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 10354 for NEMO/trunk/doc/latex/NEMO/subfiles/chap_ZDF.tex – NEMO

Ignore:
Timestamp:
2018-11-21T17:59:55+01:00 (5 years ago)
Author:
nicolasmartin
Message:

Vast edition of LaTeX subfiles to improve the readability by cutting sentences in a more suitable way
Every sentence begins in a new line and if necessary is splitted around 110 characters lenght for side-by-side visualisation,
this setting may not be adequate for everyone but something has to be set.
The punctuation was the primer trigger for the cutting process, otherwise subordinators and coordinators, in order to mostly keep a meaning for each line

File:
1 edited

Legend:

Unmodified
Added
Removed
  • NEMO/trunk/doc/latex/NEMO/subfiles/chap_ZDF.tex

    r10146 r10354  
    2121\label{sec:ZDF_zdf} 
    2222 
    23 The discrete form of the ocean subgrid scale physics has been presented in  
    24 \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}. At the surface and bottom boundaries,  
    25 the turbulent fluxes of momentum, heat and salt have to be defined. At the  
    26 surface they are prescribed from the surface forcing (see \autoref{chap:SBC}),  
    27 while at the bottom they are set to zero for heat and salt, unless a geothermal  
    28 flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl}  
    29 defined, see \autoref{subsec:TRA_bbc}), and specified through a bottom friction  
    30 parameterisation for momentum (see \autoref{sec:ZDF_bfr}). 
    31  
    32 In this section we briefly discuss the various choices offered to compute  
    33 the vertical eddy viscosity and diffusivity coefficients, $A_u^{vm}$ ,  
    34 $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$-  
    35 points, respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}). These  
    36 coefficients can be assumed to be either constant, or a function of the local  
    37 Richardson number, or computed from a turbulent closure model (either  
    38 TKE or GLS formulation). The computation of these coefficients is initialized  
    39 in the \mdl{zdfini} module and performed in the \mdl{zdfric}, \mdl{zdftke} or  
    40 \mdl{zdfgls} modules. The trends due to the vertical momentum and tracer  
    41 diffusion, including the surface forcing, are computed and added to the  
    42 general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.  
    43 These trends can be computed using either a forward time stepping scheme  
    44 (namelist parameter \np{ln\_zdfexp}\forcode{ = .true.}) or a backward time stepping  
    45 scheme (\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing  
    46 coefficients, and thus of the formulation used (see \autoref{chap:STP}). 
     23The discrete form of the ocean subgrid scale physics has been presented in 
     24\autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}. 
     25At the surface and bottom boundaries, the turbulent fluxes of momentum, heat and salt have to be defined. 
     26At the surface they are prescribed from the surface forcing (see \autoref{chap:SBC}), 
     27while at the bottom they are set to zero for heat and salt, 
     28unless a geothermal flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl} defined, 
     29see \autoref{subsec:TRA_bbc}), and specified through a bottom friction parameterisation for momentum 
     30(see \autoref{sec:ZDF_bfr}). 
     31 
     32In this section we briefly discuss the various choices offered to compute the vertical eddy viscosity and 
     33diffusivity coefficients, $A_u^{vm}$ , $A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$- points, 
     34respectively (see \autoref{sec:TRA_zdf} and \autoref{sec:DYN_zdf}). 
     35These coefficients can be assumed to be either constant, or a function of the local Richardson number, 
     36or computed from a turbulent closure model (either TKE or GLS formulation). 
     37The computation of these coefficients is initialized in the \mdl{zdfini} module and performed in 
     38the \mdl{zdfric}, \mdl{zdftke} or \mdl{zdfgls} modules. 
     39The trends due to the vertical momentum and tracer diffusion, including the surface forcing, 
     40are computed and added to the general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.  
     41These trends can be computed using either a forward time stepping scheme 
     42(namelist parameter \np{ln\_zdfexp}\forcode{ = .true.}) or a backward time stepping scheme 
     43(\np{ln\_zdfexp}\forcode{ = .false.}) depending on the magnitude of the mixing coefficients, 
     44and thus of the formulation used (see \autoref{chap:STP}). 
    4745 
    4846% ------------------------------------------------------------------------------------------------------------- 
     
    5654%-------------------------------------------------------------------------------------------------------------- 
    5755 
    58 Options are defined through the  \ngn{namzdf} namelist variables. 
    59 When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients  
    60 are set to constant values over the whole ocean. This is the crudest way to define  
    61 the vertical ocean physics. It is recommended that this option is only used in  
    62 process studies, not in basin scale simulations. Typical values used in this case are: 
     56Options are defined through the \ngn{namzdf} namelist variables. 
     57When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients are set to 
     58constant values over the whole ocean. 
     59This is the crudest way to define the vertical ocean physics. 
     60It is recommended that this option is only used in process studies, not in basin scale simulations. 
     61Typical values used in this case are: 
    6362\begin{align*}  
    6463A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\ 
     
    6766 
    6867These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.  
    69 In all cases, do not use values smaller that those associated with the molecular  
    70 viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum,  
    71 $\sim10^{-7}~m^2.s^{-1}$ for temperature and $\sim10^{-9}~m^2.s^{-1}$ for salinity. 
     68In all cases, do not use values smaller that those associated with the molecular viscosity and diffusivity, 
     69that is $\sim10^{-6}~m^2.s^{-1}$ for momentum, $\sim10^{-7}~m^2.s^{-1}$ for temperature and 
     70$\sim10^{-9}~m^2.s^{-1}$ for salinity. 
    7271 
    7372 
     
    8382%-------------------------------------------------------------------------------------------------------------- 
    8483 
    85 When \key{zdfric} is defined, a local Richardson number dependent formulation  
    86 for the vertical momentum and tracer eddy coefficients is set through the  \ngn{namzdf\_ric}  
    87 namelist variables.The vertical mixing  
    88 coefficients are diagnosed from the large scale variables computed by the model.  
    89 \textit{In situ} measurements have been used to link vertical turbulent activity to  
    90 large scale ocean structures. The hypothesis of a mixing mainly maintained by the  
    91 growth of Kelvin-Helmholtz like instabilities leads to a dependency between the  
    92 vertical eddy coefficients and the local Richardson number ($i.e.$ the  
    93 ratio of stratification to vertical shear). Following \citet{Pacanowski_Philander_JPO81}, the following  
    94 formulation has been implemented: 
     84When \key{zdfric} is defined, a local Richardson number dependent formulation for the vertical momentum and 
     85tracer eddy coefficients is set through the \ngn{namzdf\_ric} namelist variables. 
     86The vertical mixing coefficients are diagnosed from the large scale variables computed by the model.  
     87\textit{In situ} measurements have been used to link vertical turbulent activity to large scale ocean structures. 
     88The hypothesis of a mixing mainly maintained by the growth of Kelvin-Helmholtz like instabilities leads to 
     89a dependency between the vertical eddy coefficients and the local Richardson number 
     90($i.e.$ the ratio of stratification to vertical shear). 
     91Following \citet{Pacanowski_Philander_JPO81}, the following formulation has been implemented: 
    9592\begin{equation} \label{eq:zdfric} 
    9693   \left\{      \begin{aligned} 
     
    9996   \end{aligned}    \right. 
    10097\end{equation} 
    101 where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson  
    102 number, $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),  
    103 $A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the  
    104 constant case (see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$  
    105 is the maximum value that can be reached by the coefficient when $Ri\leq 0$,  
    106 $a=5$ and $n=2$. The last three values can be modified by setting the  
    107 \np{rn\_avmri}, \np{rn\_alp} and \np{nn\_ric} namelist parameters, respectively. 
    108  
    109 A simple mixing-layer model to transfer and dissipate the atmospheric 
    110  forcings (wind-stress and buoyancy fluxes) can be activated setting  
    111 the \np{ln\_mldw}\forcode{ = .true.} in the namelist. 
    112  
    113 In this case, the local depth of turbulent wind-mixing or "Ekman depth" 
    114  $h_{e}(x,y,t)$ is evaluated and the vertical eddy coefficients prescribed within this layer. 
     98where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson number, 
     99$N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),  
     100$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the constant case 
     101(see \autoref{subsec:ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ is the maximum value that 
     102can be reached by the coefficient when $Ri\leq 0$, $a=5$ and $n=2$. 
     103The last three values can be modified by setting the \np{rn\_avmri}, \np{rn\_alp} and 
     104\np{nn\_ric} namelist parameters, respectively. 
     105 
     106A simple mixing-layer model to transfer and dissipate the atmospheric forcings 
     107(wind-stress and buoyancy fluxes) can be activated setting the \np{ln\_mldw}\forcode{ = .true.} in the namelist. 
     108 
     109In this case, the local depth of turbulent wind-mixing or "Ekman depth" $h_{e}(x,y,t)$ is evaluated and 
     110the vertical eddy coefficients prescribed within this layer. 
    115111 
    116112This depth is assumed proportional to the "depth of frictional influence" that is limited by rotation: 
    117113\begin{equation} 
    118          h_{e} = Ek \frac {u^{*}} {f_{0}}    \\ 
    119 \end{equation} 
    120 where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis  
    121 parameter. 
     114h_{e} = Ek \frac {u^{*}} {f_{0}} 
     115\end{equation} 
     116where, $Ek$ is an empirical parameter, $u^{*}$ is the friction velocity and $f_{0}$ is the Coriolis parameter. 
    122117 
    123118In this similarity height relationship, the turbulent friction velocity: 
    124119\begin{equation} 
    125          u^{*} = \sqrt \frac {|\tau|} {\rho_o}     \\ 
    126 \end{equation} 
    127  
     120u^{*} = \sqrt \frac {|\tau|} {\rho_o} 
     121\end{equation} 
    128122is computed from the wind stress vector $|\tau|$ and the reference density $ \rho_o$. 
    129123The final $h_{e}$ is further constrained by the adjustable bounds \np{rn\_mldmin} and \np{rn\_mldmax}. 
    130 Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to  
     124Once $h_{e}$ is computed, the vertical eddy coefficients within $h_{e}$ are set to 
    131125the empirical values \np{rn\_wtmix} and \np{rn\_wvmix} \citep{Lermusiaux2001}. 
    132126 
     
    142136%-------------------------------------------------------------------------------------------------------------- 
    143137 
    144 The vertical eddy viscosity and diffusivity coefficients are computed from a TKE  
    145 turbulent closure model based on a prognostic equation for $\bar{e}$, the turbulent  
    146 kinetic energy, and a closure assumption for the turbulent length scales. This  
    147 turbulent closure model has been developed by \citet{Bougeault1989} in the  
    148 atmospheric case, adapted by \citet{Gaspar1990} for the oceanic case, and  
    149 embedded in OPA, the ancestor of NEMO, by \citet{Blanke1993} for equatorial Atlantic  
    150 simulations. Since then, significant modifications have been introduced by  
    151 \citet{Madec1998} in both the implementation and the formulation of the mixing  
    152 length scale. The time evolution of $\bar{e}$ is the result of the production of  
    153 $\bar{e}$ through vertical shear, its destruction through stratification, its vertical  
    154 diffusion, and its dissipation of \citet{Kolmogorov1942} type: 
     138The vertical eddy viscosity and diffusivity coefficients are computed from a TKE turbulent closure model based on 
     139a prognostic equation for $\bar{e}$, the turbulent kinetic energy, 
     140and a closure assumption for the turbulent length scales. 
     141This turbulent closure model has been developed by \citet{Bougeault1989} in the atmospheric case, 
     142adapted by \citet{Gaspar1990} for the oceanic case, and embedded in OPA, the ancestor of NEMO, 
     143by \citet{Blanke1993} for equatorial Atlantic simulations. 
     144Since then, significant modifications have been introduced by \citet{Madec1998} in both the implementation and 
     145the formulation of the mixing length scale. 
     146The time evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical shear, 
     147its destruction through stratification, its vertical diffusion, and its dissipation of \citet{Kolmogorov1942} type: 
    155148\begin{equation} \label{eq:zdftke_e} 
    156149\frac{\partial \bar{e}}{\partial t} =  
     
    170163where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}),  
    171164$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,  
    172 $P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity  
    173 and diffusivity coefficients. The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$   
    174 $\approx 0.7$ are designed to deal with vertical mixing at any depth \citep{Gaspar1990}.  
    175 They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}.  
    176 $P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function  
    177 of the local Richardson number, $R_i$: 
     165$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity and diffusivity coefficients. 
     166The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$ $\approx 0.7$ are designed to deal with 
     167vertical mixing at any depth \citep{Gaspar1990}.  
     168They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}. 
     169$P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function of the local Richardson number, $R_i$: 
    178170\begin{align*} \label{eq:prt} 
    179171P_{rt} = \begin{cases} 
     
    186178The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist variable. 
    187179 
    188 At the sea surface, the value of $\bar{e}$ is prescribed from the wind  
    189 stress field as $\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb}  
    190 namelist parameter. The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}),  
    191 however a much larger value can be used when taking into account the  
    192 surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}).  
    193 The bottom value of TKE is assumed to be equal to the value of the level just above.  
    194 The time integration of the $\bar{e}$ equation may formally lead to negative values  
    195 because the numerical scheme does not ensure its positivity. To overcome this  
    196 problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin}  
    197 namelist parameter). Following \citet{Gaspar1990}, the cut-off value is set  
    198 to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. This allows the subsequent formulations  
    199 to match that of \citet{Gargett1984} for the diffusion in the thermocline and  
    200 deep ocean :  $K_\rho = 10^{-3} / N$.  
    201 In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical  
    202 instabilities associated with too weak vertical diffusion. They must be  
    203 specified at least larger than the molecular values, and are set through  
    204 \np{rn\_avm0} and \np{rn\_avt0} (namzdf namelist, see \autoref{subsec:ZDF_cst}). 
     180At the sea surface, the value of $\bar{e}$ is prescribed from the wind stress field as 
     181$\bar{e}_o = e_{bb} |\tau| / \rho_o$, with $e_{bb}$ the \np{rn\_ebb} namelist parameter. 
     182The default value of $e_{bb}$ is 3.75. \citep{Gaspar1990}), however a much larger value can be used when 
     183taking into account the surface wave breaking (see below Eq. \autoref{eq:ZDF_Esbc}). 
     184The bottom value of TKE is assumed to be equal to the value of the level just above. 
     185The time integration of the $\bar{e}$ equation may formally lead to negative values because 
     186the numerical scheme does not ensure its positivity. 
     187To overcome this problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin} namelist parameter). 
     188Following \citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. 
     189This allows the subsequent formulations to match that of \citet{Gargett1984} for the diffusion in 
     190the thermocline and deep ocean :  $K_\rho = 10^{-3} / N$. 
     191In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical instabilities associated with 
     192too weak vertical diffusion. 
     193They must be specified at least larger than the molecular values, and are set through \np{rn\_avm0} and 
     194\np{rn\_avt0} (namzdf namelist, see \autoref{subsec:ZDF_cst}). 
    205195 
    206196\subsubsection{Turbulent length scale} 
    207 For computational efficiency, the original formulation of the turbulent length  
    208 scales proposed by \citet{Gaspar1990} has been simplified. Four formulations  
    209 are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist  
    210 parameter. The first two are based on the following first order approximation  
    211 \citep{Blanke1993}: 
     197For computational efficiency, the original formulation of the turbulent length scales proposed by 
     198\citet{Gaspar1990} has been simplified. 
     199Four formulations are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist parameter. 
     200The first two are based on the following first order approximation \citep{Blanke1993}: 
    212201\begin{equation} \label{eq:tke_mxl0_1} 
    213202l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N 
    214203\end{equation} 
    215 which is valid in a stable stratified region with constant values of the Brunt- 
    216 Vais\"{a}l\"{a} frequency. The resulting length scale is bounded by the distance  
    217 to the surface or to the bottom (\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor  
    218 (\np{nn\_mxl}\forcode{ = 1}). \citet{Blanke1993} notice that this simplification has two major  
    219 drawbacks: it makes no sense for locally unstable stratification and the  
    220 computation no longer uses all the information contained in the vertical density  
    221 profile. To overcome these drawbacks, \citet{Madec1998} introduces the  
    222 \np{nn\_mxl}\forcode{ = 2..3} cases, which add an extra assumption concerning the vertical  
    223 gradient of the computed length scale. So, the length scales are first evaluated  
    224 as in \autoref{eq:tke_mxl0_1} and then bounded such that: 
     204which is valid in a stable stratified region with constant values of the Brunt-Vais\"{a}l\"{a} frequency. 
     205The resulting length scale is bounded by the distance to the surface or to the bottom 
     206(\np{nn\_mxl}\forcode{ = 0}) or by the local vertical scale factor (\np{nn\_mxl}\forcode{ = 1}). 
     207\citet{Blanke1993} notice that this simplification has two major drawbacks: 
     208it makes no sense for locally unstable stratification and the computation no longer uses all 
     209the information contained in the vertical density profile. 
     210To overcome these drawbacks, \citet{Madec1998} introduces the \np{nn\_mxl}\forcode{ = 2..3} cases, 
     211which add an extra assumption concerning the vertical gradient of the computed length scale. 
     212So, the length scales are first evaluated as in \autoref{eq:tke_mxl0_1} and then bounded such that: 
    225213\begin{equation} \label{eq:tke_mxl_constraint} 
    226214\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1 
    227215\qquad \text{with }\  l =  l_k = l_\epsilon 
    228216\end{equation} 
    229 \autoref{eq:tke_mxl_constraint} means that the vertical variations of the length  
    230 scale cannot be larger than the variations of depth. It provides a better  
    231 approximation of the \citet{Gaspar1990} formulation while being much less  
    232 time consuming. In particular, it allows the length scale to be limited not only  
    233 by the distance to the surface or to the ocean bottom but also by the distance  
    234 to a strongly stratified portion of the water column such as the thermocline  
    235 (\autoref{fig:mixing_length}). In order to impose the \autoref{eq:tke_mxl_constraint}  
    236 constraint, we introduce two additional length scales: $l_{up}$ and $l_{dwn}$,  
    237 the upward and downward length scales, and evaluate the dissipation and  
    238 mixing length scales as (and note that here we use numerical indexing): 
     217\autoref{eq:tke_mxl_constraint} means that the vertical variations of the length scale cannot be larger than 
     218the variations of depth. 
     219It provides a better approximation of the \citet{Gaspar1990} formulation while being much less  
     220time consuming. 
     221In particular, it allows the length scale to be limited not only by the distance to the surface or 
     222to the ocean bottom but also by the distance to a strongly stratified portion of the water column such as 
     223the thermocline (\autoref{fig:mixing_length}). 
     224In order to impose the \autoref{eq:tke_mxl_constraint} constraint, we introduce two additional length scales: 
     225$l_{up}$ and $l_{dwn}$, the upward and downward length scales, and 
     226evaluate the dissipation and mixing length scales as 
     227(and note that here we use numerical indexing): 
    239228%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    240229\begin{figure}[!t] \begin{center} 
     
    253242\end{aligned} 
    254243\end{equation} 
    255 where $l^{(k)}$ is computed using \autoref{eq:tke_mxl0_1},  
    256 $i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
    257  
    258 In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same  
    259 value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the  
    260 \np{nn\_mxl}\forcode{ = 3} case, the dissipation and mixing turbulent length scales are give  
    261 as in \citet{Gaspar1990}: 
     244where $l^{(k)}$ is computed using \autoref{eq:tke_mxl0_1}, $i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
     245 
     246In the \np{nn\_mxl}\forcode{ = 2} case, the dissipation and mixing length scales take the same value: 
     247$ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the \np{nn\_mxl}\forcode{ = 3} case, 
     248the dissipation and mixing turbulent length scales are give as in \citet{Gaspar1990}: 
    262249\begin{equation} \label{eq:tke_mxl_gaspar} 
    263250\begin{aligned} 
     
    267254\end{equation} 
    268255 
    269 At the ocean surface, a non zero length scale is set through the  \np{rn\_mxl0} namelist  
    270 parameter. Usually the surface scale is given by $l_o = \kappa \,z_o$  
    271 where $\kappa = 0.4$ is von Karman's constant and $z_o$ the roughness  
    272 parameter of the surface. Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94}  
    273 leads to a 0.04~m, the default value of \np{rn\_mxl0}. In the ocean interior  
    274 a minimum length scale is set to recover the molecular viscosity when $\bar{e}$  
    275 reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ). 
     256At the ocean surface, a non zero length scale is set through the  \np{rn\_mxl0} namelist parameter. 
     257Usually the surface scale is given by $l_o = \kappa \,z_o$ where $\kappa = 0.4$ is von Karman's constant and 
     258$z_o$ the roughness parameter of the surface. 
     259Assuming $z_o=0.1$~m \citep{Craig_Banner_JPO94} leads to a 0.04~m, the default value of \np{rn\_mxl0}. 
     260In the ocean interior a minimum length scale is set to recover the molecular viscosity when 
     261$\bar{e}$ reach its minimum value ($1.10^{-6}= C_k\, l_{min} \,\sqrt{\bar{e}_{min}}$ ). 
    276262 
    277263 
    278264\subsubsection{Surface wave breaking parameterization} 
    279265%-----------------------------------------------------------------------% 
    280 Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified  
    281 to include the effect of surface wave breaking energetics. This results in a reduction of summertime  
    282 surface temperature when the mixed layer is relatively shallow. The \citet{Mellor_Blumberg_JPO04}  
    283 modifications acts on surface length scale and TKE values and air-sea drag coefficient.  
     266Following \citet{Mellor_Blumberg_JPO04}, the TKE turbulence closure model has been modified to 
     267include the effect of surface wave breaking energetics. 
     268This results in a reduction of summertime surface temperature when the mixed layer is relatively shallow. 
     269The \citet{Mellor_Blumberg_JPO04} modifications acts on surface length scale and TKE values and 
     270air-sea drag coefficient.  
    284271The latter concerns the bulk formulea and is not discussed here.  
    285272 
     
    288275\bar{e}_o = \frac{1}{2}\,\left(  15.8\,\alpha_{CB} \right)^{2/3} \,\frac{|\tau|}{\rho_o} 
    289276\end{equation} 
    290 where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality  
    291 which depends on the ''wave age'', ranging from 57 for mature waves to 146 for  
    292 younger waves \citep{Mellor_Blumberg_JPO04}.  
     277where $\alpha_{CB}$ is the \citet{Craig_Banner_JPO94} constant of proportionality which depends on the ''wave age'', 
     278ranging from 57 for mature waves to 146 for younger waves \citep{Mellor_Blumberg_JPO04}.  
    293279The boundary condition on the turbulent length scale follows the Charnock's relation: 
    294280\begin{equation} \label{eq:ZDF_Lsbc} 
     
    296282\end{equation} 
    297283where $\kappa=0.40$ is the von Karman constant, and $\beta$ is the Charnock's constant. 
    298 \citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by \citet{Stacey_JPO99} 
    299 citing observation evidence, and $\alpha_{CB} = 100$ the Craig and Banner's value. 
    300 As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$,  
     284\citet{Mellor_Blumberg_JPO04} suggest $\beta = 2.10^{5}$ the value chosen by 
     285\citet{Stacey_JPO99} citing observation evidence, and 
     286$\alpha_{CB} = 100$ the Craig and Banner's value. 
     287As the surface boundary condition on TKE is prescribed through $\bar{e}_o = e_{bb} |\tau| / \rho_o$, 
    301288with $e_{bb}$ the \np{rn\_ebb} namelist parameter, setting \np{rn\_ebb}\forcode{ = 67.83} corresponds  
    302 to $\alpha_{CB} = 100$. Further setting  \np{ln\_mxl0} to true applies \autoref{eq:ZDF_Lsbc}  
    303 as surface boundary condition on length scale, with $\beta$ hard coded to the Stacey's value. 
    304 Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters)  
    305 is applied on surface $\bar{e}$ value. 
     289to $\alpha_{CB} = 100$. 
     290Further setting  \np{ln\_mxl0} to true applies \autoref{eq:ZDF_Lsbc} as surface boundary condition on length scale, 
     291with $\beta$ hard coded to the Stacey's value. 
     292Note that a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist parameters) is applied on 
     293surface $\bar{e}$ value. 
    306294 
    307295 
    308296\subsubsection{Langmuir cells} 
    309297%--------------------------------------% 
    310 Langmuir circulations (LC) can be described as ordered large-scale vertical motions  
    311 in the surface layer of the oceans. Although LC have nothing to do with convection,  
    312 the circulation pattern is rather similar to so-called convective rolls in the atmospheric  
    313 boundary layer. The detailed physics behind LC is described in, for example,  
    314 \citet{Craik_Leibovich_JFM76}. The prevailing explanation is that LC arise from  
    315 a nonlinear interaction between the Stokes drift and wind drift currents.  
    316  
    317 Here we introduced in the TKE turbulent closure the simple parameterization of  
    318 Langmuir circulations proposed by \citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure.  
    319 The parameterization, tuned against large-eddy simulation, includes the whole effect 
    320 of LC in an extra source terms of TKE, $P_{LC}$. 
    321 The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled  
    322 by setting \np{ln\_lc} to \forcode{.true.} in the namtke namelist. 
     298Langmuir circulations (LC) can be described as ordered large-scale vertical motions in 
     299the surface layer of the oceans. 
     300Although LC have nothing to do with convection, the circulation pattern is rather similar to 
     301so-called convective rolls in the atmospheric boundary layer. 
     302The detailed physics behind LC is described in, for example, \citet{Craik_Leibovich_JFM76}. 
     303The prevailing explanation is that LC arise from a nonlinear interaction between the Stokes drift and 
     304wind drift currents.  
     305 
     306Here we introduced in the TKE turbulent closure the simple parameterization of Langmuir circulations proposed by 
     307\citep{Axell_JGR02} for a $k-\epsilon$ turbulent closure. 
     308The parameterization, tuned against large-eddy simulation, includes the whole effect of LC in 
     309an extra source terms of TKE, $P_{LC}$. 
     310The presence of $P_{LC}$ in \autoref{eq:zdftke_e}, the TKE equation, is controlled by setting \np{ln\_lc} to 
     311\forcode{.true.} in the namtke namelist. 
    323312  
    324 By making an analogy with the characteristic convective velocity scale  
    325 ($e.g.$, \citet{D'Alessio_al_JPO98}), $P_{LC}$ is assumed to be :  
     313By making an analogy with the characteristic convective velocity scale ($e.g.$, \citet{D'Alessio_al_JPO98}), 
     314$P_{LC}$ is assumed to be :  
    326315\begin{equation} 
    327316P_{LC}(z) = \frac{w_{LC}^3(z)}{H_{LC}} 
     
    330319With no information about the wave field, $w_{LC}$ is assumed to be proportional to  
    331320the Stokes drift $u_s = 0.377\,\,|\tau|^{1/2}$, where $|\tau|$ is the surface wind stress module  
    332 \footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity 
    333 may be expressed as $u_s =  0.016 \,|U_{10m}|$. Assuming an air density of  
    334 $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of $1.5~10^{-3}$ give the expression  
    335 used of $u_s$ as a function of the module of surface stress}.  
    336 For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as  
    337 at a finite depth $H_{LC}$ (which is often close to the mixed layer depth), and simply  
    338 varies as a sine function in between (a first-order profile for the Langmuir cell structures).  
     321\footnote{Following \citet{Li_Garrett_JMR93}, the surface Stoke drift velocity may be expressed as 
     322  $u_s =  0.016 \,|U_{10m}|$. 
     323  Assuming an air density of $\rho_a=1.22 \,Kg/m^3$ and a drag coefficient of 
     324  $1.5~10^{-3}$ give the expression used of $u_s$ as a function of the module of surface stress}. 
     325For the vertical variation, $w_{LC}$ is assumed to be zero at the surface as well as at 
     326a finite depth $H_{LC}$ (which is often close to the mixed layer depth), 
     327and simply varies as a sine function in between (a first-order profile for the Langmuir cell structures).  
    339328The resulting expression for $w_{LC}$ is : 
    340329\begin{equation} 
     
    344333                 \end{cases} 
    345334\end{equation} 
    346 where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise  
    347 to fit LES data. The chosen value yields maximum vertical velocities $w_{LC}$ of the order  
    348 of a few centimeters per second. The value of $c_{LC}$ is set through the \np{rn\_lc}  
    349 namelist parameter, having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.  
     335where $c_{LC} = 0.15$ has been chosen by \citep{Axell_JGR02} as a good compromise to fit LES data. 
     336The chosen value yields maximum vertical velocities $w_{LC}$ of the order of a few centimeters per second. 
     337The value of $c_{LC}$ is set through the \np{rn\_lc} namelist parameter, 
     338having in mind that it should stay between 0.15 and 0.54 \citep{Axell_JGR02}.  
    350339 
    351340The $H_{LC}$ is estimated in a similar way as the turbulent length scale of TKE equations: 
    352 $H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift 
    353 can reach on its own by converting its kinetic energy to potential energy, according to  
     341$H_{LC}$ is depth to which a water parcel with kinetic energy due to Stoke drift can reach on its own by 
     342converting its kinetic energy to potential energy, according to  
    354343\begin{equation} 
    355344- \int_{-H_{LC}}^0 { N^2\;z  \;dz} = \frac{1}{2} u_s^2 
     
    360349%--------------------------------------------------------------% 
    361350 
    362 Vertical mixing parameterizations commonly used in ocean general circulation models  
    363 tend to produce mixed-layer depths that are too shallow during summer months and windy conditions. 
    364 This bias is particularly acute over the Southern Ocean.  
    365 To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme  \cite{Rodgers_2014}.  
    366 The parameterization is an empirical one, $i.e.$ not derived from theoretical considerations,  
     351Vertical mixing parameterizations commonly used in ocean general circulation models tend to 
     352produce mixed-layer depths that are too shallow during summer months and windy conditions. 
     353This bias is particularly acute over the Southern Ocean. 
     354To overcome this systematic bias, an ad hoc parameterization is introduced into the TKE scheme \cite{Rodgers_2014}.  
     355The parameterization is an empirical one, $i.e.$ not derived from theoretical considerations, 
    367356but rather is meant to account for observed processes that affect the density structure of  
    368357the ocean’s planetary boundary layer that are not explicitly captured by default in the TKE scheme  
    369358($i.e.$ near-inertial oscillations and ocean swells and waves). 
    370359 
    371 When using this parameterization ($i.e.$ when \np{nn\_etau}\forcode{ = 1}), the TKE input to the ocean ($S$)  
    372 imposed by the winds in the form of near-inertial oscillations, swell and waves is parameterized  
    373 by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition, plus a depth depend one given by: 
     360When using this parameterization ($i.e.$ when \np{nn\_etau}\forcode{ = 1}), 
     361the TKE input to the ocean ($S$) imposed by the winds in the form of near-inertial oscillations, 
     362swell and waves is parameterized by \autoref{eq:ZDF_Esbc} the standard TKE surface boundary condition, 
     363plus a depth depend one given by: 
    374364\begin{equation}  \label{eq:ZDF_Ehtau} 
    375365S = (1-f_i) \; f_r \; e_s \; e^{-z / h_\tau}  
    376366\end{equation} 
    377 where  
    378 $z$ is the depth,   
    379 $e_s$ is TKE surface boundary condition,  
    380 $f_r$ is the fraction of the surface TKE that penetrate in the ocean,  
    381 $h_\tau$ is a vertical mixing length scale that controls exponential shape of the penetration,  
    382 and $f_i$ is the ice concentration (no penetration if $f_i=1$, that is if the ocean is entirely  
    383 covered by sea-ice). 
    384 The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter.  
    385 The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0})  
    386 or a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m  
    387 at high latitudes (\np{nn\_etau}\forcode{ = 1}).  
    388  
    389 Note that two other option existe, \np{nn\_etau}\forcode{ = 2..3}. They correspond to applying  
    390 \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer, or to using the high frequency part  
    391 of the stress to evaluate the fraction of TKE that penetrate the ocean.  
     367where $z$ is the depth, $e_s$ is TKE surface boundary condition, $f_r$ is the fraction of the surface TKE that 
     368penetrate in the ocean, $h_\tau$ is a vertical mixing length scale that controls exponential shape of 
     369the penetration, and $f_i$ is the ice concentration 
     370(no penetration if $f_i=1$, that is if the ocean is entirely covered by sea-ice). 
     371The value of $f_r$, usually a few percents, is specified through \np{rn\_efr} namelist parameter. 
     372The vertical mixing length scale, $h_\tau$, can be set as a 10~m uniform value (\np{nn\_etau}\forcode{ = 0}) or 
     373a latitude dependent value (varying from 0.5~m at the Equator to a maximum value of 30~m at high latitudes 
     374(\np{nn\_etau}\forcode{ = 1}).  
     375 
     376Note that two other option existe, \np{nn\_etau}\forcode{ = 2..3}. 
     377They correspond to applying \autoref{eq:ZDF_Ehtau} only at the base of the mixed layer, 
     378or to using the high frequency part of the stress to evaluate the fraction of TKE that penetrate the ocean.  
    392379Those two options are obsolescent features introduced for test purposes. 
    393380They will be removed in the next release.  
     
    420407%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    421408 
    422 The production of turbulence by vertical shear (the first term of the right hand side  
    423 of \autoref{eq:zdftke_e}) should balance the loss of kinetic energy associated with 
    424 the vertical momentum diffusion (first line in \autoref{eq:PE_zdf}). To do so a special care  
    425 have to be taken for both the time and space discretization of the TKE equation  
    426 \citep{Burchard_OM02,Marsaleix_al_OM08}. 
    427  
    428 Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows  
    429 how the two-level Leap-Frog time stepping of the momentum and tracer equations interplays  
    430 with the one-level forward time stepping of TKE equation. With this framework, the total loss  
    431 of kinetic energy (in 1D for the demonstration) due to the vertical momentum diffusion is  
    432 obtained by multiplying this quantity by $u^t$ and summing the result vertically:    
     409The production of turbulence by vertical shear (the first term of the right hand side of 
     410\autoref{eq:zdftke_e}) should balance the loss of kinetic energy associated with the vertical momentum diffusion 
     411(first line in \autoref{eq:PE_zdf}). 
     412To do so a special care have to be taken for both the time and space discretization of 
     413the TKE equation \citep{Burchard_OM02,Marsaleix_al_OM08}. 
     414 
     415Let us first address the time stepping issue. \autoref{fig:TKE_time_scheme} shows how 
     416the two-level Leap-Frog time stepping of the momentum and tracer equations interplays with 
     417the one-level forward time stepping of TKE equation. 
     418With this framework, the total loss of kinetic energy (in 1D for the demonstration) due to 
     419the vertical momentum diffusion is obtained by multiplying this quantity by $u^t$ and 
     420summing the result vertically:    
    433421\begin{equation} \label{eq:energ1} 
    434422\begin{split} 
     
    438426\end{split} 
    439427\end{equation} 
    440 Here, the vertical diffusion of momentum is discretized backward in time  
    441 with a coefficient, $K_m$, known at time $t$ (\autoref{fig:TKE_time_scheme}),  
    442 as it is required when using the TKE scheme (see \autoref{sec:STP_forward_imp}).  
    443 The first term of the right hand side of \autoref{eq:energ1} represents the kinetic energy  
    444 transfer at the surface (atmospheric forcing) and at the bottom (friction effect).  
    445 The second term is always negative. It is the dissipation rate of kinetic energy,  
    446 and thus minus the shear production rate of $\bar{e}$. \autoref{eq:energ1}  
    447 implies that, to be energetically consistent, the production rate of $\bar{e}$  
    448 used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as  
    449 ${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ (and not by the more straightforward  
    450 $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$). 
    451  
    452 A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification  
    453 (second term of the right hand side of \autoref{eq:zdftke_e}). This term  
    454 must balance the input of potential energy resulting from vertical mixing.  
    455 The rate of change of potential energy (in 1D for the demonstration) due vertical  
    456 mixing is obtained by multiplying vertical density diffusion  
    457 tendency by $g\,z$ and and summing the result vertically: 
     428Here, the vertical diffusion of momentum is discretized backward in time with a coefficient, $K_m$, 
     429known at time $t$ (\autoref{fig:TKE_time_scheme}), as it is required when using the TKE scheme 
     430(see \autoref{sec:STP_forward_imp}). 
     431The first term of the right hand side of \autoref{eq:energ1} represents the kinetic energy transfer at 
     432the surface (atmospheric forcing) and at the bottom (friction effect). 
     433The second term is always negative. 
     434It is the dissipation rate of kinetic energy, and thus minus the shear production rate of $\bar{e}$. 
     435\autoref{eq:energ1} implies that, to be energetically consistent, 
     436the production rate of $\bar{e}$ used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as 
     437${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ 
     438(and not by the more straightforward $K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$). 
     439 
     440A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification 
     441(second term of the right hand side of \autoref{eq:zdftke_e}). 
     442This term must balance the input of potential energy resulting from vertical mixing. 
     443The rate of change of potential energy (in 1D for the demonstration) due vertical mixing is obtained by 
     444multiplying vertical density diffusion tendency by $g\,z$ and and summing the result vertically: 
    458445\begin{equation} \label{eq:energ2} 
    459446\begin{split} 
     
    466453\end{equation} 
    467454where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.  
    468 The first term of the right hand side of \autoref{eq:energ2}  is always zero  
    469 because there is no diffusive flux through the ocean surface and bottom).  
    470 The second term is minus the destruction rate of  $\bar{e}$ due to stratification.  
    471 Therefore \autoref{eq:energ1} implies that, to be energetically consistent, the product  
    472 ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e}, the TKE equation. 
    473  
    474 Let us now address the space discretization issue.  
    475 The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity  
    476 components are in the centre of the side faces of a $t$-box in staggered C-grid  
    477 (\autoref{fig:cell}). A space averaging is thus required to obtain the shear TKE production term. 
    478 By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of  
    479 eddy coefficient by the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
    480 Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into  
    481 account. 
    482  
    483 The above energetic considerations leads to  
    484 the following final discrete form for the TKE equation: 
     455The first term of the right hand side of \autoref{eq:energ2} is always zero because 
     456there is no diffusive flux through the ocean surface and bottom). 
     457The second term is minus the destruction rate of  $\bar{e}$ due to stratification. 
     458Therefore \autoref{eq:energ1} implies that, to be energetically consistent, 
     459the product ${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \autoref{eq:zdftke_e}, the TKE equation. 
     460 
     461Let us now address the space discretization issue. 
     462The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity components are in 
     463the centre of the side faces of a $t$-box in staggered C-grid (\autoref{fig:cell}). 
     464A space averaging is thus required to obtain the shear TKE production term. 
     465By redoing the \autoref{eq:energ1} in the 3D case, it can be shown that the product of eddy coefficient by 
     466the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
     467Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into account. 
     468 
     469The above energetic considerations leads to the following final discrete form for the TKE equation: 
    485470\begin{equation} \label{eq:zdftke_ene} 
    486471\begin{split} 
     
    500485\end{split} 
    501486\end{equation} 
    502 where the last two terms in \autoref{eq:zdftke_ene} (vertical diffusion and Kolmogorov dissipation)  
    503 are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}).  
    504 Note that the Kolmogorov term has been linearized in time in order to render  
    505 the implicit computation possible. The restart of the TKE scheme  
    506 requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as they all appear in  
    507 the right hand side of \autoref{eq:zdftke_ene}. For the latter, it is in fact  
    508 the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
     487where the last two terms in \autoref{eq:zdftke_ene} (vertical diffusion and Kolmogorov dissipation) 
     488are time stepped using a backward scheme (see\autoref{sec:STP_forward_imp}). 
     489Note that the Kolmogorov term has been linearized in time in order to render the implicit computation possible. 
     490The restart of the TKE scheme requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as 
     491they all appear in the right hand side of \autoref{eq:zdftke_ene}. 
     492For the latter, it is in fact the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
    509493 
    510494% ------------------------------------------------------------------------------------------------------------- 
     
    519503%-------------------------------------------------------------------------------------------------------------- 
    520504 
    521 The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on  
    522 two prognostic equations: one for the turbulent kinetic energy $\bar {e}$, and another  
    523 for the generic length scale, $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}.  
    524 This later variable is defined as : $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
    525 where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover  
    526 a number of well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982},  
    527 $k$-$\epsilon$ \citep{Rodi_1987}, $k$-$\omega$ \citep{Wilcox_1988}  
    528 among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).  
     505The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on two prognostic equations: 
     506one for the turbulent kinetic energy $\bar {e}$, and another for the generic length scale, 
     507$\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}. 
     508This later variable is defined as: $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
     509where the triplet $(p, m, n)$ value given in Tab.\autoref{tab:GLS} allows to recover a number of 
     510well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982}, $k$-$\epsilon$ \citep{Rodi_1987}, 
     511$k$-$\omega$ \citep{Wilcox_1988} among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).  
    529512The GLS scheme is given by the following set of equations: 
    530513\begin{equation} \label{eq:zdfgls_e} 
     
    558541{\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
    559542\end{equation} 
    560 where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2})  
    561 and $\epsilon$ the dissipation rate.  
    562 The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$)  
    563 depends of the choice of the turbulence model. Four different turbulent models are pre-defined  
    564 (Tab.\autoref{tab:GLS}). They are made available through the \np{nn\_clo} namelist parameter.  
     543where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}) and 
     544$\epsilon$ the dissipation rate.  
     545The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$) depends of 
     546the choice of the turbulence model. 
     547Four different turbulent models are pre-defined (Tab.\autoref{tab:GLS}). 
     548They are made available through the \np{nn\_clo} namelist parameter.  
    565549 
    566550%--------------------------------------------------TABLE-------------------------------------------------- 
     
    584568\end{tabular} 
    585569\caption{   \protect\label{tab:GLS}  
    586 Set of predefined GLS parameters, or equivalently predefined turbulence models available  
    587 with \protect\key{zdfgls} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\ngn{namzdf\_gls} .} 
     570  Set of predefined GLS parameters, or equivalently predefined turbulence models available with 
     571  \protect\key{zdfgls} and controlled by the \protect\np{nn\_clos} namelist variable in \protect\ngn{namzdf\_gls}.} 
    588572\end{center}   \end{table} 
    589573%-------------------------------------------------------------------------------------------------------------- 
    590574 
    591 In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force 
    592 the convergence of the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length)  
    593 value near physical boundaries (logarithmic boundary layer law). $C_{\mu}$ and $C_{\mu'}$  
    594 are calculated from stability function proposed by \citet{Galperin_al_JAS88}, or by \citet{Kantha_Clayson_1994}  
    595 or one of the two functions suggested by \citet{Canuto_2001}  (\np{nn\_stab\_func}\forcode{ = 0..3}, resp.).  
     575In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force the convergence of 
     576the mixing length towards $K z_b$ ($K$: Kappa and $z_b$: rugosity length) value near physical boundaries 
     577(logarithmic boundary layer law). 
     578$C_{\mu}$ and $C_{\mu'}$ are calculated from stability function proposed by \citet{Galperin_al_JAS88}, 
     579or by \citet{Kantha_Clayson_1994} or one of the two functions suggested by \citet{Canuto_2001} 
     580(\np{nn\_stab\_func}\forcode{ = 0..3}, resp.).  
    596581The value of $C_{0\mu}$ depends of the choice of the stability function. 
    597582 
    598 The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated  
    599 thanks to Dirichlet or Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp.  
    600 As for TKE closure , the wave effect on the mixing is considered when \np{ln\_crban}\forcode{ = .true.} 
    601 \citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. The \np{rn\_crban} namelist parameter  
    602 is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and \np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.  
    603  
    604 The $\psi$ equation is known to fail in stably stratified flows, and for this reason  
    605 almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.  
    606 With this clipping, the maximum permissible length scale is determined by  
    607 $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. A value of $c_{lim} = 0.53$ is often used  
    608 \citep{Galperin_al_JAS88}. \cite{Umlauf_Burchard_CSR05} show that the value of  
    609 the clipping factor is of crucial importance for the entrainment depth predicted in  
    610 stably stratified situations, and that its value has to be chosen in accordance  
    611 with the algebraic model for the turbulent fluxes. The clipping is only activated  
    612 if \np{ln\_length\_lim}\forcode{ = .true.}, and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
    613  
    614 The time and space discretization of the GLS equations follows the same energetic  
    615 consideration as for the TKE case described in \autoref{subsec:ZDF_tke_ene}  \citep{Burchard_OM02}.  
     583The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated thanks to Dirichlet or 
     584Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp. 
     585As for TKE closure, the wave effect on the mixing is considered when 
     586\np{ln\_crban}\forcode{ = .true.} \citep{Craig_Banner_JPO94, Mellor_Blumberg_JPO04}. 
     587The \np{rn\_crban} namelist parameter is $\alpha_{CB}$ in \autoref{eq:ZDF_Esbc} and 
     588\np{rn\_charn} provides the value of $\beta$ in \autoref{eq:ZDF_Lsbc}.  
     589 
     590The $\psi$ equation is known to fail in stably stratified flows, and for this reason 
     591almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy. 
     592With this clipping, the maximum permissible length scale is determined by $l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. 
     593A value of $c_{lim} = 0.53$ is often used \citep{Galperin_al_JAS88}. 
     594\cite{Umlauf_Burchard_CSR05} show that the value of the clipping factor is of crucial importance for 
     595the entrainment depth predicted in stably stratified situations, 
     596and that its value has to be chosen in accordance with the algebraic model for the turbulent fluxes. 
     597The clipping is only activated if \np{ln\_length\_lim}\forcode{ = .true.}, 
     598and the $c_{lim}$ is set to the \np{rn\_clim\_galp} value. 
     599 
     600The time and space discretization of the GLS equations follows the same energetic consideration as for 
     601the TKE case described in \autoref{subsec:ZDF_tke_ene} \citep{Burchard_OM02}. 
    616602Examples of performance of the 4 turbulent closure scheme can be found in \citet{Warner_al_OM05}. 
    617603 
     
    640626%-------------------------------------------------------------------------------------------------------------- 
    641627 
    642 Static instabilities (i.e. light potential densities under heavy ones) may  
    643 occur at particular ocean grid points. In nature, convective processes  
    644 quickly re-establish the static stability of the water column. These  
    645 processes have been removed from the model via the hydrostatic  
    646 assumption so they must be parameterized. Three parameterisations  
    647 are available to deal with convective processes: a non-penetrative  
    648 convective adjustment or an enhanced vertical diffusion, or/and the  
    649 use of a turbulent closure scheme. 
     628Static instabilities (i.e. light potential densities under heavy ones) may occur at particular ocean grid points. 
     629In nature, convective processes quickly re-establish the static stability of the water column. 
     630These processes have been removed from the model via the hydrostatic assumption so they must be parameterized. 
     631Three parameterisations are available to deal with convective processes: 
     632a non-penetrative convective adjustment or an enhanced vertical diffusion, 
     633or/and the use of a turbulent closure scheme. 
    650634 
    651635% ------------------------------------------------------------------------------------------------------------- 
     
    665649\includegraphics[width=0.90\textwidth]{Fig_npc} 
    666650\caption{  \protect\label{fig:npc}  
    667 Example of an unstable density profile treated by the non penetrative  
    668 convective adjustment algorithm. $1^{st}$ step: the initial profile is checked from  
    669 the surface to the bottom. It is found to be unstable between levels 3 and 4.  
    670 They are mixed. The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5  
    671 are mixed. The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are  
    672 mixed. The $1^{st}$ step ends since the density profile is then stable below  
    673 the level 3. $2^{nd}$ step: the new $\rho$ profile is checked following the same  
    674 procedure as in $1^{st}$ step: levels 2 to 5 are mixed. The new density profile  
    675 is checked. It is found stable: end of algorithm.} 
     651  Example of an unstable density profile treated by the non penetrative convective adjustment algorithm. 
     652  $1^{st}$ step: the initial profile is checked from the surface to the bottom. 
     653  It is found to be unstable between levels 3 and 4. 
     654  They are mixed. 
     655  The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5 are mixed. 
     656  The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are mixed. 
     657  The $1^{st}$ step ends since the density profile is then stable below the level 3. 
     658  $2^{nd}$ step: the new $\rho$ profile is checked following the same procedure as in $1^{st}$ step: 
     659  levels 2 to 5 are mixed. 
     660  The new density profile is checked. 
     661  It is found stable: end of algorithm.} 
    676662\end{center}   \end{figure} 
    677663%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    678664 
    679 Options are defined through the  \ngn{namzdf} namelist variables. 
    680 The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}\forcode{ = .true.}.  
    681 It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously  
    682 the statically unstable portion of the water column, but only until the density  
    683 structure becomes neutrally stable ($i.e.$ until the mixed portion of the water  
    684 column has \textit{exactly} the density of the water just below) \citep{Madec_al_JPO91}.  
    685 The associated algorithm is an iterative process used in the following way  
    686 (\autoref{fig:npc}): starting from the top of the ocean, the first instability is  
    687 found. Assume in the following that the instability is located between levels  
    688 $k$ and $k+1$. The temperature and salinity in the two levels are  
    689 vertically mixed, conserving the heat and salt contents of the water column.  
    690 The new density is then computed by a linear approximation. If the new  
    691 density profile is still unstable between levels $k+1$ and $k+2$, levels $k$,  
    692 $k+1$ and $k+2$ are then mixed. This process is repeated until stability is  
    693 established below the level $k$ (the mixing process can go down to the  
    694 ocean bottom). The algorithm is repeated to check if the density profile  
    695 between level $k-1$ and $k$ is unstable and/or if there is no deeper instability. 
    696  
    697 This algorithm is significantly different from mixing statically unstable levels  
    698 two by two. The latter procedure cannot converge with a finite number  
    699 of iterations for some vertical profiles while the algorithm used in \NEMO  
    700 converges for any profile in a number of iterations which is less than the  
    701 number of vertical levels. This property is of paramount importance as  
    702 pointed out by \citet{Killworth1989}: it avoids the existence of permanent  
    703 and unrealistic static instabilities at the sea surface. This non-penetrative  
    704 convective algorithm has been proved successful in studies of the deep  
    705 water formation in the north-western Mediterranean Sea  
    706 \citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}. 
    707  
    708 The current implementation has been modified in order to deal with any non linear  
    709 equation of seawater (L. Brodeau, personnal communication).  
    710 Two main differences have been introduced compared to the original algorithm:  
     665Options are defined through the \ngn{namzdf} namelist variables. 
     666The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}\forcode{ = .true.}. 
     667It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously the statically unstable portion of 
     668the water column, but only until the density structure becomes neutrally stable 
     669($i.e.$ until the mixed portion of the water column has \textit{exactly} the density of the water just below) 
     670\citep{Madec_al_JPO91}. 
     671The associated algorithm is an iterative process used in the following way (\autoref{fig:npc}): 
     672starting from the top of the ocean, the first instability is found. 
     673Assume in the following that the instability is located between levels $k$ and $k+1$. 
     674The temperature and salinity in the two levels are vertically mixed, conserving the heat and salt contents of 
     675the water column. 
     676The new density is then computed by a linear approximation. 
     677If the new density profile is still unstable between levels $k+1$ and $k+2$, 
     678levels $k$, $k+1$ and $k+2$ are then mixed. 
     679This process is repeated until stability is established below the level $k$ 
     680(the mixing process can go down to the ocean bottom). 
     681The algorithm is repeated to check if the density profile between level $k-1$ and $k$ is unstable and/or 
     682if there is no deeper instability. 
     683 
     684This algorithm is significantly different from mixing statically unstable levels two by two. 
     685The latter procedure cannot converge with a finite number of iterations for some vertical profiles while 
     686the algorithm used in \NEMO converges for any profile in a number of iterations which is less than 
     687the number of vertical levels. 
     688This property is of paramount importance as pointed out by \citet{Killworth1989}: 
     689it avoids the existence of permanent and unrealistic static instabilities at the sea surface. 
     690This non-penetrative convective algorithm has been proved successful in studies of the deep water formation in 
     691the north-western Mediterranean Sea \citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}. 
     692 
     693The current implementation has been modified in order to deal with any non linear equation of seawater 
     694(L. Brodeau, personnal communication). 
     695Two main differences have been introduced compared to the original algorithm: 
    711696$(i)$ the stability is now checked using the Brunt-V\"{a}is\"{a}l\"{a} frequency  
    712 (not the the difference in potential density) ;  
    713 $(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients  
    714 are vertically mixed in the same way their temperature and salinity has been mixed. 
    715 These two modifications allow the algorithm to perform properly and accurately  
    716 with TEOS10 or EOS-80 without having to recompute the expansion coefficients at each  
    717 mixing iteration. 
     697(not the the difference in potential density);  
     698$(ii)$ when two levels are found unstable, their thermal and haline expansion coefficients are vertically mixed in 
     699the same way their temperature and salinity has been mixed. 
     700These two modifications allow the algorithm to perform properly and accurately with TEOS10 or EOS-80 without 
     701having to recompute the expansion coefficients at each mixing iteration. 
    718702 
    719703% ------------------------------------------------------------------------------------------------------------- 
     
    729713 
    730714Options are defined through the  \ngn{namzdf} namelist variables. 
    731 The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}.  
    732 In this case, the vertical eddy mixing coefficients are assigned very large values  
    733 (a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable  
    734 ($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative)  
    735 \citep{Lazar_PhD97, Lazar_al_JPO99}. This is done either on tracers only  
    736 (\np{nn\_evdm}\forcode{ = 0}) or on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}). 
    737  
    738 In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and  
    739 if \np{nn\_evdm}\forcode{ = 1}, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$  
    740 values also, are set equal to the namelist parameter \np{rn\_avevd}. A typical value  
    741 for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterisation of  
    742 convective processes is less time consuming than the convective adjustment  
    743 algorithm presented above when mixing both tracers and momentum in the  
    744 case of static instabilities. It requires the use of an implicit time stepping on  
    745 vertical diffusion terms (i.e. \np{ln\_zdfexp}\forcode{ = .false.}).  
    746  
    747 Note that the stability test is performed on both \textit{before} and \textit{now}  
    748 values of $N^2$. This removes a potential source of divergence of odd and 
    749 even time step in a leapfrog environment \citep{Leclair_PhD2010} (see \autoref{sec:STP_mLF}). 
     715The enhanced vertical diffusion parameterisation is used when \np{ln\_zdfevd}\forcode{ = .true.}. 
     716In this case, the vertical eddy mixing coefficients are assigned very large values 
     717(a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable 
     718($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar_PhD97, Lazar_al_JPO99}. 
     719This is done either on tracers only (\np{nn\_evdm}\forcode{ = 0}) or 
     720on both momentum and tracers (\np{nn\_evdm}\forcode{ = 1}). 
     721 
     722In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and if \np{nn\_evdm}\forcode{ = 1}, 
     723the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ values also, are set equal to 
     724the namelist parameter \np{rn\_avevd}. 
     725A typical value for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. 
     726This parameterisation of convective processes is less time consuming than 
     727the convective adjustment algorithm presented above when mixing both tracers and 
     728momentum in the case of static instabilities. 
     729It requires the use of an implicit time stepping on vertical diffusion terms 
     730(i.e. \np{ln\_zdfexp}\forcode{ = .false.}). 
     731 
     732Note that the stability test is performed on both \textit{before} and \textit{now} values of $N^2$. 
     733This removes a potential source of divergence of odd and even time step in 
     734a leapfrog environment \citep{Leclair_PhD2010} (see \autoref{sec:STP_mLF}). 
    750735 
    751736% ------------------------------------------------------------------------------------------------------------- 
     
    755740\label{subsec:ZDF_tcs} 
    756741 
    757 The turbulent closure scheme presented in \autoref{subsec:ZDF_tke} and \autoref{subsec:ZDF_gls}  
    758 (\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically  
    759 unstable density profiles. In such a case, the term corresponding to the  
    760 destruction of turbulent kinetic energy through stratification in \autoref{eq:zdftke_e}  
    761 or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative.  
    762 It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring  
    763 $A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1})$. These large values  
    764 restore the static stability of the water column in a way similar to that of the  
    765 enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}). However,  
    766 in the vicinity of the sea surface (first ocean layer), the eddy coefficients  
    767 computed by the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$,  
    768 because the mixing length scale is bounded by the distance to the sea surface.  
    769 It can thus be useful to combine the enhanced vertical  
    770 diffusion with the turbulent closure scheme, $i.e.$ setting the \np{ln\_zdfnpc}  
    771 namelist parameter to true and defining the turbulent closure CPP key all together. 
    772  
    773 The KPP turbulent closure scheme already includes enhanced vertical diffusion  
    774 in the case of convection, as governed by the variables $bvsqcon$ and $difcon$  
    775 found in \mdl{zdfkpp}, therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the KPP  
    776 scheme. %gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
     742The turbulent closure scheme presented in \autoref{subsec:ZDF_tke} and \autoref{subsec:ZDF_gls} 
     743(\key{zdftke} or \key{zdftke} is defined) in theory solves the problem of statically unstable density profiles. 
     744In such a case, the term corresponding to the destruction of turbulent kinetic energy through stratification in 
     745\autoref{eq:zdftke_e} or \autoref{eq:zdfgls_e} becomes a source term, since $N^2$ is negative.  
     746It results in large values of $A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring $A_u^{vm} {and}\;A_v^{vm}$ 
     747(up to $1\;m^2s^{-1}$). 
     748These large values restore the static stability of the water column in a way similar to that of 
     749the enhanced vertical diffusion parameterisation (\autoref{subsec:ZDF_evd}). 
     750However, in the vicinity of the sea surface (first ocean layer), the eddy coefficients computed by 
     751the turbulent closure scheme do not usually exceed $10^{-2}m.s^{-1}$, 
     752because the mixing length scale is bounded by the distance to the sea surface. 
     753It can thus be useful to combine the enhanced vertical diffusion with the turbulent closure scheme, 
     754$i.e.$ setting the \np{ln\_zdfnpc} namelist parameter to true and 
     755defining the turbulent closure CPP key all together. 
     756 
     757The KPP turbulent closure scheme already includes enhanced vertical diffusion in the case of convection, 
     758as governed by the variables $bvsqcon$ and $difcon$ found in \mdl{zdfkpp}, 
     759therefore \np{ln\_zdfevd}\forcode{ = .false.} should be used with the KPP scheme. 
     760% gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
    777761 
    778762% ================================================================ 
     
    788772 
    789773Options are defined through the  \ngn{namzdf\_ddm} namelist variables. 
    790 Double diffusion occurs when relatively warm, salty water overlies cooler, fresher  
    791 water, or vice versa. The former condition leads to salt fingering and the latter  
    792 to diffusive convection. Double-diffusive phenomena contribute to diapycnal  
    793 mixing in extensive regions of the ocean.  \citet{Merryfield1999} include a  
    794 parameterisation of such phenomena in a global ocean model and show that  
    795 it leads to relatively minor changes in circulation but exerts significant regional  
    796 influences on temperature and salinity. This parameterisation has been  
    797 introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key. 
     774Double diffusion occurs when relatively warm, salty water overlies cooler, fresher water, or vice versa. 
     775The former condition leads to salt fingering and the latter to diffusive convection. 
     776Double-diffusive phenomena contribute to diapycnal mixing in extensive regions of the ocean. 
     777\citet{Merryfield1999} include a parameterisation of such phenomena in a global ocean model and show that  
     778it leads to relatively minor changes in circulation but exerts significant regional influences on 
     779temperature and salinity. 
     780This parameterisation has been introduced in \mdl{zdfddm} module and is controlled by the \key{zdfddm} CPP key. 
    798781 
    799782Diapycnal mixing of S and T are described by diapycnal diffusion coefficients  
     
    802785    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS} 
    803786\end{align*} 
    804 where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,  
    805 and $o$ by processes other than double diffusion. The rates of double-diffusive  
    806 mixing depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,  
    807 where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline  
    808 contraction (see \autoref{subsec:TRA_eos}). To represent mixing of $S$ and $T$ by salt  
    809 fingering, we adopt the diapycnal diffusivities suggested by Schmitt (1981): 
     787where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection, 
     788and $o$ by processes other than double diffusion. 
     789The rates of double-diffusive mixing depend on the buoyancy ratio 
     790$R_\rho = \alpha \partial_z T / \beta \partial_z S$, where $\alpha$ and $\beta$ are coefficients of 
     791thermal expansion and saline contraction (see \autoref{subsec:TRA_eos}). 
     792To represent mixing of $S$ and $T$ by salt fingering, we adopt the diapycnal diffusivities suggested by Schmitt 
     793(1981): 
    810794\begin{align} \label{eq:zdfddm_f} 
    811795A_f^{vS} &=    \begin{cases} 
     
    821805\includegraphics[width=0.99\textwidth]{Fig_zdfddm} 
    822806\caption{  \protect\label{fig:zdfddm} 
    823 From \citet{Merryfield1999} : (a) Diapycnal diffusivities $A_f^{vT}$  
    824 and $A_f^{vS}$ for temperature and salt in regions of salt fingering. Heavy  
    825 curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves  
    826 $A^{\ast v} = 10^{-4}~m^2.s^{-1}$ ; (b) diapycnal diffusivities $A_d^{vT}$ and  
    827 $A_d^{vS}$ for temperature and salt in regions of diffusive convection. Heavy  
    828 curves denote the Federov parameterisation and thin curves the Kelley  
    829 parameterisation. The latter is not implemented in \NEMO. } 
     807  From \citet{Merryfield1999} : 
     808  (a) Diapycnal diffusivities $A_f^{vT}$ and $A_f^{vS}$ for temperature and salt in regions of salt fingering. 
     809  Heavy curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves $A^{\ast v} = 10^{-4}~m^2.s^{-1}$; 
     810  (b) diapycnal diffusivities $A_d^{vT}$ and $A_d^{vS}$ for temperature and salt in regions of diffusive convection. 
     811  Heavy curves denote the Federov parameterisation and thin curves the Kelley parameterisation. 
     812  The latter is not implemented in \NEMO. } 
    830813\end{center}    \end{figure} 
    831814%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    832815 
    833 The factor 0.7 in \autoref{eq:zdfddm_f_T} reflects the measured ratio  
    834 $\alpha F_T /\beta F_S \approx  0.7$ of buoyancy flux of heat to buoyancy  
    835 flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}). Following  \citet{Merryfield1999},  
    836 we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
    837  
    838 To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:  
     816The factor 0.7 in \autoref{eq:zdfddm_f_T} reflects the measured ratio $\alpha F_T /\beta F_S \approx  0.7$ of 
     817buoyancy flux of heat to buoyancy flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}). 
     818Following  \citet{Merryfield1999}, we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
     819 
     820To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by 
     821Federov (1988) is used:  
    839822\begin{align}  \label{eq:zdfddm_d} 
    840823A_d^{vT} &=    \begin{cases} 
     
    853836\end{align} 
    854837 
    855 The dependencies of \autoref{eq:zdfddm_f} to \autoref{eq:zdfddm_d_S} on $R_\rho$  
    856 are illustrated in \autoref{fig:zdfddm}. Implementing this requires computing  
    857 $R_\rho$ at each grid point on every time step. This is done in \mdl{eosbn2} at the  
    858 same time as $N^2$ is computed. This avoids duplication in the computation of  
    859 $\alpha$ and $\beta$ (which is usually quite expensive). 
     838The dependencies of \autoref{eq:zdfddm_f} to \autoref{eq:zdfddm_d_S} on $R_\rho$ are illustrated in 
     839\autoref{fig:zdfddm}. 
     840Implementing this requires computing $R_\rho$ at each grid point on every time step. 
     841This is done in \mdl{eosbn2} at the same time as $N^2$ is computed. 
     842This avoids duplication in the computation of $\alpha$ and $\beta$ (which is usually quite expensive). 
    860843 
    861844% ================================================================ 
     
    870853%-------------------------------------------------------------------------------------------------------------- 
    871854 
    872 Options to define the top and bottom friction are defined through the  \ngn{nambfr} namelist variables. 
    873 The bottom friction represents the friction generated by the bathymetry.  
    874 The top friction represents the friction generated by the ice shelf/ocean interface.  
    875 As the friction processes at the top and bottom are treated in similar way,  
     855Options to define the top and bottom friction are defined through the \ngn{nambfr} namelist variables. 
     856The bottom friction represents the friction generated by the bathymetry. 
     857The top friction represents the friction generated by the ice shelf/ocean interface. 
     858As the friction processes at the top and bottom are treated in similar way, 
    876859only the bottom friction is described in detail below. 
    877860 
    878861 
    879 Both the surface momentum flux (wind stress) and the bottom momentum  
    880 flux (bottom friction) enter the equations as a condition on the vertical  
    881 diffusive flux. For the bottom boundary layer, one has: 
     862Both the surface momentum flux (wind stress) and the bottom momentum flux (bottom friction) enter the equations as 
     863a condition on the vertical diffusive flux. 
     864For the bottom boundary layer, one has: 
    882865\begin{equation} \label{eq:zdfbfr_flux} 
    883866A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
    884867\end{equation} 
    885 where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum  
    886 outside the logarithmic turbulent boundary layer (thickness of the order of  
    887 1~m in the ocean). How ${\cal F}_h^{\textbf U}$ influences the interior depends on the  
    888 vertical resolution of the model near the bottom relative to the Ekman layer  
    889 depth. For example, in order to obtain an Ekman layer depth  
    890 $d = \sqrt{2\;A^{vm}} / f = 50$~m, one needs a vertical diffusion coefficient  
    891 $A^{vm} = 0.125$~m$^2$s$^{-1}$ (for a Coriolis frequency  
    892 $f = 10^{-4}$~m$^2$s$^{-1}$). With a background diffusion coefficient  
    893 $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.  
    894 When the vertical mixing coefficient is this small, using a flux condition is  
    895 equivalent to entering the viscous forces (either wind stress or bottom friction)  
    896 as a body force over the depth of the top or bottom model layer. To illustrate  
    897 this, consider the equation for $u$ at $k$, the last ocean level: 
     868where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum outside 
     869the logarithmic turbulent boundary layer (thickness of the order of 1~m in the ocean). 
     870How ${\cal F}_h^{\textbf U}$ influences the interior depends on the vertical resolution of the model near 
     871the bottom relative to the Ekman layer depth. 
     872For example, in order to obtain an Ekman layer depth $d = \sqrt{2\;A^{vm}} / f = 50$~m, 
     873one needs a vertical diffusion coefficient $A^{vm} = 0.125$~m$^2$s$^{-1}$ 
     874(for a Coriolis frequency $f = 10^{-4}$~m$^2$s$^{-1}$). 
     875With a background diffusion coefficient $A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.  
     876When the vertical mixing coefficient is this small, using a flux condition is equivalent to 
     877entering the viscous forces (either wind stress or bottom friction) as a body force over the depth of the top or 
     878bottom model layer. 
     879To illustrate this, consider the equation for $u$ at $k$, the last ocean level: 
    898880\begin{equation} \label{eq:zdfbfr_flux2} 
    899881\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}} 
    900882\end{equation} 
    901 If the bottom layer thickness is 200~m, the Ekman transport will  
    902 be distributed over that depth. On the other hand, if the vertical resolution  
    903 is high (1~m or less) and a turbulent closure model is used, the turbulent  
    904 Ekman layer will be represented explicitly by the model. However, the  
    905 logarithmic layer is never represented in current primitive equation model  
    906 applications: it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.  
    907 Two choices are available in \NEMO: a linear and a quadratic bottom friction.  
    908 Note that in both cases, the rotation between the interior velocity and the  
    909 bottom friction is neglected in the present release of \NEMO. 
    910  
    911 In the code, the bottom friction is imposed by adding the trend due to the bottom  
    912 friction to the general momentum trend in \mdl{dynbfr}. For the time-split surface  
    913 pressure gradient algorithm, the momentum trend due to the barotropic component  
    914 needs to be handled separately. For this purpose it is convenient to compute and  
    915 store coefficients which can be simply combined with bottom velocities and geometric  
    916 values to provide the momentum trend due to bottom friction.  
    917 These coefficients are computed in \mdl{zdfbfr} and generally take the form  
    918 $c_b^{\textbf U}$ where: 
     883If the bottom layer thickness is 200~m, the Ekman transport will be distributed over that depth. 
     884On the other hand, if the vertical resolution is high (1~m or less) and a turbulent closure model is used, 
     885the turbulent Ekman layer will be represented explicitly by the model. 
     886However, the logarithmic layer is never represented in current primitive equation model applications: 
     887it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $. 
     888Two choices are available in \NEMO: a linear and a quadratic bottom friction. 
     889Note that in both cases, the rotation between the interior velocity and the bottom friction is neglected in 
     890the present release of \NEMO. 
     891 
     892In the code, the bottom friction is imposed by adding the trend due to the bottom friction to 
     893the general momentum trend in \mdl{dynbfr}. 
     894For the time-split surface pressure gradient algorithm, the momentum trend due to 
     895the barotropic component needs to be handled separately. 
     896For this purpose it is convenient to compute and store coefficients which can be simply combined with 
     897bottom velocities and geometric values to provide the momentum trend due to bottom friction. 
     898These coefficients are computed in \mdl{zdfbfr} and generally take the form $c_b^{\textbf U}$ where: 
    919899\begin{equation} \label{eq:zdfbfr_bdef} 
    920900\frac{\partial {\textbf U_h}}{\partial t} =  
     
    929909\label{subsec:ZDF_bfr_linear} 
    930910 
    931 The linear bottom friction parameterisation (including the special case  
    932 of a free-slip condition) assumes that the bottom friction  
    933 is proportional to the interior velocity (i.e. the velocity of the last  
    934 model level): 
     911The linear bottom friction parameterisation (including the special case of a free-slip condition) assumes that 
     912the bottom friction is proportional to the interior velocity (i.e. the velocity of the last model level): 
    935913\begin{equation} \label{eq:zdfbfr_linear} 
    936914{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
    937915\end{equation} 
    938 where $r$ is a friction coefficient expressed in ms$^{-1}$.  
    939 This coefficient is generally estimated by setting a typical decay time  
    940 $\tau$ in the deep ocean,  
    941 and setting $r = H / \tau$, where $H$ is the ocean depth. Commonly accepted  
    942 values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}.  
    943 A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used  
    944 in quasi-geostrophic models. One may consider the linear friction as an  
    945 approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982},  
    946 Eq. 9.6.6). For example, with a drag coefficient $C_D = 0.002$, a typical speed  
    947 of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, and assuming an ocean depth  
    948 $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.  
    949 This is the default value used in \NEMO. It corresponds to a decay time scale  
    950 of 115~days. It can be changed by specifying \np{rn\_bfri1} (namelist parameter). 
    951  
    952 For the linear friction case the coefficients defined in the general  
    953 expression \autoref{eq:zdfbfr_bdef} are:  
     916where $r$ is a friction coefficient expressed in ms$^{-1}$. 
     917This coefficient is generally estimated by setting a typical decay time $\tau$ in the deep ocean,  
     918and setting $r = H / \tau$, where $H$ is the ocean depth. 
     919Commonly accepted values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}. 
     920A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used in quasi-geostrophic models. 
     921One may consider the linear friction as an approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ 
     922(\citet{Gill1982}, Eq. 9.6.6). 
     923For example, with a drag coefficient $C_D = 0.002$, a typical speed of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, 
     924and assuming an ocean depth $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$. 
     925This is the default value used in \NEMO. It corresponds to a decay time scale of 115~days. 
     926It can be changed by specifying \np{rn\_bfri1} (namelist parameter). 
     927 
     928For the linear friction case the coefficients defined in the general expression \autoref{eq:zdfbfr_bdef} are:  
    954929\begin{equation} \label{eq:zdfbfr_linbfr_b} 
    955930\begin{split} 
     
    958933\end{split} 
    959934\end{equation} 
    960 When \np{nn\_botfr}\forcode{ = 1}, the value of $r$ used is \np{rn\_bfri1}.  
    961 Setting \np{nn\_botfr}\forcode{ = 0} is equivalent to setting $r=0$ and leads to a free-slip  
    962 bottom boundary condition. These values are assigned in \mdl{zdfbfr}.  
    963 From v3.2 onwards there is support for local enhancement of these values  
    964 via an externally defined 2D mask array (\np{ln\_bfr2d}\forcode{ = .true.}) given 
    965 in the \ifile{bfr\_coef} input NetCDF file. The mask values should vary from 0 to 1.  
    966 Locations with a non-zero mask value will have the friction coefficient increased  
    967 by $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri1}. 
     935When \np{nn\_botfr}\forcode{ = 1}, the value of $r$ used is \np{rn\_bfri1}. 
     936Setting \np{nn\_botfr}\forcode{ = 0} is equivalent to setting $r=0$ and 
     937leads to a free-slip bottom boundary condition. 
     938These values are assigned in \mdl{zdfbfr}. 
     939From v3.2 onwards there is support for local enhancement of these values via an externally defined 2D mask array 
     940(\np{ln\_bfr2d}\forcode{ = .true.}) given in the \ifile{bfr\_coef} input NetCDF file. 
     941The mask values should vary from 0 to 1. 
     942Locations with a non-zero mask value will have the friction coefficient increased by 
     943$mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri1}. 
    968944 
    969945% ------------------------------------------------------------------------------------------------------------- 
     
    973949\label{subsec:ZDF_bfr_nonlinear} 
    974950 
    975 The non-linear bottom friction parameterisation assumes that the bottom  
    976 friction is quadratic:  
     951The non-linear bottom friction parameterisation assumes that the bottom friction is quadratic:  
    977952\begin{equation} \label{eq:zdfbfr_nonlinear} 
    978953{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h  
    979954}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b  
    980955\end{equation} 
    981 where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy  
    982 due to tides, internal waves breaking and other short time scale currents.  
    983 A typical value of the drag coefficient is $C_D = 10^{-3} $. As an example,  
    984 the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and  
    985 $e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992}  
    986 uses $C_D = 1.4\;10^{-3}$ and $e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.  
    987 The CME choices have been set as default values (\np{rn\_bfri2} and \np{rn\_bfeb2}  
    988 namelist parameters). 
    989  
    990 As for the linear case, the bottom friction is imposed in the code by  
    991 adding the trend due to the bottom friction to the general momentum trend  
    992 in \mdl{dynbfr}. 
    993 For the non-linear friction case the terms 
    994 computed in \mdl{zdfbfr}  are:  
     956where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy due to tides, 
     957internal waves breaking and other short time scale currents. 
     958A typical value of the drag coefficient is $C_D = 10^{-3} $. 
     959As an example, the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and 
     960$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992} uses $C_D = 1.4\;10^{-3}$ and 
     961$e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$. 
     962The CME choices have been set as default values (\np{rn\_bfri2} and \np{rn\_bfeb2} namelist parameters). 
     963 
     964As for the linear case, the bottom friction is imposed in the code by adding the trend due to 
     965the bottom friction to the general momentum trend in \mdl{dynbfr}. 
     966For the non-linear friction case the terms computed in \mdl{zdfbfr} are: 
    995967\begin{equation} \label{eq:zdfbfr_nonlinbfr} 
    996968\begin{split} 
     
    1000972\end{equation} 
    1001973 
    1002 The coefficients that control the strength of the non-linear bottom friction are 
    1003 initialised as namelist parameters: $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}. 
    1004 Note for applications which treat tides explicitly a low or even zero value of 
    1005 \np{rn\_bfeb2} is recommended. From v3.2 onwards a local enhancement of $C_D$ is possible 
    1006 via an externally defined 2D mask array (\np{ln\_bfr2d}\forcode{ = .true.}).  This works in the same way 
    1007 as for the linear bottom friction case with non-zero masked locations increased by 
     974The coefficients that control the strength of the non-linear bottom friction are initialised as namelist parameters: 
     975$C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}. 
     976Note for applications which treat tides explicitly a low or even zero value of \np{rn\_bfeb2} is recommended. 
     977From v3.2 onwards a local enhancement of $C_D$ is possible via an externally defined 2D mask array 
     978(\np{ln\_bfr2d}\forcode{ = .true.}). 
     979This works in the same way as for the linear bottom friction case with non-zero masked locations increased by 
    1008980$mask\_value$*\np{rn\_bfrien}*\np{rn\_bfri2}. 
    1009981 
     
    1015987\label{subsec:ZDF_bfr_loglayer} 
    1016988 
    1017 In the non-linear bottom friction case, the drag coefficient, $C_D$, can be optionally 
    1018 enhanced using a "law of the wall" scaling. If  \np{ln\_loglayer} = .true., $C_D$ is no 
    1019 longer constant but is related to the thickness of the last wet layer in each column by: 
    1020  
     989In the non-linear bottom friction case, the drag coefficient, $C_D$, can be optionally enhanced using 
     990a "law of the wall" scaling. 
     991If  \np{ln\_loglayer} = .true., $C_D$ is no longer constant but is related to the thickness of 
     992the last wet layer in each column by: 
    1021993\begin{equation} 
    1022994C_D = \left ( {\kappa \over {\rm log}\left ( 0.5e_{3t}/rn\_bfrz0 \right ) } \right )^2 
    1023995\end{equation} 
    1024996 
    1025 \noindent where $\kappa$ is the von-Karman constant and \np{rn\_bfrz0} is a roughness 
    1026 length provided via the namelist. 
     997\noindent where $\kappa$ is the von-Karman constant and \np{rn\_bfrz0} is a roughness length provided via 
     998the namelist. 
    1027999 
    10281000For stability, the drag coefficient is bounded such that it is kept greater or equal to 
    1029 the base \np{rn\_bfri2} value and it is not allowed to exceed the value of an additional 
    1030 namelist parameter: \np{rn\_bfri2\_max}, i.e.: 
    1031  
     1001the base \np{rn\_bfri2} value and it is not allowed to exceed the value of an additional namelist parameter: 
     1002\np{rn\_bfri2\_max}, i.e.: 
    10321003\begin{equation} 
    10331004rn\_bfri2 \leq C_D \leq rn\_bfri2\_max 
    10341005\end{equation} 
    10351006 
    1036 \noindent Note also that a log-layer enhancement can also be applied to the top boundary 
    1037 friction if under ice-shelf cavities are in use (\np{ln\_isfcav}\forcode{ = .true.}).  In this case, the 
    1038 relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} 
    1039 and \np{rn\_tfri2\_max}. 
     1007\noindent Note also that a log-layer enhancement can also be applied to the top boundary friction if 
     1008under ice-shelf cavities are in use (\np{ln\_isfcav}\forcode{ = .true.}). 
     1009In this case, the relevant namelist parameters are \np{rn\_tfrz0}, \np{rn\_tfri2} and \np{rn\_tfri2\_max}. 
    10401010 
    10411011% ------------------------------------------------------------------------------------------------------------- 
     
    10451015\label{subsec:ZDF_bfr_stability} 
    10461016 
    1047 Some care needs to exercised over the choice of parameters to ensure that the 
    1048 implementation of bottom friction does not induce numerical instability. For  
    1049 the purposes of stability analysis, an approximation to \autoref{eq:zdfbfr_flux2} 
    1050 is: 
     1017Some care needs to exercised over the choice of parameters to ensure that the implementation of 
     1018bottom friction does not induce numerical instability. 
     1019For the purposes of stability analysis, an approximation to \autoref{eq:zdfbfr_flux2} is: 
    10511020\begin{equation} \label{eq:Eqn_bfrstab} 
    10521021\begin{split} 
     
    10551024\end{split} 
    10561025\end{equation} 
    1057 \noindent where linear bottom friction and a leapfrog timestep have been assumed.  
     1026\noindent where linear bottom friction and a leapfrog timestep have been assumed. 
    10581027To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have: 
    10591028\begin{equation} 
     
    10641033r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\ 
    10651034\end{equation} 
    1066 This same inequality can also be derived in the non-linear bottom friction case  
    1067 if a velocity of 1 m.s$^{-1}$ is assumed. Alternatively, this criterion can be  
    1068 rearranged to suggest a minimum bottom box thickness to ensure stability: 
     1035This same inequality can also be derived in the non-linear bottom friction case if 
     1036a velocity of 1 m.s$^{-1}$ is assumed. 
     1037Alternatively, this criterion can be rearranged to suggest a minimum bottom box thickness to ensure stability: 
    10691038\begin{equation} 
    10701039e_{3u} > 2\;r\;\rdt 
    10711040\end{equation} 
    1072 \noindent which it may be necessary to impose if partial steps are being used.  
    1073 For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then 
    1074 $e_{3u}$ should be greater than 3.6 m. For most applications, with physically 
    1075 sensible parameters these restrictions should not be of concern. But  
    1076 caution may be necessary if attempts are made to locally enhance the bottom 
    1077 friction parameters.  
    1078 To ensure stability limits are imposed on the bottom friction coefficients both during  
    1079 initialisation and at each time step. Checks at initialisation are made in \mdl{zdfbfr}  
    1080 (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
    1081 The number of breaches of the stability criterion are reported as well as the minimum  
    1082 and maximum values that have been set. The criterion is also checked at each time step,  
    1083 using the actual velocity, in \mdl{dynbfr}. Values of the bottom friction coefficient are  
    1084 reduced as necessary to ensure stability; these changes are not reported. 
     1041\noindent which it may be necessary to impose if partial steps are being used. 
     1042For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then $e_{3u}$ should be greater than 3.6 m. 
     1043For most applications, with physically sensible parameters these restrictions should not be of concern. 
     1044But caution may be necessary if attempts are made to locally enhance the bottom friction parameters.  
     1045To ensure stability limits are imposed on the bottom friction coefficients both 
     1046during initialisation and at each time step. 
     1047Checks at initialisation are made in \mdl{zdfbfr} (assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
     1048The number of breaches of the stability criterion are reported as well as 
     1049the minimum and maximum values that have been set. 
     1050The criterion is also checked at each time step, using the actual velocity, in \mdl{dynbfr}. 
     1051Values of the bottom friction coefficient are reduced as necessary to ensure stability; 
     1052these changes are not reported. 
    10851053 
    10861054Limits on the bottom friction coefficient are not imposed if the user has elected to 
    1087 handle the bottom friction implicitly (see \autoref{subsec:ZDF_bfr_imp}). The number of potential 
    1088 breaches of the explicit stability criterion are still reported for information purposes. 
     1055handle the bottom friction implicitly (see \autoref{subsec:ZDF_bfr_imp}). 
     1056The number of potential breaches of the explicit stability criterion are still reported for information purposes. 
    10891057 
    10901058% ------------------------------------------------------------------------------------------------------------- 
     
    10941062\label{subsec:ZDF_bfr_imp} 
    10951063 
    1096 An optional implicit form of bottom friction has been implemented to improve 
    1097 model stability. We recommend this option for shelf sea and coastal ocean applications, especially  
    1098 for split-explicit time splitting. This option can be invoked by setting \np{ln\_bfrimp}  
    1099 to \forcode{.true.} in the \textit{nambfr} namelist. This option requires \np{ln\_zdfexp} to be \forcode{.false.}  
    1100 in the \textit{namzdf} namelist.  
    1101  
    1102 This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, the  
    1103 bottom boundary condition is implemented implicitly. 
     1064An optional implicit form of bottom friction has been implemented to improve model stability. 
     1065We recommend this option for shelf sea and coastal ocean applications, especially for split-explicit time splitting. 
     1066This option can be invoked by setting \np{ln\_bfrimp} to \forcode{.true.} in the \textit{nambfr} namelist. 
     1067This option requires \np{ln\_zdfexp} to be \forcode{.false.} in the \textit{namzdf} namelist.  
     1068 
     1069This implementation is realised in \mdl{dynzdf\_imp} and \mdl{dynspg\_ts}. In \mdl{dynzdf\_imp}, 
     1070the bottom boundary condition is implemented implicitly. 
    11041071 
    11051072\begin{equation} \label{eq:dynzdf_bfr} 
     
    11081075\end{equation} 
    11091076 
    1110 where $mbk$ is the layer number of the bottom wet layer. superscript $n+1$ means the velocity used in the 
    1111 friction formula is to be calculated, so, it is implicit. 
    1112  
    1113 If split-explicit time splitting is used, care must be taken to avoid the double counting of 
    1114 the bottom friction in the 2-D barotropic momentum equations. As NEMO only updates the barotropic  
    1115 pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, we need to remove 
    1116 the bottom friction induced by these two terms which has been included in the 3-D momentum trend  
    1117 and update it with the latest value. On the other hand, the bottom friction contributed by the 
    1118 other terms (e.g. the advection term, viscosity term) has been included in the 3-D momentum equations 
    1119 and should not be added in the 2-D barotropic mode. 
    1120  
    1121 The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the 
    1122 following: 
     1077where $mbk$ is the layer number of the bottom wet layer. 
     1078Superscript $n+1$ means the velocity used in the friction formula is to be calculated, so, it is implicit. 
     1079 
     1080If split-explicit time splitting is used, care must be taken to avoid the double counting of the bottom friction in 
     1081the 2-D barotropic momentum equations. 
     1082As NEMO only updates the barotropic pressure gradient and Coriolis' forcing terms in the 2-D barotropic calculation, 
     1083we need to remove the bottom friction induced by these two terms which has been included in the 3-D momentum trend  
     1084and update it with the latest value. 
     1085On the other hand, the bottom friction contributed by the other terms 
     1086(e.g. the advection term, viscosity term) has been included in the 3-D momentum equations and 
     1087should not be added in the 2-D barotropic mode. 
     1088 
     1089The implementation of the implicit bottom friction in \mdl{dynspg\_ts} is done in two steps as the following: 
    11231090 
    11241091\begin{equation} \label{eq:dynspg_ts_bfr1} 
     
    11321099\end{equation} 
    11331100 
    1134 where $\textbf{T}$ is the vertical integrated 3-D momentum trend. We assume the leap-frog time-stepping 
    1135 is used here. $\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step. 
    1136  $c_{b}$ is the friction coefficient. $\eta$ is the sea surface level calculated in the barotropic loops 
    1137 while $\eta^{'}$ is the sea surface level used in the 3-D baroclinic mode. $\textbf{u}_{b}$ is the bottom 
    1138 layer horizontal velocity. 
     1101where $\textbf{T}$ is the vertical integrated 3-D momentum trend. 
     1102We assume the leap-frog time-stepping is used here. 
     1103$\Delta t$ is the barotropic mode time step and $\Delta t_{bc}$ is the baroclinic mode time step. 
     1104$c_{b}$ is the friction coefficient. 
     1105$\eta$ is the sea surface level calculated in the barotropic loops while $\eta^{'}$ is the sea surface level used in 
     1106the 3-D baroclinic mode. 
     1107$\textbf{u}_{b}$ is the bottom layer horizontal velocity. 
    11391108 
    11401109 
     
    11481117\label{subsec:ZDF_bfr_ts} 
    11491118 
    1150 When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, the 
    1151 bottom velocity at the before time step is used. This velocity includes both the 
    1152 baroclinic and barotropic components which is appropriate when using either the 
    1153 explicit or filtered surface pressure gradient algorithms (\key{dynspg\_exp} or  
    1154 \key{dynspg\_flt}). Extra attention is required, however, when using  
    1155 split-explicit time stepping (\key{dynspg\_ts}). In this case the free surface  
    1156 equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, while the three  
    1157 dimensional prognostic variables are solved with the longer time step  
    1158 of \np{rn\_rdt} seconds. The trend in the barotropic momentum due to bottom  
    1159 friction appropriate to this method is that given by the selected parameterisation  
    1160 ($i.e.$ linear or non-linear bottom friction) computed with the evolving velocities  
    1161 at each barotropic timestep.  
    1162  
    1163 In the case of non-linear bottom friction, we have elected to partially linearise  
    1164 the problem by keeping the coefficients fixed throughout the barotropic  
    1165 time-stepping to those computed in \mdl{zdfbfr} using the now timestep.  
     1119When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, 
     1120the bottom velocity at the before time step is used. 
     1121This velocity includes both the baroclinic and barotropic components which is appropriate when 
     1122using either the explicit or filtered surface pressure gradient algorithms 
     1123(\key{dynspg\_exp} or \key{dynspg\_flt}). 
     1124Extra attention is required, however, when using split-explicit time stepping (\key{dynspg\_ts}). 
     1125In this case the free surface equation is solved with a small time step \np{rn\_rdt}/\np{nn\_baro}, 
     1126while the three dimensional prognostic variables are solved with the longer time step of \np{rn\_rdt} seconds. 
     1127The trend in the barotropic momentum due to bottom friction appropriate to this method is that given by 
     1128the selected parameterisation ($i.e.$ linear or non-linear bottom friction) computed with 
     1129the evolving velocities at each barotropic timestep.  
     1130 
     1131In the case of non-linear bottom friction, we have elected to partially linearise the problem by 
     1132keeping the coefficients fixed throughout the barotropic time-stepping to those computed in 
     1133\mdl{zdfbfr} using the now timestep. 
    11661134This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to: 
    11671135 
    11681136\begin{enumerate} 
    1169 \item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before 
    1170 barotropic velocity to the bottom friction component of the vertically 
    1171 integrated momentum trend. Note the same stability check that is carried out  
    1172 on the bottom friction coefficient in \mdl{dynbfr} has to be applied here to 
    1173 ensure that the trend removed matches that which was added in \mdl{dynbfr}. 
    1174 \item At each barotropic step, compute the contribution of the current barotropic 
    1175 velocity to the trend due to bottom friction. Add this contribution to the 
    1176 vertically integrated momentum trend. This contribution is handled implicitly which 
    1177 eliminates the need to impose a stability criteria on the values of the bottom friction 
    1178 coefficient within the barotropic loop.  
     1137\item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before barotropic velocity to 
     1138  the bottom friction component of the vertically integrated momentum trend. 
     1139  Note the same stability check that is carried out on the bottom friction coefficient in \mdl{dynbfr} has to 
     1140  be applied here to ensure that the trend removed matches that which was added in \mdl{dynbfr}. 
     1141\item At each barotropic step, compute the contribution of the current barotropic velocity to 
     1142  the trend due to bottom friction. 
     1143  Add this contribution to the vertically integrated momentum trend. 
     1144  This contribution is handled implicitly which eliminates the need to impose a stability criteria on 
     1145  the values of the bottom friction coefficient within the barotropic loop.  
    11791146\end{enumerate} 
    11801147 
    1181 Note that the use of an implicit formulation within the barotropic loop 
    1182 for the bottom friction trend means that any limiting of the bottom friction coefficient  
    1183 in \mdl{dynbfr} does not adversely affect the solution when using split-explicit time  
    1184 splitting. This is because the major contribution to bottom friction is likely to come from  
    1185 the barotropic component which uses the unrestricted value of the coefficient. However, if the 
    1186 limiting is thought to be having a major effect (a more likely prospect in coastal and shelf seas 
    1187 applications) then the fully implicit form of the bottom friction should be used (see \autoref{subsec:ZDF_bfr_imp} )  
     1148Note that the use of an implicit formulation within the barotropic loop for the bottom friction trend means that 
     1149any limiting of the bottom friction coefficient in \mdl{dynbfr} does not adversely affect the solution when 
     1150using split-explicit time splitting. 
     1151This is because the major contribution to bottom friction is likely to come from the barotropic component which 
     1152uses the unrestricted value of the coefficient. 
     1153However, if the limiting is thought to be having a major effect 
     1154(a more likely prospect in coastal and shelf seas applications) then 
     1155the fully implicit form of the bottom friction should be used (see \autoref{subsec:ZDF_bfr_imp}) 
    11881156which can be selected by setting \np{ln\_bfrimp} $=$ \forcode{.true.}. 
    11891157 
     
    11931161\end{equation} 
    11941162where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),  
    1195 $c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and $RHS$ represents  
    1196 all the components to the vertically integrated momentum trend except for that due to bottom friction. 
     1163$c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and 
     1164$RHS$ represents all the components to the vertically integrated momentum trend except for 
     1165that due to bottom friction. 
    11971166 
    11981167 
     
    12181187 
    12191188Options are defined through the  \ngn{namzdf\_tmx} namelist variables. 
    1220 The parameterization of tidal mixing follows the general formulation for  
    1221 the vertical eddy diffusivity proposed by \citet{St_Laurent_al_GRL02} and  
    1222 first introduced in an OGCM by \citep{Simmons_al_OM04}.  
    1223 In this formulation an additional vertical diffusivity resulting from internal tide breaking,  
    1224 $A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, the energy transfer from barotropic  
    1225 tides to baroclinic tides :  
     1189The parameterization of tidal mixing follows the general formulation for the vertical eddy diffusivity proposed by 
     1190\citet{St_Laurent_al_GRL02} and first introduced in an OGCM by \citep{Simmons_al_OM04}.  
     1191In this formulation an additional vertical diffusivity resulting from internal tide breaking, 
     1192$A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, 
     1193the energy transfer from barotropic tides to baroclinic tides: 
    12261194\begin{equation} \label{eq:Ktides} 
    12271195A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 } 
    12281196\end{equation} 
    1229 where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency  
    1230 (see \autoref{subsec:TRA_bn2}), $\rho$ the density, $q$ the tidal dissipation efficiency,  
    1231 and $F(z)$ the vertical structure function.  
    1232  
    1233 The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) 
    1234 and is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).  
    1235 The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter)  
    1236 represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally,  
    1237 with the remaining $1-q$ radiating away as low mode internal waves and  
    1238 contributing to the background internal wave field. A value of $q=1/3$ is typically used   
    1239 \citet{St_Laurent_al_GRL02}. 
    1240 The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical.  
    1241 It is implemented as a simple exponential decaying upward away from the bottom,  
    1242 with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},  
     1197where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency (see \autoref{subsec:TRA_bn2}), 
     1198$\rho$ the density, $q$ the tidal dissipation efficiency, and $F(z)$ the vertical structure function.  
     1199 
     1200The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) and 
     1201is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).  
     1202The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter) 
     1203represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally, 
     1204with the remaining $1-q$ radiating away as low mode internal waves and 
     1205contributing to the background internal wave field. 
     1206A value of $q=1/3$ is typically used \citet{St_Laurent_al_GRL02}. 
     1207The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical. 
     1208It is implemented as a simple exponential decaying upward away from the bottom, 
     1209with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, 
     1210with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},  
    12431211\begin{equation} \label{eq:Fz} 
    12441212F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) } 
     
    12461214and is normalized so that vertical integral over the water column is unity.  
    12471215 
    1248 The associated vertical viscosity is calculated from the vertical  
    1249 diffusivity assuming a Prandtl number of 1, $i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$.  
    1250 In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity  
    1251 is capped at $300\,cm^2/s$ and impose a lower limit on $N^2$ of \np{rn\_n2min}  
    1252 usually set to $10^{-8} s^{-2}$. These bounds are usually rarely encountered. 
    1253  
    1254 The internal wave energy map, $E(x, y)$ in \autoref{eq:Ktides}, is derived  
    1255 from a barotropic model of the tides utilizing a parameterization of the  
    1256 conversion of barotropic tidal energy into internal waves.  
    1257 The essential goal of the parameterization is to represent the momentum  
    1258 exchange between the barotropic tides and the unrepresented internal waves  
    1259 induced by the tidal flow over rough topography in a stratified ocean.  
    1260 In the current version of \NEMO, the map is built from the output of  
     1216The associated vertical viscosity is calculated from the vertical diffusivity assuming a Prandtl number of 1, 
     1217$i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$. 
     1218In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity is capped at $300\,cm^2/s$ and 
     1219impose a lower limit on $N^2$ of \np{rn\_n2min} usually set to $10^{-8} s^{-2}$. 
     1220These bounds are usually rarely encountered. 
     1221 
     1222The internal wave energy map, $E(x, y)$ in \autoref{eq:Ktides}, is derived from a barotropic model of 
     1223the tides utilizing a parameterization of the conversion of barotropic tidal energy into internal waves. 
     1224The essential goal of the parameterization is to represent the momentum exchange between the barotropic tides and 
     1225the unrepresented internal waves induced by the tidal flow over rough topography in a stratified ocean. 
     1226In the current version of \NEMO, the map is built from the output of 
    12611227the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}. 
    1262 This model provides the dissipation associated with internal wave energy for the M2 and K1  
    1263 tides component (\autoref{fig:ZDF_M2_K1_tmx}). The S2 dissipation is simply approximated 
    1264 as being $1/4$ of the M2 one. The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$.  
    1265 Its global mean value is $1.1$ TW, in agreement with independent estimates  
    1266 \citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.  
     1228This model provides the dissipation associated with internal wave energy for the M2 and K1 tides component 
     1229(\autoref{fig:ZDF_M2_K1_tmx}). 
     1230The S2 dissipation is simply approximated as being $1/4$ of the M2 one. 
     1231The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$. 
     1232Its global mean value is $1.1$ TW, 
     1233in agreement with independent estimates \citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.  
    12671234 
    12681235%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     
    12811248 
    12821249When the Indonesian Through Flow (ITF) area is included in the model domain, 
    1283 a specific treatment of tidal induced mixing in this area can be used.  
    1284 It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide 
    1285 an input NetCDF file, \ifile{mask\_itf}, which contains a mask array defining the ITF area 
    1286 where the specific treatment is applied. 
    1287  
    1288 When \np{ln\_tmx\_itf}\forcode{ = .true.}, the two key parameters $q$ and $F(z)$ are adjusted following  
     1250a specific treatment of tidal induced mixing in this area can be used. 
     1251It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide an input NetCDF file, 
     1252\ifile{mask\_itf}, which contains a mask array defining the ITF area where the specific treatment is applied. 
     1253 
     1254When \np{ln\_tmx\_itf}\forcode{ = .true.}, the two key parameters $q$ and $F(z)$ are adjusted following 
    12891255the parameterisation developed by \citet{Koch-Larrouy_al_GRL07}: 
    12901256 
    1291 First, the Indonesian archipelago is a complex geographic region  
    1292 with a series of large, deep, semi-enclosed basins connected via  
    1293 numerous narrow straits. Once generated, internal tides remain  
    1294 confined within this semi-enclosed area and hardly radiate away.  
    1295 Therefore all the internal tides energy is consumed within this area.  
     1257First, the Indonesian archipelago is a complex geographic region with a series of 
     1258large, deep, semi-enclosed basins connected via numerous narrow straits. 
     1259Once generated, internal tides remain confined within this semi-enclosed area and hardly radiate away. 
     1260Therefore all the internal tides energy is consumed within this area. 
    12961261So it is assumed that $q = 1$, $i.e.$ all the energy generated is available for mixing. 
    1297 Note that for test purposed, the ITF tidal dissipation efficiency is a  
    1298 namelist parameter (\np{rn\_tfe\_itf}). A value of $1$ or close to is 
    1299 this recommended for this parameter. 
    1300  
    1301 Second, the vertical structure function, $F(z)$, is no more associated 
    1302 with a bottom intensification of the mixing, but with a maximum of  
    1303 energy available within the thermocline. \citet{Koch-Larrouy_al_GRL07}  
    1304 have suggested that the vertical distribution of the energy dissipation  
    1305 proportional to $N^2$ below the core of the thermocline and to $N$ above.  
     1262Note that for test purposed, the ITF tidal dissipation efficiency is a namelist parameter (\np{rn\_tfe\_itf}). 
     1263A value of $1$ or close to is this recommended for this parameter. 
     1264 
     1265Second, the vertical structure function, $F(z)$, is no more associated with a bottom intensification of the mixing, 
     1266but with a maximum of energy available within the thermocline. 
     1267\citet{Koch-Larrouy_al_GRL07} have suggested that the vertical distribution of 
     1268the energy dissipation proportional to $N^2$ below the core of the thermocline and to $N$ above.  
    13061269The resulting $F(z)$ is: 
    13071270\begin{equation} \label{eq:Fz_itf} 
     
    13131276 
    13141277Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,  
    1315 which agrees with the independent estimates inferred from observations.  
    1316 Introduced in a regional OGCM, the parameterization improves the water mass  
    1317 characteristics in the different Indonesian seas, suggesting that the horizontal  
    1318 and vertical distributions of the mixing are adequately prescribed  
    1319 \citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}. 
    1320 Note also that such a parameterisation has a significant impact on the behaviour  
    1321 of global coupled GCMs \citep{Koch-Larrouy_al_CD10}. 
     1278which agrees with the independent estimates inferred from observations. 
     1279Introduced in a regional OGCM, the parameterization improves the water mass characteristics in 
     1280the different Indonesian seas, suggesting that the horizontal and vertical distributions of 
     1281the mixing are adequately prescribed \citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}. 
     1282Note also that such a parameterisation has a significant impact on the behaviour of 
     1283global coupled GCMs \citep{Koch-Larrouy_al_CD10}. 
    13221284 
    13231285 
     
    13331295%-------------------------------------------------------------------------------------------------------------- 
    13341296 
    1335 The parameterization of mixing induced by breaking internal waves is a generalization  
    1336 of the approach originally proposed by \citet{St_Laurent_al_GRL02}.  
    1337 A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed,  
     1297The parameterization of mixing induced by breaking internal waves is a generalization of 
     1298the approach originally proposed by \citet{St_Laurent_al_GRL02}. 
     1299A three-dimensional field of internal wave energy dissipation $\epsilon(x,y,z)$ is first constructed, 
    13381300and the resulting diffusivity is obtained as  
    13391301\begin{equation} \label{eq:Kwave} 
    13401302A^{vT}_{wave} =  R_f \,\frac{ \epsilon }{ \rho \, N^2 } 
    13411303\end{equation} 
    1342 where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution  
    1343 of the energy available for mixing. If the \np{ln\_mevar} namelist parameter is set to false,  
    1344 the mixing efficiency is taken as constant and equal to 1/6 \citep{Osborn_JPO80}.  
    1345 In the opposite (recommended) case, $R_f$ is instead a function of the turbulence intensity parameter  
    1346 $Re_b = \frac{ \epsilon}{\nu \, N^2}$, with $\nu$ the molecular viscosity of seawater,  
    1347 following the model of \cite{Bouffard_Boegman_DAO2013}  
    1348 and the implementation of \cite{de_lavergne_JPO2016_efficiency}. 
    1349 Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when the mixing efficiency is constant. 
     1304where $R_f$ is the mixing efficiency and $\epsilon$ is a specified three dimensional distribution of 
     1305the energy available for mixing. 
     1306If the \np{ln\_mevar} namelist parameter is set to false, the mixing efficiency is taken as constant and 
     1307equal to 1/6 \citep{Osborn_JPO80}. 
     1308In the opposite (recommended) case, $R_f$ is instead a function of 
     1309the turbulence intensity parameter $Re_b = \frac{ \epsilon}{\nu \, N^2}$, 
     1310with $\nu$ the molecular viscosity of seawater, following the model of \cite{Bouffard_Boegman_DAO2013} and 
     1311the implementation of \cite{de_lavergne_JPO2016_efficiency}. 
     1312Note that $A^{vT}_{wave}$ is bounded by $10^{-2}\,m^2/s$, a limit that is often reached when 
     1313the mixing efficiency is constant. 
    13501314 
    13511315In addition to the mixing efficiency, the ratio of salt to heat diffusivities can chosen to vary  
    1352 as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to true, a recommended choice).  
    1353 This parameterization of differential mixing, due to \cite{Jackson_Rehmann_JPO2014},  
     1316as a function of $Re_b$ by setting the \np{ln\_tsdiff} parameter to true, a recommended choice.  
     1317This parameterization of differential mixing, due to \cite{Jackson_Rehmann_JPO2014}, 
    13541318is implemented as in \cite{de_lavergne_JPO2016_efficiency}. 
    13551319 
    1356 The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$, is constructed  
    1357 from three static maps of column-integrated internal wave energy dissipation, $E_{cri}(i,j)$,  
    1358 $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures  
     1320The three-dimensional distribution of the energy available for mixing, $\epsilon(i,j,k)$, 
     1321is constructed from three static maps of column-integrated internal wave energy dissipation, 
     1322$E_{cri}(i,j)$, $E_{pyc}(i,j)$, and $E_{bot}(i,j)$, combined to three corresponding vertical structures 
    13591323(de Lavergne et al., in prep): 
    13601324\begin{align*} 
     
    13631327F_{bot}(i,j,k) &\propto N^2 \, e^{- h_{wkb} / h_{bot} } 
    13641328\end{align*}  
    1365 In the above formula, $h_{ab}$ denotes the height above bottom,  
     1329In the above formula, $h_{ab}$ denotes the height above bottom, 
    13661330$h_{wkb}$ denotes the WKB-stretched height above bottom, defined by 
    13671331\begin{equation*} 
    13681332h_{wkb} = H \, \frac{ \int_{-H}^{z} N \, dz' } { \int_{-H}^{\eta} N \, dz'  } \; , 
    13691333\end{equation*} 
    1370 The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_tmx\_new} namelist)  controls the stratification-dependence of the pycnocline-intensified dissipation.  
     1334The $n_p$ parameter (given by \np{nn\_zpyc} in \ngn{namzdf\_tmx\_new} namelist) 
     1335controls the stratification-dependence of the pycnocline-intensified dissipation. 
    13711336It can take values of 1 (recommended) or 2. 
    1372 Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of  
    1373 the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps.  
    1374 $h_{cri}$ is related to the large-scale topography of the ocean (etopo2)  
    1375 and $h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of  
     1337Finally, the vertical structures $F_{cri}$ and $F_{bot}$ require the specification of 
     1338the decay scales $h_{cri}(i,j)$ and $h_{bot}(i,j)$, which are defined by two additional input maps. 
     1339$h_{cri}$ is related to the large-scale topography of the ocean (etopo2) and 
     1340$h_{bot}$ is a function of the energy flux $E_{bot}$, the characteristic horizontal scale of 
    13761341the abyssal hill topography \citep{Goff_JGR2010} and the latitude. 
    13771342 
Note: See TracChangeset for help on using the changeset viewer.