1 |
% $Header: $ |
2 |
% $Name: $ |
3 |
|
4 |
\section{Spatial discretization of the dynamical equations} |
5 |
|
6 |
\subsection{C grid staggering of variables} |
7 |
|
8 |
\begin{figure} |
9 |
\centerline{ \resizebox{!}{2in}{ \includegraphics{part2/cgrid3d.eps}} } |
10 |
\label{fig-cgrid3d} |
11 |
\caption{Three dimensional staggering of velocity components. This |
12 |
facilitates the natural discretization of the continuity and tracer |
13 |
equations. } |
14 |
\end{figure} |
15 |
|
16 |
\subsection{Horizontal grid} |
17 |
|
18 |
\begin{figure} |
19 |
\centerline{ \begin{tabular}{cc} |
20 |
\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Ac.eps}} |
21 |
& \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Az.eps}} |
22 |
\\ |
23 |
\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Au.eps}} |
24 |
& \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Av.eps}} |
25 |
\end{tabular} } |
26 |
\label{fig-hgrid} |
27 |
\caption{Three dimensional staggering of velocity components. This |
28 |
facilitates the natural discretization of the continuity and tracer |
29 |
equations. } |
30 |
\end{figure} |
31 |
|
32 |
\subsection{Vertical grid} |
33 |
|
34 |
\begin{figure} |
35 |
\centerline{ \begin{tabular}{cc} |
36 |
\raisebox{4in}{a)} |
37 |
\resizebox{!}{4in}{ \includegraphics{part2/vgrid-cellcentered.eps}} |
38 |
& |
39 |
\raisebox{4in}{b)} |
40 |
\resizebox{!}{4in}{ \includegraphics{part2/vgrid-accurate.eps}} |
41 |
\end{tabular} } |
42 |
\label{fig-vgrid} |
43 |
\caption{Two versions of the vertical grid. a) The cell centered |
44 |
approach where the interface depths are specified and the tracer |
45 |
points centered in between the interfaces. b) The interface centered |
46 |
approach where tracer levels are specified and the w-interfaces are |
47 |
centered in between.} |
48 |
\end{figure} |
49 |
|
50 |
|
51 |
\subsection{Continuity and horizontal pressure gradient terms} |
52 |
|
53 |
The core algorithm is based on the ``C grid'' discretization of the |
54 |
continuity equation which can be summarized as: |
55 |
\begin{eqnarray} |
56 |
\partial_t u + \frac{1}{\Delta x_c} \delta_i \left. \frac{ \partial \Phi}{\partial r}\right|_{s} \eta + \frac{\epsilon_{nh}}{\Delta x_c} \delta_i \Phi_{nh}' & = & G_u - \frac{1}{\Delta x_c} \delta_i \Phi_h' \\ |
57 |
\partial_t v + \frac{1}{\Delta y_c} \delta_j \left. \frac{ \partial \Phi}{\partial r}\right|_{s} \eta + \frac{\epsilon_{nh}}{\Delta y_c} \delta_j \Phi_{nh}' & = & G_v - \frac{1}{\Delta y_c} \delta_j \Phi_h' \\ |
58 |
\epsilon_{nh} \left( \partial_t w + \frac{1}{\Delta r_c} \delta_k \Phi_{nh}' \right) & = & \epsilon_{nh} G_w + \overline{b}^k - \frac{1}{\Delta r_c} \delta_k \Phi_{h}' \\ |
59 |
\delta_i \Delta y_g \Delta r_f h_w u + |
60 |
\delta_j \Delta x_g \Delta r_f h_s v + |
61 |
\delta_k {\cal A}_c w & = & {\cal A}_c \delta_k (P-E)_{r=0} |
62 |
\end{eqnarray} |
63 |
where the continuity equation has been most naturally discretized by |
64 |
staggering the three components of velocity as shown in |
65 |
Fig.~\ref{fig-cgrid3d}. The grid lengths $\Delta x_c$ and $\Delta y_c$ |
66 |
are the lengths between tracer points (cell centers). The grid lengths |
67 |
$\Delta x_g$, $\Delta y_g$ are the grid lengths between cell |
68 |
corners. $\Delta r_f$ and $\Delta r_c$ are the distance (in units of |
69 |
$r$) between level interfaces (w-level) and level centers (tracer |
70 |
level). The surface area presented in the vertical is denoted ${\cal |
71 |
A}_c$. The factors $h_w$ and $h_s$ are non-dimensional fractions |
72 |
(between 0 and 1) that represent the fraction cell depth that is |
73 |
``open'' for fluid flow. |
74 |
\marginpar{$h_w$: {\bf hFacW}} |
75 |
\marginpar{$h_s$: {\bf hFacS}} |
76 |
|
77 |
The last equation, the discrete continuity equation, can be summed in |
78 |
the vertical to yeild the free-surface equation: |
79 |
\begin{equation} |
80 |
{\cal A}_c \partial_t \eta + \delta_i \sum_k \Delta y_g \Delta r_f h_w u + \delta_j \sum_k \Delta x_g \Delta r_f h_s v = |
81 |
{\cal A}_c(P-E)_{r=0} |
82 |
\end{equation} |
83 |
The source term $P-E$ on the rhs of continuity accounts for the local |
84 |
addition of volume due to excess precipitation and run-off over |
85 |
evaporation and only enters the top-level of the {\em ocean} model. |
86 |
|
87 |
\subsection{Hydrostatic balance} |
88 |
|
89 |
The vertical momentum equation has the hydrostatic or |
90 |
quasi-hydrostatic balance on the right hand side. This discretization |
91 |
guarantees that the conversion of potential to kinetic energy as |
92 |
derived from the buoyancy equation exactly matches the form derived |
93 |
from the pressure gradient terms when forming the kinetic energy |
94 |
equation. |
95 |
|
96 |
In the ocean, using z-ccordinates, the hydrostatic balance terms are |
97 |
discretized: |
98 |
\begin{equation} |
99 |
\epsilon_{nh} \partial_t w |
100 |
+ g \overline{\rho'}^k + \frac{1}{\Delta z} \delta_k \Phi_h' = \ldots |
101 |
\end{equation} |
102 |
|
103 |
In the atmosphere, using p-coordinates, hydrostatic balance is |
104 |
discretized: |
105 |
\begin{equation} |
106 |
\overline{\theta'}^k + \frac{1}{\Delta \Pi} \delta_k \Phi_h' = 0 |
107 |
\end{equation} |
108 |
where $\Delta \Pi$ is the difference in Exner function between the |
109 |
pressure points. The non-hydrostatic equations are not available in |
110 |
the atmosphere. |
111 |
|
112 |
The difference in approach between ocean and atmosphere occurs because |
113 |
of the direct use of the ideal gas equation in forming the potential |
114 |
energy conversion term $\alpha \omega$. The form of these consversion |
115 |
terms is discussed at length in \cite{Adcroft01}. |
116 |
|
117 |
Because of the different representation of hydrostatic balance between |
118 |
ocean and atmosphere there is no elegant way to represent both systems |
119 |
using an arbitrary coordinate. |
120 |
|
121 |
The integration for hydrostatic pressure is made in the positive $r$ |
122 |
direction (increasing k-index). For the ocean, this is from the |
123 |
free-surface down and for the atmosphere this is from the ground up. |
124 |
|
125 |
The calculations are made in the subroutine {\em |
126 |
CALC\_PHI\_HYD}. Inside this routine, one of other of the |
127 |
atmospheric/oceanic form is selected based on the string variable {\bf |
128 |
buoyancyRelation}. |
129 |
|
130 |
\subsection{Flux-form momentum equations} |
131 |
|
132 |
The original finite volume model was based on the Eulerian flux form |
133 |
momentum equations. This is the default though the vector invariant |
134 |
form is optionally available (and recommended in some cases). |
135 |
|
136 |
The ``G's'' (our colloquial name for all terms on rhs!) are broken |
137 |
into the various advective, Coriolis, horizontal dissipation, vertical |
138 |
dissipation and metric forces: |
139 |
\marginpar{$G_u$: {\bf Gu} } |
140 |
\marginpar{$G_v$: {\bf Gv} } |
141 |
\marginpar{$G_w$: {\bf Gw} } |
142 |
\begin{eqnarray} |
143 |
G_u & = & G_u^{adv} + G_u^{cor} + G_u^{h-diss} + G_u^{v-diss} + |
144 |
G_u^{metric} + G_u^{nh-metric} \\ |
145 |
G_v & = & G_v^{adv} + G_v^{cor} + G_v^{h-diss} + G_v^{v-diss} + |
146 |
G_v^{metric} + G_v^{nh-metric} \\ |
147 |
G_w & = & G_w^{adv} + G_w^{cor} + G_w^{h-diss} + G_w^{v-diss} + |
148 |
G_w^{metric} + G_w^{nh-metric} |
149 |
\end{eqnarray} |
150 |
In the hydrostatic limit, $G_w=0$ and $\epsilon_{nh}=0$, reducing the |
151 |
vertical momentum to hydrostatic balance. |
152 |
|
153 |
These terms are calculated in routines called from subroutine {\em |
154 |
CALC\_MOM\_RHS} a collected into the global arrays {\bf Gu}, {\bf Gv}, |
155 |
and {\bf Gw}. |
156 |
|
157 |
\fbox{ \begin{minipage}{4.25in} |
158 |
{\em S/R CALC\_MOM\_RHS} ({\em pkg/mom\_fluxform/calc\_mom\_rhs.F}) |
159 |
|
160 |
$G_u$: {\bf Gu} ({\em DYNVARS.h}) |
161 |
|
162 |
$G_v$: {\bf Gv} ({\em DYNVARS.h}) |
163 |
|
164 |
$G_w$: {\bf Gw} ({\em DYNVARS.h}) |
165 |
\end{minipage} } |
166 |
|
167 |
|
168 |
\subsubsection{Advection of momentum} |
169 |
|
170 |
The advective operator is second order accurate in space: |
171 |
\begin{eqnarray} |
172 |
{\cal A}_w \Delta r_f h_w G_u^{adv} & = & |
173 |
\delta_i \overline{ U }^i \overline{ u }^i |
174 |
+ \delta_j \overline{ V }^i \overline{ u }^j |
175 |
+ \delta_k \overline{ W }^i \overline{ u }^k \\ |
176 |
{\cal A}_s \Delta r_f h_s G_v^{adv} & = & |
177 |
\delta_i \overline{ U }^j \overline{ v }^i |
178 |
+ \delta_j \overline{ V }^j \overline{ v }^j |
179 |
+ \delta_k \overline{ W }^j \overline{ v }^k \\ |
180 |
{\cal A}_c \Delta r_c G_w^{adv} & = & |
181 |
\delta_i \overline{ U }^k \overline{ w }^i |
182 |
+ \delta_j \overline{ V }^k \overline{ w }^j |
183 |
+ \delta_k \overline{ W }^k \overline{ w }^k \\ |
184 |
\end{eqnarray} |
185 |
and because of the flux form does not contribute to the global budget |
186 |
of linear momentum. The quantities $U$, $V$ and $W$ are volume fluxes |
187 |
defined: |
188 |
\marginpar{$U$: {\bf uTrans} } |
189 |
\marginpar{$V$: {\bf vTrans} } |
190 |
\marginpar{$W$: {\bf rTrans} } |
191 |
\begin{eqnarray} |
192 |
U & = & \Delta y_g \Delta r_f h_w u \\ |
193 |
V & = & \Delta x_g \Delta r_f h_s v \\ |
194 |
W & = & {\cal A}_c w |
195 |
\end{eqnarray} |
196 |
The advection of momentum takes the same form as the advection of |
197 |
tracers but by a translated advective flow. Consequently, the |
198 |
conservation of second moments, derived for tracers later, applies to |
199 |
$u^2$ and $v^2$ and $w^2$ so that advection of momentum correctly |
200 |
conserves kinetic energy. |
201 |
|
202 |
\fbox{ \begin{minipage}{4.25in} |
203 |
{\em S/R MOM\_U\_ADV\_UU} ({\em mom\_u\_adv\_uu.F}) |
204 |
|
205 |
{\em S/R MOM\_U\_ADV\_VU} ({\em mom\_u\_adv\_vu.F}) |
206 |
|
207 |
{\em S/R MOM\_U\_ADV\_WU} ({\em mom\_u\_adv\_wu.F}) |
208 |
|
209 |
{\em S/R MOM\_U\_ADV\_UV} ({\em mom\_u\_adv\_uv.F}) |
210 |
|
211 |
{\em S/R MOM\_U\_ADV\_VV} ({\em mom\_u\_adv\_vv.F}) |
212 |
|
213 |
{\em S/R MOM\_U\_ADV\_WV} ({\em mom\_u\_adv\_wv.F}) |
214 |
|
215 |
$uu$, $uv$, $vu$, $vv$: {\bf aF} (local to {\em calc\_mom\_rhs.F}) |
216 |
\end{minipage} } |
217 |
|
218 |
|
219 |
|
220 |
\subsubsection{Coriolis terms} |
221 |
|
222 |
The ``pure C grid'' Coriolis terms (i.e. in absence of C-D scheme) are |
223 |
discretized: |
224 |
\begin{eqnarray} |
225 |
{\cal A}_w \Delta r_f h_w G_u^{Cor} & = & |
226 |
\overline{ f {\cal A}_c \Delta r_f h_c \overline{ v }^j }^i |
227 |
- \epsilon_{nh} \overline{ f' {\cal A}_c \Delta r_f h_c \overline{ w }^k }^i \\ |
228 |
{\cal A}_s \Delta r_f h_s G_v^{Cor} & = & |
229 |
- \overline{ f {\cal A}_c \Delta r_f h_c \overline{ u }^i }^j \\ |
230 |
{\cal A}_c \Delta r_c G_w^{Cor} & = & |
231 |
\epsilon_{nh} \overline{ f' {\cal A}_c \Delta r_f h_c \overline{ u }^i }^k |
232 |
\end{eqnarray} |
233 |
where the Coriolis parameters $f$ and $f'$ are defined: |
234 |
\begin{eqnarray} |
235 |
f & = & 2 \Omega \sin{\phi} \\ |
236 |
f' & = & 2 \Omega \cos{\phi} |
237 |
\end{eqnarray} |
238 |
when using spherical geometry, otherwise the $\beta$-plane definition is used: |
239 |
\begin{eqnarray} |
240 |
f & = & f_o + \beta y \\ |
241 |
f' & = & 0 |
242 |
\end{eqnarray} |
243 |
|
244 |
This discretization globally conserves kinetic energy. It should be |
245 |
noted that despite the use of this discretization in former |
246 |
publications, all calculations to date have used the following |
247 |
different discretization: |
248 |
\begin{eqnarray} |
249 |
G_u^{Cor} & = & |
250 |
f_u \overline{ v }^{ji} |
251 |
- \epsilon_{nh} f_u' \overline{ w }^{ik} \\ |
252 |
G_v^{Cor} & = & |
253 |
- f_v \overline{ u }^{ij} \\ |
254 |
G_w^{Cor} & = & |
255 |
\epsilon_{nh} f_w' \overline{ u }^{ik} |
256 |
\end{eqnarray} |
257 |
\marginpar{Need to change the default in code to match this} |
258 |
where the subscripts on $f$ and $f'$ indicate evaluation of the |
259 |
Coriolis parameters at the appropriate points in space. The above |
260 |
discretization does {\em not} conserve anything, especially energy. An |
261 |
option to recover this discretization has been retained for backward |
262 |
compatibility testing (set run-time logical {\bf |
263 |
useNonconservingCoriolis} to {\em true} which otherwise defaults to |
264 |
{\em false}). |
265 |
|
266 |
\fbox{ \begin{minipage}{4.25in} |
267 |
{\em S/R MOM\_CDSCHEME} ({\em mom\_cdscheme.F}) |
268 |
|
269 |
{\em S/R MOM\_U\_CORIOLIS} ({\em mom\_u\_coriolis.F}) |
270 |
|
271 |
{\em S/R MOM\_V\_CORIOLIS} ({\em mom\_v\_coriolis.F}) |
272 |
|
273 |
$G_u^{Cor}$, $G_v^{Cor}$: {\bf cF} (local to {\em calc\_mom\_rhs.F}) |
274 |
\end{minipage} } |
275 |
|
276 |
|
277 |
\subsubsection{Curvature metric terms} |
278 |
|
279 |
The most commonly used coordinate system on the sphere is the |
280 |
geographic system $(\lambda,\phi)$. The curvilinear nature of these |
281 |
coordinates on the sphere lead to some ``metric'' terms in the |
282 |
component momentum equations. Under the thin-atmosphere and |
283 |
hydrostatic approximations these terms are discretized: |
284 |
\begin{eqnarray} |
285 |
{\cal A}_w \Delta r_f h_w G_u^{metric} & = & |
286 |
\overline{ \frac{ \overline{u}^i }{a} \tan{\phi} {\cal A}_c \Delta r_f h_c \overline{ v }^j }^i \\ |
287 |
{\cal A}_s \Delta r_f h_s G_v^{metric} & = & |
288 |
- \overline{ \frac{ \overline{u}^i }{a} \tan{\phi} {\cal A}_c \Delta r_f h_c \overline{ u }^i }^j \\ |
289 |
G_w^{metric} & = & 0 |
290 |
\end{eqnarray} |
291 |
where $a$ is the radius of the planet (sphericity is assumed) or the |
292 |
radial distance of the particle (i.e. a function of height). It is |
293 |
easy to see that this discretization satisfies all the properties of |
294 |
the discrete Coriolis terms since the metric factor $\frac{u}{a} |
295 |
\tan{\phi}$ can be viewed as a modification of the vertical Coriolis |
296 |
parameter: $f \rightarrow f+\frac{u}{a} \tan{\phi}$. |
297 |
|
298 |
However, as for the Coriolis terms, a non-energy conserving form has |
299 |
exclusively been used to date: |
300 |
\begin{eqnarray} |
301 |
G_u^{metric} & = & \frac{u \overline{v}^{ij} }{a} \tan{\phi} \\ |
302 |
G_v^{metric} & = & \frac{ \overline{u}^{ij} \overline{u}^{ij}}{a} \tan{\phi} |
303 |
\end{eqnarray} |
304 |
where $\tan{\phi}$ is evaluated at the $u$ and $v$ points |
305 |
respectively. |
306 |
|
307 |
\fbox{ \begin{minipage}{4.25in} |
308 |
{\em S/R MOM\_U\_METRIC\_SPHERE} ({\em mom\_u\_metric\_sphere.F}) |
309 |
|
310 |
{\em S/R MOM\_V\_METRIC\_SPHERE} ({\em mom\_v\_metric\_sphere.F}) |
311 |
|
312 |
$G_u^{metric}$, $G_v^{metric}$: {\bf mT} (local to {\em calc\_mom\_rhs.F}) |
313 |
\end{minipage} } |
314 |
|
315 |
|
316 |
|
317 |
\subsubsection{Non-hydrostatic metric terms} |
318 |
|
319 |
For the non-hydrostatic equations, dropping the thin-atmosphere |
320 |
approximation re-introduces metric terms involving $w$ and are |
321 |
required to conserve anglular momentum: |
322 |
\begin{eqnarray} |
323 |
{\cal A}_w \Delta r_f h_w G_u^{metric} & = & |
324 |
- \overline{ \frac{ \overline{u}^i \overline{w}^k }{a} {\cal A}_c \Delta r_f h_c }^i \\ |
325 |
{\cal A}_s \Delta r_f h_s G_v^{metric} & = & |
326 |
- \overline{ \frac{ \overline{v}^j \overline{w}^k }{a} {\cal A}_c \Delta r_f h_c}^j \\ |
327 |
{\cal A}_c \Delta r_c G_w^{metric} & = & |
328 |
\overline{ \frac{ {\overline{u}^i}^2 + {\overline{v}^j}^2}{a} {\cal A}_c \Delta r_f h_c }^k |
329 |
\end{eqnarray} |
330 |
|
331 |
Because we are always consistent, even if consistently wrong, we have, |
332 |
in the past, used a different discretization in the model which is: |
333 |
\begin{eqnarray} |
334 |
G_u^{metric} & = & |
335 |
- \frac{u}{a} \overline{w}^{ik} \\ |
336 |
G_v^{metric} & = & |
337 |
- \frac{v}{a} \overline{w}^{jk} \\ |
338 |
G_w^{metric} & = & |
339 |
\frac{1}{a} ( {\overline{u}^{ik}}^2 + {\overline{v}^{jk}}^2 ) |
340 |
\end{eqnarray} |
341 |
|
342 |
\fbox{ \begin{minipage}{4.25in} |
343 |
{\em S/R MOM\_U\_METRIC\_NH} ({\em mom\_u\_metric\_nh.F}) |
344 |
|
345 |
{\em S/R MOM\_V\_METRIC\_NH} ({\em mom\_v\_metric\_nh.F}) |
346 |
|
347 |
$G_u^{metric}$, $G_v^{metric}$: {\bf mT} (local to {\em calc\_mom\_rhs.F}) |
348 |
\end{minipage} } |
349 |
|
350 |
|
351 |
\subsubsection{Lateral dissipation} |
352 |
|
353 |
Historically, we have represented the SGS Reynolds stresses as simply |
354 |
down gradient momentum fluxes, ignoring constraints on the stress |
355 |
tensor such as symmetry. |
356 |
\begin{eqnarray} |
357 |
{\cal A}_w \Delta r_f h_w G_u^{h-diss} & = & |
358 |
\delta_i \Delta y_f \Delta r_f h_c \tau_{11} |
359 |
+ \delta_j \Delta x_v \Delta r_f h_\zeta \tau_{12} \\ |
360 |
{\cal A}_s \Delta r_f h_s G_v^{h-diss} & = & |
361 |
\delta_i \Delta y_u \Delta r_f h_\zeta \tau_{21} |
362 |
+ \delta_j \Delta x_f \Delta r_f h_c \tau_{22} |
363 |
\end{eqnarray} |
364 |
\marginpar{Check signs of stress definitions} |
365 |
|
366 |
The lateral viscous stresses are discretized: |
367 |
\begin{eqnarray} |
368 |
\tau_{11} & = & A_h c_{11\Delta}(\phi) \frac{1}{\Delta x_f} \delta_i u |
369 |
-A_4 c_{11\Delta^2}(\phi) \frac{1}{\Delta x_f} \delta_i \nabla^2 u \\ |
370 |
\tau_{12} & = & A_h c_{12\Delta}(\phi) \frac{1}{\Delta y_u} \delta_j u |
371 |
-A_4 c_{12\Delta^2}(\phi)\frac{1}{\Delta y_u} \delta_j \nabla^2 u \\ |
372 |
\tau_{21} & = & A_h c_{21\Delta}(\phi) \frac{1}{\Delta x_v} \delta_i v |
373 |
-A_4 c_{21\Delta^2}(\phi) \frac{1}{\Delta x_v} \delta_i \nabla^2 v \\ |
374 |
\tau_{22} & = & A_h c_{22\Delta}(\phi) \frac{1}{\Delta y_f} \delta_j v |
375 |
-A_4 c_{22\Delta^2}(\phi) \frac{1}{\Delta y_f} \delta_j \nabla^2 v |
376 |
\end{eqnarray} |
377 |
where the non-dimensional factors $c_{lm\Delta^n}(\phi), \{l,m,n\} \in |
378 |
\{1,2\}$ define the ``cosine'' scaling with latitude which can be |
379 |
applied in various ad-hoc ways. For instance, $c_{11\Delta} = |
380 |
c_{21\Delta} = (\cos{\phi})^{3/2}$, $c_{12\Delta}=c_{22\Delta}=0$ would |
381 |
represent the an-isotropic cosine scaling typically used on the |
382 |
``lat-lon'' grid for Laplacian viscosity. |
383 |
\marginpar{Need to tidy up method for controlling this in code} |
384 |
|
385 |
It should be noted that dispite the ad-hoc nature of the scaling, some |
386 |
scaling must be done since on a lat-lon grid the converging meridians |
387 |
make it very unlikely that a stable viscosity parameter exists across |
388 |
the entire model domain. |
389 |
|
390 |
The Laplacian viscosity coefficient, $A_h$ ({\bf viscAh}), has units |
391 |
of $m^2 s^{-1}$. The bi-harmonic viscosity coefficient, $A_4$ ({\bf |
392 |
viscA4}), has units of $m^4 s^{-1}$. |
393 |
|
394 |
\fbox{ \begin{minipage}{4.25in} |
395 |
{\em S/R MOM\_U\_XVISCFLUX} ({\em mom\_u\_xviscflux.F}) |
396 |
|
397 |
{\em S/R MOM\_U\_YVISCFLUX} ({\em mom\_u\_yviscflux.F}) |
398 |
|
399 |
{\em S/R MOM\_V\_XVISCFLUX} ({\em mom\_v\_xviscflux.F}) |
400 |
|
401 |
{\em S/R MOM\_V\_YVISCFLUX} ({\em mom\_v\_yviscflux.F}) |
402 |
|
403 |
$\tau_{11}$, $\tau_{12}$, $\tau_{22}$, $\tau_{22}$: {\bf vF}, {\bf |
404 |
v4F} (local to {\em calc\_mom\_rhs.F}) |
405 |
\end{minipage} } |
406 |
|
407 |
Two types of lateral boundary condition exist for the lateral viscous |
408 |
terms, no-slip and free-slip. |
409 |
|
410 |
The free-slip condition is most convenient to code since it is |
411 |
equivalent to zero-stress on boundaries. Simple masking of the stress |
412 |
components sets them to zero. The fractional open stress is properly |
413 |
handled using the lopped cells. |
414 |
|
415 |
The no-slip condition defines the normal gradient of a tangential flow |
416 |
such that the flow is zero on the boundary. Rather than modify the |
417 |
stresses by using complicated functions of the masks and ``ghost'' |
418 |
points (see \cite{Adcroft+Marshall98}) we add the boundary stresses as |
419 |
an additional source term in cells next to solid boundaries. This has |
420 |
the advantage of being able to cope with ``thin walls'' and also makes |
421 |
the interior stress calculation (code) independent of the boundary |
422 |
conditions. The ``body'' force takes the form: |
423 |
\begin{eqnarray} |
424 |
G_u^{side-drag} & = & |
425 |
\frac{4}{\Delta z_f} \overline{ (1-h_\zeta) \frac{\Delta x_v}{\Delta y_u} }^j |
426 |
\left( A_h c_{12\Delta}(\phi) u - A_4 c_{12\Delta^2}(\phi) \nabla^2 u \right) |
427 |
\\ |
428 |
G_v^{side-drag} & = & |
429 |
\frac{4}{\Delta z_f} \overline{ (1-h_\zeta) \frac{\Delta y_u}{\Delta x_v} }^i |
430 |
\left( A_h c_{21\Delta}(\phi) v - A_4 c_{21\Delta^2}(\phi) \nabla^2 v \right) |
431 |
\end{eqnarray} |
432 |
|
433 |
In fact, the above discretization is not quite complete because it |
434 |
assumes that the bathymetry at velocity points is deeper than at |
435 |
neighbouring vorticity points, e.g. $1-h_w < 1-h_\zeta$ |
436 |
|
437 |
\fbox{ \begin{minipage}{4.25in} |
438 |
{\em S/R MOM\_U\_SIDEDRAG} ({\em mom\_u\_sidedrag.F}) |
439 |
|
440 |
{\em S/R MOM\_V\_SIDEDRAG} ({\em mom\_v\_sidedrag.F}) |
441 |
|
442 |
$G_u^{side-drag}$, $G_v^{side-drag}$: {\bf vF} (local to {\em calc\_mom\_rhs.F}) |
443 |
\end{minipage} } |
444 |
|
445 |
|
446 |
\subsubsection{Vertical dissipation} |
447 |
|
448 |
Vertical viscosity terms are discretized with only partial adherence |
449 |
to the variable grid lengths introduced by the finite volume |
450 |
formulation. This reduces the formal accuracy of these terms to just |
451 |
first order but only next to boundaries; exactly where other terms |
452 |
appear such as linar and quadratic bottom drag. |
453 |
\begin{eqnarray} |
454 |
G_u^{v-diss} & = & |
455 |
\frac{1}{\Delta r_f h_w} \delta_k \tau_{13} \\ |
456 |
G_v^{v-diss} & = & |
457 |
\frac{1}{\Delta r_f h_s} \delta_k \tau_{23} \\ |
458 |
G_w^{v-diss} & = & \epsilon_{nh} |
459 |
\frac{1}{\Delta r_f h_d} \delta_k \tau_{33} |
460 |
\end{eqnarray} |
461 |
represents the general discrete form of the vertical dissipation terms. |
462 |
|
463 |
In the interior the vertical stresses are discretized: |
464 |
\begin{eqnarray} |
465 |
\tau_{13} & = & A_v \frac{1}{\Delta r_c} \delta_k u \\ |
466 |
\tau_{23} & = & A_v \frac{1}{\Delta r_c} \delta_k v \\ |
467 |
\tau_{33} & = & A_v \frac{1}{\Delta r_f} \delta_k w |
468 |
\end{eqnarray} |
469 |
It should be noted that in the non-hydrostatic form, the stress tensor |
470 |
is even less consistent than for the hydrostatic (see Wazjowicz |
471 |
\cite{Waojz}). It is well known how to do this properly (see Griffies |
472 |
\cite{Griffies}) and is on the list of to-do's. |
473 |
|
474 |
\fbox{ \begin{minipage}{4.25in} |
475 |
{\em S/R MOM\_U\_RVISCLFUX} ({\em mom\_u\_rviscflux.F}) |
476 |
|
477 |
{\em S/R MOM\_V\_RVISCLFUX} ({\em mom\_v\_rviscflux.F}) |
478 |
|
479 |
$\tau_{13}$: {\bf urf} (local to {\em calc\_mom\_rhs.F}) |
480 |
|
481 |
$\tau_{23}$: {\bf vrf} (local to {\em calc\_mom\_rhs.F}) |
482 |
\end{minipage} } |
483 |
|
484 |
|
485 |
As for the lateral viscous terms, the free-slip condition is |
486 |
equivalent to simply setting the stress to zero on boundaries. The |
487 |
no-slip condition is implemented as an additional term acting on top |
488 |
of the interior and free-slip stresses. Bottom drag represents |
489 |
additional friction, in addition to that imposed by the no-slip |
490 |
condition at the bottom. The drag is cast as a stress expressed as a |
491 |
linear or quadratic function of the mean flow in the layer above the |
492 |
topography: |
493 |
\begin{eqnarray} |
494 |
\tau_{13}^{bottom-drag} & = & |
495 |
\left( |
496 |
2 A_v \frac{1}{\Delta r_c} |
497 |
+ r_b |
498 |
+ C_d \sqrt{ \overline{2 KE}^i } |
499 |
\right) u \\ |
500 |
\tau_{23}^{bottom-drag} & = & |
501 |
\left( |
502 |
2 A_v \frac{1}{\Delta r_c} |
503 |
+ r_b |
504 |
+ C_d \sqrt{ \overline{2 KE}^j } |
505 |
\right) v |
506 |
\end{eqnarray} |
507 |
where these terms are only evaluated immediately above topography. |
508 |
$r_b$ ({\bf bottomDragLinear}) has units of $m s^{-1}$ and a typical value |
509 |
of the order 0.0002 $m s^{-1}$. $C_d$ ({\bf bottomDragQuadratic}) is |
510 |
dimensionless with typical values in the range 0.001--0.003. |
511 |
|
512 |
\fbox{ \begin{minipage}{4.25in} |
513 |
{\em S/R MOM\_U\_BOTTOMDRAG} ({\em mom\_u\_bottomdrag.F}) |
514 |
|
515 |
{\em S/R MOM\_V\_BOTTOMDRAG} ({\em mom\_v\_bottomdrag.F}) |
516 |
|
517 |
$\tau_{13}^{bottom-drag}$, $\tau_{23}^{bottom-drag}$: {\bf vf} (local to {\em calc\_mom\_rhs.F}) |
518 |
\end{minipage} } |
519 |
|
520 |
|
521 |
|
522 |
|
523 |
|
524 |
\subsection{Tracer equations} |
525 |
|
526 |
The tracer equations are discretized consistantly with the continuity |
527 |
equation to facilitate conservation properties analogous to the |
528 |
continuum: |
529 |
\begin{equation} |
530 |
{\cal A}_c \Delta r_f h_c \partial_\theta |
531 |
+ \delta_i U \overline{ \theta }^i |
532 |
+ \delta_j V \overline{ \theta }^j |
533 |
+ \delta_k W \overline{ \theta }^k |
534 |
= {\cal A}_c \Delta r_f h_c {\cal S}_\theta + \theta {\cal A}_c \delta_k (P-E)_{r=0} |
535 |
\end{equation} |
536 |
The quantities $U$, $V$ and $W$ are volume fluxes defined: |
537 |
\marginpar{$U$: {\bf uTrans} } |
538 |
\marginpar{$V$: {\bf vTrans} } |
539 |
\marginpar{$W$: {\bf rTrans} } |
540 |
\begin{eqnarray} |
541 |
U & = & \Delta y_g \Delta r_f h_w u \\ |
542 |
V & = & \Delta x_g \Delta r_f h_s v \\ |
543 |
W & = & {\cal A}_c w |
544 |
\end{eqnarray} |
545 |
${\cal S}$ represents the ``parameterized'' SGS processes and |
546 |
physics associated with the tracer. For instance, potential |
547 |
temperature equation in the ocean has is forced by surface and |
548 |
partially penetrating heat fluxes: |
549 |
\begin{equation} |
550 |
{\cal A}_c \Delta r_f h_c {\cal S}_\theta = \frac{1}{c_p \rho_o} \delta_k {\cal A}_c {\cal Q} |
551 |
\end{equation} |
552 |
while the salt equation has no real sources, ${\cal S}=0$, which |
553 |
leaves just the $P-E$ term. |
554 |
|
555 |
The continuity equation can be recovered by setting ${\cal Q}=0$ and |
556 |
$\theta=1$. The term $\theta (P-E)_{r=0}$ is required to retain local |
557 |
conservation of $\theta$. Global conservation is not possible using |
558 |
the flux-form (as here) and a linearized free-surface |
559 |
(\cite{Griffies00,Campin02}). |
560 |
|
561 |
|
562 |
|
563 |
|
564 |
\subsection{Derivation of discrete energy conservation} |
565 |
|
566 |
These discrete equations conserve kinetic plus potential energy using the |
567 |
following definitions: |
568 |
\begin{equation} |
569 |
KE = \frac{1}{2} \left( \overline{ u^2 }^i + \overline{ v^2 }^j + |
570 |
\epsilon_{nh} \overline{ w^2 }^k \right) |
571 |
\end{equation} |
572 |
|
573 |
|
574 |
\subsection{Vector invariant momentum equations} |
575 |
|
576 |
The finite volume method lends itself to describing the continuity and |
577 |
tracer equations in curvilinear coordinate systems but the appearance |
578 |
of new metric terms in the flux-form momentum equations makes |
579 |
generalizing them far from elegant. The vector invariant form of the |
580 |
momentum equations are exactly that; invariant under coordinate |
581 |
transformations. |
582 |
|
583 |
The non-hydrostatic vector invariant equations read: |
584 |
\begin{equation} |
585 |
\partial_t \vec{v} + ( 2\vec{\Omega} + \vec{\zeta}) \wedge \vec{v} |
586 |
- b \hat{r} |
587 |
+ \vec{\nabla} B = \vec{\nabla} \cdot \vec{\bf \tau} |
588 |
\end{equation} |
589 |
which describe motions in any orthogonal curvilinear coordinate |
590 |
system. Here, $B$ is the Bernoulli function and $\vec{\zeta}=\nabla |
591 |
\wedge \vec{v}$ is the vorticity vector. We can take advantage of the |
592 |
elegance of these equations when discretizing them and use the |
593 |
discrete definitions of the grad, curl and divergence operators to |
594 |
satisfy constraints. We can also consider the analogy to forming |
595 |
derived equations, such as the vorticity equation, and examine how the |
596 |
discretization can be adjusted to give suitable vorticity advection |
597 |
among other things. |
598 |
|
599 |
The underlying algorithm is the same as for the flux form |
600 |
equations. All that has changed is the contents of the ``G's''. For |
601 |
the time-being, only the hydrostatic terms have been coded but we will |
602 |
indicate the points where non-hydrostatic contributions will enter: |
603 |
\begin{eqnarray} |
604 |
G_u & = & G_u^{fv} + G_u^{\zeta_3 v} + G_u^{\zeta_2 w} + G_u^{\partial_x B} |
605 |
+ G_u^{\partial_z \tau^x} + G_u^{h-dissip} + G_u^{v-dissip} \\ |
606 |
G_v & = & G_v^{fu} + G_v^{\zeta_3 u} + G_v^{\zeta_1 w} + G_v^{\partial_y B} |
607 |
+ G_v^{\partial_z \tau^y} + G_v^{h-dissip} + G_v^{v-dissip} \\ |
608 |
G_w & = & G_w^{fu} + G_w^{\zeta_1 v} + G_w^{\zeta_2 u} + G_w^{\partial_z B} |
609 |
+ G_w^{h-dissip} + G_w^{v-dissip} |
610 |
\end{eqnarray} |
611 |
|
612 |
\fbox{ \begin{minipage}{4.25in} |
613 |
{\em S/R CALC\_MOM\_RHS} ({\em pkg/mom\_vecinv/calc\_mom\_rhs.F}) |
614 |
|
615 |
$G_u$: {\bf Gu} ({\em DYNVARS.h}) |
616 |
|
617 |
$G_v$: {\bf Gv} ({\em DYNVARS.h}) |
618 |
|
619 |
$G_w$: {\bf Gw} ({\em DYNVARS.h}) |
620 |
\end{minipage} } |
621 |
|
622 |
\subsubsection{Relative vorticity} |
623 |
|
624 |
The vertical component of relative vorticity is explicitly calculated |
625 |
and use in the discretization. The particular form is crucial for |
626 |
numerical stablility; alternative definitions break the conservation |
627 |
properties of the discrete equations. |
628 |
|
629 |
Relative vorticity is defined: |
630 |
\begin{equation} |
631 |
\zeta_3 = \frac{\Gamma}{A_\zeta} |
632 |
= \frac{1}{{\cal A}_\zeta} ( \delta_i \Delta y_c v - \delta_j \Delta x_c u ) |
633 |
\end{equation} |
634 |
where ${\cal A}_\zeta$ is the area of the vorticity cell presented in |
635 |
the vertical and $\Gamma$ is the circulation about that cell. |
636 |
|
637 |
\fbox{ \begin{minipage}{4.25in} |
638 |
{\em S/R MOM\_VI\_CALC\_RELVORT3} ({\em mom\_vi\_calc\_relvort3.F}) |
639 |
|
640 |
$\zeta_3$: {\bf vort3} (local to {\em calc\_mom\_rhs.F}) |
641 |
\end{minipage} } |
642 |
|
643 |
|
644 |
\subsubsection{Kinetic energy} |
645 |
|
646 |
The kinetic energy, denoted $KE$, is defined: |
647 |
\begin{equation} |
648 |
KE = \frac{1}{2} ( \overline{ u^2 }^i + \overline{ v^2 }^j |
649 |
+ \epsilon_{nh} \overline{ w^2 }^k ) |
650 |
\end{equation} |
651 |
|
652 |
\fbox{ \begin{minipage}{4.25in} |
653 |
{\em S/R MOM\_VI\_CALC\_KE} ({\em mom\_vi\_calc\_ke.F}) |
654 |
|
655 |
$KE$: {\bf KE} (local to {\em calc\_mom\_rhs.F}) |
656 |
\end{minipage} } |
657 |
|
658 |
|
659 |
\subsubsection{Coriolis terms} |
660 |
|
661 |
The potential enstrophy conserving form of the linear Coriolis terms |
662 |
are written: |
663 |
\begin{eqnarray} |
664 |
G_u^{fv} & = & |
665 |
\frac{1}{\Delta x_c} |
666 |
\overline{ \frac{f}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\ |
667 |
G_v^{fu} & = & - |
668 |
\frac{1}{\Delta y_c} |
669 |
\overline{ \frac{f}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j |
670 |
\end{eqnarray} |
671 |
Here, the Coriolis parameter $f$ is defined at vorticity (corner) |
672 |
points. |
673 |
\marginpar{$f$: {\bf fCoriG}} |
674 |
\marginpar{$h_\zeta$: {\bf hFacZ}} |
675 |
|
676 |
The potential enstrophy conserving form of the non-linear Coriolis |
677 |
terms are written: |
678 |
\begin{eqnarray} |
679 |
G_u^{\zeta_3 v} & = & |
680 |
\frac{1}{\Delta x_c} |
681 |
\overline{ \frac{\zeta_3}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\ |
682 |
G_v^{\zeta_3 u} & = & - |
683 |
\frac{1}{\Delta y_c} |
684 |
\overline{ \frac{\zeta_3}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j |
685 |
\end{eqnarray} |
686 |
\marginpar{$\zeta_3$: {\bf vort3}} |
687 |
|
688 |
The Coriolis terms can also be evaluated together and expressed in |
689 |
terms of absolute vorticity $f+\zeta_3$. The potential enstrophy |
690 |
conserving form using the absolute vorticity is written: |
691 |
\begin{eqnarray} |
692 |
G_u^{fv} + G_u^{\zeta_3 v} & = & |
693 |
\frac{1}{\Delta x_c} |
694 |
\overline{ \frac{f + \zeta_3}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\ |
695 |
G_v^{fu} + G_v^{\zeta_3 u} & = & - |
696 |
\frac{1}{\Delta y_c} |
697 |
\overline{ \frac{f + \zeta_3}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j |
698 |
\end{eqnarray} |
699 |
|
700 |
\marginpar{Run-time control needs to be added for these options} The |
701 |
disctinction between using absolute vorticity or relative vorticity is |
702 |
useful when constructing higher order advection schemes; monotone |
703 |
advection of relative vorticity behaves differently to monotone |
704 |
advection of absolute vorticity. Currently the choice of |
705 |
relative/absolute vorticity, centered/upwind/high order advection is |
706 |
available only through commented subroutine calls. |
707 |
|
708 |
\fbox{ \begin{minipage}{4.25in} |
709 |
{\em S/R MOM\_VI\_CORIOLIS} ({\em mom\_vi\_coriolis.F}) |
710 |
|
711 |
{\em S/R MOM\_VI\_U\_CORIOLIS} ({\em mom\_vi\_u\_coriolis.F}) |
712 |
|
713 |
{\em S/R MOM\_VI\_V\_CORIOLIS} ({\em mom\_vi\_v\_coriolis.F}) |
714 |
|
715 |
$G_u^{fv}$, $G_u^{\zeta_3 v}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F}) |
716 |
|
717 |
$G_v^{fu}$, $G_v^{\zeta_3 u}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F}) |
718 |
\end{minipage} } |
719 |
|
720 |
|
721 |
\subsubsection{Shear terms} |
722 |
|
723 |
The shear terms ($\zeta_2w$ and $\zeta_1w$) are are discretized to |
724 |
guarantee that no spurious generation of kinetic energy is possible; |
725 |
the horizontal gradient of Bernoulli function has to be consistent |
726 |
with the vertical advection of shear: |
727 |
\marginpar{N-H terms have not been tried!} |
728 |
\begin{eqnarray} |
729 |
G_u^{\zeta_2 w} & = & |
730 |
\frac{1}{ {\cal A}_w \Delta r_f h_w } \overline{ |
731 |
\overline{ {\cal A}_c w }^i ( \delta_k u - \epsilon_{nh} \delta_j w ) |
732 |
}^k \\ |
733 |
G_v^{\zeta_1 w} & = & |
734 |
\frac{1}{ {\cal A}_s \Delta r_f h_s } \overline{ |
735 |
\overline{ {\cal A}_c w }^i ( \delta_k u - \epsilon_{nh} \delta_j w ) |
736 |
}^k |
737 |
\end{eqnarray} |
738 |
|
739 |
\fbox{ \begin{minipage}{4.25in} |
740 |
{\em S/R MOM\_VI\_U\_VERTSHEAR} ({\em mom\_vi\_u\_vertshear.F}) |
741 |
|
742 |
{\em S/R MOM\_VI\_V\_VERTSHEAR} ({\em mom\_vi\_v\_vertshear.F}) |
743 |
|
744 |
$G_u^{\zeta_2 w}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F}) |
745 |
|
746 |
$G_v^{\zeta_1 w}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F}) |
747 |
\end{minipage} } |
748 |
|
749 |
|
750 |
|
751 |
\subsubsection{Gradient of Bernoulli function} |
752 |
|
753 |
\begin{eqnarray} |
754 |
G_u^{\partial_x B} & = & |
755 |
\frac{1}{\Delta x_c} \delta_i ( \phi' + KE ) \\ |
756 |
G_v^{\partial_y B} & = & |
757 |
\frac{1}{\Delta x_y} \delta_j ( \phi' + KE ) |
758 |
%G_w^{\partial_z B} & = & |
759 |
%\frac{1}{\Delta r_c} h_c \delta_k ( \phi' + KE ) |
760 |
\end{eqnarray} |
761 |
|
762 |
\fbox{ \begin{minipage}{4.25in} |
763 |
{\em S/R MOM\_VI\_U\_GRAD\_KE} ({\em mom\_vi\_u\_grad\_ke.F}) |
764 |
|
765 |
{\em S/R MOM\_VI\_V\_GRAD\_KE} ({\em mom\_vi\_v\_grad\_ke.F}) |
766 |
|
767 |
$G_u^{\partial_x KE}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F}) |
768 |
|
769 |
$G_v^{\partial_y KE}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F}) |
770 |
\end{minipage} } |
771 |
|
772 |
|
773 |
|
774 |
\subsubsection{Horizontal dissipation} |
775 |
|
776 |
The horizontal divergence, a complimentary quantity to relative |
777 |
vorticity, is used in parameterizing the Reynolds stresses and is |
778 |
discretized: |
779 |
\begin{equation} |
780 |
D = \frac{1}{{\cal A}_c h_c} ( |
781 |
\delta_i \Delta y_g h_w u |
782 |
+ \delta_j \Delta x_g h_s v ) |
783 |
\end{equation} |
784 |
|
785 |
\fbox{ \begin{minipage}{4.25in} |
786 |
{\em S/R MOM\_VI\_CALC\_HDIV} ({\em mom\_vi\_calc\_hdiv.F}) |
787 |
|
788 |
$D$: {\bf hDiv} (local to {\em calc\_mom\_rhs.F}) |
789 |
\end{minipage} } |
790 |
|
791 |
|
792 |
\subsubsection{Horizontal dissipation} |
793 |
|
794 |
The following discretization of horizontal dissipation conserves |
795 |
potential vorticity (thickness weighted relative vorticity) and |
796 |
divergence and dissipates energy, enstrophy and divergence squared: |
797 |
\begin{eqnarray} |
798 |
G_u^{h-dissip} & = & |
799 |
\frac{1}{\Delta x_c} \delta_i ( A_D D - A_{D4} D^*) |
800 |
- \frac{1}{\Delta y_u h_w} \delta_j h_\zeta ( A_\zeta \zeta - A_{\zeta4} \zeta^* ) |
801 |
\\ |
802 |
G_v^{h-dissip} & = & |
803 |
\frac{1}{\Delta x_v h_s} \delta_i h_\zeta ( A_\zeta \zeta - A_\zeta \zeta^* ) |
804 |
+ \frac{1}{\Delta y_c} \delta_j ( A_D D - A_{D4} D^* ) |
805 |
\end{eqnarray} |
806 |
where |
807 |
\begin{eqnarray} |
808 |
D^* & = & \frac{1}{{\cal A}_c h_c} ( |
809 |
\delta_i \Delta y_g h_w \nabla^2 u |
810 |
+ \delta_j \Delta x_g h_s \nabla^2 v ) \\ |
811 |
\zeta^* & = & \frac{1}{{\cal A}_\zeta} ( |
812 |
\delta_i \Delta y_c \nabla^2 v |
813 |
- \delta_j \Delta x_c \nabla^2 u ) |
814 |
\end{eqnarray} |
815 |
|
816 |
\fbox{ \begin{minipage}{4.25in} |
817 |
{\em S/R MOM\_VI\_HDISSIP} ({\em mom\_vi\_hdissip.F}) |
818 |
|
819 |
$G_u^{h-dissip}$: {\bf uDiss} (local to {\em calc\_mom\_rhs.F}) |
820 |
|
821 |
$G_v^{h-dissip}$: {\bf vDiss} (local to {\em calc\_mom\_rhs.F}) |
822 |
\end{minipage} } |
823 |
|
824 |
|
825 |
\subsubsection{Vertical dissipation} |
826 |
|
827 |
Currently, this is exactly the same code as the flux form equations. |
828 |
\begin{eqnarray} |
829 |
G_u^{v-diss} & = & |
830 |
\frac{1}{\Delta r_f h_w} \delta_k \tau_{13} \\ |
831 |
G_v^{v-diss} & = & |
832 |
\frac{1}{\Delta r_f h_s} \delta_k \tau_{23} |
833 |
\end{eqnarray} |
834 |
represents the general discrete form of the vertical dissipation terms. |
835 |
|
836 |
In the interior the vertical stresses are discretized: |
837 |
\begin{eqnarray} |
838 |
\tau_{13} & = & A_v \frac{1}{\Delta r_c} \delta_k u \\ |
839 |
\tau_{23} & = & A_v \frac{1}{\Delta r_c} \delta_k v |
840 |
\end{eqnarray} |
841 |
|
842 |
\fbox{ \begin{minipage}{4.25in} |
843 |
{\em S/R MOM\_U\_RVISCLFUX} ({\em mom\_u\_rviscflux.F}) |
844 |
|
845 |
{\em S/R MOM\_V\_RVISCLFUX} ({\em mom\_v\_rviscflux.F}) |
846 |
|
847 |
$\tau_{13}$: {\bf urf} (local to {\em calc\_mom\_rhs.F}) |
848 |
|
849 |
$\tau_{23}$: {\bf vrf} (local to {\em calc\_mom\_rhs.F}) |
850 |
\end{minipage} } |