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.
Chap_ZDF.tex in trunk/DOC/TexFiles/Chapters – NEMO

source: trunk/DOC/TexFiles/Chapters/Chap_ZDF.tex @ 994

Last change on this file since 994 was 994, checked in by gm, 16 years ago

trunk - add steven correction + several other things + rename BETA into TexFiles?

  • Property svn:executable set to *
File size: 34.5 KB
Line 
1% ================================================================
2% Chapter Ñ Vertical Ocean Physics (ZDF)
3% ================================================================
4\chapter{Vertical Ocean Physics (ZDF)}
5\label{ZDF}
6\minitoc
7
8%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 : 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!
10Gurvan :  I kept "turbulent closure"...}
11\gmcomment{Steven bis : parameterization is the american spelling, parameterisation is the british}
12
13
14% ================================================================
15% Vertical Mixing
16% ================================================================
17\section{Vertical Mixing}
18\label{ZDF_zdf}
19
20The discrete form of the ocean subgrid scale physics has been presented in
21\S\ref{TRA_zdf} and \S\ref{DYN_zdf}. At the surface and bottom boundaries,
22the turbulent fluxes of momentum, heat and salt have to be defined. At the
23surface they are prescribed from the surface forcing (see Chap.~\ref{SBC}),
24while at the bottom they are set to zero for heat and salt, unless a geothermal
25flux forcing is prescribed as a bottom boundary condition ($i.e.$ \key{trabbl} 
26defined, see \S\ref{TRA_bbc}), and specified through a bottom friction
27parameterization for momentum (see \S\ref{ZDF_bfr}).
28
29In this section we briefly discuss the various choices offered to compute
30the vertical eddy viscosity and diffusivity coefficients, $A_u^{vm}$ ,
31$A_v^{vm}$ and $A^{vT}$ ($A^{vS}$), defined at $uw$-, $vw$- and $w$-
32points, respectively (see \S\ref{TRA_zdf} and \S\ref{DYN_zdf}). These
33coefficients can be assumed to be either constant, or a function of the local
34Richardson number, or computed from a turbulent closure model (either
35TKE or KPP formulation). The computation of these coefficients is initialized
36in the \mdl{zdfini} module and performed in the \mdl{zdfric}, \mdl{zdftke} or
37\mdl{zdfkpp} modules. The trends due to the vertical momentum and tracer
38diffusion, including the surface forcing, are computed and added to the
39general trend in the \mdl{dynzdf} and \mdl{trazdf} modules, respectively.
40These trends can be computed using either a forward time stepping scheme
41(namelist parameter \np{np\_zdfexp}=true) or a backward time stepping
42scheme (\np{np\_zdfexp}=false) depending on the magnitude of the mixing
43coefficients, and thus of the formulation used (see \S\ref{DOM_nxt}).
44
45% -------------------------------------------------------------------------------------------------------------
46%        Constant
47% -------------------------------------------------------------------------------------------------------------
48\subsection{Constant (\key{zdfcst})}
49\label{ZDF_cst}
50%--------------------------------------------namzdf---------------------------------------------------------
51\namdisplay{namzdf}
52%--------------------------------------------------------------------------------------------------------------
53
54When \key{zdfcst} is defined, the momentum and tracer vertical eddy coefficients
55are set to constant values over the whole ocean. This is the crudest way to define
56the vertical ocean physics. It is recommended that this option is only used in
57process studies, not in basin scale simulations. Typical values used in this case are:
58\begin{align*} 
59A_u^{vm} = A_v^{vm} &= 1.2\ 10^{-4}~m^2.s^{-1}  \\
60\\
61A^{vT} = A^{vS} &= 1.2\ 10^{-5}~m^2.s^{-1}
62\end{align*}
63
64These values are set through the \np{avm0} and \np{avt0} namelist parameters.
65In all cases, do not use values smaller that those associated with the molecular
66viscosity and diffusivity, that is $\sim10^{-6}~m^2.s^{-1}$ for momentum,
67$\sim10^{-7}~m^2.s^{-1}$ for temperature and $\sim10^{-9}~m^2.s^{-1}$ for salinity.
68
69
70% -------------------------------------------------------------------------------------------------------------
71%        Richardson Number Dependent
72% -------------------------------------------------------------------------------------------------------------
73\subsection{Richardson Number Dependent (\key{zdfric})}
74\label{ZDF_ric}
75
76%--------------------------------------------namric---------------------------------------------------------
77\namdisplay{namric}
78%--------------------------------------------------------------------------------------------------------------
79
80When \key{zdfric} is defined, a local Richardson number dependent formulation
81for the vertical momentum and tracer eddy coefficients is set. The vertical mixing
82coefficients are diagnosed from the large scale variables computed by the model.
83\textit{In situ} measurements have been used to link vertical turbulent activity to
84large scale ocean structures. The hypothesis of a mixing mainly maintained by the
85growth of Kelvin-Helmholtz like instabilities leads to a dependency between the
86vertical turbulence eddy coefficients and the local Richardson number ($i.e.$ the
87ratio of stratification to vertical shear). Following \citet{PacPhil1981}, the following
88formulation has been implemented:
89\begin{equation} \label{Eq_zdfric}
90   \left\{      \begin{aligned}
91         A^{vT} &= \frac {A_{ric}^{vT}}{\left( 1+a \; Ri \right)^n} + A_b^{vT}       \\
92         \\
93         A^{vm} &= \frac{A^{vT}        }{\left( 1+ a \;Ri  \right)   } + A_b^{vm}
94   \end{aligned}    \right.
95\end{equation}
96where $Ri = N^2 / \left(\partial_z \textbf{U}_h \right)^2$ is the local Richardson
97number, $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
98$A_b^{vT} $ and $A_b^{vm}$ are the constant background values set as in the
99constant case (see \S\ref{ZDF_cst}), and $A_{ric}^{vT} = 10^{-4}~m^2.s^{-1}$ 
100is the maximum value that can be reached by the coefficient when $Ri\leq 0$,
101$a=5$ and $n=2$. The last three values can be modified by setting the
102\np{avmri}, \np{alp} and \np{nric} namelist parameters, respectively.
103
104% -------------------------------------------------------------------------------------------------------------
105%        TKE Turbulent Closure Scheme
106% -------------------------------------------------------------------------------------------------------------
107\subsection{TKE Turbulent Closure Scheme (\key{zdftke})}
108\label{ZDF_tke}
109
110%--------------------------------------------namtke---------------------------------------------------------
111\namdisplay{nam_tke}
112%--------------------------------------------------------------------------------------------------------------
113
114The vertical eddy viscosity and diffusivity coefficients are computed from a TKE
115turbulent closure model based on a prognostic equation for $\bar {e}$, the turbulent
116kinetic energy, and a closure assumption for the turbulence length scales. This
117turbulent closure model has been developed by \citet{Bougeault1989} in the
118atmospheric case, adapted by \citet{Gaspar1990} for the oceanic case, and
119embedded in OPA by \citet{Blanke1993} for equatorial Atlantic simulations. Since
120then, significant modifications have been introduced by \citet{Madec1998} in both
121the implementation and the formulation of the mixing length scale. The time
122evolution of $\bar{e}$ is the result of the production of $\bar{e}$ through vertical
123shear, its destruction through stratification, its vertical diffusion, and its dissipation
124of \citet{Kolmogorov1942} type:
125\begin{equation} \label{Eq_zdftke_e}
126\frac{\partial \bar{e}}{\partial t} =
127\frac{A^{vm}}{e_3 }\;\left[ {\left( {\frac{\partial u}{\partial k}} \right)^2
128                     +\left( {\frac{\partial v}{\partial k}} \right)^2} \right]
129-A^{vT}\,N^2
130+\frac{1}{e_3} \;\frac{\partial }{\partial k}\left[ {\frac{A^{vm}}{e_3 }
131            \;\frac{\partial \bar{e}}{\partial k}} \right]
132- c_\epsilon \;\frac{\bar {e}^{3/2}}{l_\epsilon }
133\end{equation}
134\begin{equation} \label{Eq_zdftke_kz}
135   \begin{split}
136         A^{vm} &= C_k\  l_k\  \sqrt {\bar{e}}     \\
137         A^{vT} &= A^{vm} / P_{rt}
138   \end{split}
139\end{equation}
140where $N$ is the local Brunt-Vais\"{a}l\"{a} frequency (see \S\ref{TRA_bn2}),
141$l_{\epsilon }$ and $l_{\kappa }$ are the dissipation and mixing length scales,
142$P_{rt} $ is the Prandtl number. The constants $C_k = \sqrt {2} /2$ and
143$C_\epsilon = 0.1$ are designed to deal with vertical mixing at any depth
144\citep{Gaspar1990}. They are set through namelist parameters \np{ediff} 
145and \np{ediss}. $P_{rt} $ can be set to unity or, following \citet{Blanke1993},
146be a function of the local Richardson number, $R_i $:
147\begin{align*} \label{Eq_prt}
148P_{rt} = \begin{cases}
149                    \ \ \ 1 &      \text{if $\ R_i \leq 0.2$}  \\
150                    5\,R_i &      \text{if $\ 0.2 \leq R_i \leq 2$}  \\
151                    \ \ 10 &      \text{if $\ 2 \leq R_i$} 
152            \end{cases}
153\end{align*}
154Note that a horizontal Shapiro filter can optionally be applied to $R_i$.
155However it is an obsolescent option that is not recommended. 
156The choice of $P_{rt} $ is controlled by the \np{npdl} namelist parameter.
157
158For computational efficiency, the original formulation of the turbulence length
159scales proposed by \citet{Gaspar1990} has been simplified. Four formulations
160are proposed, the choice of which is controlled by the \np{nmxl} namelist
161parameter. The first two are based on the following first order approximation
162\citep{Blanke1993}:
163\begin{equation} \label{Eq_tke_mxl0_1}
164l_k = l_\epsilon = \sqrt {2 \bar e} / N
165\end{equation}
166which is valid in a stable stratified region with constant values of the brunt-
167Vais\"{a}l\"{a} frequency. The resulting length scale is bounded by the distance
168to 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
169drawbacks: it makes no sense for locally unstable stratification and the
170computation no longer uses all the information contained in the vertical density
171profile. To overcome these drawbacks, \citet{Madec1998} introduces the
172\np{nmxl}=2 or 3 cases, which add an extra assumption concerning the vertical
173gradient of the computed length scale. So, the length scales are first evaluated
174as in \eqref{Eq_tke_mxl0_1} and then bounded such that:
175\begin{equation} \label{Eq_tke_mxl_constraint}
176\frac{1}{e_3 }\left| {\frac{\partial l}{\partial k}} \right| \leq 1
177\qquad \text{with }\  l =  l_k = l_\epsilon
178\end{equation}
179\eqref{Eq_tke_mxl_constraint} means that the vertical variations of the length
180scale cannot be larger than the variations of depth. It provides a better
181approximation of the \citet{Gaspar1990} formulation while being much less
182time consuming. In particular, it allows the length scale to be limited not only
183by the distance to the surface or to the ocean bottom but also by the distance
184to a strongly stratified portion of the water column such as the thermocline
185(Fig.~\ref{Fig_mixing_length}). In order to impose the \eqref{Eq_tke_mxl_constraint} 
186constraint, we introduce two additional length scales: $l_{up}$ and $l_{dwn}$,
187the upward and downward length scales, and evaluate the dissipation and
188mixing turbulence length scales as (and note that here we use numerical
189indexing):
190%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
191\begin{figure}[!t] \label{Fig_mixing_length}  \begin{center}
192\includegraphics[width=1.00\textwidth]{./Figures/Fig_mixing_length.pdf}
193\caption {Illustration of the mixing length computation. }
194\end{center} 
195\end{figure}
196%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
197\begin{equation} \label{Eq_tke_mxl2}
198\begin{aligned}
199 l_{up\ \ }^{(k)} &= \min \left(  l^{(k)} \ , \ l_{up}^{(k+1)} + e_{3T}^{(k)}\ \ \ \;  \right)
200    \quad &\text{ from $k=1$ to $jpk$ }\ \\
201 l_{dwn}^{(k)} &= \min \left(  l^{(k)} \ , \ l_{dwn}^{(k-1)} + e_{3T}^{(k-1)}  \right)   
202    \quad &\text{ from $k=jpk$ to $1$ }\ \\
203\end{aligned}
204\end{equation}
205where $l^{(k)}$ is computed using \eqref{Eq_tke_mxl0_1},
206$i.e.$ $l^{(k)} = \sqrt {2 \bar e^{(k)} / N^{(k)} }$.
207
208In the \np{nmxl}=2 case, the dissipation and mixing length scales take the same
209value: $ l_k=  l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)$, while in the
210\np{nmxl}=2 case, the dissipation and mixing turbulence length scales are give
211as in \citet{Gaspar1990}:
212\begin{equation} \label{Eq_tke_mxl_gaspar}
213\begin{aligned}
214& l_k          = \sqrt{\  l_{up} \ \ l_{dwn}\ }    \\
215& l_\epsilon = \min \left(\ l_{up} \;,\;  l_{dwn}\ \right)
216\end{aligned}
217\end{equation}
218
219At the sea surface the value of $\bar{e}$ is prescribed from the wind
220stress field: $\bar{e}=ebb\;\left| \tau \right|$ ($ebb=60$ by default)
221with a minimal threshold of $emin0=10^{-4}~m^2.s^{-2}$ (namelist
222parameters). Its value at the bottom of the ocean is assumed to be
223equal to the value of the level just above. The time integration of the
224$\bar{e}$ equation may formally lead to negative values because the
225numerical scheme does not ensure its positivity. To overcome this
226problem, a cut-off in the minimum value of $\bar{e}$ is used. Following
227\citet{Gaspar1990}, the cut-off value is set to $\sqrt{2}/2~10^{-6}~m^2.s^{-2}$.
228This allows the subsequent formulations to match that of\citet{Gargett1984} 
229for the diffusion in the thermocline and deep ocean :  $(A^{vT} = 10^{-3} / N)$.
230In addition, a cut-off is applied on $A^{vm}$ and $A^{vT}$ to avoid numerical
231instabilities associated with too weak vertical diffusion. They must be
232specified at least larger than the molecular values, and are set through
233\textit{avm0} and \textit{avt0} (\textbf{namelist} parameters).
234
235% -------------------------------------------------------------------------------------------------------------
236%        K Profile Parametrisation (KPP)
237% -------------------------------------------------------------------------------------------------------------
238\subsection{K Profile Parametrisation (KPP) (\key{zdfkpp}) }
239\label{ZDF_kpp}
240
241%--------------------------------------------namkpp--------------------------------------------------------
242\namdisplay{namkpp}
243%--------------------------------------------------------------------------------------------------------------
244
245The KKP scheme has been implemented by J. Chanut ...
246
247\colorbox{yellow}{Add a description of KPP here.}
248
249
250% ================================================================
251% Convection
252% ================================================================
253\section{Convection}
254\label{ZDF_conv}
255
256%--------------------------------------------namzdf--------------------------------------------------------
257\namdisplay{namzdf}
258%--------------------------------------------------------------------------------------------------------------
259
260Static instabilities (i.e. light potential densities under heavy ones) may
261occur at particular ocean grid points. In nature, convective processes
262quickly re-establish the static stability of the water column. These
263processes have been removed from the model via the hydrostatic
264assumption so they must be parameterized. Three parameterizations
265are available to deal with convective processes: a non-penetrative
266convective adjustment or an enhanced vertical diffusion, or/and the
267use of a turbulent closure scheme.
268
269% -------------------------------------------------------------------------------------------------------------
270%       Non-Penetrative Convective Adjustment
271% -------------------------------------------------------------------------------------------------------------
272\subsection   [Non-Penetrative Convective Adjustment (\np{ln\_tranpc}) ]
273         {Non-Penetrative Convective Adjustment (\np{ln\_tranpc}=.true.) }
274\label{ZDF_npc}
275
276%--------------------------------------------namnpc--------------------------------------------------------
277\namdisplay{namnpc}
278%--------------------------------------------------------------------------------------------------------------
279
280
281%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
282\begin{figure}[!htb] \label{Fig_npc}   \begin{center}
283\includegraphics[width=0.90\textwidth]{./Figures/Fig_npc.pdf}
284\caption {Example of an unstable density profile treated by the non penetrative
285convective adjustment algorithm. $1^{st}$ step: the initial profile is checked from
286the surface to the bottom. It is found to be unstable between levels 3 and 4.
287They are mixed. The resulting $\rho$ is still larger than $\rho$(5): levels 3 to 5
288are mixed. The resulting $\rho$ is still larger than $\rho$(6): levels 3 to 6 are
289mixed. The $1^{st}$ step ends since the density profile is then stable below
290the level 3. $2^{nd}$ step: the new $\rho$ profile is checked following the same
291procedure as in $1^{st}$ step: levels 2 to 5 are mixed. The new density profile
292is checked. It is found stable: end of algorithm.}
293\end{center}   \end{figure}
294%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
295
296The non-penetrative convective adjustment is used when \np{ln\_zdfnpc}=true.
297It is applied at each \np{nnpc1} time step and mixes downwards instantaneously
298the statically unstable portion of the water column, but only until the density
299structure becomes neutrally stable ($i.e.$ until the mixed portion of the water
300column has \textit{exactly} the density of the water just below) \citep{Madec1991a}.
301The associated algorithm is an iterative process used in the following way
302(Fig. \ref{Fig_npc}): starting from the top of the ocean, the first instability is
303found. Assume in the following that the instability is located between levels
304$k$ and $k+1$. The potential temperature and salinity in the two levels are
305vertically mixed, conserving the heat and salt contents of the water column.
306The new density is then computed by a linear approximation. If the new
307density profile is still unstable between levels $k+1$ and $k+2$, levels $k$,
308$k+1$ and $k+2$ are then mixed. This process is repeated until stability is
309established below the level $k$ (the mixing process can go down to the
310ocean bottom). The algorithm is repeated to check if the density profile
311between level $k-1$ and $k$ is unstable and/or if there is no deeper instability.
312
313This algorithm is significantly different from mixing statically unstable levels
314two by two. The latter procedure cannot converge with a finite number
315of iterations for some vertical profiles while the algorithm used in \NEMO 
316converges for any profile in a number of iterations which is less than the
317number of vertical levels. This property is of paramount importance as
318pointed out by \citet{Killworth1989}: it avoids the existence of permanent
319and unrealistic static instabilities at the sea surface. This non-penetrative
320convective algorithm has been proved successful in studies of the deep
321water formation in the north-western Mediterranean Sea
322\citep{Madec1991a, Madec1991b, Madec1991c}.
323
324Note that in the current implementation of this algorithm presents several
325limitations. First, potential density referenced to the sea surface is used to
326check whether the density profile is stable or not. This is a strong
327simplification which leads to large errors for realistic ocean simulations.
328Indeed, many water masses of the world ocean, especially Antarctic Bottom
329Water, are unstable when represented in surface-referenced potential density.
330The scheme will erroneously mix them up. Second, the mixing of potential
331density is assumed to be linear. This assures the convergence of the algorithm
332even when the equation of state is non-linear. Small static instabilities can thus
333persist due to cabbeling: they will be treated at the next time step.
334Third, temperature and salinity, and thus density, are mixed, but the
335corresponding velocity fields remain unchanged. When using a Richardson
336Number dependent eddy viscosity, the mixing of momentum is done through
337the vertical diffusion: after a static adjustment, the Richardson Number is zero
338and thus the eddy viscosity coefficient is at a maximum. When this convective
339adjustment algorithm is used with constant vertical eddy viscosity, spurious
340solutions can occur since the vertical momentum diffusion remains small even
341after a static adjustment. In that case, we recommend the addition of momentum
342mixing in a manner that mimics the mixing in temperature and salinity
343\citep{Speich1992, Speich1996}.
344
345% -------------------------------------------------------------------------------------------------------------
346%       Enhanced Vertical Diffusion
347% -------------------------------------------------------------------------------------------------------------
348\subsection   [Enhanced Vertical Diffusion (\np{ln\_zdfevd})]
349         {Enhanced Vertical Diffusion (\np{ln\_zdfevd}=.true.)}
350\label{ZDF_evd}
351
352%--------------------------------------------namzdf--------------------------------------------------------
353\namdisplay{namzdf}
354%--------------------------------------------------------------------------------------------------------------
355
356The enhanced vertical diffusion parameterization is used when \np{ln\_zdfevd}=true.
357In this case, the vertical eddy mixing coefficients are assigned very large values
358(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}.
360This is done either on tracers only (\np{n\_evdm}=0) or on both momentum and
361tracers (\np{n\_evdm}=1).
362
363In practice, where $N^2\leq 10^{-12}$, $A_T^{vT}$ and $A_T^{vS}$, and
364if \np{n\_evdm}=1, the four neighbouring $A_u^{vm} \;\mbox{and}\;A_v^{vm}$ 
365values also, are set equal to the namelist parameter \np{avevd}. A typical value
366for $avevd$ is between 1 and $100~m^2.s^{-1}$. This parameterization of
367convective processes is less time consuming than the convective adjustment
368algorithm presented above when mixing both tracers and momentum in the
369case of static instabilities. It requires the use of an implicit time stepping on
370vertical diffusion terms (i.e. \np{ln\_zdfexp}=false).
371
372% -------------------------------------------------------------------------------------------------------------
373%       Turbulent Closure Scheme
374% -------------------------------------------------------------------------------------------------------------
375\subsection{Turbulent Closure Scheme (\key{zdftke})}
376\label{ZDF_tcs}
377
378The TKE turbulent closure scheme presented in \S\ref{ZDF_tke} and used
379when the \key{zdftke} is defined, in theory solves the problem of statically
380unstable density profiles. In such a case, the term corresponding to the
381destruction of turbulent kinetic energy through stratification in \eqref{Eq_zdftke_e} 
382becomes a source term, since $N^2$ is negative. It results in large values of
383$A_T^{vT}$ and  $A_T^{vT}$, and also the four neighbouring
384$A_u^{vm} {and}\;A_v^{vm}$ (up to $1\;m^2s^{-1})$. These large values
385restore the static stability of the water column in a way similar to that of the
386enhanced vertical diffusion parameterization (\S\ref{ZDF_evd}). However,
387in the vicinity of the sea surface (first ocean layer), the eddy coefficients
388computed by the turbulence scheme do not usually exceed $10^{-2}m.s^{-1}$,
389because the mixing length scale is bounded by the distance to the sea surface
390(see \S\ref{ZDF_tke}). It can thus be useful to combine the enhanced vertical
391diffusion with the turbulent closure scheme, $i.e.$ setting the \np{ln\_zdfnpc} 
392namelist parameter to true and defining the \key{zdftke} CPP key all together.
393
394The KPP turbulent closure scheme already includes enhanced vertical diffusion
395in the case of convection, as governed by the variables $bvsqcon$ and $difcon$ 
396found in \mdl{zdfkpp}, therefore \np{np\_zdfevd} should not be used with the KPP
397scheme. %gm%  + one word on non local flux with KPP scheme trakpp.F90 module...
398
399% ================================================================
400% Double Diffusion Mixing
401% ================================================================
402\section  [Double Diffusion Mixing (\textit{zdfddm} - \key{zdfddm})]
403      {Double Diffusion Mixing (\mdl{zdfddm} module - \key{zdfddm})}
404\label{ZDF_ddm}
405
406%-------------------------------------------namddm--------------------------------------------------------
407\namdisplay{namddm}
408%--------------------------------------------------------------------------------------------------------------
409
410Double diffusion occurs when relatively warm, salty water overlies cooler, fresher
411water, or vice versa. The former condition leads to salt fingering and the latter
412to diffusive convection. Double-diffusive phenomena contribute to diapycnal
413mixing in extensive regions of the ocean.  \citet{Merryfield1999} include a
414parameterization of such phenomena in a global ocean model and show that
415it leads to relatively minor changes in circulation but exerts significant regional
416influences on temperature and salinity.
417
418Diapycnal mixing of S and T are described by diapycnal diffusion coefficients
419\begin{align*} % \label{Eq_zdfddm_Kz}
420    &A^{vT} = A_o^{vT}+A_f^{vT}+A_d^{vT}  \\
421    &A^{vS} = A_o^{vS}+A_f^{vS}+A_d^{vS}
422\end{align*}
423where subscript $f$ represents mixing by salt fingering, $d$ by diffusive convection,
424and $o$ by processes other than double diffusion. The rates of double-diffusive mixing depend on the buoyancy ratio $R_\rho = \alpha \partial_z T / \beta \partial_z S$,
425where $\alpha$ and $\beta$ are coefficients of thermal expansion and saline
426contraction (see \S\ref{TRA_eos}). To represent mixing of $S$ and $T$ by salt
427fingering, we adopt the diapycnal diffusivities suggested by Schmitt (1981):
428\begin{align} \label{Eq_zdfddm_f}
429A_f^{vS} &=    \begin{cases}
430   \frac{A^{\ast v}}{1+(R_\rho / R_c)^n   } &\text{if  $R_\rho > 1$ and $N^2>0$ } \\
431   0                              &\text{otherwise} 
432            \end{cases}   
433\\           \label{Eq_zdfddm_f_T}
434A_f^{vT} &= 0.7 \ A_f^{vS} / R_\rho 
435\end{align}
436
437%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
438\begin{figure}[!t] \label{Fig_zdfddm}  \begin{center}
439\includegraphics[width=0.99\textwidth]{./Figures/Fig_zdfddm.pdf}
440\caption {From \citet{Merryfield1999} : (a) Diapycnal diffusivities $A_f^{vT}$ 
441and $A_f^{vS}$ for temperature and salt in regions of salt fingering. Heavy
442curves denote $A^{\ast v} = 10^{-3}~m^2.s^{-1}$ and thin curves
443$A^{\ast v} = 10^{-4}~m^2.s^{-1}$ ; (b) diapycnal diffusivities $A_d^{vT}$ and
444$A_d^{vS}$ for temperature and salt in regions of diffusive convection. Heavy
445curves denote the Federov parameterization and thin curves the Kelley
446parameterization. The latter is not implemented in \NEMO. }
447\end{center}    \end{figure}
448%>>>>>>>>>>>>>>>>>>>>>>>>>>>>
449
450The factor 0.7 in \eqref{Eq_zdfddm_f_T} reflects the measured ratio
451$\alpha F_T /\beta F_S \approx  0.7$ of buoyancy flux of heat to buoyancy
452flux of salt ($e.g.$, \citet{McDougall_Taylor_JMR84}). Following  \citet{Merryfield1999},
453we adopt $R_c = 1.6$, $n = 6$, and $A^{\ast v} = 10^{-4}~m^2.s^{-1}$.
454
455To represent mixing of S and T by diffusive layering,  the diapycnal diffusivities suggested by Federov (1988) is used:
456\begin{align}  \label{Eq_zdfddm_d}
457A_d^{vT} &=    \begin{cases}
458   1.3635 \, \exp{\left( 4.6\, \exp{ \left[  -0.54\,( R_{\rho}^{-1} - 1 )  \right] }    \right)}
459                           &\text{if  $0<R_\rho < 1$ and $N^2>0$ } \\
460   0                       &\text{otherwise} 
461            \end{cases}   
462\\          \label{Eq_zdfddm_d_S}
463A_d^{vS} &=    \begin{cases}
464   A_d^{vT}\ \left( 1.85\,R_{\rho} - 0.85 \right)
465                           &\text{if  $0.5 \leq R_\rho<1$ and $N^2>0$ } \\
466   A_d^{vT} \ 0.15 \ R_\rho
467                           &\text{if  $\ \ 0 < R_\rho<0.5$ and $N^2>0$ } \\
468   0                       &\text{otherwise} 
469            \end{cases}   
470\end{align}
471
472The dependencies of \eqref{Eq_zdfddm_f} to \eqref{Eq_zdfddm_d_S} on $R_\rho$ 
473are illustrated in Fig.~\ref{Fig_zdfddm}. Implementing this requires computing
474$R_\rho$ at each grid point on every time step. This is done in \mdl{eosbn2} at the
475same time as $N^2$ is computed. This avoids duplication in the computation of
476$\alpha$ and $\beta$ (which is usually quite expensive).
477
478% ================================================================
479% Bottom Friction
480% ================================================================
481\section  [Bottom Friction (\textit{zdfbfr})]
482      {Bottom Friction (\mdl{zdfbfr} module)}
483\label{ZDF_bfr}
484
485%--------------------------------------------nambfr--------------------------------------------------------
486\namdisplay{nambfr}
487%--------------------------------------------------------------------------------------------------------------
488
489Both the surface momentum flux (wind stress) and the bottom momentum
490flux (bottom friction) enter the equations as a condition on the vertical
491diffusive flux. For the bottom boundary layer, one has:
492\begin{equation} \label{Eq_zdfbfr_flux}
493A^{vm} \left( \partial \textbf{U}_h / \partial z \right) = \textbf{F}_h
494\end{equation}
495where $\textbf{F}_h$ is supposed to represent the horizontal momentum flux
496outside the logarithmic turbulent boundary layer (thickness of the order of
4971~m in the ocean). How $\textbf{F}_h$ influences the interior depends on the
498vertical resolution of the model near the bottom relative to the Ekman layer
499depth. For example, in order to obtain an Ekman layer depth
500$d = \sqrt{2\;A^{vm}} / f = 50$~m, one needs a vertical diffusion coefficient
501$A^{vm} = 0.125$~m$^2$s$^{-1}$ (for a Coriolis frequency
502$f = 10^{-4}$~m$^2$s$^{-1}$). With a background diffusion coefficient
503$A^{vm} = 10^{-4}$~m$^2$s$^{-1}$, the Ekman layer depth is only 1.4~m.
504When the vertical mixing coefficient is this small, using a flux condition is
505equivalent to entering the viscous forces (either wind stress or bottom friction)
506as a body force over the depth of the top or bottom model layer. To illustrate
507this, consider the equation for $u$ at $k$, the last ocean level:
508\begin{equation} \label{Eq_zdfbfr_flux2}
509\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}}
510\end{equation}
511For example, if the bottom layer thickness is 200~m, the Ekman transport will
512be distributed over that depth. On the other hand, if the vertical resolution
513is high (1~m or less) and a turbulent closure model is used, the turbulent
514Ekman layer will be represented explicitly by the model. However, the
515logarithmic layer is never represented in current primitive equation model
516applications: it is \emph{necessary} to parameterize the flux $\textbf{F}_h $.
517Two choices are available in \NEMO: a linear and a quadratic bottom friction.
518Note that in both cases, the rotation between the interior velocity and the
519bottom friction is neglected in the present release of \NEMO.
520
521% -------------------------------------------------------------------------------------------------------------
522%       Linear Bottom Friction
523% -------------------------------------------------------------------------------------------------------------
524\subsection{Linear Bottom Friction (\np{nbotfr} = 1) }
525\label{ZDF_bfr_linear}
526
527The linear bottom friction parameterization assumes that the bottom friction
528is proportional to the interior velocity (i.e. the velocity of the last model level):
529\begin{equation} \label{Eq_zdfbfr_linear}
530\textbf{F}_h = \frac{A^{vm}}{e_3} \; \frac{\partial \textbf{U}_h}{\partial k} = r \; \textbf{U}_h^b
531\end{equation}
532where $\textbf{U}_h^b$ is the horizontal velocity vector of the bottom ocean
533layer and $r$ is a friction coefficient expressed in m.s$^{-1}$. This coefficient
534is generally estimated by setting a typical decay time $\tau$ in the deep ocean,
535and setting $r = H / \tau$, where $H$ is the ocean depth. Commonly accepted
536values of $\tau$ are of the order of 100 to 200 days \citep{Weatherly1984}.
537A value $\tau^{-1} = 10^{-7}$~s$^{-1}$ equivalent to 115 days, is usually used
538in quasi-geostrophic models. One may consider the linear friction as an
539approximation of quadratic friction, $r \approx 2\;C_D\;U_{av}$ (\citet{Gill1982},
540Eq. 9.6.6). For example, with a drag coefficient $C_D = 0.002$, a typical speed
541of tidal currents of $U_{av} =0.1$~m.s$^{-1}$, and assuming an ocean depth
542$H = 4000$~m, the resulting friction coefficient is $r = 4\;10^{-4}$~m.s$^{-1}$.
543This is the default value used in \NEMO. It corresponds to a decay time scale
544of 115~days. It can be changed by specifying \np{bfric1} (namelist parameter).
545
546In the code, the bottom friction is imposed by updating the value of the
547vertical eddy coefficient at the bottom level. Indeed, the discrete formulation
548of (\ref{Eq_zdfbfr_linear}) at the last ocean $T-$level, using the fact that
549$\textbf {U}_h =0$ below the ocean floor, leads to
550\begin{equation} \label{Eq_zdfbfr_linKz}
551\begin{split}
552A_u^{vm} &= r\;e_{3uw}\\
553A_v^{vm} &= r\;e_{3vw}\\
554\end{split}
555\end{equation}
556
557This update is done in \mdl{zdfbfr} when \np{nbotfr}=1. The value of $r$ 
558used is \np{bfric1}. Setting \np{nbotfr}=3 is equivalent to setting $r=0$ and
559leads to a free-slip bottom boundary condition. Setting \np{nbotfr}=0 sets
560$r=2\;A_{vb}^{\rm {\bf U}} $, where $A_{vb}^{\rm {\bf U}} $ is the background
561vertical eddy coefficient, and a no-slip boundary condition is imposed.
562Note that this latter choice generally leads to an underestimation of the
563bottom friction: for example with a deepest level thickness of $200~m$ 
564and $A_{vb}^{\rm {\bf U}} =10^{-4}$m$^2$.s$^{-1}$, the friction coefficient
565is only $r=10^{-6}$m.s$^{-1}$.
566
567% -------------------------------------------------------------------------------------------------------------
568%       Non-Linear Bottom Friction
569% -------------------------------------------------------------------------------------------------------------
570\subsection{Non-Linear Bottom Friction (\np{nbotfr} = 2)}
571\label{ZDF_bfr_nonlinear}
572
573The non-linear bottom friction parameterization assumes that the bottom
574friction is quadratic:
575\begin{equation} \label{Eq_zdfbfr_nonlinear}
576\textbf {F}_h = \frac{A^{vm}}{e_3 }\frac{\partial \textbf {U}_h
577}{\partial k}=C_D \;\sqrt {u_b ^2+v_b ^2+e_b } \;\; \textbf {U}_h^b
578\end{equation}
579
580with $\textbf{U}_h^b = (u_b\;,\;v_b)$ the horizontal interior velocity ($i.e.$ 
581the horizontal velocity of the bottom ocean layer), $C_D$ a drag coefficient,
582and $e_b $ a bottom turbulent kinetic energy due to tides, internal waves
583breaking and other short time scale currents. A typical value of the drag
584coefficient is $C_D = 10^{-3} $. As an example, the CME experiment
585\citep{Treguier1992} uses $C_D = 10^{-3}$ and $e_b = 2.5\;10^{-3}$m$^2$.s$^{-2}$,
586while the FRAM experiment \citep{Killworth1992} uses $e_b =0$ 
587and $e_b =2.5\;\;10^{-3}$m$^2$.s$^{-2}$. The FRAM choices have been
588set as default values (\np{bfric2} and \np{bfeb2} namelist parameters).
589
590As for the linear case, the bottom friction is imposed in the code by
591updating the value of the vertical eddy coefficient at the bottom level:
592\begin{equation} \label{Eq_zdfbfr_nonlinKz}
593\begin{split}
594A_u^{vm} &=C_D\; e_{3uw} \left[ u^2 + \left(\bar{\bar{v}}^{i+1,j}\right)^2 + e_b \right]^
595{1/2}\\
596A_v^{vm} &=C_D\; e_{3uw} \left[  \left(\bar{\bar{u}}^{i,j+1}\right)^2 + v^2 + e_b \right]^{1/2}\\
597\end{split}
598\end{equation}
599
600This update is done in \mdl{zdfbfr}. The coefficients that control the strength of the
601non-linear bottom friction are initialized as namelist parameters: $C_D$= \np{bfri2}, and $e_b$ =\np{bfeb2}.
602
603% ================================================================
Note: See TracBrowser for help on using the repository browser.