/[MITgcm]/manual/s_algorithm/text/spatial-discrete.tex
ViewVC logotype

Diff of /manual/s_algorithm/text/spatial-discrete.tex

Parent Directory Parent Directory | Revision Log Revision Log | View Revision Graph Revision Graph | View Patch Patch

revision 1.1 by adcroft, Wed Aug 8 16:15:21 2001 UTC revision 1.7 by adcroft, Wed Sep 26 21:59:33 2001 UTC
# Line 3  Line 3 
3    
4  \section{Spatial discretization of the dynamical equations}  \section{Spatial discretization of the dynamical equations}
5    
6    Spatial discretization is carried out using the finite volume
7    method. This amounts to a grid-point method (namely second-order
8    centered finite difference) in the fluid interior but allows
9    boundaries to intersect a regular grid allowing a more accurate
10    representation of the position of the boundary. We treat the
11    horizontal and veritical directions as seperable and differently.
12    
13    \input{part2/notation}
14    
15    
16    \subsection{The finite volume method: finite volumes versus finite difference}
17    
18    The finite volume method is used to discretize the equations in
19    space. The expression ``finite volume'' actually has two meanings; one
20    is the method of cut or instecting boundaries (shaved or lopped cells
21    in our terminology) and the other is non-linear interpolation methods
22    that can deal with non-smooth solutions such as shocks (i.e. flux
23    limiters for advection). Both make use of the integral form of the
24    conservation laws to which the {\it weak solution} is a solution on
25    each finite volume of (sub-domain). The weak solution can be
26    constructed outof piece-wise constant elements or be
27    differentiable. The differentiable equations can not be satisfied by
28    piece-wise constant functions.
29    
30    As an example, the 1-D constant coefficient advection-diffusion
31    equation:
32    \begin{displaymath}
33    \partial_t \theta + \partial_x ( u \theta - \kappa \partial_x \theta ) = 0
34    \end{displaymath}
35    can be discretized by integrating over finite sub-domains, i.e.
36    the lengths $\Delta x_i$:
37    \begin{displaymath}
38    \Delta x \partial_t \theta + \delta_i ( F ) = 0
39    \end{displaymath}
40    is exact if $\theta(x)$ is peice-wise constant over the interval
41    $\Delta x_i$ or more generally if $\theta_i$ is defined as the average
42    over the interval $\Delta x_i$.
43    
44    The flux, $F_{i-1/2}$, must be approximated:
45    \begin{displaymath}
46    F = u \overline{\theta} - \frac{\kappa}{\Delta x_c} \partial_i \theta
47    \end{displaymath}
48    and this is where truncation errors can enter the solution. The
49    method for obtaining $\overline{\theta}$ is unspecified and a wide
50    range of possibilities exist including centered and upwind
51    interpolation, polynomial fits based on the the volume average
52    definitions of quantities and non-linear interpolation such as
53    flux-limiters.
54    
55    Choosing simple centered second-order interpolation and differencing
56    recovers the same ODE's resulting from finite differencing for the
57    interior of a fluid. Differences arise at boundaries where a boundary
58    is not positioned on a regular or smoothly varying grid. This method
59    is used to represent the topography using lopped cell, see
60    \cite{Adcroft98}. Subtle difference also appear in more than one
61    dimension away from boundaries. This happens because the each
62    direction is discretized independantly in the finite difference method
63    while the integrating over finite volume implicitly treats all
64    directions simultaneously. Illustration of this is given in
65    \cite{Adcroft02}.
66    
67  \subsection{C grid staggering of variables}  \subsection{C grid staggering of variables}
68    
69  \begin{figure}  \begin{figure}
70  \centerline{ \resizebox{!}{2in}{ \includegraphics{part2/cgrid3d.eps}} }  \begin{center}
71  \label{fig-cgrid3d}  \resizebox{!}{2in}{ \includegraphics{part2/cgrid3d.eps}}
72    \end{center}
73  \caption{Three dimensional staggering of velocity components. This  \caption{Three dimensional staggering of velocity components. This
74  facilitates the natural discretization of the continuity and tracer  facilitates the natural discretization of the continuity and tracer
75  equations. }  equations. }
76    \label{fig:cgrid3d}
77  \end{figure}  \end{figure}
78    
79    The basic algorithm employed for stepping forward the momentum
80    equations is based on retaining non-divergence of the flow at all
81    times. This is most naturally done if the components of flow are
82    staggered in space in the form of an Arakawa C grid \cite{Arakawa70}.
83    
84    Fig. \ref{fig:cgrid3d} shows the components of flow ($u$,$v$,$w$)
85    staggered in space such that the zonal component falls on the
86    interface between continiuty cells in the zonal direction. Similarly
87    for the meridional and vertical directions.  The continiuty cell is
88    synonymous with tracer cells (they are one and the same).
89    
90    
91    
92    \subsection{Grid initialization and data}
93    
94    Initialization of grid data is controlled by subroutine {\em
95    INI\_GRID} which in calls {\em INI\_VERTICAL\_GRID} to initialize the
96    vertical grid, and then either of {\em INI\_CARTESIAN\_GRID}, {\em
97    INI\_SPHERICAL\_POLAR\_GRID} or {\em INI\_CURV\-ILINEAR\_GRID} to
98    initialize the horizontal grid for cartesian, spherical-polar or
99    curvilinear coordinates respectively.
100    
101    The reciprocals of all grid quantities are pre-calculated and this is
102    done in subroutine {\em INI\_MASKS\_ETC} which is called later by
103    subroutine {\em INITIALIZE\_FIXED}.
104    
105    All grid descriptors are global arrays and stored in common blocks in
106    {\em GRID.h} and a generally declared as {\em \_RS}.
107    
108    \fbox{ \begin{minipage}{4.75in}
109    {\em S/R INI\_GRID} ({\em model/src/ini\_grid.F})
110    
111    {\em S/R INI\_MASKS\_ETC} ({\em model/src/ini\_masks\_etc.F})
112    
113    grid data: ({\em model/inc/GRID.h})
114    \end{minipage} }
115    
116    
117  \subsection{Horizontal grid}  \subsection{Horizontal grid}
118    
119  \begin{figure}  \begin{figure}
120  \centerline{ \begin{tabular}{cc}  \begin{center}
121    \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Ac.eps}}  \begin{tabular}{cc}
122  & \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Az.eps}}    \raisebox{1.5in}{a)}\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Ac.eps}}
123    & \raisebox{1.5in}{b)}\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Az.eps}}
124  \\  \\
125    \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Au.eps}}    \raisebox{1.5in}{c)}\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Au.eps}}
126  & \resizebox{!}{2in}{ \includegraphics{part2/hgrid-Av.eps}}  & \raisebox{1.5in}{d)}\resizebox{!}{2in}{ \includegraphics{part2/hgrid-Av.eps}}
127  \end{tabular} }  \end{tabular}
128  \label{fig-hgrid}  \end{center}
129  \caption{Three dimensional staggering of velocity components. This  \caption{
130  facilitates the natural discretization of the continuity and tracer  Staggering of horizontal grid descriptors (lengths and areas). The
131  equations. }  grid lines indicate the tracer cell boundaries and are the reference
132    grid for all panels. a) The area of a tracer cell, $A_c$, is bordered
133    by the lengths $\Delta x_g$ and $\Delta y_g$. b) The area of a
134    vorticity cell, $A_\zeta$, is bordered by the lengths $\Delta x_c$ and
135    $\Delta y_c$. c) The area of a u cell, $A_c$, is bordered by the
136    lengths $\Delta x_v$ and $\Delta y_f$. d) The area of a v cell, $A_c$,
137    is bordered by the lengths $\Delta x_f$ and $\Delta y_u$.}
138    \label{fig:hgrid}
139  \end{figure}  \end{figure}
140    
141    The model domain is decomposed into tiles and within each tile a
142    quasi-regular grid is used. A tile is the basic unit of domain
143    decomposition for parallelization but may be used whether parallized
144    or not; see section \ref{sect:tiles} for more details. Although the
145    tiles may be patched together in an unstructured manner
146    (i.e. irregular or non-tessilating pattern), the interior of tiles is
147    a structered grid of quadrilateral cells. The horizontal coordinate
148    system is orthogonal curvilinear meaning we can not necessarily treat
149    the two horizontal directions as seperable. Instead, each cell in the
150    horizontal grid is described by the length of it's sides and it's
151    area.
152    
153    The grid information is quite general and describes any of the
154    available coordinates systems, cartesian, spherical-polar or
155    curvilinear. All that is necessary to distinguish between the
156    coordinate systems is to initialize the grid data (discriptors)
157    appropriately.
158    
159    In the following, we refer to the orientation of quantities on the
160    computational grid using geographic terminology such as points of the
161    compass.
162    \marginpar{Caution!}
163    This is purely for convenience but should note be confused
164    with the actual geographic orientation of model quantities.
165    
166    Fig.~\ref{fig:hgrid}a shows the tracer cell (synonymous with the
167    continuity cell). The length of the southern edge, $\Delta x_g$,
168    western edge, $\Delta y_g$ and surface area, $A_c$, presented in the
169    vertical are stored in arrays {\bf DXg}, {\bf DYg} and {\bf rAc}.
170    \marginpar{$A_c$: {\bf rAc}}
171    \marginpar{$\Delta x_g$: {\bf DXg}}
172    \marginpar{$\Delta y_g$: {\bf DYg}}
173    The ``g'' suffix indicates that the lengths are along the defining
174    grid boundaries. The ``c'' suffix associates the quantity with the
175    cell centers. The quantities are staggered in space and the indexing
176    is such that {\bf DXg(i,j)} is positioned to the south of {\bf
177    rAc(i,j)} and {\bf DYg(i,j)} positioned to the west.
178    
179    Fig.~\ref{fig:hgrid}b shows the vorticity cell. The length of the
180    southern edge, $\Delta x_c$, western edge, $\Delta y_c$ and surface
181    area, $A_\zeta$, presented in the vertical are stored in arrays {\bf
182    DXg}, {\bf DYg} and {\bf rAz}.
183    \marginpar{$A_\zeta$: {\bf rAz}}
184    \marginpar{$\Delta x_c$: {\bf DXc}}
185    \marginpar{$\Delta y_c$: {\bf DYc}}
186    The ``z'' suffix indicates that the lengths are measured between the
187    cell centers and the ``$\zeta$'' suffix associates points with the
188    vorticity points. The quantities are staggered in space and the
189    indexing is such that {\bf DXc(i,j)} is positioned to the north of
190    {\bf rAc(i,j)} and {\bf DYc(i,j)} positioned to the east.
191    
192    Fig.~\ref{fig:hgrid}c shows the ``u'' or western (w) cell. The length of
193    the southern edge, $\Delta x_v$, eastern edge, $\Delta y_f$ and
194    surface area, $A_w$, presented in the vertical are stored in arrays
195    {\bf DXv}, {\bf DYf} and {\bf rAw}.
196    \marginpar{$A_w$: {\bf rAw}}
197    \marginpar{$\Delta x_v$: {\bf DXv}}
198    \marginpar{$\Delta y_f$: {\bf DYf}}
199    The ``v'' suffix indicates that the length is measured between the
200    v-points, the ``f'' suffix indicates that the length is measured
201    between the (tracer) cell faces and the ``w'' suffix associates points
202    with the u-points (w stands for west). The quantities are staggered in
203    space and the indexing is such that {\bf DXv(i,j)} is positioned to
204    the south of {\bf rAw(i,j)} and {\bf DYf(i,j)} positioned to the east.
205    
206    Fig.~\ref{fig:hgrid}d shows the ``v'' or southern (s) cell. The length of
207    the northern edge, $\Delta x_f$, western edge, $\Delta y_u$ and
208    surface area, $A_s$, presented in the vertical are stored in arrays
209    {\bf DXf}, {\bf DYu} and {\bf rAs}.
210    \marginpar{$A_s$: {\bf rAs}}
211    \marginpar{$\Delta x_f$: {\bf DXf}}
212    \marginpar{$\Delta y_u$: {\bf DYu}}
213    The ``u'' suffix indicates that the length is measured between the
214    u-points, the ``f'' suffix indicates that the length is measured
215    between the (tracer) cell faces and the ``s'' suffix associates points
216    with the v-points (s stands for south). The quantities are staggered
217    in space and the indexing is such that {\bf DXf(i,j)} is positioned to
218    the north of {\bf rAs(i,j)} and {\bf DYu(i,j)} positioned to the west.
219    
220    \fbox{ \begin{minipage}{4.75in}
221    {\em S/R INI\_CARTESIAN\_GRID} ({\em
222    model/src/ini\_cartesian\_grid.F})
223    
224    {\em S/R INI\_SPHERICAL\_POLAR\_GRID} ({\em
225    model/src/ini\_spherical\_polar\_grid.F})
226    
227    {\em S/R INI\_CURVILINEAR\_GRID} ({\em
228    model/src/ini\_curvilinear\_grid.F})
229    
230    $A_c$, $A_\zeta$, $A_w$, $A_s$: {\bf rAc}, {\bf rAz}, {\bf rAw}, {\bf rAs}
231    ({\em GRID.h})
232    
233    $\Delta x_g$, $\Delta y_g$: {\bf DXg}, {\bf DYg} ({\em GRID.h})
234    
235    $\Delta x_c$, $\Delta y_c$: {\bf DXc}, {\bf DYc} ({\em GRID.h})
236    
237    $\Delta x_f$, $\Delta y_f$: {\bf DXf}, {\bf DYf} ({\em GRID.h})
238    
239    $\Delta x_v$, $\Delta y_u$: {\bf DXv}, {\bf DYu} ({\em GRID.h})
240    
241    \end{minipage} }
242    
243    \subsubsection{Reciprocals of horizontal grid descriptors}
244    
245    %\marginpar{$A_c^{-1}$: {\bf RECIP\_rAc}}
246    %\marginpar{$A_\zeta^{-1}$: {\bf RECIP\_rAz}}
247    %\marginpar{$A_w^{-1}$: {\bf RECIP\_rAw}}
248    %\marginpar{$A_s^{-1}$: {\bf RECIP\_rAs}}
249    Lengths and areas appear in the denominator of expressions as much as
250    in the numerator. For efficiency and portability, we pre-calculate the
251    reciprocal of the horizontal grid quantities so that in-line divisions
252    can be avoided.
253    
254    %\marginpar{$\Delta x_g^{-1}$: {\bf RECIP\_DXg}}
255    %\marginpar{$\Delta y_g^{-1}$: {\bf RECIP\_DYg}}
256    %\marginpar{$\Delta x_c^{-1}$: {\bf RECIP\_DXc}}
257    %\marginpar{$\Delta y_c^{-1}$: {\bf RECIP\_DYc}}
258    %\marginpar{$\Delta x_f^{-1}$: {\bf RECIP\_DXf}}
259    %\marginpar{$\Delta y_f^{-1}$: {\bf RECIP\_DYf}}
260    %\marginpar{$\Delta x_v^{-1}$: {\bf RECIP\_DXv}}
261    %\marginpar{$\Delta y_u^{-1}$: {\bf RECIP\_DYu}}
262    For each grid descriptor (array) there is a reciprocal named using the
263    prefix {\bf RECIP\_}. This doubles the amount of storage in {\em
264    GRID.h} but they are all only 2-D descriptors.
265    
266    \fbox{ \begin{minipage}{4.75in}
267    {\em S/R INI\_MASKS\_ETC} ({\em
268    model/src/ini\_masks\_etc.F})
269    
270    $A_c^{-1}$: {\bf RECIP\_Ac} ({\em GRID.h})
271    
272    $A_\zeta^{-1}$: {\bf RECIP\_Az} ({\em GRID.h})
273    
274    $A_w^{-1}$: {\bf RECIP\_Aw} ({\em GRID.h})
275    
276    $A_s^{-1}$: {\bf RECIP\_As} ({\em GRID.h})
277    
278    $\Delta x_g^{-1}$, $\Delta y_g^{-1}$: {\bf RECIP\_DXg}, {\bf RECIP\_DYg} ({\em GRID.h})
279    
280    $\Delta x_c^{-1}$, $\Delta y_c^{-1}$: {\bf RECIP\_DXc}, {\bf RECIP\_DYc} ({\em GRID.h})
281    
282    $\Delta x_f^{-1}$, $\Delta y_f^{-1}$: {\bf RECIP\_DXf}, {\bf RECIP\_DYf} ({\em GRID.h})
283    
284    $\Delta x_v^{-1}$, $\Delta y_u^{-1}$: {\bf RECIP\_DXv}, {\bf RECIP\_DYu} ({\em GRID.h})
285    
286    \end{minipage} }
287    
288    \subsubsection{Cartesian coordinates}
289    
290    Cartesian coordinates are selected when the logical flag {\bf
291    using\-Cartes\-ianGrid} in namelist {\em PARM04} is set to true. The grid
292    spacing can be set to uniform via scalars {\bf dXspacing} and {\bf
293    dYspacing} in namelist {\em PARM04} or to variable resolution by the
294    vectors {\bf DELX} and {\bf DELY}. Units are normally
295    meters. Non-dimensional coordinates can be used by interpretting the
296    gravitational constant as the Rayleigh number.
297    
298    \subsubsection{Spherical-polar coordinates}
299    
300    Spherical coordinates are selected when the logical flag {\bf
301    using\-Spherical\-PolarGrid} in namelist {\em PARM04} is set to true. The
302    grid spacing can be set to uniform via scalars {\bf dXspacing} and
303    {\bf dYspacing} in namelist {\em PARM04} or to variable resolution by
304    the vectors {\bf DELX} and {\bf DELY}. Units of these namelist
305    variables are alway degrees. The horizontal grid descriptors are
306    calculated from these namelist variables have units of meters.
307    
308    \subsubsection{Curvilinear coordinates}
309    
310    Curvilinear coordinates are selected when the logical flag {\bf
311    using\-Curvil\-inear\-Grid} in namelist {\em PARM04} is set to true. The
312    grid spacing can not be set via the namelist. Instead, the grid
313    descriptors are read from data files, one for each descriptor. As for
314    other grids, the horizontal grid descriptors have units of meters.
315    
316    
317  \subsection{Vertical grid}  \subsection{Vertical grid}
318    
319  \begin{figure}  \begin{figure}
320  \centerline{ \begin{tabular}{cc}  \begin{center}
321    \raisebox{4in}{a)}    \begin{tabular}{cc}
322    \resizebox{!}{4in}{ \includegraphics{part2/vgrid-cellcentered.eps}}    \raisebox{4in}{a)} \resizebox{!}{4in}{
323  &    \includegraphics{part2/vgrid-cellcentered.eps}} & \raisebox{4in}{b)}
324   \raisebox{4in}{b)}    \resizebox{!}{4in}{ \includegraphics{part2/vgrid-accurate.eps}}
325   \resizebox{!}{4in}{ \includegraphics{part2/vgrid-accurate.eps}}  \end{tabular}
326  \end{tabular} }  \end{center}
 \label{fig-vgrid}  
327  \caption{Two versions of the vertical grid. a) The cell centered  \caption{Two versions of the vertical grid. a) The cell centered
328  approach where the interface depths are specified and the tracer  approach where the interface depths are specified and the tracer
329  points centered in between the interfaces. b) The interface centered  points centered in between the interfaces. b) The interface centered
330  approach where tracer levels are specified and the w-interfaces are  approach where tracer levels are specified and the w-interfaces are
331  centered in between.}  centered in between.}
332    \label{fig:vgrid}
333  \end{figure}  \end{figure}
334    
335    As for the horizontal grid, we use the suffixes ``c'' and ``f'' to
336    indicates faces and centers. Fig.~\ref{fig:vgrid}a shows the default
337    vertical grid used by the model.
338    \marginpar{$\Delta r_f$: {\bf DRf}}
339    \marginpar{$\Delta r_c$: {\bf DRc}}
340    $\Delta r_f$ is the difference in $r$
341    (vertical coordinate) between the faces (i.e. $\Delta r_f \equiv -
342    \delta_k r$ where the minus sign appears due to the convention that the
343    surface layer has index $k=1$.).
344    
345    The vertical grid is calculated in subroutine {\em
346    INI\_VERTICAL\_GRID} and specified via the vector {\bf DELR} in
347    namelist {\em PARM04}. The units of ``r'' are either meters or Pascals
348    depending on the isomorphism being used which in turn is dependent
349    only on the choise of equation of state.
350    
351    There are alternative namelist vectors {\bf DELZ} and {\bf DELP} which
352    dictate whether z- or
353    \marginpar{Caution!}
354    p- coordinates are to be used but we intend to
355    phase this out since they are redundant.
356    
357    The reciprocals $\Delta r_f^{-1}$ and $\Delta r_c^{-1}$ are
358    pre-calculated (also in subroutine {\em INI\_VERTICAL\_GRID}). All
359    vertical grid descriptors are stored in common blocks in {\em GRID.h}.
360    
361    The above grid (Fig.~\ref{fig:vgrid}a) is known as the cell centered
362    approach because the tracer points are at cell centers; the cell
363    centers are mid-way between the cell interfaces. An alternative, the
364    vertex or interface centered approach, is shown in
365    Fig.~\ref{fig:vgrid}b. Here, the interior interfaces are positioned
366    mid-way between the tracer nodes (no longer cell centers). This
367    approach is formally more accurate for evaluation of hydrostatic
368    pressure and vertical advection but historically the cell centered
369    approach has been used. An alternative form of subroutine {\em
370    INI\_VERTICAL\_GRID} is used to select the interface centered approach
371    but no run time option is currently available.
372    
373    \fbox{ \begin{minipage}{4.75in}
374    {\em S/R INI\_VERTICAL\_GRID} ({\em
375    model/src/ini\_vertical\_grid.F})
376    
377    $\Delta r_f$: {\bf DRf} ({\em GRID.h})
378    
379    $\Delta r_c$: {\bf DRc} ({\em GRID.h})
380    
381    $\Delta r_f^{-1}$: {\bf RECIP\_DRf} ({\em GRID.h})
382    
383    $\Delta r_c^{-1}$: {\bf RECIP\_DRc} ({\em GRID.h})
384    
385    \end{minipage} }
386    
387    
388    \subsection{Topography: partially filled cells}
389    
390    \begin{figure}
391    \begin{center}
392    \resizebox{4.5in}{!}{\includegraphics{part2/vgrid-xz.eps}}
393    \end{center}
394    \caption{
395    A schematic of the x-r plane showing the location of the
396    non-dimensional fractions $h_c$ and $h_w$. The physical thickness of a
397    tracer cell is given by $h_c(i,j,k) \Delta r_f(k)$ and the physical
398    thickness of the open side is given by $h_w(i,j,k) \Delta r_f(k)$.}
399    \label{fig:hfacs}
400    \end{figure}
401    
402    \cite{Adcroft97} presented two alternatives to the step-wise finite
403    difference representation of topography. The method is known to the
404    engineering community as {\em intersecting boundary method}. It
405    involves allowing the boundary to intersect a grid of cells thereby
406    modifying the shape of those cells intersected. We suggested allowing
407    the topgoraphy to take on a peice-wise linear representation (shaved
408    cells) or a simpler piecewise constant representaion (partial step).
409    Both show dramatic improvements in solution compared to the
410    traditional full step representation, the piece-wise linear being the
411    best. However, the storage requirements are excessive so the simpler
412    piece-wise constant or partial-step method is all that is currently
413    supported.
414    
415    Fig.~\ref{fig:hfacs} shows a schematic of the x-r plane indicating how
416    the thickness of a level is determined at tracer and u points.
417    \marginpar{$h_c$: {\bf hFacC}}
418    \marginpar{$h_w$: {\bf hFacW}}
419    \marginpar{$h_s$: {\bf hFacS}}
420    The physical thickness of a tracer cell is given by $h_c(i,j,k) \Delta
421    r_f(k)$ and the physical thickness of the open side is given by
422    $h_w(i,j,k) \Delta r_f(k)$. Three 3-D discriptors $h_c$, $h_w$ and
423    $h_s$ are used to describe the geometry: {\bf hFacC}, {\bf hFacW} and
424    {\bf hFacS} respectively. These are calculated in subroutine {\em
425    INI\_MASKS\_ETC} along with there reciprocals {\bf RECIP\_hFacC}, {\bf
426    RECIP\_hFacW} and {\bf RECIP\_hFacS}.
427    
428    The non-dimensional fractions (or h-facs as we call them) are
429    calculated from the model depth array and then processed to avoid tiny
430    volumes. The rule is that if a fraction is less than {\bf hFacMin}
431    then it is rounded to the nearer of $0$ or {\bf hFacMin} or if the
432    physical thickness is less than {\bf hFacMinDr} then it is similarly
433    rounded. The larger of the two methods is used when there is a
434    conflict. By setting {\bf hFacMinDr} equal to or larger than the
435    thinnest nominal layers, $\min{(\Delta z_f)}$, but setting {\bf
436    hFacMin} to some small fraction then the model will only lop thick
437    layers but retain stability based on the thinnest unlopped thickness;
438    $\min{(\Delta z_f,\mbox{\bf hFacMinDr})}$.
439    
440    \fbox{ \begin{minipage}{4.75in}
441    {\em S/R INI\_MASKS\_ETC} ({\em model/src/ini\_masks\_etc.F})
442    
443    $h_c$: {\bf hFacC} ({\em GRID.h})
444    
445    $h_w$: {\bf hFacW} ({\em GRID.h})
446    
447  \subsection{Continuity and horizontal pressure gradient terms}  $h_s$: {\bf hFacS} ({\em GRID.h})
448    
449    $h_c^{-1}$: {\bf RECIP\_hFacC} ({\em GRID.h})
450    
451    $h_w^{-1}$: {\bf RECIP\_hFacW} ({\em GRID.h})
452    
453    $h_s^{-1}$: {\bf RECIP\_hFacS} ({\em GRID.h})
454    
455    \end{minipage} }
456    
457    
458    \section{Continuity and horizontal pressure gradient terms}
459    
460  The core algorithm is based on the ``C grid'' discretization of the  The core algorithm is based on the ``C grid'' discretization of the
461  continuity equation which can be summarized as:  continuity equation which can be summarized as:
462  \begin{eqnarray}  \begin{eqnarray}
463  \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' \\  \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' \label{eq:discrete-momu} \\
464  \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' \\  \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' \label{eq:discrete-momv} \\
465  \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}' \\  \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}' \label{eq:discrete-momw} \\
466  \delta_i \Delta y_g \Delta r_f h_w u +  \delta_i \Delta y_g \Delta r_f h_w u +
467  \delta_j \Delta x_g \Delta r_f h_s v +  \delta_j \Delta x_g \Delta r_f h_s v +
468  \delta_k {\cal A}_c w & = & {\cal A}_c \delta_k (P-E)_{r=0}  \delta_k {\cal A}_c w & = & {\cal A}_c \delta_k (P-E)_{r=0}
469    \label{eq:discrete-continuity}
470  \end{eqnarray}  \end{eqnarray}
471  where the continuity equation has been most naturally discretized by  where the continuity equation has been most naturally discretized by
472  staggering the three components of velocity as shown in  staggering the three components of velocity as shown in
# Line 77  A}_c$.  The factors $h_w$ and $h_s$ are Line 485  A}_c$.  The factors $h_w$ and $h_s$ are
485  The last equation, the discrete continuity equation, can be summed in  The last equation, the discrete continuity equation, can be summed in
486  the vertical to yeild the free-surface equation:  the vertical to yeild the free-surface equation:
487  \begin{equation}  \begin{equation}
488  {\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 =  {\cal A}_c \partial_t \eta + \delta_i \sum_k \Delta y_g \Delta r_f h_w
489  {\cal A}_c(P-E)_{r=0}  u + \delta_j \sum_k \Delta x_g \Delta r_f h_s v = {\cal
490    A}_c(P-E)_{r=0} \label{eq:discrete-freesurface}
491  \end{equation}  \end{equation}
492  The source term $P-E$ on the rhs of continuity accounts for the local  The source term $P-E$ on the rhs of continuity accounts for the local
493  addition of volume due to excess precipitation and run-off over  addition of volume due to excess precipitation and run-off over
494  evaporation and only enters the top-level of the {\em ocean} model.  evaporation and only enters the top-level of the {\em ocean} model.
495    
496  \subsection{Hydrostatic balance}  \section{Hydrostatic balance}
497    
498  The vertical momentum equation has the hydrostatic or  The vertical momentum equation has the hydrostatic or
499  quasi-hydrostatic balance on the right hand side. This discretization  quasi-hydrostatic balance on the right hand side. This discretization
# Line 98  discretized: Line 507  discretized:
507  \begin{equation}  \begin{equation}
508  \epsilon_{nh} \partial_t w  \epsilon_{nh} \partial_t w
509  + g \overline{\rho'}^k + \frac{1}{\Delta z} \delta_k \Phi_h' = \ldots  + g \overline{\rho'}^k + \frac{1}{\Delta z} \delta_k \Phi_h' = \ldots
510    \label{eq:discrete_hydro_ocean}
511  \end{equation}  \end{equation}
512    
513  In the atmosphere, using p-coordinates, hydrostatic balance is  In the atmosphere, using p-coordinates, hydrostatic balance is
514  discretized:  discretized:
515  \begin{equation}  \begin{equation}
516  \overline{\theta'}^k + \frac{1}{\Delta \Pi} \delta_k \Phi_h' = 0  \overline{\theta'}^k + \frac{1}{\Delta \Pi} \delta_k \Phi_h' = 0
517    \label{eq:discrete_hydro_atmos}
518  \end{equation}  \end{equation}
519  where $\Delta \Pi$ is the difference in Exner function between the  where $\Delta \Pi$ is the difference in Exner function between the
520  pressure points. The non-hydrostatic equations are not available in  pressure points. The non-hydrostatic equations are not available in
# Line 127  CALC\_PHI\_HYD}. Inside this routine, on Line 538  CALC\_PHI\_HYD}. Inside this routine, on
538  atmospheric/oceanic form is selected based on the string variable {\bf  atmospheric/oceanic form is selected based on the string variable {\bf
539  buoyancyRelation}.  buoyancyRelation}.
540    
 \subsection{Flux-form momentum equations}  
   
 The original finite volume model was based on the Eulerian flux form  
 momentum equations. This is the default though the vector invariant  
 form is optionally available (and recommended in some cases).  
   
 The ``G's'' (our colloquial name for all terms on rhs!) are broken  
 into the various advective, Coriolis, horizontal dissipation, vertical  
 dissipation and metric forces:  
 \marginpar{$G_u$: {\bf Gu} }  
 \marginpar{$G_v$: {\bf Gv} }  
 \marginpar{$G_w$: {\bf Gw} }  
 \begin{eqnarray}  
 G_u & = & G_u^{adv} + G_u^{cor} + G_u^{h-diss} + G_u^{v-diss} +  
 G_u^{metric} + G_u^{nh-metric} \\  
 G_v & = & G_v^{adv} + G_v^{cor} + G_v^{h-diss} + G_v^{v-diss} +  
 G_v^{metric} + G_v^{nh-metric} \\  
 G_w & = & G_w^{adv} + G_w^{cor} + G_w^{h-diss} + G_w^{v-diss} +  
 G_w^{metric} + G_w^{nh-metric}  
 \end{eqnarray}  
 In the hydrostatic limit, $G_w=0$ and $\epsilon_{nh}=0$, reducing the  
 vertical momentum to hydrostatic balance.  
   
 These terms are calculated in routines called from subroutine {\em  
 CALC\_MOM\_RHS} a collected into the global arrays {\bf Gu}, {\bf Gv},  
 and {\bf Gw}.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R CALC\_MOM\_RHS} ({\em pkg/mom\_fluxform/calc\_mom\_rhs.F})  
   
 $G_u$: {\bf Gu} ({\em DYNVARS.h})  
   
 $G_v$: {\bf Gv} ({\em DYNVARS.h})  
   
 $G_w$: {\bf Gw} ({\em DYNVARS.h})  
 \end{minipage} }  
   
   
 \subsubsection{Advection of momentum}  
   
 The advective operator is second order accurate in space:  
 \begin{eqnarray}  
 {\cal A}_w \Delta r_f h_w G_u^{adv} & = &  
   \delta_i \overline{ U }^i \overline{ u }^i  
 + \delta_j \overline{ V }^i \overline{ u }^j  
 + \delta_k \overline{ W }^i \overline{ u }^k \\  
 {\cal A}_s \Delta r_f h_s G_v^{adv} & = &  
   \delta_i \overline{ U }^j \overline{ v }^i  
 + \delta_j \overline{ V }^j \overline{ v }^j  
 + \delta_k \overline{ W }^j \overline{ v }^k \\  
 {\cal A}_c \Delta r_c G_w^{adv} & = &  
   \delta_i \overline{ U }^k \overline{ w }^i  
 + \delta_j \overline{ V }^k \overline{ w }^j  
 + \delta_k \overline{ W }^k \overline{ w }^k \\  
 \end{eqnarray}  
 and because of the flux form does not contribute to the global budget  
 of linear momentum. The quantities $U$, $V$ and $W$ are volume fluxes  
 defined:  
 \marginpar{$U$: {\bf uTrans} }  
 \marginpar{$V$: {\bf vTrans} }  
 \marginpar{$W$: {\bf rTrans} }  
 \begin{eqnarray}  
 U & = & \Delta y_g \Delta r_f h_w u \\  
 V & = & \Delta x_g \Delta r_f h_s v \\  
 W & = & {\cal A}_c w  
 \end{eqnarray}  
 The advection of momentum takes the same form as the advection of  
 tracers but by a translated advective flow. Consequently, the  
 conservation of second moments, derived for tracers later, applies to  
 $u^2$ and $v^2$ and $w^2$ so that advection of momentum correctly  
 conserves kinetic energy.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_ADV\_UU} ({\em mom\_u\_adv\_uu.F})  
   
 {\em S/R MOM\_U\_ADV\_VU} ({\em mom\_u\_adv\_vu.F})  
   
 {\em S/R MOM\_U\_ADV\_WU} ({\em mom\_u\_adv\_wu.F})  
   
 {\em S/R MOM\_U\_ADV\_UV} ({\em mom\_u\_adv\_uv.F})  
   
 {\em S/R MOM\_U\_ADV\_VV} ({\em mom\_u\_adv\_vv.F})  
   
 {\em S/R MOM\_U\_ADV\_WV} ({\em mom\_u\_adv\_wv.F})  
   
 $uu$, $uv$, $vu$, $vv$: {\bf aF} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
   
 \subsubsection{Coriolis terms}  
   
 The ``pure C grid'' Coriolis terms (i.e. in absence of C-D scheme) are  
 discretized:  
 \begin{eqnarray}  
 {\cal A}_w \Delta r_f h_w G_u^{Cor} & = &  
   \overline{ f {\cal A}_c \Delta r_f h_c \overline{ v }^j }^i  
 - \epsilon_{nh} \overline{ f' {\cal A}_c \Delta r_f h_c \overline{ w }^k }^i \\  
 {\cal A}_s \Delta r_f h_s G_v^{Cor} & = &  
 - \overline{ f {\cal A}_c \Delta r_f h_c \overline{ u }^i }^j \\  
 {\cal A}_c \Delta r_c G_w^{Cor} & = &  
  \epsilon_{nh} \overline{ f' {\cal A}_c \Delta r_f h_c \overline{ u }^i }^k  
 \end{eqnarray}  
 where the Coriolis parameters $f$ and $f'$ are defined:  
 \begin{eqnarray}  
 f & = & 2 \Omega \sin{\phi} \\  
 f' & = & 2 \Omega \cos{\phi}  
 \end{eqnarray}  
 when using spherical geometry, otherwise the $\beta$-plane definition is used:  
 \begin{eqnarray}  
 f & = & f_o + \beta y \\  
 f' & = & 0  
 \end{eqnarray}  
   
 This discretization globally conserves kinetic energy. It should be  
 noted that despite the use of this discretization in former  
 publications, all calculations to date have used the following  
 different discretization:  
 \begin{eqnarray}  
 G_u^{Cor} & = &  
   f_u \overline{ v }^{ji}  
 - \epsilon_{nh} f_u' \overline{ w }^{ik} \\  
 G_v^{Cor} & = &  
 - f_v \overline{ u }^{ij} \\  
 G_w^{Cor} & = &  
  \epsilon_{nh} f_w' \overline{ u }^{ik}  
 \end{eqnarray}  
 \marginpar{Need to change the default in code to match this}  
 where the subscripts on $f$ and $f'$ indicate evaluation of the  
 Coriolis parameters at the appropriate points in space. The above  
 discretization does {\em not} conserve anything, especially energy. An  
 option to recover this discretization has been retained for backward  
 compatibility testing (set run-time logical {\bf  
 useNonconservingCoriolis} to {\em true} which otherwise defaults to  
 {\em false}).  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_CDSCHEME} ({\em mom\_cdscheme.F})  
   
 {\em S/R MOM\_U\_CORIOLIS} ({\em mom\_u\_coriolis.F})  
   
 {\em S/R MOM\_V\_CORIOLIS} ({\em mom\_v\_coriolis.F})  
   
 $G_u^{Cor}$, $G_v^{Cor}$: {\bf cF} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Curvature metric terms}  
   
 The most commonly used coordinate system on the sphere is the  
 geographic system $(\lambda,\phi)$. The curvilinear nature of these  
 coordinates on the sphere lead to some ``metric'' terms in the  
 component momentum equations. Under the thin-atmosphere and  
 hydrostatic approximations these terms are discretized:  
 \begin{eqnarray}  
 {\cal A}_w \Delta r_f h_w G_u^{metric} & = &  
   \overline{ \frac{ \overline{u}^i }{a} \tan{\phi} {\cal A}_c \Delta r_f h_c \overline{ v }^j }^i \\  
 {\cal A}_s \Delta r_f h_s G_v^{metric} & = &  
 - \overline{ \frac{ \overline{u}^i }{a} \tan{\phi} {\cal A}_c \Delta r_f h_c \overline{ u }^i }^j \\  
 G_w^{metric} & = & 0  
 \end{eqnarray}  
 where $a$ is the radius of the planet (sphericity is assumed) or the  
 radial distance of the particle (i.e. a function of height).  It is  
 easy to see that this discretization satisfies all the properties of  
 the discrete Coriolis terms since the metric factor $\frac{u}{a}  
 \tan{\phi}$ can be viewed as a modification of the vertical Coriolis  
 parameter: $f \rightarrow f+\frac{u}{a} \tan{\phi}$.  
   
 However, as for the Coriolis terms, a non-energy conserving form has  
 exclusively been used to date:  
 \begin{eqnarray}  
 G_u^{metric} & = & \frac{u \overline{v}^{ij} }{a} \tan{\phi} \\  
 G_v^{metric} & = & \frac{ \overline{u}^{ij} \overline{u}^{ij}}{a} \tan{\phi}  
 \end{eqnarray}  
 where $\tan{\phi}$ is evaluated at the $u$ and $v$ points  
 respectively.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_METRIC\_SPHERE} ({\em mom\_u\_metric\_sphere.F})  
   
 {\em S/R MOM\_V\_METRIC\_SPHERE} ({\em mom\_v\_metric\_sphere.F})  
   
 $G_u^{metric}$, $G_v^{metric}$: {\bf mT} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
   
 \subsubsection{Non-hydrostatic metric terms}  
   
 For the non-hydrostatic equations, dropping the thin-atmosphere  
 approximation re-introduces metric terms involving $w$ and are  
 required to conserve anglular momentum:  
 \begin{eqnarray}  
 {\cal A}_w \Delta r_f h_w G_u^{metric} & = &  
 - \overline{ \frac{ \overline{u}^i \overline{w}^k }{a} {\cal A}_c \Delta r_f h_c }^i \\  
 {\cal A}_s \Delta r_f h_s G_v^{metric} & = &  
 - \overline{ \frac{ \overline{v}^j \overline{w}^k }{a} {\cal A}_c \Delta r_f h_c}^j \\  
 {\cal A}_c \Delta r_c G_w^{metric} & = &  
   \overline{ \frac{ {\overline{u}^i}^2 + {\overline{v}^j}^2}{a} {\cal A}_c \Delta r_f h_c }^k  
 \end{eqnarray}  
   
 Because we are always consistent, even if consistently wrong, we have,  
 in the past, used a different discretization in the model which is:  
 \begin{eqnarray}  
 G_u^{metric} & = &  
 - \frac{u}{a} \overline{w}^{ik} \\  
 G_v^{metric} & = &  
 - \frac{v}{a} \overline{w}^{jk} \\  
 G_w^{metric} & = &  
   \frac{1}{a} ( {\overline{u}^{ik}}^2 + {\overline{v}^{jk}}^2 )  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_METRIC\_NH} ({\em mom\_u\_metric\_nh.F})  
   
 {\em S/R MOM\_V\_METRIC\_NH} ({\em mom\_v\_metric\_nh.F})  
   
 $G_u^{metric}$, $G_v^{metric}$: {\bf mT} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Lateral dissipation}  
   
 Historically, we have represented the SGS Reynolds stresses as simply  
 down gradient momentum fluxes, ignoring constraints on the stress  
 tensor such as symmetry.  
 \begin{eqnarray}  
 {\cal A}_w \Delta r_f h_w G_u^{h-diss} & = &  
   \delta_i  \Delta y_f \Delta r_f h_c \tau_{11}  
 + \delta_j  \Delta x_v \Delta r_f h_\zeta \tau_{12} \\  
 {\cal A}_s \Delta r_f h_s G_v^{h-diss} & = &  
   \delta_i  \Delta y_u \Delta r_f h_\zeta \tau_{21}  
 + \delta_j  \Delta x_f \Delta r_f h_c \tau_{22}  
 \end{eqnarray}  
 \marginpar{Check signs of stress definitions}  
   
 The lateral viscous stresses are discretized:  
 \begin{eqnarray}  
 \tau_{11} & = & A_h c_{11\Delta}(\phi) \frac{1}{\Delta x_f} \delta_i u  
                -A_4 c_{11\Delta^2}(\phi) \frac{1}{\Delta x_f} \delta_i \nabla^2 u \\  
 \tau_{12} & = & A_h c_{12\Delta}(\phi) \frac{1}{\Delta y_u} \delta_j u  
                -A_4 c_{12\Delta^2}(\phi)\frac{1}{\Delta y_u} \delta_j \nabla^2 u \\  
 \tau_{21} & = & A_h c_{21\Delta}(\phi) \frac{1}{\Delta x_v} \delta_i v  
                -A_4 c_{21\Delta^2}(\phi) \frac{1}{\Delta x_v} \delta_i \nabla^2 v \\  
 \tau_{22} & = & A_h c_{22\Delta}(\phi) \frac{1}{\Delta y_f} \delta_j v  
                -A_4 c_{22\Delta^2}(\phi) \frac{1}{\Delta y_f} \delta_j \nabla^2 v  
 \end{eqnarray}  
 where the non-dimensional factors $c_{lm\Delta^n}(\phi), \{l,m,n\} \in  
 \{1,2\}$ define the ``cosine'' scaling with latitude which can be  
 applied in various ad-hoc ways. For instance, $c_{11\Delta} =  
 c_{21\Delta} = (\cos{\phi})^{3/2}$, $c_{12\Delta}=c_{22\Delta}=0$ would  
 represent the an-isotropic cosine scaling typically used on the  
 ``lat-lon'' grid for Laplacian viscosity.  
 \marginpar{Need to tidy up method for controlling this in code}  
   
 It should be noted that dispite the ad-hoc nature of the scaling, some  
 scaling must be done since on a lat-lon grid the converging meridians  
 make it very unlikely that a stable viscosity parameter exists across  
 the entire model domain.  
   
 The Laplacian viscosity coefficient, $A_h$ ({\bf viscAh}), has units  
 of $m^2 s^{-1}$. The bi-harmonic viscosity coefficient, $A_4$ ({\bf  
 viscA4}), has units of $m^4 s^{-1}$.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_XVISCFLUX} ({\em mom\_u\_xviscflux.F})  
   
 {\em S/R MOM\_U\_YVISCFLUX} ({\em mom\_u\_yviscflux.F})  
   
 {\em S/R MOM\_V\_XVISCFLUX} ({\em mom\_v\_xviscflux.F})  
   
 {\em S/R MOM\_V\_YVISCFLUX} ({\em mom\_v\_yviscflux.F})  
   
 $\tau_{11}$, $\tau_{12}$, $\tau_{22}$, $\tau_{22}$: {\bf vF}, {\bf  
 v4F} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
 Two types of lateral boundary condition exist for the lateral viscous  
 terms, no-slip and free-slip.  
   
 The free-slip condition is most convenient to code since it is  
 equivalent to zero-stress on boundaries. Simple masking of the stress  
 components sets them to zero. The fractional open stress is properly  
 handled using the lopped cells.  
   
 The no-slip condition defines the normal gradient of a tangential flow  
 such that the flow is zero on the boundary. Rather than modify the  
 stresses by using complicated functions of the masks and ``ghost''  
 points (see \cite{Adcroft+Marshall98}) we add the boundary stresses as  
 an additional source term in cells next to solid boundaries. This has  
 the advantage of being able to cope with ``thin walls'' and also makes  
 the interior stress calculation (code) independent of the boundary  
 conditions. The ``body'' force takes the form:  
 \begin{eqnarray}  
 G_u^{side-drag} & = &  
 \frac{4}{\Delta z_f} \overline{ (1-h_\zeta) \frac{\Delta x_v}{\Delta y_u} }^j  
 \left( A_h c_{12\Delta}(\phi) u - A_4 c_{12\Delta^2}(\phi) \nabla^2 u \right)  
 \\  
 G_v^{side-drag} & = &  
 \frac{4}{\Delta z_f} \overline{ (1-h_\zeta) \frac{\Delta y_u}{\Delta x_v} }^i  
 \left( A_h c_{21\Delta}(\phi) v - A_4 c_{21\Delta^2}(\phi) \nabla^2 v \right)  
 \end{eqnarray}  
   
 In fact, the above discretization is not quite complete because it  
 assumes that the bathymetry at velocity points is deeper than at  
 neighbouring vorticity points, e.g. $1-h_w < 1-h_\zeta$  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_SIDEDRAG} ({\em mom\_u\_sidedrag.F})  
   
 {\em S/R MOM\_V\_SIDEDRAG} ({\em mom\_v\_sidedrag.F})  
   
 $G_u^{side-drag}$, $G_v^{side-drag}$: {\bf vF} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Vertical dissipation}  
   
 Vertical viscosity terms are discretized with only partial adherence  
 to the variable grid lengths introduced by the finite volume  
 formulation. This reduces the formal accuracy of these terms to just  
 first order but only next to boundaries; exactly where other terms  
 appear such as linar and quadratic bottom drag.  
 \begin{eqnarray}  
 G_u^{v-diss} & = &  
 \frac{1}{\Delta r_f h_w} \delta_k \tau_{13} \\  
 G_v^{v-diss} & = &  
 \frac{1}{\Delta r_f h_s} \delta_k \tau_{23} \\  
 G_w^{v-diss} & = & \epsilon_{nh}  
 \frac{1}{\Delta r_f h_d} \delta_k \tau_{33}  
 \end{eqnarray}  
 represents the general discrete form of the vertical dissipation terms.  
   
 In the interior the vertical stresses are discretized:  
 \begin{eqnarray}  
 \tau_{13} & = & A_v \frac{1}{\Delta r_c} \delta_k u \\  
 \tau_{23} & = & A_v \frac{1}{\Delta r_c} \delta_k v \\  
 \tau_{33} & = & A_v \frac{1}{\Delta r_f} \delta_k w  
 \end{eqnarray}  
 It should be noted that in the non-hydrostatic form, the stress tensor  
 is even less consistent than for the hydrostatic (see Wazjowicz  
 \cite{Waojz}). It is well known how to do this properly (see Griffies  
 \cite{Griffies}) and is on the list of to-do's.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_RVISCLFUX} ({\em mom\_u\_rviscflux.F})  
   
 {\em S/R MOM\_V\_RVISCLFUX} ({\em mom\_v\_rviscflux.F})  
   
 $\tau_{13}$: {\bf urf} (local to {\em calc\_mom\_rhs.F})  
   
 $\tau_{23}$: {\bf vrf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 As for the lateral viscous terms, the free-slip condition is  
 equivalent to simply setting the stress to zero on boundaries.  The  
 no-slip condition is implemented as an additional term acting on top  
 of the interior and free-slip stresses. Bottom drag represents  
 additional friction, in addition to that imposed by the no-slip  
 condition at the bottom. The drag is cast as a stress expressed as a  
 linear or quadratic function of the mean flow in the layer above the  
 topography:  
 \begin{eqnarray}  
 \tau_{13}^{bottom-drag} & = &  
 \left(  
 2 A_v \frac{1}{\Delta r_c}  
 + r_b  
 + C_d \sqrt{ \overline{2 KE}^i }  
 \right) u \\  
 \tau_{23}^{bottom-drag} & = &  
 \left(  
 2 A_v \frac{1}{\Delta r_c}  
 + r_b  
 + C_d \sqrt{ \overline{2 KE}^j }  
 \right) v  
 \end{eqnarray}  
 where these terms are only evaluated immediately above topography.  
 $r_b$ ({\bf bottomDragLinear}) has units of $m s^{-1}$ and a typical value  
 of the order 0.0002 $m s^{-1}$. $C_d$ ({\bf bottomDragQuadratic}) is  
 dimensionless with typical values in the range 0.001--0.003.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_BOTTOMDRAG} ({\em mom\_u\_bottomdrag.F})  
   
 {\em S/R MOM\_V\_BOTTOMDRAG} ({\em mom\_v\_bottomdrag.F})  
   
 $\tau_{13}^{bottom-drag}$, $\tau_{23}^{bottom-drag}$: {\bf vf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
   
   
   
 \subsection{Tracer equations}  
   
 The tracer equations are discretized consistantly with the continuity  
 equation to facilitate conservation properties analogous to the  
 continuum:  
 \begin{equation}  
 {\cal A}_c \Delta r_f h_c \partial_\theta  
 + \delta_i U \overline{ \theta }^i  
 + \delta_j V \overline{ \theta }^j  
 + \delta_k W \overline{ \theta }^k  
 = {\cal A}_c \Delta r_f h_c {\cal S}_\theta + \theta {\cal A}_c \delta_k (P-E)_{r=0}  
 \end{equation}  
 The quantities $U$, $V$ and $W$ are volume fluxes defined:  
 \marginpar{$U$: {\bf uTrans} }  
 \marginpar{$V$: {\bf vTrans} }  
 \marginpar{$W$: {\bf rTrans} }  
 \begin{eqnarray}  
 U & = & \Delta y_g \Delta r_f h_w u \\  
 V & = & \Delta x_g \Delta r_f h_s v \\  
 W & = & {\cal A}_c w  
 \end{eqnarray}  
 ${\cal S}$ represents the ``parameterized'' SGS processes and  
 physics associated with the tracer. For instance, potential  
 temperature equation in the ocean has is forced by surface and  
 partially penetrating heat fluxes:  
 \begin{equation}  
 {\cal A}_c \Delta r_f h_c {\cal S}_\theta = \frac{1}{c_p \rho_o} \delta_k {\cal A}_c {\cal Q}  
 \end{equation}  
 while the salt equation has no real sources, ${\cal S}=0$, which  
 leaves just the $P-E$ term.  
   
 The continuity equation can be recovered by setting ${\cal Q}=0$ and  
 $\theta=1$. The term $\theta (P-E)_{r=0}$ is required to retain local  
 conservation of $\theta$. Global conservation is not possible using  
 the flux-form (as here) and a linearized free-surface  
 (\cite{Griffies00,Campin02}).  
   
   
   
   
 \subsection{Derivation of discrete energy conservation}  
   
 These discrete equations conserve kinetic plus potential energy using the  
 following definitions:  
 \begin{equation}  
 KE = \frac{1}{2} \left( \overline{ u^2 }^i + \overline{ v^2 }^j +  
 \epsilon_{nh} \overline{ w^2 }^k \right)  
 \end{equation}  
   
   
 \subsection{Vector invariant momentum equations}  
   
 The finite volume method lends itself to describing the continuity and  
 tracer equations in curvilinear coordinate systems but the appearance  
 of new metric terms in the flux-form momentum equations makes  
 generalizing them far from elegant. The vector invariant form of the  
 momentum equations are exactly that; invariant under coordinate  
 transformations.  
   
 The non-hydrostatic vector invariant equations read:  
 \begin{equation}  
 \partial_t \vec{v} + ( 2\vec{\Omega} + \vec{\zeta}) \wedge \vec{v}  
 - b \hat{r}  
 + \vec{\nabla} B = \vec{\nabla} \cdot \vec{\bf \tau}  
 \end{equation}  
 which describe motions in any orthogonal curvilinear coordinate  
 system. Here, $B$ is the Bernoulli function and $\vec{\zeta}=\nabla  
 \wedge \vec{v}$ is the vorticity vector. We can take advantage of the  
 elegance of these equations when discretizing them and use the  
 discrete definitions of the grad, curl and divergence operators to  
 satisfy constraints. We can also consider the analogy to forming  
 derived equations, such as the vorticity equation, and examine how the  
 discretization can be adjusted to give suitable vorticity advection  
 among other things.  
   
 The underlying algorithm is the same as for the flux form  
 equations. All that has changed is the contents of the ``G's''. For  
 the time-being, only the hydrostatic terms have been coded but we will  
 indicate the points where non-hydrostatic contributions will enter:  
 \begin{eqnarray}  
 G_u & = & G_u^{fv} + G_u^{\zeta_3 v} + G_u^{\zeta_2 w} + G_u^{\partial_x B}  
 + G_u^{\partial_z \tau^x} + G_u^{h-dissip} + G_u^{v-dissip} \\  
 G_v & = & G_v^{fu} + G_v^{\zeta_3 u} + G_v^{\zeta_1 w} + G_v^{\partial_y B}  
 + G_v^{\partial_z \tau^y} + G_v^{h-dissip} + G_v^{v-dissip} \\  
 G_w & = & G_w^{fu} + G_w^{\zeta_1 v} + G_w^{\zeta_2 u} + G_w^{\partial_z B}  
 + G_w^{h-dissip} + G_w^{v-dissip}  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R CALC\_MOM\_RHS} ({\em pkg/mom\_vecinv/calc\_mom\_rhs.F})  
   
 $G_u$: {\bf Gu} ({\em DYNVARS.h})  
   
 $G_v$: {\bf Gv} ({\em DYNVARS.h})  
   
 $G_w$: {\bf Gw} ({\em DYNVARS.h})  
 \end{minipage} }  
   
 \subsubsection{Relative vorticity}  
   
 The vertical component of relative vorticity is explicitly calculated  
 and use in the discretization. The particular form is crucial for  
 numerical stablility; alternative definitions break the conservation  
 properties of the discrete equations.  
   
 Relative vorticity is defined:  
 \begin{equation}  
 \zeta_3 = \frac{\Gamma}{A_\zeta}  
 = \frac{1}{{\cal A}_\zeta} ( \delta_i \Delta y_c v - \delta_j \Delta x_c u )  
 \end{equation}  
 where ${\cal A}_\zeta$ is the area of the vorticity cell presented in  
 the vertical and $\Gamma$ is the circulation about that cell.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_CALC\_RELVORT3} ({\em mom\_vi\_calc\_relvort3.F})  
   
 $\zeta_3$: {\bf vort3} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Kinetic energy}  
   
 The kinetic energy, denoted $KE$, is defined:  
 \begin{equation}  
 KE = \frac{1}{2} ( \overline{ u^2 }^i + \overline{ v^2 }^j  
 + \epsilon_{nh} \overline{ w^2 }^k )  
 \end{equation}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_CALC\_KE} ({\em mom\_vi\_calc\_ke.F})  
   
 $KE$: {\bf KE} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Coriolis terms}  
   
 The potential enstrophy conserving form of the linear Coriolis terms  
 are written:  
 \begin{eqnarray}  
 G_u^{fv} & = &  
 \frac{1}{\Delta x_c}  
 \overline{ \frac{f}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\  
 G_v^{fu} & = & -  
 \frac{1}{\Delta y_c}  
 \overline{ \frac{f}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j  
 \end{eqnarray}  
 Here, the Coriolis parameter $f$ is defined at vorticity (corner)  
 points.  
 \marginpar{$f$: {\bf fCoriG}}  
 \marginpar{$h_\zeta$: {\bf hFacZ}}  
   
 The potential enstrophy conserving form of the non-linear Coriolis  
 terms are written:  
 \begin{eqnarray}  
 G_u^{\zeta_3 v} & = &  
 \frac{1}{\Delta x_c}  
 \overline{ \frac{\zeta_3}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\  
 G_v^{\zeta_3 u} & = & -  
 \frac{1}{\Delta y_c}  
 \overline{ \frac{\zeta_3}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j  
 \end{eqnarray}  
 \marginpar{$\zeta_3$: {\bf vort3}}  
   
 The Coriolis terms can also be evaluated together and expressed in  
 terms of absolute vorticity $f+\zeta_3$. The potential enstrophy  
 conserving form using the absolute vorticity is written:  
 \begin{eqnarray}  
 G_u^{fv} + G_u^{\zeta_3 v} & = &  
 \frac{1}{\Delta x_c}  
 \overline{ \frac{f + \zeta_3}{h_\zeta} }^j \overline{ \overline{ \Delta x_g h_s v }^j }^i \\  
 G_v^{fu} + G_v^{\zeta_3 u} & = & -  
 \frac{1}{\Delta y_c}  
 \overline{ \frac{f + \zeta_3}{h_\zeta} }^i \overline{ \overline{ \Delta y_g h_w u }^i }^j  
 \end{eqnarray}  
   
 \marginpar{Run-time control needs to be added for these options} The  
 disctinction between using absolute vorticity or relative vorticity is  
 useful when constructing higher order advection schemes; monotone  
 advection of relative vorticity behaves differently to monotone  
 advection of absolute vorticity. Currently the choice of  
 relative/absolute vorticity, centered/upwind/high order advection is  
 available only through commented subroutine calls.  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_CORIOLIS} ({\em mom\_vi\_coriolis.F})  
   
 {\em S/R MOM\_VI\_U\_CORIOLIS} ({\em mom\_vi\_u\_coriolis.F})  
   
 {\em S/R MOM\_VI\_V\_CORIOLIS} ({\em mom\_vi\_v\_coriolis.F})  
   
 $G_u^{fv}$, $G_u^{\zeta_3 v}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F})  
   
 $G_v^{fu}$, $G_v^{\zeta_3 u}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Shear terms}  
   
 The shear terms ($\zeta_2w$ and $\zeta_1w$) are are discretized to  
 guarantee that no spurious generation of kinetic energy is possible;  
 the horizontal gradient of Bernoulli function has to be consistent  
 with the vertical advection of shear:  
 \marginpar{N-H terms have not been tried!}  
 \begin{eqnarray}  
 G_u^{\zeta_2 w} & = &  
 \frac{1}{ {\cal A}_w \Delta r_f h_w } \overline{  
 \overline{ {\cal A}_c w }^i ( \delta_k u - \epsilon_{nh} \delta_j w )  
 }^k \\  
 G_v^{\zeta_1 w} & = &  
 \frac{1}{ {\cal A}_s \Delta r_f h_s } \overline{  
 \overline{ {\cal A}_c w }^i ( \delta_k u - \epsilon_{nh} \delta_j w )  
 }^k  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_U\_VERTSHEAR} ({\em mom\_vi\_u\_vertshear.F})  
   
 {\em S/R MOM\_VI\_V\_VERTSHEAR} ({\em mom\_vi\_v\_vertshear.F})  
   
 $G_u^{\zeta_2 w}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F})  
   
 $G_v^{\zeta_1 w}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
   
 \subsubsection{Gradient of Bernoulli function}  
   
 \begin{eqnarray}  
 G_u^{\partial_x B} & = &  
 \frac{1}{\Delta x_c} \delta_i ( \phi' + KE ) \\  
 G_v^{\partial_y B} & = &  
 \frac{1}{\Delta x_y} \delta_j ( \phi' + KE )  
 %G_w^{\partial_z B} & = &  
 %\frac{1}{\Delta r_c} h_c \delta_k ( \phi' + KE )  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_U\_GRAD\_KE} ({\em mom\_vi\_u\_grad\_ke.F})  
   
 {\em S/R MOM\_VI\_V\_GRAD\_KE} ({\em mom\_vi\_v\_grad\_ke.F})  
   
 $G_u^{\partial_x KE}$: {\bf uCf} (local to {\em calc\_mom\_rhs.F})  
   
 $G_v^{\partial_y KE}$: {\bf vCf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
   
 \subsubsection{Horizontal dissipation}  
   
 The horizontal divergence, a complimentary quantity to relative  
 vorticity, is used in parameterizing the Reynolds stresses and is  
 discretized:  
 \begin{equation}  
 D = \frac{1}{{\cal A}_c h_c} (  
   \delta_i \Delta y_g h_w u  
 + \delta_j \Delta x_g h_s v )  
 \end{equation}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_CALC\_HDIV} ({\em mom\_vi\_calc\_hdiv.F})  
   
 $D$: {\bf hDiv} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Horizontal dissipation}  
   
 The following discretization of horizontal dissipation conserves  
 potential vorticity (thickness weighted relative vorticity) and  
 divergence and dissipates energy, enstrophy and divergence squared:  
 \begin{eqnarray}  
 G_u^{h-dissip} & = &  
   \frac{1}{\Delta x_c} \delta_i ( A_D D - A_{D4} D^*)  
 - \frac{1}{\Delta y_u h_w} \delta_j h_\zeta ( A_\zeta \zeta - A_{\zeta4} \zeta^* )  
 \\  
 G_v^{h-dissip} & = &  
   \frac{1}{\Delta x_v h_s} \delta_i h_\zeta ( A_\zeta \zeta - A_\zeta \zeta^* )  
 + \frac{1}{\Delta y_c} \delta_j ( A_D D - A_{D4} D^* )  
 \end{eqnarray}  
 where  
 \begin{eqnarray}  
 D^* & = & \frac{1}{{\cal A}_c h_c} (  
   \delta_i \Delta y_g h_w \nabla^2 u  
 + \delta_j \Delta x_g h_s \nabla^2 v ) \\  
 \zeta^* & = & \frac{1}{{\cal A}_\zeta} (  
   \delta_i \Delta y_c \nabla^2 v  
 - \delta_j \Delta x_c \nabla^2 u )  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_VI\_HDISSIP} ({\em mom\_vi\_hdissip.F})  
   
 $G_u^{h-dissip}$: {\bf uDiss} (local to {\em calc\_mom\_rhs.F})  
   
 $G_v^{h-dissip}$: {\bf vDiss} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  
   
   
 \subsubsection{Vertical dissipation}  
   
 Currently, this is exactly the same code as the flux form equations.  
 \begin{eqnarray}  
 G_u^{v-diss} & = &  
 \frac{1}{\Delta r_f h_w} \delta_k \tau_{13} \\  
 G_v^{v-diss} & = &  
 \frac{1}{\Delta r_f h_s} \delta_k \tau_{23}  
 \end{eqnarray}  
 represents the general discrete form of the vertical dissipation terms.  
   
 In the interior the vertical stresses are discretized:  
 \begin{eqnarray}  
 \tau_{13} & = & A_v \frac{1}{\Delta r_c} \delta_k u \\  
 \tau_{23} & = & A_v \frac{1}{\Delta r_c} \delta_k v  
 \end{eqnarray}  
   
 \fbox{ \begin{minipage}{4.25in}  
 {\em S/R MOM\_U\_RVISCLFUX} ({\em mom\_u\_rviscflux.F})  
   
 {\em S/R MOM\_V\_RVISCLFUX} ({\em mom\_v\_rviscflux.F})  
   
 $\tau_{13}$: {\bf urf} (local to {\em calc\_mom\_rhs.F})  
   
 $\tau_{23}$: {\bf vrf} (local to {\em calc\_mom\_rhs.F})  
 \end{minipage} }  

Legend:
Removed from v.1.1  
changed lines
  Added in v.1.7

  ViewVC Help
Powered by ViewVC 1.1.22