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 2282 for branches/nemo_v3_3_beta/DOC/TexFiles/Chapters/Chap_ZDF.tex – NEMO

Ignore:
Timestamp:
2010-10-15T16:42:00+02:00 (14 years ago)
Author:
gm
Message:

ticket:#658 merge DOC of all the branches that form the v3.3 beta

File:
1 edited

Legend:

Unmodified
Added
Removed
  • branches/nemo_v3_3_beta/DOC/TexFiles/Chapters/Chap_ZDF.tex

    r1225 r2282  
    77 
    88%gm% Add here a small introduction to ZDF and naming of the different physics (similar to what have been written for TRA and DYN. 
    9 \gmcomment{Steven remark (not taken into account : problem here with turbulent vs turbulence.  I've changed "turbulent closure" to "turbulence closure", "turbulent mixing length" to "turbulence mixing length", but I've left "turbulent kinetic energy" alone - though I think it is an historical abberation! 
    10 Gurvan :  I kept "turbulent closure etc "...} 
     9 
     10 
     11\newpage 
     12$\ $\newline    % force a new ligne 
    1113 
    1214 
     
    3840general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.  
    3941These trends can be computed using either a forward time stepping scheme  
    40 (namelist parameter \np{np\_zdfexp}=true) or a backward time stepping  
    41 scheme (\np{np\_zdfexp}=false) depending on the magnitude of the mixing  
    42 coefficients, and thus of the formulation used (see \S\ref{DOM_nxt}). 
     42(namelist parameter \np{ln\_zdfexp}=true) or a backward time stepping  
     43scheme (\np{ln\_zdfexp}=false) depending on the magnitude of the mixing  
     44coefficients, and thus of the formulation used (see \S\ref{STP}). 
    4345 
    4446% ------------------------------------------------------------------------------------------------------------- 
     
    5759\begin{align*}  
    5860A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\ 
    59 \\ 
    6061A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1} 
    6162\end{align*} 
    6263 
    63 These values are set through the \np{avm0} and \np{avt0} namelist parameters.  
     64These values are set through the \np{rn\_avm0} and \np{rn\_avt0} namelist parameters.  
    6465In all cases, do not use values smaller that those associated with the molecular  
    6566viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum,  
     
    7475 
    7576%--------------------------------------------namric--------------------------------------------------------- 
    76 \namdisplay{namric} 
     77\namdisplay{namzdf_ric} 
    7778%-------------------------------------------------------------------------------------------------------------- 
    7879 
     
    8485growth of Kelvin-Helmholtz like instabilities leads to a dependency between the  
    8586vertical eddy coefficients and the local Richardson number ($i.e.$ the  
    86 ratio of stratification to vertical shear). Following \citet{PacPhil1981}, the following  
     87ratio of stratification to vertical shear). Following \citet{Pacanowski_Philander_JPO81}, the following  
    8788formulation has been implemented: 
    8889\begin{equation} \label{Eq_zdfric} 
    8990   \left\{      \begin{aligned} 
    9091         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\ 
    91          \\ 
    9292         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm} 
    9393   \end{aligned}    \right. 
     
    9999is the maximum value that can be reached by the coefficient when $Ri\leq 0$,  
    100100$a=5$ and $n=2$. The last three values can be modified by setting the  
    101 \np{avmri}, \np{alp} and \np{nric} namelist parameters, respectively. 
     101\np{rn\_avmri}, \np{rn\_alp} and \np{nn\_ric} namelist parameters, respectively. 
    102102 
    103103% ------------------------------------------------------------------------------------------------------------- 
     
    107107\label{ZDF_tke} 
    108108 
    109 %--------------------------------------------namtke--------------------------------------------------------- 
    110 \namdisplay{namtke} 
     109%--------------------------------------------namzdf_tke-------------------------------------------------- 
     110\namdisplay{namzdf_tke} 
    111111%-------------------------------------------------------------------------------------------------------------- 
    112112 
    113113The vertical eddy viscosity and diffusivity coefficients are computed from a TKE  
    114 turbulent closure model based on a prognostic equation for $\bar {e}$, the turbulent  
     114turbulent closure model based on a prognostic equation for $\bar{e}$, the turbulent  
    115115kinetic energy, and a closure assumption for the turbulent length scales. This  
    116116turbulent closure model has been developed by \citet{Bougeault1989} in the  
    117117atmospheric case, adapted by \citet{Gaspar1990} for the oceanic case, and  
    118 embedded in OPA by \citet{Blanke1993} for equatorial Atlantic simulations. Since  
    119 then, significant modifications have been introduced by \citet{Madec1998} in both  
    120 the implementation and the formulation of the mixing length scale. The time  
    121 evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical  
    122 shear, its destruction through stratification, its vertical diffusion, and its dissipation  
    123 of \citet{Kolmogorov1942} type: 
     118embedded in OPA, the ancestor of NEMO, by \citet{Blanke1993} for equatorial Atlantic  
     119simulations. Since then, significant modifications have been introduced by  
     120\citet{Madec1998} in both the implementation and the formulation of the mixing  
     121length scale. The time evolution of $\bar{e}$ is the result of the production of  
     122$\bar{e}$ through vertical shear, its destruction through stratification, its vertical  
     123diffusion, and its dissipation of \citet{Kolmogorov1942} type: 
    124124\begin{equation} \label{Eq_zdftke_e} 
    125125\frac{\partial \bar{e}}{\partial t} =  
    126 \frac{A^{vm}}{e_3 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2 
    127                      +\left( {\frac{\partial v}{\partial k}} \right)^2} \right] 
    128 -A^{vT}\,N^2 
     126\frac{K_m}{{e_3}^2 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2 
     127                    +\left( {\frac{\partial v}{\partial k}} \right)^2} \right] 
     128-K_\rho\,N^2 
    129129+\frac{1}{e_3} \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 } 
    130130            \;\frac{\partial \bar{e}}{\partial k}} \right] 
     
    133133\begin{equation} \label{Eq_zdftke_kz} 
    134134   \begin{split} 
    135          A^{vm} &= C_k\  l_k\  \sqrt {\bar{e}}     \\ 
    136          A^{vT} &= A^{vm} / P_{rt} 
     135         K_m &= C_k\  l_k\  \sqrt {\bar{e}\; }     \\ 
     136         K_\rho &= A^{vm} / P_{rt} 
    137137   \end{split} 
    138138\end{equation} 
    139139where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),  
    140140$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,  
    141 $P_{rt} $ is the Prandtl number. The constants $C_k = \sqrt {2} /2$ and  
    142 $C_\epsilon = 0.1$ are designed to deal with vertical mixing at any depth  
    143 \citep{Gaspar1990}. They are set through namelist parameters \np{ediff}  
    144 and \np{ediss}. $P_{rt} $ can be set to unity or, following \citet{Blanke1993},  
    145 be a function of the local Richardson number, $R_i $: 
     141$P_{rt}$ is the Prandtl number, $K_m$ and $K_\rho$ are the vertical eddy viscosity  
     142and diffusivity coefficients. The constants $C_k =  0.1$ and $C_\epsilon = \sqrt {2} /2$   
     143$\approx 0.7$ are designed to deal with vertical mixing at any depth \citep{Gaspar1990}.  
     144They are set through namelist parameters \np{nn\_ediff} and \np{nn\_ediss}.  
     145$P_{rt}$ can be set to unity or, following \citet{Blanke1993}, be a function  
     146of the local Richardson number, $R_i$: 
    146147\begin{align*} \label{Eq_prt} 
    147148P_{rt} = \begin{cases} 
     
    151152            \end{cases} 
    152153\end{align*} 
    153 Note that a horizontal Shapiro filter can optionally be applied to $R_i$.  
    154 However it is an obsolescent option that is not recommended.   
    155 The choice of $P_{rt} $ is controlled by the \np{npdl} namelist parameter. 
     154The choice of $P_{rt}$ is controlled by the \np{nn\_pdl} namelist parameter. 
    156155 
    157156For computational efficiency, the original formulation of the turbulent length  
    158157scales proposed by \citet{Gaspar1990} has been simplified. Four formulations  
    159 are proposed, the choice of which is controlled by the \np{nmxl} namelist  
     158are proposed, the choice of which is controlled by the \np{nn\_mxl} namelist  
    160159parameter. The first two are based on the following first order approximation  
    161160\citep{Blanke1993}: 
    162161\begin{equation} \label{Eq_tke_mxl0_1} 
    163 l_k = l_\epsilon = \sqrt {2 \bar e} / N 
    164 \end{equation} 
    165 which is valid in a stable stratified region with constant values of the brunt- 
     162l_k = l_\epsilon = \sqrt {2 \bar{e}\; } / N 
     163\end{equation} 
     164which is valid in a stable stratified region with constant values of the Brunt- 
    166165Vais\"{a}l\"{a} frequency. The resulting length scale is bounded by the distance  
    167 to the surface or to the bottom (\np{nmxl}=0) or by the local vertical scale factor (\np{nmxl}=1). \citet{Blanke1993} notice that this simplification has two major  
     166to the surface or to the bottom (\np{nn\_mxl} = 0) or by the local vertical scale factor  
     167(\np{nn\_mxl} = 1). \citet{Blanke1993} notice that this simplification has two major  
    168168drawbacks: it makes no sense for locally unstable stratification and the  
    169169computation no longer uses all the information contained in the vertical density  
    170170profile. To overcome these drawbacks, \citet{Madec1998} introduces the  
    171 \np{nmxl}=2 or 3 cases, which add an extra assumption concerning the vertical  
     171\np{nn\_mxl} = 2 or 3 cases, which add an extra assumption concerning the vertical  
    172172gradient of the computed length scale. So, the length scales are first evaluated  
    173173as in \eqref{Eq_tke_mxl0_1} and then bounded such that: 
     
    185185constraint, we introduce two additional length scales: $l_{up}$ and $l_{dwn}$,  
    186186the upward and downward length scales, and evaluate the dissipation and  
    187 mixing turbulent length scales as (and note that here we use numerical  
    188 indexing): 
     187mixing length scales as (and note that here we use numerical indexing): 
    189188%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    190189\begin{figure}[!t] \label{Fig_mixing_length}  \begin{center} 
     
    196195\begin{equation} \label{Eq_tke_mxl2} 
    197196\begin{aligned} 
    198  l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3T}^{(k)}\ \ \ \;  \right) 
     197 l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3t}^{(k)}\ \ \ \;  \right) 
    199198    \quad &\text{ from $k=1$ to $jpk$ }\ \\ 
    200  l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3T}^{(k-1)}  \right)    
     199 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3t}^{(k-1)}  \right)    
    201200    \quad &\text{ from $k=jpk$ to $1$ }\ \\ 
    202201\end{aligned} 
    203202\end{equation} 
    204203where $l^{(k)}$ is computed using \eqref{Eq_tke_mxl0_1},  
    205 $i.e.$ $l^{(k)} = \sqrt {2 \bar e^{(k)} / N^{(k)} }$. 
    206  
    207 In the \np{nmxl}=2 case, the dissipation and mixing length scales take the same  
     204$i.e.$ $l^{(k)} = \sqrt {2 {\bar e}^{(k)} / {N^2}^{(k)} }$. 
     205 
     206In the \np{nn\_mxl}=2 case, the dissipation and mixing length scales take the same  
    208207value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the  
    209 \np{nmxl}=2 case, the dissipation and mixing length scales are give  
     208\np{nn\_mxl}=2 case, the dissipation and mixing turbulent length scales are give  
    210209as in \citet{Gaspar1990}: 
    211210\begin{equation} \label{Eq_tke_mxl_gaspar} 
     
    217216 
    218217At the sea surface the value of $\bar{e}$ is prescribed from the wind  
    219 stress field: $\bar{e}=ebb\;\left| \tau \right|$ ($ebb=60$ by default)  
    220 with a minimal threshold of $emin0=10^{-4}~m^2.s^{-2}$ (namelist 
     218stress field: $\bar{e}=rn\_ebb\;\left| \tau \right|$ (\np{rn\_ebb}=60 by default)  
     219with a minimal threshold of \np{rn\_emin0}$=10^{-4}~m^2.s^{-2}$ (namelist 
    221220parameters). Its value at the bottom of the ocean is assumed to be  
    222221equal to the value of the level just above. The time integration of the  
    223222$\bar{e}$ equation may formally lead to negative values because the  
    224223numerical scheme does not ensure its positivity. To overcome this  
    225 problem, a cut-off in the minimum value of $\bar{e}$ is used. Following  
    226 \citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$.  
    227 This allows the subsequent formulations to match that of\citet{Gargett1984}  
    228 for the diffusion in the thermocline and deep ocean :  $(A^{vT} = 10^{-3} / N)$.  
    229 In addition, a cut-off is applied on $A^{vm}$ and $A^{vT}$ to avoid numerical  
     224problem, a cut-off in the minimum value of $\bar{e}$ is used (\np{rn\_emin}  
     225namelist parameter). Following \citet{Gaspar1990}, the cut-off value is set  
     226to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$. This allows the subsequent formulations  
     227to match that of \citet{Gargett1984} for the diffusion in the thermocline and  
     228deep ocean :  $K_\rho = 10^{-3} / N$.  
     229In addition, a cut-off is applied on $K_m$ and $K_\rho$ to avoid numerical  
    230230instabilities associated with too weak vertical diffusion. They must be  
    231231specified at least larger than the molecular values, and are set through  
    232 \textit{avm0} and \textit{avt0} (\textbf{namelist} parameters). 
     232\np{rn\_avm0} and \np{rn\_avt0} (namzdf namelist, see \S\ref{ZDF_cst}). 
     233 
     234% ------------------------------------------------------------------------------------------------------------- 
     235%        TKE Turbulent Closure Scheme : new organization to energetic considerations 
     236% ------------------------------------------------------------------------------------------------------------- 
     237\subsection{TKE discretization considerations (\key{zdftke})} 
     238\label{ZDF_tke_ene} 
     239 
     240%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     241\begin{figure}[!t] \label{Fig_TKE_time_scheme}  \begin{center} 
     242\includegraphics[width=1.00\textwidth]{./TexFiles/Figures/Fig_ZDF_TKE_time_scheme.pdf} 
     243\caption {Illustration of the TKE time integration and its links to the momentum and tracer time integration. } 
     244\end{center}   
     245\end{figure} 
     246%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     247 
     248The production of turbulence by vertical shear (the first term of the right hand side  
     249of \eqref{Eq_zdftke_e}) should balance the loss of kinetic energy associated with 
     250the vertical momentum diffusion (first line in \eqref{Eq_PE_zdf}). To do so a special care  
     251have to be taken for both the time and space discretization of the TKE equation \citep{Burchard_OM02}. 
     252 
     253Let us first address the time stepping issue. Fig.~\ref{Fig_TKE_time_scheme} shows  
     254how the two-level Leap-Frog time stepping of the momentum and tracer equations interplays  
     255with the one-level forward time stepping of TKE equation. With this framework, the total loss  
     256of kinetic energy (in 1D for the demonstration) due to the vertical momentum diffusion is  
     257obtained by multiplying this quantity by $u^t$ and summing the result vertically:    
     258\begin{equation} \label{Eq_energ1} 
     259\begin{split} 
     260\int_{-H}^{\eta}  u^t \,\partial_z &\left( {K_m}^t \,(\partial_z u)^{t+\rdt}  \right) \,dz   \\ 
     261&= \Bigl[  u^t \,{K_m}^t \,(\partial_z u)^{t+\rdt} \Bigr]_{-H}^{\eta}           
     262 - \int_{-H}^{\eta}{ {K_m}^t \,\partial_z{u^t} \,\partial_z u^{t+\rdt} \,dz } 
     263\end{split} 
     264\end{equation} 
     265Here, the vertical diffusion of momentum is discretized backward in time  
     266with a coefficient, $K_m$, known at time $t$ (Fig.~\ref{Fig_TKE_time_scheme}),  
     267as it is required when using the TKE scheme (see \S\ref{STP_forward_imp}).  
     268The first term of the right hand side of \eqref{Eq_energ1} represents the kinetic energy  
     269transfer at the surface (atmospheric forcing) and at the bottom (friction effect).  
     270The second term is always negative. It is the dissipation rate of kinetic energy,  
     271and thus minus the shear production rate of $\bar{e}$. \eqref{Eq_energ1}  
     272implies that, to be energetically consistent, the production rate of $\bar{e}$  
     273used to compute $(\bar{e})^t$ (and thus ${K_m}^t$) should be expressed as  
     274${K_m}^{t-\rdt}\,(\partial_z u)^{t-\rdt} \,(\partial_z u)^t$ (and not by the more straightforward  
     275$K_m \left( \partial_z u \right)^2$ expression taken at time $t$ or $t-\rdt$). 
     276 
     277A similar consideration applies on the destruction rate of $\bar{e}$ due to stratification  
     278(second term of the right hand side of \eqref{Eq_zdftke_e}). This term  
     279must balance the input of potential energy resulting from vertical mixing.  
     280The rate of change of potential energy (in 1D for the demonstration) due vertical  
     281mixing is obtained by multiplying vertical density diffusion  
     282tendency by $g\,z$ and and summing the result vertically: 
     283\begin{equation} \label{Eq_energ2} 
     284\begin{split} 
     285\int_{-H}^{\eta} g\,z\,\partial_z &\left( {K_\rho}^t \,(\partial_k \rho)^{t+\rdt}   \right) \,dz    \\ 
     286&= \Bigl[  g\,z \,{K_\rho}^t \,(\partial_z \rho)^{t+\rdt} \Bigr]_{-H}^{\eta}   
     287   - \int_{-H}^{\eta}{ g \,{K_\rho}^t \,(\partial_k \rho)^{t+\rdt} } \,dz   \\ 
     288&= - \Bigl[  z\,{K_\rho}^t \,(N^2)^{t+\rdt} \Bigr]_{-H}^{\eta} 
     289+ \int_{-H}^{\eta}{  \rho^{t+\rdt} \, {K_\rho}^t \,(N^2)^{t+\rdt} \,dz  } 
     290\end{split} 
     291\end{equation} 
     292where we use $N^2 = -g \,\partial_k \rho / (e_3 \rho)$.  
     293The first term of the right hand side of \eqref{Eq_energ2}  is always zero  
     294because there is no diffusive flux through the ocean surface and bottom).  
     295The second term is minus the destruction rate of  $\bar{e}$ due to stratification.  
     296Therefore \eqref{Eq_energ1} implies that, to be energetically consistent, the product  
     297${K_\rho}^{t-\rdt}\,(N^2)^t$ should be used in \eqref{Eq_zdftke_e}, the TKE equation. 
     298 
     299Let us now address the space discretization issue.  
     300The vertical eddy coefficients are defined at $w$-point whereas the horizontal velocity  
     301components are in the centre of the side faces of a $t$-box in staggered C-grid  
     302(Fig.\ref{Fig_cell}). A space averaging is thus required to obtain the shear TKE production term. 
     303By redoing the \eqref{Eq_energ1} in the 3D case, it can be shown that the product of  
     304eddy coefficient by the shear at $t$ and $t-\rdt$ must be performed prior to the averaging. 
     305Furthermore, the possible time variation of $e_3$ (\key{vvl} case) have to be taken into  
     306account. 
     307 
     308The above energetic considerations leads to  
     309the following final discrete form for the TKE equation: 
     310\begin{equation} \label{Eq_zdftke_ene} 
     311\begin{split} 
     312\frac { (\bar{e})^t - (\bar{e})^{t-\rdt} } {\rdt}  \equiv   
     313\Biggl\{ \Biggr. 
     314  &\overline{ \left( \left(\overline{K_m}^{\,i+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[u^{t+\rdt}]}{{e_3u}^{t+\rdt} }  
     315                                                                              \ \frac{\delta_{k+1/2}[u^ t         ]}{{e_3u}^ t          }  \right) }^{\,i} \\ 
     316+&\overline{  \left( \left(\overline{K_m}^{\,j+1/2}\right)^{t-\rdt} \,\frac{\delta_{k+1/2}[v^{t+\rdt}]}{{e_3v}^{t+\rdt} }  
     317                                                                               \ \frac{\delta_{k+1/2}[v^ t         ]}{{e_3v}^ t          }  \right) }^{\,j}  
     318\Biggr. \Biggr\}   \\ 
     319% 
     320- &{K_\rho}^{t-\rdt}\,{(N^2)^t}    \\ 
     321% 
     322+&\frac{1}{{e_3w}^{t+\rdt}}  \;\delta_{k+1/2} \left[   {K_m}^{t-\rdt} \,\frac{\delta_{k}[(\bar{e})^{t+\rdt}]} {{e_3w}^{t+\rdt}}   \right]   \\ 
     323% 
     324- &c_\epsilon \; \left( \frac{\sqrt{\bar {e}}}{l_\epsilon}\right)^{t-\rdt}\,(\bar {e})^{t+\rdt} 
     325\end{split} 
     326\end{equation} 
     327where the last two terms in \eqref{Eq_zdftke_ene} (vertical diffusion and Kolmogorov dissipation)  
     328are time stepped using a backward scheme (see\S\ref{STP_forward_imp}).  
     329Note that the Kolmogorov term has been linearized in time in order to render  
     330the implicit computation possible. The restart of the TKE scheme  
     331requires the storage of $\bar {e}$, $K_m$, $K_\rho$ and $l_\epsilon$ as they all appear in  
     332the right hand side of \eqref{Eq_zdftke_ene}. For the latter, it is in fact  
     333the ratio $\sqrt{\bar{e}}/l_\epsilon$ which is stored.  
     334 
     335% ------------------------------------------------------------------------------------------------------------- 
     336%        GLS Generic Length Scale Scheme  
     337% ------------------------------------------------------------------------------------------------------------- 
     338\subsection{GLS Generic Length Scale (\key{zdfgls})} 
     339\label{ZDF_gls} 
     340 
     341%--------------------------------------------namgls--------------------------------------------------------- 
     342\namdisplay{namgls} 
     343%-------------------------------------------------------------------------------------------------------------- 
     344 
     345The Generic Length Scale (GLS) scheme is a turbulent closure scheme based on  
     346two prognostic equations: one for the turbulent kinetic energy $\bar {e}$, and another  
     347for the generic length scale, $\psi$ \citep{Umlauf_Burchard_JMS03, Umlauf_Burchard_CSR05}.  
     348This later variable is defined as : $\psi = {C_{0\mu}}^{p} \ {\bar{e}}^{m} \ l^{n}$,  
     349where the triplet $(p, m, n)$ value given in Tab.\ref{Tab_GLS} allows to recover  
     350a number of well-known turbulent closures ($k$-$kl$ \citep{Mellor_Yamada_1982},  
     351$k$-$\epsilon$ \citep{Rodi_1987}, $k$-$\omega$ \citep{Wilcox_1988}  
     352among others \citep{Umlauf_Burchard_JMS03,Kantha_Carniel_CSR05}).  
     353The GLS scheme is given by the following set of equations: 
     354\begin{equation} \label{Eq_zdfgls_e} 
     355\frac{\partial \bar{e}}{\partial t} =  
     356\frac{K_m}{\sigma_e e_3 }\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     357                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     358-K_\rho \,N^2 
     359+\frac{1}{e_3}\,\frac{\partial}{\partial k} \left[ \frac{K_m}{e_3}\,\frac{\partial \bar{e}}{\partial k} \right] 
     360- \epsilon 
     361\end{equation} 
     362 
     363\begin{equation} \label{Eq_zdfgls_psi} 
     364   \begin{split} 
     365\frac{\partial \psi}{\partial t} =& \frac{\psi}{\bar{e}} \left\{ 
     366\frac{C_1\,K_m}{\sigma_{\psi} {e_3}}\;\left[ {\left( \frac{\partial u}{\partial k} \right)^2 
     367                                                                   +\left( \frac{\partial v}{\partial k} \right)^2} \right] 
     368- C_3 \,K_\rho\,N^2   - C_2 \,\epsilon \,Fw   \right\}             \\ 
     369&+\frac{1}{e_3}  \;\frac{\partial }{\partial k}\left[ {\frac{K_m}{e_3 } 
     370                                  \;\frac{\partial \psi}{\partial k}} \right]\; 
     371   \end{split} 
     372\end{equation} 
     373 
     374\begin{equation} \label{Eq_zdfgls_kz} 
     375   \begin{split} 
     376         K_m    &= C_{\mu} \ \sqrt {\bar{e}} \ l         \\ 
     377         K_\rho &= C_{\mu'}\ \sqrt {\bar{e}} \ l 
     378   \end{split} 
     379\end{equation} 
     380 
     381\begin{equation} \label{Eq_zdfgls_eps} 
     382{\epsilon} = C_{0\mu} \,\frac{\bar {e}^{3/2}}{l} \; 
     383\end{equation} 
     384where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2})  
     385and $\epsilon$ the dissipation rate.  
     386The constants $C_1$, $C_2$, $C_3$, ${\sigma_e}$, ${\sigma_{\psi}}$ and the wall function ($Fw$)  
     387depends of the choice of the turbulence model. Four different turbulent models are pre-defined  
     388(Tab.\ref{Tab_GLS}). They are made available through th \np{gls} namelist parameter.  
     389 
     390%--------------------------------------------------TABLE-------------------------------------------------- 
     391\begin{table}[htbp]  \label{Tab_GLS} 
     392\begin{center} 
     393%\begin{tabular}{cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}cp{70pt}c} 
     394\begin{tabular}{ccccc} 
     395                         &   $k-kl$   & $k-\epsilon$ & $k-\omega$ &   generic   \\   
     396%                        & \citep{Mellor_Yamada_1982} &  \citep{Rodi_1987}       & \citep{Wilcox_1988} &                 \\   
     397\hline  \hline  
     398\np{nn\_clo}     & \textbf{0} &   \textbf{1}  &   \textbf{2}   &    \textbf{3}   \\   
     399\hline  
     400$( p , n , m )$          &   ( 0 , 1 , 1 )   & ( 3 , 1.5 , -1 )   & ( -1 , 0.5 , -1 )    &  ( 2 , 1 , -0.67 )  \\ 
     401$\sigma_k$      &    2.44         &     1.              &      2.                &      0.8          \\ 
     402$\sigma_\psi$  &    2.44         &     1.3            &      2.                 &       1.07       \\ 
     403$C_1$              &      0.9         &     1.44          &      0.555          &       1.           \\ 
     404$C_2$              &      0.5         &     1.92          &      0.833          &       1.22       \\ 
     405$C_3$              &      1.           &     1.              &      1.                &       1.           \\ 
     406$F_{wall}$        &      Yes        &       --             &     --                  &      --          \\ 
     407\hline 
     408\hline 
     409\end{tabular} 
     410\caption {Set of predefined GLS parameters, or equivalently predefined turbulence models available with \key{gls} and controlled by the \np{nn\_clos} namelist parameter.} 
     411\end{center} 
     412\end{table} 
     413%-------------------------------------------------------------------------------------------------------------- 
     414 
     415In the Mellor-Yamada model, the negativity of $n$ allows to use a wall function to force 
     416the convergence of the mixing length towards $K\,z_b$ ($K$: Kappa and $z_b$: rugosity length)  
     417value near physical boundaries (logarithmic boundary layer law). $C_{\mu}$ and $C_{\mu'}$  
     418are calculated from stability function proposed by \citet{Galperin_al_JAS88}, or by \citet{Kantha_Clayson_1994}  
     419or one of the two functions suggested by \citet{Canuto_2001}  (\np{nn\_stab\_func} = 0, 1, 2 or 3, resp.}). The value of $C_{0\mu}$ depends of the choice of the stability function. 
     420 
     421The surface and bottom boundary condition on both $\bar{e}$ and $\psi$ can be calculated  
     422thanks to Dirichlet or Neumann condition through \np{nn\_tkebc\_surf} and \np{nn\_tkebc\_bot}, resp.  
     423The wave effect on the mixing could be also being considered \citep{Craig_Banner_1994}. 
     424 
     425The $\psi$ equation is known to fail in stably stratified flows, and for this reason  
     426almost all authors apply a clipping of the length scale as an \textit{ad hoc} remedy.  
     427With this clipping, the maximum permissible length scale is determined by  
     428$l_{max} = c_{lim} \sqrt{2\bar{e}}/ N$. A value of $c_{lim} = 0.53$ is often used  
     429\citep{Galperin_al_JAS88}. \cite{Umlauf_Burchard_CSR05} show that the value of  
     430the clipping factor is of crucial importance for the entrainment depth predicted in  
     431stably stratified situations, and that its value has to be chosen in accordance  
     432with the algebraic model for the turbulent ßuxes. The clipping is only activated  
     433if \np{ln\_length\_lim}=true, and the $c_{lim}$ is set to the \np{clim\_galp} value. 
    233434 
    234435% ------------------------------------------------------------------------------------------------------------- 
     
    239440 
    240441%--------------------------------------------namkpp-------------------------------------------------------- 
    241 \namdisplay{namkpp} 
    242 %-------------------------------------------------------------------------------------------------------------- 
    243  
    244 The K-Profile Parametrization (KKP) developed by \cite{Large_al_RG94} has been  
    245 implemented in \NEMO by J. Chanut (PhD reference to be added here!). 
     442\namdisplay{namzdf_kpp} 
     443%-------------------------------------------------------------------------------------------------------------- 
     444 
     445The KKP scheme has been implemented by J. Chanut ... 
    246446 
    247447\colorbox{yellow}{Add a description of KPP here.} 
     
    274474\label{ZDF_npc} 
    275475 
    276 %--------------------------------------------namnpc-------------------------------------------------------- 
    277 \namdisplay{namnpc} 
    278 %-------------------------------------------------------------------------------------------------------------- 
    279  
     476%--------------------------------------------namzdf-------------------------------------------------------- 
     477\namdisplay{namzdf} 
     478%-------------------------------------------------------------------------------------------------------------- 
    280479 
    281480%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     
    295494 
    296495The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}=true.  
    297 It is applied at each \np{nnpc1} time step and mixes downwards instantaneously  
     496It is applied at each \np{nn\_npc} time step and mixes downwards instantaneously  
    298497the statically unstable portion of the water column, but only until the density  
    299498structure becomes neutrally stable ($i.e.$ until the mixed portion of the water  
    300 column has \textit{exactly} the density of the water just below) \citep{Madec1991a}.  
     499column has \textit{exactly} the density of the water just below) \citep{Madec_al_JPO91}.  
    301500The associated algorithm is an iterative process used in the following way  
    302501(Fig. \ref{Fig_npc}): starting from the top of the ocean, the first instability is  
     
    320519convective algorithm has been proved successful in studies of the deep  
    321520water formation in the north-western Mediterranean Sea  
    322 \citep{Madec1991a, Madec1991b, Madec1991c}. 
     521\citep{Madec_al_JPO91, Madec_al_DAO91, Madec_Crepon_Bk91}. 
    323522 
    324523Note that in the current implementation of this algorithm presents several  
     
    341540after a static adjustment. In that case, we recommend the addition of momentum  
    342541mixing in a manner that mimics the mixing in temperature and salinity  
    343 \citep{Speich1992, Speich1996}. 
     542\citep{Speich_PhD92, Speich_al_JPO96}. 
    344543 
    345544% ------------------------------------------------------------------------------------------------------------- 
     
    347546% ------------------------------------------------------------------------------------------------------------- 
    348547\subsection   [Enhanced Vertical Diffusion (\np{ln\_zdfevd})] 
    349          {Enhanced Vertical Diffusion (\np{ln\_zdfevd}=.true.)} 
     548         {Enhanced Vertical Diffusion (\np{ln\_zdfevd}=true)} 
    350549\label{ZDF_evd} 
    351550 
     
    357556In this case, the vertical eddy mixing coefficients are assigned very large values  
    358557(a typical value is $10\;m^2s^{-1})$ in regions where the stratification is unstable  
    359 ($i.e.$ when the Brunt-Vais\"{a}l\"{a} frequency is negative) \citep{Lazar1997, Lazar1999}.  
    360 This is done either on tracers only (\np{n\_evdm}=0) or on both momentum and  
    361 tracers (\np{n\_evdm}=1). 
     558($i.e.$ when $N^2$ the Brunt-Vais\"{a}l\"{a} frequency is negative)  
     559\citep{Lazar_PhD97, Lazar_al_JPO99}. This is done either on tracers only  
     560(\np{nn\_evdm}=0) or on both momentum and tracers (\np{nn\_evdm}=1). 
    362561 
    363562In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and  
    364 if \np{n\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$  
    365 values also, are set equal to the namelist parameter \np{avevd}. A typical value  
    366 for $avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterisation of  
     563if \np{nn\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$  
     564values also, are set equal to the namelist parameter \np{rn\_avevd}. A typical value  
     565for $rn\_avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterisation of  
    367566convective processes is less time consuming than the convective adjustment  
    368567algorithm presented above when mixing both tracers and momentum in the  
    369568case of static instabilities. It requires the use of an implicit time stepping on  
    370569vertical diffusion terms (i.e. \np{ln\_zdfexp}=false).  
     570 
     571Note that the stability test is performed on both \textit{before} and \textit{now}  
     572values of $N^2$. This removes a potential source of divergence of odd and 
     573even time step in a leapfrog environment \citep{Leclair_PhD2010} (see \S\ref{STP_mLF}). 
    371574 
    372575% ------------------------------------------------------------------------------------------------------------- 
     
    394597The KPP turbulent closure scheme already includes enhanced vertical diffusion  
    395598in the case of convection, as governed by the variables $bvsqcon$ and $difcon$  
    396 found in \mdl{zdfkpp}, therefore \np{np\_zdfevd} should not be used with the KPP  
     599found in \mdl{zdfkpp}, therefore \np{ln\_zdfevd}=false should be used with the KPP  
    397600scheme. %gm%  + one word on non local flux with KPP scheme trakpp.F90 module... 
    398601 
     
    404607\label{ZDF_ddm} 
    405608 
    406 %-------------------------------------------namddm-------------------------------------------------------- 
    407 \namdisplay{namddm} 
     609%-------------------------------------------namzdf_ddm------------------------------------------------- 
     610\namdisplay{namzdf_ddm} 
    408611%-------------------------------------------------------------------------------------------------------------- 
    409612 
     
    422625\end{align*} 
    423626where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,  
    424 and $o$ by processes other than double diffusion. The rates of double-diffusive mixing  
    425 depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,  
     627and $o$ by processes other than double diffusion. The rates of double-diffusive mixing depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,  
    426628where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline  
    427629contraction (see \S\ref{TRA_eos}). To represent mixing of $S$ and $T$ by salt  
     
    454656we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$. 
    455657 
    456 To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested  
    457 by Federov (1988) is used:  
     658To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:  
    458659\begin{align}  \label{Eq_zdfddm_d} 
    459660A_d^{vT} &=    \begin{cases} 
     
    481682% Bottom Friction 
    482683% ================================================================ 
    483 \section  [Bottom Friction (\textit{zdfbfr})] 
    484       {Bottom Friction (\mdl{zdfbfr} module)} 
     684\section  [Bottom Friction (\textit{zdfbfr})]   {Bottom Friction (\mdl{zdfbfr} module)} 
    485685\label{ZDF_bfr} 
    486686 
     
    493693diffusive flux. For the bottom boundary layer, one has: 
    494694\begin{equation} \label{Eq_zdfbfr_flux} 
    495 A^{vm} \left( \partial \textbf{U}_h / \partial z \right) = \textbf{F}_h 
    496 \end{equation} 
    497 where $\textbf{F}_h$ is supposed to represent the horizontal momentum flux  
     695A^{vm} \left( \partial {\textbf U}_h / \partial z \right) = {{\cal F}}_h^{\textbf U} 
     696\end{equation} 
     697where ${\cal F}_h^{\textbf U}$ is represents the downward flux of horizontal momentum  
    498698outside the logarithmic turbulent boundary layer (thickness of the order of  
    499 1~m in the ocean). How $\textbf{F}_h$ influences the interior depends on the  
     6991~m in the ocean). How ${\cal F}_h^{\textbf U}$ influences the interior depends on the  
    500700vertical resolution of the model near the bottom relative to the Ekman layer  
    501701depth. For example, in order to obtain an Ekman layer depth  
     
    509709this, consider the equation for $u$ at $k$, the last ocean level: 
    510710\begin{equation} \label{Eq_zdfbfr_flux2} 
    511 \frac{\partial u \; (k)}{\partial t} = \frac{1}{e_{3u}} \left[ A^{vm} \; (k) \frac{U \; (k-1) - U \; (k)}{e_{3uw} \; (k-1)} - F_u \right] \approx - \frac{F_u}{e_{3u}} 
    512 \end{equation} 
    513 For example, if the bottom layer thickness is 200~m, the Ekman transport will  
     711\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}} 
     712\end{equation} 
     713If the bottom layer thickness is 200~m, the Ekman transport will  
    514714be distributed over that depth. On the other hand, if the vertical resolution  
    515715is high (1~m or less) and a turbulent closure model is used, the turbulent  
    516716Ekman layer will be represented explicitly by the model. However, the  
    517717logarithmic layer is never represented in current primitive equation model  
    518 applications: it is \emph{necessary} to parameterize the flux $\textbf{F}_h $.  
     718applications: it is \emph{necessary} to parameterize the flux ${\cal F}^u_h $.  
    519719Two choices are available in \NEMO: a linear and a quadratic bottom friction.  
    520720Note that in both cases, the rotation between the interior velocity and the  
    521721bottom friction is neglected in the present release of \NEMO. 
    522722 
     723In the code, the bottom friction is imposed by adding the trend due to the bottom  
     724friction to the general momentum trend in \mdl{dynbfr}. For the time-split surface  
     725pressure gradient algorithm, the momentum trend due to the barotropic component  
     726needs to be handled separately. For this purpose it is convenient to compute and  
     727store coefficients which can be simply combined with bottom velocities and geometric  
     728values to provide the momentum trend due to bottom friction.  
     729These coefficients are computed in \mdl{zdfbfr} and generally take the form  
     730$c_b^{\textbf U}$ where: 
     731\begin{equation} \label{Eq_zdfbfr_bdef} 
     732\frac{\partial {\textbf U_h}}{\partial t} =  
     733  - \frac{{\cal F}^{\textbf U}_{h}}{e_{3u}} = \frac{c_b^{\textbf U}}{e_{3u}} \;{\textbf U}_h^b 
     734\end{equation} 
     735where $\textbf{U}_h^b = (u_b\;,\;v_b)$ is the near-bottom, horizontal, ocean velocity. 
     736 
    523737% ------------------------------------------------------------------------------------------------------------- 
    524738%       Linear Bottom Friction 
    525739% ------------------------------------------------------------------------------------------------------------- 
    526 \subsection{Linear Bottom Friction (\np{nbotfr} = 1) } 
     740\subsection{Linear Bottom Friction (\np{nn\_botfr} = 0 or 1) } 
    527741\label{ZDF_bfr_linear} 
    528742 
    529 The linear bottom friction parameterisation assumes that the bottom friction  
    530 is proportional to the interior velocity (i.e. the velocity of the last model level): 
     743The linear bottom friction parameterisation (including the special case  
     744of a free-slip condition) assumes that the bottom friction  
     745is proportional to the interior velocity (i.e. the velocity of the last  
     746model level): 
    531747\begin{equation} \label{Eq_zdfbfr_linear} 
    532 \textbf{F}_h = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
    533 \end{equation} 
    534 where $\textbf{U}_h^b$ is the horizontal velocity vector of the bottom ocean  
    535 layer and $r$ is a friction coefficient expressed in m.s$^{-1}$. This coefficient  
    536 is generally estimated by setting a typical decay time $\tau$ in the deep ocean,  
     748{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b 
     749\end{equation} 
     750where $r$ is a friction coefficient expressed in ms$^{-1}$.  
     751This coefficient is generally estimated by setting a typical decay time  
     752$\tau$ in the deep ocean,  
    537753and setting $r = H / \tau$, where $H$ is the ocean depth. Commonly accepted  
    538 values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly1984}.  
     754values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly_JMR84}.  
    539755A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used  
    540756in quasi-geostrophic models. One may consider the linear friction as an  
    541757approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982},  
    542758Eq. 9.6.6). For example, with a drag coefficient $C_D = 0.002$, a typical speed  
    543 of tidal currents of $U_{av} =0.1$~m.s$^{-1}$, and assuming an ocean depth  
    544 $H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m.s$^{-1}$.  
     759of tidal currents of $U_{av} =0.1$~m\;s$^{-1}$, and assuming an ocean depth  
     760$H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m\;s$^{-1}$.  
    545761This is the default value used in \NEMO. It corresponds to a decay time scale  
    546 of 115~days. It can be changed by specifying \np{bfric1} (namelist parameter). 
    547  
    548 In the code, the bottom friction is imposed by updating the value of the  
    549 vertical eddy coefficient at the bottom level. Indeed, the discrete formulation  
    550 of (\ref{Eq_zdfbfr_linear}) at the last ocean $T-$level, using the fact that  
    551 $\textbf {U}_h =0$ below the ocean floor, leads to 
    552 \begin{equation} \label{Eq_zdfbfr_linKz} 
     762of 115~days. It can be changed by specifying \np{rn\_bfric1} (namelist parameter). 
     763 
     764For the linear friction case the coefficients defined in the general  
     765expression \eqref{Eq_zdfbfr_bdef} are:  
     766\begin{equation} \label{Eq_zdfbfr_linbfr_b} 
    553767\begin{split} 
    554 A_u^{vm} &= r\;e_{3uw}\\ 
    555 A_v^{vm} &= r\;e_{3vw}\\ 
     768 c_b^u &= - r\\ 
     769 c_b^v &= - r\\ 
    556770\end{split} 
    557771\end{equation} 
    558  
    559 This update is done in \mdl{zdfbfr} when \np{nbotfr}=1. The value of $r$  
    560 used is \np{bfric1}. Setting \np{nbotfr}=3 is equivalent to setting $r=0$ and  
    561 leads to a free-slip bottom boundary condition. Setting \np{nbotfr}=0 sets  
    562 $r=2\;A_{vb}^{\rm {\bf U}} $, where $A_{vb}^{\rm {\bf U}} $ is the background  
    563 vertical eddy coefficient, and a no-slip boundary condition is imposed.  
    564 Note that this latter choice generally leads to an underestimation of the  
    565 bottom friction: for example with a deepest level thickness of $200~m$  
    566 and $A_{vb}^{\rm {\bf U}} =10^{-4}$m$^2$.s$^{-1}$, the friction coefficient  
    567 is only $r=10^{-6}$m.s$^{-1}$. 
     772When \np{nn\_botfr}=1, the value of $r$ used is \np{rn\_bfric1}.  
     773Setting \np{nn\_botfr}=0 is equivalent to setting $r=0$ and leads to a free-slip  
     774bottom boundary condition. These values are assigned in \mdl{zdfbfr}.  
     775From v3.2 onwards there is support for local enhancement of these values  
     776via an externally defined 2D mask array (\np{ln\_bfr2d}=true) given 
     777in the \ifile{bfr\_coef} input NetCDF file. The mask values should vary from 0 to 1.  
     778Locations with a non-zero mask value will have the friction coefficient increased  
     779by $mask\_value$*\np{rn\_bfrien}*\np{rn\_bfric1}. 
    568780 
    569781% ------------------------------------------------------------------------------------------------------------- 
    570782%       Non-Linear Bottom Friction 
    571783% ------------------------------------------------------------------------------------------------------------- 
    572 \subsection{Non-Linear Bottom Friction (\np{nbotfr} = 2)} 
     784\subsection{Non-Linear Bottom Friction (\np{nn\_botfr} = 2)} 
    573785\label{ZDF_bfr_nonlinear} 
    574786 
     
    576788friction is quadratic:  
    577789\begin{equation} \label{Eq_zdfbfr_nonlinear} 
    578 \textbf {F}_h = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h  
     790{\cal F}_h^\textbf{U} = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h  
    579791}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b  
    580792\end{equation} 
    581  
    582 with $\textbf{U}_h^b = (u_b\;,\;v_b)$ the horizontal interior velocity ($i.e.$  
    583 the horizontal velocity of the bottom ocean layer), $C_D$ a drag coefficient,  
    584 and $e_b $ a bottom turbulent kinetic energy due to tides, internal waves  
    585 breaking and other short time scale currents. A typical value of the drag  
    586 coefficient is $C_D = 10^{-3} $. As an example, the CME experiment  
    587 \citep{Treguier1992} uses $C_D = 10^{-3}$ and $e_b = 2.5\;10^{-3}$m$^2$.s$^{-2}$,  
    588 while the FRAM experiment \citep{Killworth1992} uses $e_b =0$  
    589 and $e_b =2.5\;\;10^{-3}$m$^2$.s$^{-2}$. The FRAM choices have been  
    590 set as default values (\np{bfric2} and \np{bfeb2} namelist parameters). 
     793where $C_D$ is a drag coefficient, and $e_b $ a bottom turbulent kinetic energy  
     794due to tides, internal waves breaking and other short time scale currents.  
     795A typical value of the drag coefficient is $C_D = 10^{-3} $. As an example,  
     796the CME experiment \citep{Treguier_JGR92} uses $C_D = 10^{-3}$ and  
     797$e_b = 2.5\;10^{-3}$m$^2$\;s$^{-2}$, while the FRAM experiment \citep{Killworth1992}  
     798uses $C_D = 1.4\;10^{-3}$ and $e_b =2.5\;\;10^{-3}$m$^2$\;s$^{-2}$.  
     799The CME choices have been set as default values (\np{rn\_bfric2} and \np{rn\_bfeb2}  
     800namelist parameters). 
    591801 
    592802As for the linear case, the bottom friction is imposed in the code by  
    593 updating the value of the vertical eddy coefficient at the bottom level: 
    594 \begin{equation} \label{Eq_zdfbfr_nonlinKz} 
     803adding the trend due to the bottom friction to the general momentum trend  
     804in \mdl{dynbfr}. 
     805For the non-linear friction case the terms 
     806computed in \mdl{zdfbfr}  are:  
     807\begin{equation} \label{Eq_zdfbfr_nonlinbfr} 
    595808\begin{split} 
    596 A_u^{vm} &=C_D\; e_{3uw} \left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^ 
    597 {1/2}\\ 
    598 A_v^{vm} &=C_D\; e_{3uw} \left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\ 
     809 c_b^u &= - \; C_D\;\left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^{1/2}\\ 
     810 c_b^v &= - \; C_D\;\left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\ 
    599811\end{split} 
    600812\end{equation} 
    601813 
    602 This update is done in \mdl{zdfbfr}. The coefficients that control the strength of the  
    603 non-linear bottom friction are initialized as namelist parameters: $C_D$= \np{bfri2},  
    604 and $e_b$ =\np{bfeb2}. 
    605  
    606 % ================================================================ 
     814The coefficients that control the strength of the non-linear bottom friction are  
     815initialised as namelist parameters: $C_D$= \np{rn\_bfri2}, and $e_b$ =\np{rn\_bfeb2}.  
     816Note for applications which treat tides explicitly a low or even zero value of  
     817\np{rn\_bfeb2} is recommended. From v3.2 onwards a local enhancement of $C_D$  
     818is possible via an externally defined 2D mask array (\np{ln\_bfr2d}=true).  
     819See previous section for details. 
     820 
     821% ------------------------------------------------------------------------------------------------------------- 
     822%       Bottom Friction stability 
     823% ------------------------------------------------------------------------------------------------------------- 
     824\subsection{Bottom Friction stability considerations} 
     825\label{ZDF_bfr_stability} 
     826 
     827Some care needs to exercised over the choice of parameters to ensure that the 
     828implementation of bottom friction does not induce numerical instability. For  
     829the purposes of stability analysis, an approximation to \eqref{Eq_zdfbfr_flux2} 
     830is: 
     831\begin{equation} \label{Eqn_bfrstab} 
     832\begin{split} 
     833 \Delta u &= -\frac{{{\cal F}_h}^u}{e_{3u}}\;2 \rdt    \\ 
     834               &= -\frac{ru}{e_{3u}}\;2\rdt\\ 
     835\end{split} 
     836\end{equation} 
     837\noindent where linear bottom friction and a leapfrog timestep have been assumed.  
     838To ensure that the bottom friction cannot reverse the direction of flow it is necessary to have: 
     839\begin{equation} 
     840 |\Delta u| < \;|u|  
     841\end{equation} 
     842\noindent which, using \eqref{Eqn_bfrstab}, gives: 
     843\begin{equation} 
     844r\frac{2\rdt}{e_{3u}} < 1 \qquad  \Rightarrow \qquad r < \frac{e_{3u}}{2\rdt}\\ 
     845\end{equation} 
     846This same inequality can also be derived in the non-linear bottom friction case  
     847if a velocity of 1 m.s$^{-1}$ is assumed. Alternatively, this criterion can be  
     848rearranged to suggest a minimum bottom box thickness to ensure stability: 
     849\begin{equation} 
     850e_{3u} > 2\;r\;\rdt 
     851\end{equation} 
     852\noindent which it may be necessary to impose if partial steps are being used.  
     853For example, if $|u| = 1$ m.s$^{-1}$, $rdt = 1800$ s, $r = 10^{-3}$ then 
     854$e_{3u}$ should be greater than 3.6 m. For most applications, with physically 
     855sensible parameters these restrictions should not be of concern. But  
     856caution may be necessary if attempts are made to locally enhance the bottom 
     857friction parameters.  
     858To ensure stability limits are imposed on the bottom friction coefficients both during  
     859initialisation and at each time step. Checks at initialisation are made in \mdl{zdfbfr}  
     860(assuming a 1 m.s$^{-1}$ velocity in the non-linear case). 
     861The number of breaches of the stability criterion are reported as well as the minimum  
     862and maximum values that have been set. The criterion is also checked at each time step,  
     863using the actual velocity, in \mdl{dynbfr}. Values of the bottom friction coefficient are  
     864reduced as necessary to ensure stability; these changes are not reported. 
     865 
     866% ------------------------------------------------------------------------------------------------------------- 
     867%       Bottom Friction with split-explicit time splitting 
     868% ------------------------------------------------------------------------------------------------------------- 
     869\subsection{Bottom Friction with split-explicit time splitting} 
     870\label{ZDF_bfr_ts} 
     871 
     872When calculating the momentum trend due to bottom friction in \mdl{dynbfr}, the 
     873bottom velocity at the before time step is used. This velocity includes both the 
     874baroclinic and barotropic components which is appropriate when using either the 
     875explicit or filtered surface pressure gradient algorithms (\key{dynspg\_exp} or  
     876{\key{dynspg\_flt}). Extra attention is required, however, when using  
     877split-explicit time stepping (\key{dynspg\_ts}). In this case the free surface  
     878equation is solved with a small time step \np{nn\_baro}*\np{rn\_rdt}, while the three  
     879dimensional prognostic variables are solved with a longer time step that is a  
     880multiple of \np{rn\_rdt}. The trend in the barotropic momentum due to bottom  
     881friction appropriate to this method is that given by the selected parameterisation  
     882($i.e.$ linear or non-linear bottom friction) computed with the evolving velocities  
     883at each barotropic timestep.  
     884 
     885In the case of non-linear bottom friction, we have elected to partially linearise  
     886the problem by keeping the coefficients fixed throughout the barotropic  
     887time-stepping to those computed in \mdl{zdfbfr} using the now timestep.  
     888This decision allows an efficient use of the $c_b^{\vect{U}}$ coefficients to: 
     889 
     890\begin{enumerate} 
     891\item On entry to \rou{dyn\_spg\_ts}, remove the contribution of the before 
     892barotropic velocity to the bottom friction component of the vertically 
     893integrated momentum trend. Note the same stability check that is carried out  
     894on the bottom friction coefficient in \mdl{dynbfr} has to be applied here to 
     895ensure that the trend removed matches that which was added in \mdl{dynbfr}. 
     896\item At each barotropic step, compute the contribution of the current barotropic 
     897velocity to the trend due to bottom friction. Add this contribution to the 
     898vertically integrated momentum trend. This contribution is handled implicitly which 
     899eliminates the need to impose a stability criteria on the values of the bottom friction 
     900coefficient within the barotropic loop.  
     901\end{enumerate} 
     902 
     903Note that the use of an implicit formulation 
     904for the bottom friction trend means that any limiting of the bottom friction coefficient  
     905in \mdl{dynbfr} does not adversely affect the solution when using split-explicit time  
     906splitting. This is because the major contribution to bottom friction is likely to come from  
     907the barotropic component which uses the unrestricted value of the coefficient. 
     908 
     909The implicit formulation takes the form: 
     910\begin{equation} \label{Eq_zdfbfr_implicitts} 
     911 \bar{U}^{t+ \rdt} = \; \left [ \bar{U}^{t-\rdt}\; + 2 \rdt\;RHS \right ] / \left [ 1 - 2 \rdt \;c_b^{u} / H_e \right ]   
     912\end{equation} 
     913where $\bar U$ is the barotropic velocity, $H_e$ is the full depth (including sea surface height),  
     914$c_b^u$ is the bottom friction coefficient as calculated in \rou{zdf\_bfr} and $RHS$ represents  
     915all the components to the vertically integrated momentum trend except for that due to bottom friction. 
     916 
     917 
     918 
     919 
     920% ================================================================ 
     921% Tidal Mixing 
     922% ================================================================ 
     923\section{Tidal Mixing} 
     924\label{ZDF_tmx} 
     925 
     926%--------------------------------------------namzdf_tmx-------------------------------------------------- 
     927\namdisplay{namzdf_tmx} 
     928%-------------------------------------------------------------------------------------------------------------- 
     929 
     930 
     931% ------------------------------------------------------------------------------------------------------------- 
     932%        Bottom intensified tidal mixing  
     933% ------------------------------------------------------------------------------------------------------------- 
     934\subsection{Bottom intensified tidal mixing} 
     935\label{ZDF_tmx_bottom} 
     936 
     937The parameterization of tidal mixing follows the general formulation for  
     938the vertical eddy diffusivity proposed by \citet{St_Laurent_al_GRL02} and  
     939first introduced in an OGCM by \citep{Simmons_al_OM04}.  
     940In this formulation an additional vertical diffusivity resulting from internal tide breaking,  
     941$A^{vT}_{tides}$ is expressed as a function of $E(x,y)$, the energy transfer from barotropic  
     942tides to baroclinic tides :  
     943\begin{equation} \label{Eq_Ktides} 
     944A^{vT}_{tides} =  q \,\Gamma \,\frac{ E(x,y) \, F(z) }{ \rho \, N^2 } 
     945\end{equation} 
     946where $\Gamma$ is the mixing efficiency, $N$ the Brunt-Vais\"{a}l\"{a} frequency  
     947(see \S\ref{TRA_bn2}), $\rho$ the density, $q$ the tidal dissipation efficiency,  
     948and $F(z)$ the vertical structure function.  
     949 
     950The mixing efficiency of turbulence is set by $\Gamma$ (\np{rn\_me} namelist parameter) 
     951and is usually taken to be the canonical value of $\Gamma = 0.2$ (Osborn 1980).  
     952The tidal dissipation efficiency is given by the parameter $q$ (\np{rn\_tfe} namelist parameter)  
     953represents the part of the internal wave energy flux $E(x, y)$ that is dissipated locally,  
     954with the remaining $1-q$ radiating away as low mode internal waves and  
     955contributing to the background internal wave field. A value of $q=1/3$ is typically used   
     956\citet{St_Laurent_al_GRL02}. 
     957The vertical structure function $F(z)$ models the distribution of the turbulent mixing in the vertical.  
     958It is implemented as a simple exponential decaying upward away from the bottom,  
     959with a vertical scale of $h_o$ (\np{rn\_htmx} namelist parameter, with a typical value of $500\,m$) \citep{St_Laurent_Nash_DSR04},  
     960\begin{equation} \label{Eq_Fz} 
     961F(i,j,k) = \frac{ e^{ -\frac{H+z}{h_o} } }{ h_o \left( 1- e^{ -\frac{H}{h_o} } \right) } 
     962\end{equation} 
     963and is normalized so that vertical integral over the water column is unity.  
     964 
     965The associated vertical viscosity is calculated from the vertical  
     966diffusivity assuming a Prandtl number of 1, $i.e.$ $A^{vm}_{tides}=A^{vT}_{tides}$.  
     967In the limit of $N \rightarrow 0$ (or becoming negative), the vertical diffusivity  
     968is capped at $300\,cm^2/s$ and impose a lower limit on $N^2$ of \np{rn\_n2min}  
     969usually set to $10^{-8} s^{-2}$. These bounds are usually rarely encountered. 
     970 
     971The internal wave energy map, $E(x, y)$ in \eqref{Eq_Ktides}, is derived  
     972from a barotropic model of the tides utilizing a parameterization of the  
     973conversion of barotropic tidal energy into internal waves.  
     974The essential goal of the parameterization is to represent the momentum  
     975exchange between the barotropic tides and the unrepresented internal waves  
     976induced by the tidal ßow over rough topography in a stratified ocean.  
     977In the current version of \NEMO, the map is built from the output of  
     978the barotropic global ocean tide model MOG2D-G \citep{Carrere_Lyard_GRL03}. 
     979This model provides the dissipation associated with internal wave energy for the M2 and K1  
     980tides component (Fig.~\ref{Fig_ZDF_M2_K1_tmx}). The S2 dissipation is simply approximated 
     981as being $1/4$ of the M2 one. The internal wave energy is thus : $E(x, y) = 1.25 E_{M2} + E_{K1}$.  
     982Its global mean value is $1.1$ TW, in agreement with independent estimates  
     983\citep{Egbert_Ray_Nat00, Egbert_Ray_JGR01}.  
     984 
     985%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     986\begin{figure}[!t] \label{Fig_ZDF_M2_K1_tmx}  \begin{center} 
     987\includegraphics[width=0.90\textwidth]{./TexFiles/Figures/Fig_ZDF_M2_K1_tmx.pdf} 
     988\caption {(a) M2 and (b) K2 internal wave drag energy from \citet{Carrere_Lyard_GRL03} ($W/m^2$). } 
     989\end{center}   
     990\end{figure} 
     991%>>>>>>>>>>>>>>>>>>>>>>>>>>>>  
     992  
     993% ------------------------------------------------------------------------------------------------------------- 
     994%        Indonesian area specific treatment  
     995% ------------------------------------------------------------------------------------------------------------- 
     996\subsection{Indonesian area specific treatment} 
     997\label{ZDF_tmx_itf} 
     998 
     999When the Indonesian Through Flow (ITF) area is included in the model domain, 
     1000a specific treatment of tidal induced mixing in this area can be used.  
     1001It is activated through the namelist logical \np{ln\_tmx\_itf}, and the user must provide 
     1002an input NetCDF file, \ifile{mask\_itf}, which contains a mask array defining the ITF area 
     1003where the specific treatment is applied. 
     1004 
     1005When \np{ln\_tmx\_itf}=true, the two key parameters $q$ and $F(z)$ are adjusted following  
     1006the parameterisation developed by \ref{Koch-Larrouy_al_GRL07}: 
     1007 
     1008First, the Indonesian archipelago is a complex geographic region  
     1009with a series of large, deep, semi-enclosed basins connected via  
     1010numerous narrow straits. Once generated, internal tides remain  
     1011confined within this semi-enclosed area and hardly radiate away.  
     1012Therefore all the internal tides energy is consumed within this area.  
     1013So it is assumed that $q = 1$, $i.e.$ all the energy generated is available for mixing. 
     1014Note that for test purposed, the ITF tidal dissipation efficiency is a  
     1015namelist parameter (\np{rn\_tfe\_itf}). A value of $1$ or close to is 
     1016this recommended for this parameter. 
     1017 
     1018Second, the vertical structure function, $F(z)$, is no more associated 
     1019with a bottom intensification of the mixing, but with a maximum of  
     1020energy available within the thermocline. \ref{Koch-Larrouy_al_GRL07}  
     1021have suggested that the vertical distribution of the energy dissipation  
     1022proportional to $N^2$ below the core of the thermocline and to $N$ above.  
     1023The resulting $F(z)$ is: 
     1024\begin{equation} \label{Eq_Fz_itf} 
     1025F(i,j,k) \sim     \left\{ \begin{aligned} 
     1026\frac{q\,\Gamma E(i,j) } {\rho N \, \int N     dz}    \qquad \text{when $\partial_z N < 0$} \\ 
     1027\frac{q\,\Gamma E(i,j) } {\rho     \, \int N^2 dz}    \qquad \text{when $\partial_z N > 0$} 
     1028                      \end{aligned} \right. 
     1029\end{equation} 
     1030 
     1031Averaged over the ITF area, the resulting tidal mixing coefficient is $1.5\,cm^2/s$,  
     1032which agrees with the independent estimates inferred from observations.  
     1033Introduced in a regional OGCM, the parameterization improves the water mass  
     1034characteristics in the different Indonesian seas, suggesting that the horizontal  
     1035and vertical distributions of the mixing are adequately prescribed  
     1036\citep{Koch-Larrouy_al_GRL07, Koch-Larrouy_al_OD08a, Koch-Larrouy_al_OD08b}. 
     1037Note also that such a parameterisation has a sugnificant impact on the behaviour  
     1038of global coupled GCMs \citep{Koch-Larrouy_al_CD10}. 
     1039 
     1040 
     1041% ================================================================ 
Note: See TracChangeset for help on using the changeset viewer.