overlaid with  information on the gridcell i- and k- 
     1493      locations of the maximum value.  
     1494    } 
     1495  \end{center} 
     1496\end{figure} 
    13981497 
    13991498% ================================================================ 
Note: See TracChangeset for help on using the changeset viewer.